Next Article in Journal
User-Driven Climate Resilience Across Southern European Regions
Previous Article in Journal
The Impact of Climate Change on Energy Consumption on Small Tropical Islands
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Changes in Climate and Their Implications for Cattle Nutrition and Management

by
Bashiri Iddy Muzzo
*,
R. Douglas Ramsey
and
Juan J. Villalba
Department of Wildland Resources, Quinney College of Natural Resources, Utah State University, Logan, UT 84322-5230, USA
*
Author to whom correspondence should be addressed.
Climate 2025, 13(1), 1; https://doi.org/10.3390/cli13010001
Submission received: 22 November 2024 / Revised: 21 December 2024 / Accepted: 22 December 2024 / Published: 24 December 2024

Abstract

:
Climate change is a global challenge that impacts rangeland and pastureland landscapes by inducing shifts in temperature variability, precipitation patterns, and extreme weather events. These changes alter soil and plant conditions, reducing forage availability and chemical composition and leading to nutritional stress in cattle. This stress occurs when animals lack adequate water and feed sources or when these resources are insufficient in quantity, composition, or nutrient balance. Several strategies are essential to address these impacts. Genetic selection, epigenetic biomarkers, and exploration of epigenetic memories present promising avenues for enhancing the resilience of cattle populations and improving adaptation to environmental stresses. Remote sensing and GIS technologies assist in locating wet spots to establish islands of plant diversity and high forage quality for grazing amid ongoing climate change challenges. Establishing islands of functional plant diversity improves forage quality, reduces carbon and nitrogen footprints, and provides essential nutrients and bioactives, thus enhancing cattle health, welfare, and productivity. Real-time GPS collars coupled with accelerometers provide detailed data on cattle movement and activity, aiding livestock nutrition management while mitigating heat stress. Integrating these strategies may offer significant advantages to animals facing a changing world while securing the future of livestock production and the global food system.

1. Introduction

Climate change is widely considered the most significant global challenge of the 21st century, with its impact being experienced worldwide (recent long-term shifts in temperature, precipitation patterns, and the frequency and intensity of extreme weather events). The primary cause of climate change is the greenhouse effect, which occurs when gases such as carbon dioxide, methane, and nitrous oxide in the Earth’s atmosphere act like the glass in a greenhouse, trapping the sun’s heat and preventing it from escaping into space, leading to global warming [1]. Human activities, including industries, burning of fossil fuels, and deforestation, are the primary drivers of climate change, with carbon dioxide emissions being the largest contributor. Among these, the energy sector is the largest contributor, followed by transportation, manufacturing, agriculture, and forestry [2]. As of 2020, the concentration of CO2 in the atmosphere has increased by 48% above pre-industrial levels (before 1750) [2]. Although other greenhouse gases such as methane and nitrous oxide are emitted in smaller quantities by human activities such as livestock farming, they still contribute to global warming over different timescales. Moreover, non-greenhouse gas pollutants such as soot have varying warming and cooling effects and can affect air quality. Soon et al. [3] reveal that natural causes such as changes in solar radiation or volcanic activity are estimated to have contributed much less to global warming between 1890 and 2010 than human activities. Moreover, combined human and natural drivers have been shown to have a greater influence on long-term increases in global temperature [4].
The increase of atmospheric greenhouse gases not only leads to rising sea levels and more frequent extreme weather events but also leads to changes in ecological communities [5]. Climate change indirectly affects agricultural systems, particularly livestock production. Cattle, as an essential source of food and income for millions of people worldwide, are expected to face significant challenges as the global demand for beef and dairy products is projected to rise in the coming decades [6]. As a result of climate changes, cattle are expected to face significant challenges, with nutritional stress being one of the most critical factors affecting their health and productivity. Forage is an important component of the beef and dairy industry, playing a vital role in human livelihoods and food security. Climate change has the potential to adversely impact forage nutritive value and quantity through alterations in precipitation patterns, temperature, extreme weather events [7], and plant distribution and abundance [8], as well as through reductions in forage accessibility. Changes in rainfall patterns resulting from climate change, including shifts in timing and intensity, have significant implications for plant growth. These changes ultimately affect the quantity and quality of forage. In semi-arid climates, increased rainfall can provide the necessary moisture for plants to thrive, resulting in enhanced productivity and greater forage quantity. Moreover, increased rainfall in these regions has been found to promote plant diversity, which contributes to a more balanced diet for grazing animals through improvements in overall forage quality [9]. However, increased rainfall in humid climates may have the opposite effect, leading to a reduction in forage quantity and quality. Excess water can result in waterlogging, which can damage plant roots and impede nutrient uptake, decreasing productivity and forage quality [10].
Changes in temperature due to climate change can also affect the quality and quantity of cattle forage. Higher temperatures during the growing period can lead to increased plant growth and productivity, resulting in greater forage quantity [11]. However, prolonged temperatures can lead to drought, which may reduce both the quantity and quality of forage by accelerating plant maturity, stunting growth, and increasing plant die-off [12]. Additionally, high temperatures can cause heat stress in livestock, resulting in reduced feed intake and weight gain [13]. On the other hand, a decrease in temperatures due to climate change can lead to shorter growing seasons and a delay in the onset of plant growth, potentially leading to decreased forage quantity [14]. Cold temperatures can also reduce the digestibility of forage and its quality as a feed source for livestock [15]. The impact of temperature on forage quality and quantity can vary depending on the specific plant species and the region in which they are grown. For instance, the same plant species can benefit from increasing temperatures in one area, but not in a different location, whereas in other areas, the quality and quantity of forage may decline as temperatures increase [11]. In Australia, increasing temperatures may benefit certain pasture species such as kikuyu grass (Pennisetum clandestinum) in the southeastern region of the country, but it can lead to a decrease in forage quantity and quality in the northern region due to increased water stress [16]. Similarly, in the United States, certain forage species, such as Sporobolus cryptandrus and A. purpurea, may benefit from both increased temperatures and elevated carbon dioxide levels, leading to increments in forage production [17]. However, other studies have reported that increased temperatures can reduce the nutritional quality of forage species such as alfalfa (Medicago sativa) and reduce the growth of cool-season grasses [18]. More frequent extreme weather events due to climate change can lead to an increase in droughts, floods, and wildfires, as reported by [19]. These events can have significant impacts on forage availability and quality [20,21], as well as on the ability of animals to access forage. Nutritional stress occurs when an animal does not get enough nutrients to maintain its health and productivity, which can happen when an animal does not have access to enough water or high-quality feed, or when the available feed is not well-balanced in terms of nutrient content [7]. This narrative review article describes how climate change impacts cattle nutrition, synthesizing current knowledge on its effects on soil and plant water availability and quality, as well as on cattle health. Furthermore, it explores potential mitigation strategies and proposes practical management approaches to help alleviate the adverse effects of climate change on cattle production systems.

2. Discussion

2.1. Impact of Climate Change on Soil

Soil Microbes and Nutrient Cycling

Climate change has a profound indirect impact on cattle through its effects on soil microbial communities, which are vital for nutrient cycling and availability. Changes in temperature and moisture levels can significantly impact the composition and function of soil microorganisms [22]. Temperature plays a crucial role in shaping microbial communities and their metabolic activities. Research indicated that temperature influences enzyme activity, microbial growth rates, and community composition [23]. Warmer temperatures generally accelerate microbial metabolic rates by increasing enzymatic activity and enhancing organic matter decomposition and nutrient cycling. In contrast, excessively high temperatures can denature enzymes and proteins, reducing microbial activity and potentially leading to cell death [24]. Conversely, lower temperatures slow down microbial metabolic processes, decreasing enzyme activity and nutrient cycling efficiency [25]. Different microbial species have distinct optimal temperature ranges for their growth and shifts in temperature can favor temperature-tolerant species, altering the composition of the community and affecting the decomposition of organic matter and rates of nutrient cycling [26].
Climate change can increase water availability in some regions, providing optimal soil moisture conditions that support microbial activity. Optimal soil moisture levels facilitate microbial activity by ensuring sufficient water for nutrient solubilization and enzyme function [27]. In contrast, dry conditions limit microbial activity due to water scarcity, leading to decreased nutrient cycling [28]. Conversely, waterlogged soils restrict oxygen diffusion, favoring anaerobic processes and potentially shifting microbial communities towards anaerobic species such as denitrifiers and methanogens, impacting nutrient pathways and greenhouse gas production [29,30]. This balance between aerobic and anaerobic microbial processes, influenced by soil moisture levels, significantly affects overall soil function and nutrient cycling dynamics [22]. The highlighted changes in temperature and moisture levels, impacting the composition and function of soil microbes, also substantially influence the breakdown and release of essential forage nutrients such as nitrogen, phosphorus, and potassium. Research conducted by [31] demonstrated that warming and increased nitrogen deposition affected the stability and complexity of soil microbial networks, with warming promoting bacterial network stability and nitrogen addition decreasing it. Additionally, Qiu et al. [32] found that warming reduced fungal abundance but not bacterial abundance, leading to a shift in the microbial community’s composition in semi-arid grasslands. Moreover, Wang et al. [33] highlighted the importance of microbial network complexity in driving soil multifunctionality, showing that altered precipitation regimes affected soil functions mediated by microbial communities. These studies collectively emphasize the fundamental role of soil microbes in nutrient cycling processes and the potential consequences of climate change on ecosystem functioning.
Changes in soil nutrient availability, driven by altered microbial activity and plant composition, can affect the concentration of essential micronutrients and minerals in the forage. For instance, shifts in plant species can affect the availability of critical nutrients such as phosphorus and calcium, which are essential for cattle health and productivity, as different plant species have varying abilities to uptake these nutrients from the soil [34]. In conclusion, modifications of plant species composition due to climate change may impact soil nutrient availability, further influencing the nutritional quality and balance of forage accessible to cattle. Alterations in temperature and moisture levels can shift the composition and function of soil microbes, affecting the breakdown and release of essential nutrients such as nitrogen, phosphorus, and potassium [35]. These changes in nutrient cycling can influence plant growth and nutrient content, affecting the forage available to cattle. As the microbial community structure shifts, nutrient utilization and recycling efficiency in the soil ecosystem can be disrupted, potentially leading to nutrient imbalances that compromise plant health and productivity. Bainard et al. [36] reported that increased plant species diversity enhances the soil’s nutrient content, potentially mitigating some negative effects of climate change on soil fertility.

2.2. Impact of Climate Change on Forage Plants

2.2.1. Plant Growth, Establishment, and Phenology

Higher temperatures during the growing period can lead to increased plant growth and productivity, resulting in greater forage quantity. A rise in temperature causes an increase in CO2 concentration levels [37,38]. The relationship between atmospheric CO2 and temperature can be considered a “hen-or-egg” problem, where it is not always clear which is the cause and which is the effect [39]. However, recent studies suggest that the dominant direction is temperature (T) → CO2, meaning that changes in CO2 follow changes in temperature [40]. This can be attributed to biochemical reactions, such as increased soil respiration and CO2 emission at higher temperatures [39]. Photosynthesis, the fundamental process by which green plants convert carbon dioxide and water into organic compounds and oxygen, varies across different plant species and their respective photosynthetic pathways C3, C4, and CAM. C3 plants such as fescue and ryegrass grass species rely exclusively on the Calvin cycle for CO2 assimilation. In contrast, C4 plants such as Bermuda grass and crabgrass follow a typical pathway involving enzymes and a four-carbon compound, exhibiting high photosynthetic rates and a distinct advantage under conditions conducive to high transpiration rates. In the future, increased temperatures with CO2 levels are expected to favor C4 plants over C3 plants [41]. C4 plants have a more efficient photosynthetic process and are better able to concentrate CO2, which allows them to thrive in high-CO2 environments [42]. The increased temperatures are expected to have a negative impact on photosynthesis, particularly in C3 plants [41] as these plants have limited rubisco activity at high temperatures, which can decrease their photosynthetic efficiency [43]. Therefore, under increased temperatures, C4 plants may have a competitive advantage over C3 plants. Overall, the interaction between CO2 levels and temperature can have different effects on C4 and C3 plants, with elevated CO2 favoring C4 plants and increased temperatures potentially favoring C3 plants. The prolonged increased temperature during sprouting of the forage hastens plant maturity and reduces plant growth and die-off [44]. In this period, temperatures promote stem-over-leaf growth, resulting in lower digestibility of both leaves and stems due to high cell wall concentrations and low soluble sugar concentrations [45]. Moreover, the growth of various organs decreases gradually, leading to weaker plants and a slowdown in growth during summer [46]. Temperature and variety are important factors that modify tuber initiation and growth, with temperature having a considerable influence [47]. Elevated temperatures can affect seed germination and subsequent seedling vigor, but in the case of fodder tree Trichilia emetica, exposure to elevated temperatures did not disrupt metabolic and ultrastructural integrity and may have even improved germinative development and seedling growth rates [48]. While climate change is typically associated with temperature increases, localized cooling can occur in specific regions, such as higher latitudes, elevated altitudes, or as a result of particular atmospheric phenomena such as volcanic eruptions or La Niña events [39].
A decline in temperature due to climate change can result in a shorter growing season, delaying the onset of plant growth and potentially reducing the quantity of overall forage available for grazing. Cold temperatures constrain the time available for active plant growth, slowing down metabolic rates and diminishing nutrient assimilation, ultimately impacting biomass production [49]. The structural changes induced by cold stress can compromise the digestibility of forage, diminishing its quality as a feed source for cattle. These declines in temperature have multifaceted effects on various stages of plant growth and development. Cold temperatures impede metabolic processes during germination and seedling establishment, delaying germination and compromising seedling vigor [50]. In the vegetative stage, plants experience reduced photosynthesis and nutrient uptake, leading to stunted growth as cell expansion and elongation are hindered [51]. Reproductive development is hampered by cold stress during flowering, disrupting pollination and fertilization processes and diminishing seed or fruit production [49]. Cold temperature stress also affects the structural integrity of plant cells, leading to cell membrane fluidity issues and ice crystal formation, causing cellular damage [52]. Plants exhibit metabolic adaptations, such as the accumulation of compatible solutes, to serve as cryoprotectants against freezing temperatures [53]. Some plants also undergo cold acclimation, a process involving physiological changes induced by non-freezing low temperatures, enhancing their tolerance to subsequent cold temperature stress. Phenological shifts in grassland species, influenced by climate change and land use practices, have been extensively studied, revealing species-specific responses to changing environmental conditions [54,55]. These changes in flowering phenology can impact the availability and quality of forage for cattle, with late-flowering species and those with specific functional traits advancing their flowering under future climate conditions, potentially affecting nutrient uptake and storage in plants [55]. Additionally, alterations in precipitation patterns can drive phenological shifts, influencing flower and fruit production in temperate grasslands and ultimately impacting the nutritional quality of forage available to cattle [56]. Furthermore, vegetation phenology, including the start and end of the growing season, responds to climate fluctuations, affecting gross primary productivity and potentially influencing nutrient availability in the soil and the nutritional balance of forage [57].

2.2.2. Plant Species Composition, Forage Biomass, and Nutritional Quality

Climate change causes shifts in plant species composition (PSC), forage biomass production, and nutritional quality. As temperatures rise and precipitation patterns change, the abundance and distribution of plant species adjust, leading to altered competitive dynamics within plant communities [58]. For instance, Liu et al. [59] found that an increase in temperature of 1.2 °C (0.38 per decade) led to increased above-ground net primary productivity (ANPP) in grass, decreased ANPP in sedge, and relatively stable ANPP in forbs. These shifts in species composition can result in changes to the nutrient profile of forages, with forbs generally being high in protein and low in fiber relative to grasses [60]. Studies by Zhang [61] in northern Tibet and Yoshihara [62] in Mongolian drylands demonstrate that warming can lead to an increase in acid detergent fiber (ADF) content and a decrease in digestibility, affecting the overall forage quality of the diet. Petit Bon [63] further highlights how environmental perturbations can alter plant community nutrient concentrations, with varying effects on different functional plant types. Sebastià [64] reported the significant impact of management practices on forage nutritive value, demonstrating that sainfoin swards managed under different regimes by displaying varying nutrient profiles, with cattle-grazed swards exhibiting a higher nutritional quality that declines faster compared to sheep-grazed or mown swards. Habermann [65] showed that elevated CO2 levels and warming can synergistically decrease forage digestibility and modify nutrient content, illustrating the intricate interplay between climate change and forage quality in grassland ecosystems. Climate change-induced shifts in PSC directly impact the nutritional quality and abundance of forage available to cattle. Sebastià [64] explained that changes in PSC, including alterations in the relative abundance of grasses, sedges, and forbs, can significantly influence the protein and fiber content of forage, thereby affecting its suitability as a feed source. Furthermore, climate-induced alterations in plant growth cycles and species dominance can lead to variations in forage biomass production, affecting the quantity of feed available for cattle and potentially leading to nutritional stress.
On the other hand, climate change impacts water availability, causing it to become more limited in semi-arid and temperate climates as the probability of summer droughts increases [66]. Forage production is dependent on adequate water supply [67], and soil water deficits typically limit biomass production, particularly in soil types that tend to show low water holding capacities, such as coarse-textured soils [68,69]. Irrigation amends this problem but with the predicted interannual variabilities in rainfall and thus water availability [70], it becomes crucial to better understand the relationship between water inputs and the quantity and quality of the forage produced. For instance, the total alfalfa yield was 100% greater on average for irrigated compared to unirrigated treatments. However, treatments with high levels of irrigation (to bring root zone extractable soil water (ESW) levels to 15% of field capacity at 35% ESW depletion), or with 34% less irrigation than those higher levels, produced similar amounts of forage after four harvests [68]. Despite multiple studies on the negative impacts of soil water availability on forage biomass production, knowledge regarding the influence of drought on important characteristics of the nutritive value of forages is inconsistent and limited [59,71]. In general, perennial forage plants growing under moderate water deficits tend to present greater nutritive quality than those forages grown under irrigation [72], provided that the deficit was not severe and initiated early in the growth of the herbage [73,74]. In contrast, long and extreme drought events due to climate change inhibit tillering and branching, accelerate the death of tillers and the senescence of leaves, and relocate protein, nitrogen, and soluble carbohydrates from leaves to roots, reducing the nutritive value of the forage [73,75,76]. Under moderate water restrictions, in vitro dry matter digestibility (IVDMD) in alfalfa is greater and the contents of cell walls are lower under prolonged water deficits that reduce plant water potential (ψw) than in control plants grown without a water deficit [77]. Improvements in forage quality have been attributed to a greater leaf-to-stem ratio (LSR) in water-stressed plants, given that water stress has the greatest effect on reducing stem growth [77]. Alternatively, the high forage nutritive value in plants grown under a water deficit could be better explained by a retardation of the rate of plant maturation rather than by a differential effect on leaf-to-stem growth [72,73]. In fact, forage quality in alfalfa declines with maturation [78], and the mean maturity stage decreases with increasing water stress [79]. Consistent with this idea, increments of IVDMD in alfalfa have been primarily associated with plant maturity [77]. Nevertheless, the same authors showed that increments in stem crude protein (CP) contents and declines in cellulose concentration in alfalfa stems and leaves under water stress were not fully accounted for by differences in plant maturity. Research on other legumes such as birdsfoot trefoil, sainfoin, clover, and cicer milkvetch shows similar patterns in response to water stress to those described for alfalfa [69,71], although alfalfa yields tend to be less affected by drought than for the other species. Droughted alternative legumes produced biomass with lower contents of ADF, NDF, and lignin than alfalfa, with birdsfoot trefoil and cicer milkvetch producing the highest-quality forage [69]. Nevertheless, because of its superior yield (and persistence) under drought, it was concluded that alfalfa would produce more nutrients per unit of area than the alternative legumes [69].

2.2.3. Plant Secondary Compounds

Plant secondary metabolites (PSMs) are a diverse group of chemical compounds, including alkaloids, terpenoids, and phenolics, that are not directly involved in essential plant growth but are crucial for the plant’s defense against abiotic and biotic stressors [80,81]. These compounds also aid in protecting plants from pathogens, attracting pollinators, and thriving in diverse environmental conditions. PSMs play significant roles in plant communication, ecosystem functioning, and defense responses to stress [82,83]. When plants encounter stressors such as drought or herbivory, specific receptors on the plant cells are activated, signaling the stress and triggering various biochemical pathways. These pathways involve molecules such as calcium ions, reactive oxygen species (ROS), and plant hormones such as jasmonic acid and salicylic acid. These molecules activate transcription factors that upregulate genes responsible for the biosynthesis of PSMs, which are formed by converting primary metabolites into alkaloids, terpenoids, and phenolic compounds [82,84]. For instance, during drought, phenolic compounds such as flavonoids and tannins accumulate in plants as protective mechanisms against oxidative damage and to minimize water loss. Abscisic acid, another plant hormone, helps with water conservation by regulating stomatal closure. In response to herbivore feeding, some plants produce alkaloids, such as nicotine in tobacco, which act as neurotoxins to deter herbivores. Similarly, terpenoids such as menthol in mint release strong odors and flavors to discourage herbivores from feeding [85,86]. In addition to protecting plants from herbivory and stress, certain PSMs, such as condensed tannins, provide benefits beyond plant defense. These compounds can improve livestock productivity by enhancing plant protein utilization and reducing enteric methane emissions [87]. Furthermore, PSMs can benefit soils by increasing carbon sequestration and reducing nitrogen losses [88], while also improving meat and milk quality through reduced fat biohydrogenation and extended shelf life [89]. With the ongoing effects of climate change, the production and accumulation of PSMs in plants are expected to change. A meta-analysis of the effects of global climate change on medicinal and aromatic PSMs revealed that phenolic and terpenoid levels generally increase with elevated carbon dioxide but decrease with higher nitrogen concentrations. Additionally, total alkaloid concentrations increase significantly with higher nitrogen levels.
Elevated temperatures also lead to alterations in plant secondary metabolites (PSMs) in plants, affecting their interactions with herbivores and overall plant performance. Temperature increases can influence both the types and concentrations of PSMs, thereby affecting plant–herbivore interactions and plant defense strategies [90,91]. Moreover, elevated temperatures may increase the concentrations of certain toxic compounds, such as aristolochic acids, which can alter herbivore growth and nutritional intake [91,92]. Such changes in PSM concentrations can have dual effects on herbivores. On the one hand, they may enhance plant resistance to herbivores, but on the other, they may reduce the palatability and nutritional value of forage, which can limit feed intake and affect herbivore health [93,94]. Research by Balluffi-Fry [95] suggests that increased temperatures could elevate PSM levels, potentially improving plant resistance to pests and diseases, but also reduce the overall quality of forage for grazing animals. Certain PSMs can be toxic to cattle, especially when consumed in large quantities, leading to reduced feed intake and potential health problems [96,97,98]. As ambient temperatures continue to rise, Beck and Gregorini [99] emphasize the growing challenge of managing livestock diets that reduce PSM content while maintaining adequate protein levels. Furthermore, Villalba and Manteca [100] highlight the importance of identifying areas within grazing landscapes that facilitate heat dissipation, helping livestock manage temperature-related stress. They also stress that animals accustomed to complex landscapes and diverse plant types may have an advantage in adapting to environmental changes.
Changes in rainfall patterns, a critical aspect of climate change, can lead to water deficits, which in turn can affect the concentration of PSMs in plants. Water stress limits photosynthetic activity by disrupting the balance between light capture and utilization, which leads to the generation of ROS. To cope with this, plants activate mechanisms for ROS detoxification, which often involve phenolic compounds, such as tannins, known for their antioxidant properties [101,102]. Consequently, under mild water stress, plants are likely to produce more carbon-based PSMs, such as tannins, due to a higher concentration of carbohydrates. However, as water deficits become more severe, plants reduce their carbon gain, prioritizing primary metabolites over PSMs. This can lead to a decrease in the production of PSMs in more extreme drought conditions [103,104,105]. The impact of water deficits on PSMs is non-linear and varies depending on the severity of the stress. For instance, condensed tannin concentrations in tanniferous legumes such as Lotus corniculatus and Lotus uliginosus can either increase, remain stable, or decrease in response to water stress, depending on the degree of water limitation [68,106,107,108].
PSMs also have an important role in herbivore thermoregulation. The energy required for thermoregulation in mammals such as cattle is derived from their diet, which makes PSMs an important factor in determining their ability to maintain stable body temperatures under varying environmental conditions. Studies have shown that PSMs can interfere with thermoregulation by uncoupling mitochondrial oxidative phosphorylation, thus reducing the available energy for maintaining body temperature. Additionally, the detoxification processes that break down PSMs are thermogenic, meaning they generate additional heat, further challenging animals in warmer conditions. Beale [93] found that PSMs in grazing forage could negatively impact thermoregulation, especially in hotter climates. Supporting studies have shown that temperature affects the tolerance of herbivores to PSMs. For example, Windley and Shimada [109] showed that mice acclimated to cooler temperatures exhibited improved liver function and tannin tolerance when consuming tannin-rich acorns. However, at higher temperatures, liver function decreased, which impaired the detoxification of PSMs. Similarly, Beale [110] observed that temperature influenced the detoxification of PSMs in herbivorous marsupials. The findings indicate that ambient temperatures increase the detoxification burden of herbivores, reducing food intake and making it more difficult for them to manage diets high in PSMs. Longer acclimation periods may be required to fully understand these impacts.

2.3. Impact of Climate Change on Cattle

2.3.1. Heat Stress

Climate change causes an increase or decrease in the frequency and intensity of extreme temperature events, leading to both heat stress (HS) and cold stress in cattle, indirectly affecting their nutritional status. Higher temperatures cause heat stress in cattle, reducing feed intake and digestion efficiency, leading to weight loss and decreased productivity [13]. This decline in productivity during HS is not solely due to reduced feed intake. However, it is also linked to compromised gastrointestinal tract function and increased permeability, allowing antigens to enter the circulation [111]. Consequently, nutrient partitioning shifts to support immune responses, characterized by increased insulin secretion and altered metabolism in response to the activated immune system, ultimately reducing animal productivity during higher-temperature months [112]. HS has been extensively studied in dairy cows, with findings indicating a significant negative impact on productivity. For instance, Chen [113] conducted a meta-analysis using data from 31 studies (34 trials) and found that HS led to decreased dry matter intake (DMI), energy-corrected milk (ECM), and milk protein concentration, with reductions of 4.13% and 3.25% in DMI and ECM, respectively, for each unit increase in the temperature-humidity index (THI). Additionally, Souza [114] reported that pregnant nonlactating heat-stressed dairy cows experienced reduced DMI as the THI increased, emphasizing the importance of environmental conditions in influencing feed intake. Sammad [115] and Safa [116] reported contrasting results, showing that HS affects metabolic, endocrinological, and inflammatory parameters and alters milk yield, milk composition, and feed efficiency in dairy cows. Furthermore, Antanaitis [117] showed that HS negatively affected behaviors such as rumination and drinking time, showing the detrimental effects of HS on dairy cow health and productivity. Nevertheless, Chen [113] revealed that HS did not affect milk fat concentration or feed efficiency. During moderate-to-severe heat stress, when compared to a thermoneutral period, dairy cows experience a significant decrease in milk production, which can be attributed to a complex interplay of altered metabolic, endocrinological, and inflammatory parameters [13,113]. HS causes physiological changes, including hyperthermia, impacting feed intake and milk yield and ultimately compromising dairy production [118].
Regarding biological responses to HS, Baumgard and Rhoads [111] stated that energy allocation is shifted towards protective mechanisms, diminishing milk production and increasing susceptibility to diseases and mortality in dairy cattle. In beef cattle, HS also poses a significant challenge, particularly during summer in temperate regions and dry seasons in tropical regions, leading to various detrimental effects on animal productivity and well-being. Studies have shown that exposure to heat stress results in elevated rectal temperatures in beef cattle, reduced feed intake, body weight loss, and changes in behavior, such as increased stepping and altered body posture [119,120]. Additionally, heat stress negatively impacts reproductive rates in beef cattle, with pregnancy rates being lower under moderate stress conditions compared to mild stress conditions [121,122]. Heat stress exacerbates the susceptibility of animals to PSCs by disrupting detoxification processes [123], impairing liver and kidney function, and reducing the activity of enzymes involved in metabolizing toxic compounds. This disruption, coupled with heat-induced oxidative stress and increased reactive oxygen species [124], intensifies the harmful effects of PSMs such as tannins and flavonoids, leading to greater cellular damage. Furthermore, heat stress impairs gut function, alters the gut microbiome, and reduces digestion efficiency [125], which might result in higher concentrations of PSMs in the gut in cattle forage containing PSMs. In identifying HS indicators for dairy cows, Chen [126] exhibited that HS can induce changes in blood parameters such as cortisol, glucose, and interleukin-6, with heat-tolerant cows exhibiting lower levels of these indicators compared to heat-sensitive cows. Genetic studies have identified potential genetic markers associated with heat stress tolerance in Chinese Holstein cattle, highlighting genes such as PDZRN4 and PRKG1 as important for this trait [127]. Lemal [128] also explored using sensor data, such as activity and rumination sensors, to evaluate heat stress in dairy cattle, indicating the complexity of the relationship between sensor traits and production traits under heat stress conditions. Cold stress increases the energy requirements of cattle to maintain their body temperature, which can result in higher forage intake. Research on Simmental crossbred bulls exposed to long-term cold stress showed decreased body weight and average daily gain, increased dry matter intake, and altered physiological behaviors, indicating heightened energy demands [129].

2.3.2. Cold Stress

After highlighting the impact of heat stress caused by elevated temperatures, it is equally important to consider cold stress (CS) in cattle. Cold stress occurs when ambient temperatures drop below the animal’s thermal comfort zone, leading to physiological challenges in maintaining their core body temperature. Key factors contributing to cold stress include low ambient temperatures, wind chill that exacerbates heat loss, and wet conditions that diminish the insulating properties of an animal’s coat. Additionally, cattle with low body condition scores or compromised health, as well as younger animals, are more vulnerable due to insufficient fat insulation and lower metabolic heat production.
CS reduces forage intake as animals prioritize energy conservation over feeding. The mechanism behind this reduced intake involves prioritizing energy expenditure for thermoregulation rather than digestion. Research by Chase [130] indicated that cattle exposed to CS reduced their dry matter intake by approximately 30%, leading to decreased nutrient availability that impacts overall health and productivity. Similarly, a study by Kim [131] found that beef cattle subjected to CS conditions consumed 25% less feed than their counterparts in thermally neutral environments. Conversely, a study by Delfino and Mathison [132] reported that while feed intake decreased, the efficiency of nutrient utilization improved, suggesting that cattle may adapt their feeding behavior under specific CS conditions to optimize energy use. Under cold conditions, the rate of passage of feed through the digestive system may also slow down, affecting nutrient absorption and energy utilization and further diminishing digestive efficiency. For instance, Wang [133] demonstrated that cold-stressed cattle exhibited a significantly slower rate of rumen evacuation, resulting in less efficient digestion and nutrient absorption. A study by Nardone [134] suggested that some cattle could maintain rumen function under mild CS, indicating variability in response based on individual animal factors and environmental conditions. Additionally, Coloma-García [135] indicated that cattle exposed to prolonged cold stress exhibited a 15% decrease in the digestibility of dry matter, emphasizing the impact of cold stress on nutrient absorption. Moreover, cold stress compels cattle to expend more energy to maintain body temperature, leading to decreased activity levels and potential weight loss, which can adversely affect their welfare. The increased energy expenditure can result in an additional caloric requirement of 10–20% during cold weather, which cattle may not compensate for if forage intake is reduced. The consequences of CS extend to milk production, which may decline as the energy required to sustain body temperature detracts from the energy available for lactation, significantly impacting dairy operations during winter months. Berian [136] found that lactating cows exposed to CS experienced a reduction in milk yield compared to cows in thermally neutral conditions. Similarly, a study by Sahib [137] demonstrated that dairy cows subjected to CS had an average drop of 12% in milk production due to increased energy demands for thermoregulation. In beef cattle, research by Wagner [138] indicated that CS led to reduced weight gain in heifers, as those experiencing cold stress consumed approximately 20% less feed compared to control groups, impacting their overall growth rates. In another study, Cantalapiedra-Hijar et al. [139] found that steers exposed to CS conditions exhibited a significant decline in average daily gain, emphasizing the long-term implications of CS on beef production systems. Additionally, a study by Ames [140] reported that beef cattle experiencing cold stress had lower feed efficiency, with a 15% reduction in the conversion of feed to body weight gain. Physiologically, CS can alter critical parameters such as increased heart and respiratory rates, as ruminants work harder to maintain their body temperature, indicating heightened stress levels that may necessitate management interventions. Young [141] found that dairy cows exposed to prolonged CS exhibited increased heart rates, which could be linked to decreased milk yields and altered feeding behavior. Similarly, Lees et al. [142] observed that beef cattle at different growth stages experienced changes in heart rate, rectal temperature, and dry matter intake under extreme cold stress, emphasizing the increased energy requirements in such conditions. Hence, CS can elevate the energy demands of cattle to maintain body temperatures, potentially driving an increase in forage intake to meet these heightened requirements. In contrast, a study by Colditz and Hine [143] suggested that certain beef cattle breeds demonstrated resilience to CS, maintaining normal heart rates and growth performance despite environmental challenges. By comprehensively addressing these aspects, the discussion of cold stress can be effectively balanced with that of heat stress.

2.3.3. Health

Livestock, especially beef and dairy cattle, are significantly impacted by climate change-induced weather fluctuations, with factors such as rising temperatures, erratic rainfall, and intermittent drought promoting different causative agents of disease, compromising their health. Changes in precipitation patterns, particularly increased intensity and frequency, have an important role in altering the distribution and abundance of vector-borne diseases such as ticks, consequently affecting the spread of diseases such as Lyme, Anaplasmosis, Babesiosis, and Theileriosis to new areas [144,145]. Bryer [146] showed that increased humidity, often associated with higher precipitation levels, prevents ticks from desiccating. Deshpande [147] reported that this environmental factor helps ticks stay hydrated and promotes vegetation growth, which provides suitable habitats for ticks to thrive. Moreover, the rise in humidity levels contributes to an increase in the availability of hosts such as rodents, which are essential for ticks’ feeding and reproduction cycles. Ixodes ricinus, a common tick species in Europe, transmits various pathogens, including Lyme borreliosis and tick-borne encephalitis [148]. Climate change models predict potential range expansions of tick areas with milder winter conditions, leading to increased survival rates and higher probabilities of tick bites [149,150,151]. Milder winter temperatures reduce tick mortality by preventing exposure to lethal cold conditions [152]. Warmer winters allow ticks to avoid freezing, which could disrupt their bodily fluids and metabolic functions [153], ultimately improving their chances of survival during the winter months. Additionally, the decreased severity of cold conditions reduces the energy expenditure required for survival processes, further enhancing tick survival rates [154]. Waterlogging in tropical cattle boma areas due to high-intensity rainfall can lead to foot disorders in cattle, such as foot rot diseases [155]. Urban-Chmiel [154] showed that foot disorders are prevalent in cattle, with laminitis and its consequences being a significant issue, especially during the rainy season. Additionally, waterlogging can increase soil concentrations of certain elements such as iron and manganese, potentially causing leaf abnormalities, which may affect the overall health of the cattle in these areas [155]. Furthermore, rainfall has been linked to increased leptospiral infections in cattle, with higher seropositivity and urine PCR positivity during rainy seasons compared to dry seasons, indicating a correlation between rainfall and cattle health issues [156].

2.4. Impact of Climate Change on Water for Cattle

Water Availability and Quality

Climate change impacts water availability and quality for cattle through alterations in precipitation patterns, increased temperatures, and more frequent extreme weather events. Rainfall patterns, including distribution, intensity, and timing, are evidence of changing climate affecting cattle via various interconnected mechanisms. Alizadeh [157] showed that these patterns are altered due to rising global temperatures, increasing evaporation rates, loading the atmosphere with more moisture, and intensifying precipitation events in some regions, potentially causing flooding. This scenario has led to significant cattle losses, with farmers experiencing the deaths of a substantial number of cattle due to the adverse effects of climate change. Conversely, higher temperatures can exacerbate drought conditions by accelerating evapotranspiration rates [158]. For example, prolonged droughts in regions such as the western United States have reduced surface water availability, forcing ranchers to rely more on groundwater resources, which can become depleted or contaminated over time [159]. Prolonged droughts as a result of climate change lead to water scarcity in beef cattle, which affects their intake, weight gain, and reproductive success and can even lead to death in severe cases [160]. Free-grazing cattle tend to walk longer distances to find water during a drought, which has detrimental effects on their health and productivity. Maurya [161] reported that cattle walking long distances in search of water can lead to increased energy expenditure, impacting the maintenance energy requirements of cattle by up to 24% compared to stall-fed animals. Additionally, walking stress induces physiological and metabolic changes, altering hormonal concentrations and potentially reducing dairy cows’ growth, milk production, and reproductive performance [162]. However, the physiological responses to long-distance walking, including changes in adrenal and thyroid hormone concentrations, showcase the stress and adaptation mechanisms involved in local cattle breeds such as Tanzania shorthorn zebu (TSZ). The TSZ cattle withstand drought due to their small body size, skin color, and hair coats, which are adapted to reflect sunlight and reduce heat absorption [163]. Climate change-induced temperatures also increase the temperature of water sources, negatively impacting cattle. Cattle drink less hot water primarily due to body thermoregulatory and physiological reasons. He [164] demonstrated that beef cattle consuming heated water in colder seasons experienced disruptions in serum metabolism, rumen microbial fermentation, and the metabolome, leading to altered glucose and fatty acid metabolism. Conversely, Golher [165] highlighted that providing cooler water sources can aid in regulating body temperature and maintaining metabolic balance in cattle. In the hot season, due to climate change with warmer water, cattle will tend to increase their energy demands for thermoregulation, affecting glucose and fatty acid metabolism, altering rumen microbial fermentation, and causing stress-induced metabolic changes. Bewley [166] found that prolonged water consumption with higher temperatures decreased the reticular temperatures of dairy cows. Singh [167] found that cows consumed more water at 37 °C than at 25 °C and 15 °C. In contrast, Grossi et al. [168] reported no significant difference in animal performance when provided with water at different temperatures of 17 °C and 24 °C. Pereyra et al. [169] conducted three experiments. They found that water temperatures of 18 °C and 31 °C had no effect on water consumption by dairy cows but recommended providing fresh water to improve thermoregulation and animal performance under heat stress conditions. Another set of trials within the same study showed that providing a simple cloth shade over the water trough can prevent a significant effect on water temperature, which may modify water intake behavior in dairy cattle. Moreover, changes in water quality due to increased weather events and water temperature can also affect animal health and productivity [170]. The increased occurrence of weather events due to climate change, such as flooding in mining areas, reduces water quality by introducing mineral content such as selenium, sulfur, and copper into animal water sources. Water with high concentrations of these minerals has been shown to be toxic to cattle [171] and may lead to death. Generally, these changes in water quality and availability due to climate change, as Bownik and Wlodkowic [172] showed, significantly impact animal behavior and movement, reducing their nutritional intake.

2.5. Mitigation Strategies

2.5.1. Cattle Selection, Breeding, and Epigenetics

Selection, breeding, and epigenetics strategies significantly enhance cattle’s adaptability to climate change effects (Figure 1). Selection involves choosing individuals with desirable traits to serve as parents for the next generation. Breeders can improve livestock resilience by selecting individuals with traits such as heat tolerance, disease resistance, and efficient nutrient utilization [173]. Rexroad [174] pointed out that genomic analysis aids in identifying genetic markers associated with these traits, facilitating tailored breeding programs to enhance adaptability in cattle populations. A current study on genetic variations in diverse cattle populations has revealed genes such as HSPB2, HSPB3, HSP20, HSP90AB1, HSF4, HSPA1B, and CLPB linked to climate adaptation [175]. For example, in their study, they show that variations in the HSP70 gene are associated with heat tolerance in cattle such as the Brahman breed. Genetic variants in HSP70 in cattle breeds improve their ability to withstand high temperatures and thermal stress [175]. Tian [176] conducted a whole-genome sequencing study on crossbred and Mongolian cattle populations that revealed significant genetic diversity and distinct clustering based on geographic distribution, with identified candidate genes such as TRPM8, NMUR1, PRKAA2, SMTNL2, and OXR1 that are associated with immune response and cold adaptation and crucial for developing cattle populations better suited to cope with climate change challenges. By systematically selecting for these advantageous traits, breeders can ensure the sustainability of livestock production systems via breeding programs in the face of evolving environmental pressures.
Breeding involves strategically applying selected climate adaptation and resilience traits that cattle possess, enhancing the climate change stress resilience of future cattle populations. In livestock production, breeding can involve both inbreeding and crossbreeding techniques. Inbreeding, while altering genetic footprints within cattle genomes and impacting phenotypic variance, can potentially influence adaptive traits negatively due to reduced genetic diversity [177]. Conversely, crossbreeding techniques offer a promising avenue to improve resilience to climate stress by introducing genetic diversity from cattle breeds with adaptive traits into non-resilient populations [178], thereby enhancing overall climate resilience across cattle populations. In Rwanda, which has a tropical climate, Niyonzima [179] found that crossbred cows (Ankole-Friesian) with higher heat tolerances had better milk production and fertility than purebred cows. In the southwest U.S., Castaño-Sánchez [180] revealed that crossbreed cattle (Criollo × Angus) exhibited lower water use, fuel consumption, nitrogen footprints, and production costs regardless of the finishing diet. Muzzo et al. [163] emphasized that crossbreeding can produce offspring with more desirable traits of climate and disease resilience and adaptability to local conditions.
On the other hand, epigenetics studies have an important role in understanding how cattle might adapt to climate change. Holliday [181] defined epigenetics as biochemical and structural changes in chromatin without alterations in the DNA sequence. In the context of the effect of climate change on cattle, epigenetics serves as molecular mechanisms influencing cattle in terms of gene expression, physiology, and phenotypic variation to adapt to stress due to the changing climate, with the potential to cause transgenerational effects. Epigenetic mechanisms, such as DNA methylation, histone modifications, and non-coding RNAs, can provide a bridge between environmental stressors and phenotypic responses, allowing organisms to acclimate and cope with rapidly changing climates. Identifying epigenetic biomarkers and exploring epigenetic memories and diversity present promising avenues for enhancing the resilience of cattle populations, improving adaptation to environmental stresses, and facilitating the development of climate-resilient cattle breeds. For example, Creole cattle breeds in tropical environments, when compared to their Spanish ancestors, reveal differential methylation, impacting genes related to immune responses, heat resistance, and energy management, highlighting adaptation mechanisms [182]. Similarly, the methylation patterns of heat stress-related genes in Sahiwal and Frieswal cattle under heat stress conditions offer insights into the molecular mechanisms of adaptation to tropical climates, indicating similarities in their responses due to acclimatization [183]. By unraveling these epigenetic mechanisms, we can pave the way for developing climate-resilient cattle breeds that can thrive in challenging environmental conditions, ultimately ensuring sustainable cattle production in the face of climate change challenges. Concurrently, integrating selection, breeding, and epigenetics can produce a comprehensive approach that ensures that cattle are resilient, healthy, and productive in various environmental conditions, supporting sustainable cattle production systems in the face of ongoing climate changes.

2.5.2. Spatial Selection, Remote Sensing, and GIS

In the current era of climate change challenges, the use of remote sensing and geographic information systems (GISs) has an important role in assisting farmers in locating pasture areas that are greener than the surrounding area and thus inferring localized higher soil moisture (“wet spots”) that can assist in establishing forage for sustaining livestock during dry periods (Figure 1).
Remote sensing instruments mounted to satellites, unpiloted aerial systems (UASs), or piloted aircraft offer a variety of data in the form of multi-spectral imagery for real-time or near-real-time monitoring of vegetation health, soil moisture levels, and land cover transformations [184]. The normalized difference vegetation index (NDVI) and enhanced vegetation index (EVI) are remote sensing-derived vegetation indices used to measure plant density and health [185], facilitating the identification of productive regions even amidst varying climatic conditions. NDVI, obtained by normalizing near-infrared and red-light reflectance differentials, can evaluate vegetation vigor areas where the leaf area index is less than 3. The EVI provides similar capabilities in more dense vegetation areas [186]. Additionally, Bhaga [187] explained that indices such as the soil moisture index (SMI), which uses remote sensing data to estimate soil moisture content, and the normalized difference water index (NDWI), which monitors water levels in vegetation, enable the evaluation of drought resilience and identifying areas with adequate water availability.
Geographic information systems (GISs) leverage the wealth of data offered by remote sensing technologies, particularly in the realm of climate change. Combining spatial and temporal data from diverse data sources such as high- and low-resolution satellite images, UAS-captured imagery, and other spatial data such as fences, topography, and even livestock movement recorded with GPS, etc., provides a holistic view of landscape dynamics and a platform to monitor environmental conditions, including changes in vegetation health, soil moisture, and land cover patterns. GISs empower the creation of intricate maps and models crucial for informed decision-making. Historical imagery from satellites such as Landsat and the Moderate Resolution Imaging Spectroradiometer (MODIS) offer several decades of information [188] and is invaluable for analyzing long-term trends and understanding the impacts of climate change on vegetation, water, and soil conditions [189]. These and other platforms provide decades of information at a temporal repeat rate spanning 1 to 16 days, providing both long-term and high-temporal-grain data to evaluate surface conditions.
Previous research by Poitras [190] demonstrated the effectiveness of satellite imagery in identifying vegetation hotspots in rangeland ecosystems. Combining data from multiple remote sensing platforms allows a comprehensive assessment of vegetation health, phenology, and responses to climatic events over time, thus aiding in identifying areas that consistently exhibit higher levels of photosynthetically active vegetation (“green spots”), indicating forage availability and quality for a particular temporal period. Identifying periodic green areas can help ranchers or pastoralists allocate their cattle grazing regimes and stocking density based on the estimated carrying capacity of their grazing area. In addition, a suitable mix of soil and water indices can help identify “wet spots” for better establishment of seeded forage species. This approach aligns with the principles employed in developing the Islands of Diversity Location Tool (IDLT) by researchers at Utah State University. This tool assisted in identifying suitable areas for establishing islands of plant diversity “feed patches” with plants containing different phytochemicals across a 55-acre pasture of grass monoculture (Figure 2), a demonstration research field of the ongoing Smart Foodscape project [81]. Additionally, remote sensing provides a reliable and cost-effective means of measuring and evaluating mitigation and adaptation strategies against the effects of climate change, enhancing resilience and promoting sustainable agricultural intensification. Therefore, remote sensing and GIS tools significantly enhance the selection of optimal “green spots” for establishing forage in rangeland areas with limited precipitation. However, identifying the most resilient and adaptable forage species in rangeland areas using these tools requires further research. The identification of forage species adapted to specific ecosites requires the development of testing plots to gather information on species establishment, persistence, and tolerance to grazing. Regos [191] suggested that integrating ecological niche models and eco-evolutionary mechanisms through GIS and remote sensing can dynamically model geographic range shifts and assess evolutionary rescue probabilities. These approaches allow researchers to predict species responses to climate change more accurately. For instance, integrating models such as AdaptR and incorporating adaptive capacity, phenotypic plasticity, and evolutionary adaptation into species distribution modeling help to reduce projected range losses and enhance species resilience [192]. This integration supports the development of effective strategies for forage establishment and ecosystem management in the face of ongoing climate change.

2.5.3. Forage Species and Chemical Diversity to Attenuate the Negative Impacts of Climate Fluctuations on Livestock Performance

The establishment of islands of plant diversity “feed patches” within grazing landscapes facing long drought conditions enhances forage availability and quality, mitigating cattle nutrition stress (Figure 1). Wet spots identified within the landscape using the above-highlighted principles provide established forages with self-sustaining resources to grow and mature, providing essential nutrients to cattle grazing most grass monoculture landscapes even during extended dry periods. During the dry season, the nutritive value of these grasses declines while perennial legumes are revealed to fix atmospheric N2 [193], persist for years without additional planting [194], and provide a significant input of nutrients for the ruminant animal [195]. In addition, legumes and some forbs are of greater nutritional value than grasses [196]. Diverse legumes with different biochemical profiles, such as protein, non-fibrous carbohydrates, and plant secondary compounds (PSMs), such as condensed tannins (CTs) complement grasses, increase the diversity of feed resources that improve animal fitness [197] while decreasing carbon and N footprints [125]. In support of this, Lagrange and Villalba [198] and Lagrange [199] showed that cattle and sheep consuming combinations of alfalfa (ALF; Medicago sativa), birdsfoot trefoil (BFT; Lotus corniculatus), and sainfoin (SF; Onobrychis viciifolia) showed reduced urinary N excretions and improved body weight gains and N retention. This forage diversity motivates animal intake by offering numerous orosensorial and post-ingestive stimuli [200]. Moreover, allowing animals to choose pastures of different chemical and organoleptic profiles contributes to improved intake, animal welfare, and performance [201,202]. Diverse pastures grazed in specific sequences improved dry matter intake compared to repeated allocations of monotonous pastures [203]. Mote [204] reported that providing sheep with a tannin-rich ration followed by terpenes doubled their intake of food with terpenes relative to the reverse sequence, suggesting that temporal allocation of diverse forages impacts foraging behavior and potentially performance.
The availability of taxonomically diverse forage alternatives allows ruminants to select diets that meet their nutrient requirements while incorporating bioactives that are medicinal or prophylactic (e.g., antioxidants). Doses of these biochemicals in diverse diets are typically below the thresholds of toxicity [205] and thus more beneficial than consuming individual plant species [206,207]. Incorporating a diversity of forage species into grazing systems is essential for reducing the metabolic burden of herbivores associated with detoxifying PSMs. A monotonous diet dominated by a single species can elevate the metabolic rate of herbivores, leading to an increased internal heat load, particularly under changing climate conditions with rising temperatures and frequent heat waves [5]. A diverse forage landscape allows herbivores to consume various plants, each with different PSMs, effectively diluting the intake of any single toxin and reducing the detoxification load on their organs [176,207]. Moreover, the interaction between diverse PSMs can enhance detoxification efficiency, as some compounds may mitigate the toxicity of others [125]. For instance, studies have shown that sheep grazing on mixed diets experienced less heat stress and improved health compared to those on monocultures [208]. Similarly, cattle fed diverse forages exhibited lower internal temperatures and enhanced growth rates [202]. Studies by Jordán [209] further support this, showing that goats grazing on diverse Mediterranean shrublands had improved protein utilization and lower liver enzyme activity compared to those on single-species diets. Moreover, diverse forages stimulate the expression of various detoxification enzymes, enhancing the herbivore’s ability to metabolize a broader range of toxins [210]. This enzymatic adaptability is vital as different plants produce distinct PSMs, each requiring specific pathways for detoxification. Key enzymes involved in this process include cytochrome P450 enzymes, which facilitate the oxidation of organic compounds, and glutathione S-transferases, which conjugate toxins with glutathione for excretion. Additionally, the presence of antioxidant compounds in diverse forages can alleviate oxidative stress associated with PSM metabolism, as antioxidants help neutralize reactive oxygen species generated during detoxification [211]. These findings underscore that promoting plant diversity is a crucial adaptive strategy for grazing systems, enabling herbivores such as cattle to maintain better health and productivity in warmer environments, thereby addressing the challenges posed by climate change. Integrating shrubs into pasturelands represents a vital strategy for enhancing forage diversity and adapting to climate variability and climate change. The establishment of shrub islands can significantly extend the grazing season by providing evergreen forage that remains available during periods when traditional pastures are dormant, particularly in the face of increased drought frequency and changing precipitation patterns. This incorporation of shrubs enhances the nutritional profile of the forage, offering increased levels of protein, minerals, and vitamins that are often lacking in pasturelands devoid of such vegetation, thus supporting livestock health and resilience during climate-induced stressors. Research has demonstrated the potential contributions of forage shrubs to both economic returns and environmental management. For instance, Monjardino [212] pointed out that incorporating perennial forage shrubs in Australian agriculture can enhance farm profitability by an average of 24% while providing reliable “out-of-season” feed, thus reducing supplementary feed costs and improving animal production, as demonstrated using the model of an integrated dryland agricultural system (MIDAS). Additionally, these shrubs improve water use efficiency and carbon storage, addressing challenges associated with climate change. Revell [213] highlighted the importance of breeding and incorporating shrubs, such as tagasaste (Chamaecytisus palmensis) and saltbush (Atriplex spp.), into farming systems and within fences in the southwest of Western Australia to improve diet quality, maintain soil health, and support livestock nutrition amidst increasing climate variability. Further investigations by Revell [214] confirmed that Australian perennial shrub species can enhance the feed base for grazing livestock, particularly in low- and medium-rainfall zones, by providing high-quality forage options during periods of scarcity. Additionally, Robertson and [215] provided an overview of the role of perennial pastures, including shrubs, in cropping systems in southern Australia, underscoring the need for ongoing research to explore the future benefits of these systems. In summary, incorporating shrubs into pasture systems not only extends the grazing season but also improves the nutritional quality of forage available to livestock, making this diversification essential for optimizing grazing systems amid the challenges posed by climate variability. Further research is needed regarding optimal synergistic or associative benefits derived from diverse supplementary forages established in these wet spots on livestock nutrition, health, and welfare, as well as their impact on the environment. Utilizing techniques such as ground cover fabric to enhance seedbed conditions can aid in the successful establishment of native forbs, contributing to improved forage quality for livestock [216]. Additionally, the inclusion of native tree islands in silvopastoral systems can further enhance biodiversity and ecosystem functioning in tropical regions, providing alternative forage sources while restoring degraded areas for conservation purposes [217]. Understanding the functional traits of plant species on islands is crucial for predicting establishment success, biotic interactions, and evolutionary pathways, highlighting the importance of considering ecological strategies when selecting species for restoration efforts [218].

2.5.4. Water Treatment, Water Points, and Strategic Shade Allocation

Ensuring high-quality water is fundamental to cattle health, as poor water quality can lead to various health issues that adversely affect productivity and welfare (Figure 1). Implementing effective water treatment methods is essential, particularly in regions where natural water sources may be contaminated due to climate change effects. Filtration systems, including sand filters, membrane filters, and UV light treatment, are commonly employed to remove particulates and pathogens in water points. Chemical treatments such as chlorination effectively disinfect water by killing harmful bacteria and viruses. A supportive study by Kim [219] indicated that disinfectants such as chlorine and ozone can effectively reduce Escherichia coli O157:H7 levels in contaminated water, achieving significant reductions when applied in specific concentrations. Maintaining a chlorine concentration of 1–3 parts per million (ppm) is recommended for effective livestock water treatment [11]. Lactic acid combined with acidic calcium sulfate reduced Escherichia coli, while no effect was observed on the amount of water consumed by cattle drinking water that contained compounds [220]. Cattle consuming water treated with 0.1% lactic acid and 0.9% acidic calcium sulfate (pH 2.1) drank an average of 18.6 L per day, comparable to untreated water consumption [221]. In local or rural settings, biofiltration systems that utilize layers of gravel, sand, and activated carbon can effectively remove contaminants while being cost-efficient. Rainwater harvesting also presents an effective method for ensuring water availability during dry periods, provided proper treatment measures are implemented to maintain water quality. A study by Richards [222] demonstrated that implementing rainwater harvesting systems in remote areas resulted in a 30% increase in water availability for livestock, thereby reducing dependence on unreliable surface water sources. Furthermore, research by Sobsey [223] found that biofiltration systems improved water quality by reducing total coliform bacteria levels by up to 95%, ensuring that cattle had access to safe drinking water. Establishing reliable water sources and effective water treatment methods can significantly enhance cattle health, productivity, and resilience against climate-related challenges.
The impact of climate change on water availability is a pressing issue that requires immediate attention. Developing accessible water sources is essential for maintaining cattle hydration, especially during heat stress or drought. Water is vital for various physiological functions, including digestion, nutrient absorption, and thermoregulation. As climate change exacerbates water scarcity, cattle are forced to walk longer distances to access water, particularly in free rangeland grazing systems. Water points are not just sources of water but also strategic tools that can be placed within 1 to 2 km (0.6 to 1.2 miles) of grazing areas to minimize the time and energy cattle expend walking to water. A study by Hunt [224] demonstrated that cattle within 1 km of water sources exhibited improved distribution across pastures, while distances exceeding 2 km reduced grazing efficiency and increased localized overgrazing. Cattle walking distances can increase energy costs by 10% to 24% compared to stall-fed animals, highlighting the importance of minimizing travel distance to water. A study found that energy costs for locomotion varied, with horizontal movement costing approximately 2.8 to 2.9 kJ/kg of body weight per kilometer. Cattle with access to water points less than 1 km away had a 15% improvement in their intake, ultimately enhancing their performance.
Conversely, the scarcity of water resources in drought-prone areas leads to cattle congestion at water points, resulting in overgrazing and pasture nutrient depletion. This ultimately causes the degradation of grazing areas. Muzzo and Provenza [225] recommended investing in increasing the number of water points to encourage cattle to utilize a broader area, promoting better pasture health and nutrient cycling in response to climate change. In areas with water scarcity, increasing the number of water points within strategic distances of 1 to 2 km enhances cattle performance, reduces animal congestion at water points, and maintains pasture health. Furthermore, integrating water point placement with vegetation assessments could optimize grazing management, as cattle are more likely to utilize areas with better forage availability near water sources.
In addition to water availability, providing cover and shade is essential for effective thermoregulation in cattle. Shade structures, tree plantings, and other forms of cover can significantly reduce heat stress, improving animal welfare and performance. Shade reduces plasma cortisol levels, indicating lower stress, which is vital for maintaining appetite and growth rates [226]. Tree shade significantly lowers black globe temperatures (BGTs), with studies showing shaded areas averaging 30.2 °C compared to 35.5 °C outside [227]. Cattle under shade exhibited lower rectal temperatures (RTs) and respiration rates (RRs), indicating better thermoregulation. Non-shaded cows had an average RT of 40.1 °C, while shaded cows had an RT of 39.3 °C [227]. Titto [228] reported that cattle with access to shade altered their behavior, spending more time grazing and ruminating, which is crucial for their overall health and productivity. Cattle under shade minimize their heat load, reducing the energy expenditure required for thermoregulation and allocating more energy to production. For example, shade use in hot climates increased significantly; over 50% of Holstein dairy and Belgian Blue beef cattle utilized shade, maintaining normal RR and panting scores compared to those without shade [229]. Cows with access to shade exhibited higher milk yields and better milk composition, with significant differences in solids-not-fat yield and lactose percentage [230]. Yearling Hereford steers gained an additional 19 lbs over the summer when shade was available [231]. Therefore, farmers can improve resilience to climate variability by incorporating water treatment, water points, and shade allocation strategies, ultimately fostering sustainable cattle systems.

2.5.5. Real-Time GPS Coupled with Accelerometer Collars

Cattle exhibit divergent foraging and site preferences, resulting in discrete behavioral events [232]. Such events include grazing, walking, resting, or ruminating activities. Grazing activities are determined by bite size formation, the bite size rate, and time spent foraging. Cattle allocate their grazing activities across different spatial scales and this consists of a hierarchy, from a feeding patch where they graze specific forage species, to a feeding station that includes multiple patches, and ultimately to the broader landscape, which encompasses the entire grazing area and influences their long-term movement patterns, influenced by their physiological state, the pasture’s structure and composition, and past experiences [233]. Using their senses, animals perform different activities, enabling them to select more profitable patches within feeding stations. In these stations, cattle initially recognize greener patches as indicative of greater nutritional quality, while forage abundance is another variable determining where to graze. Additionally, flat terrains are preferred over steep slopes, and grazing distribution is partly influenced by proximity to water and shelter and previous experience [234]. Hence, understanding both grazing events and distribution helps determine effective cattle grazing behaviors and, indirectly, animal performance and welfare.
In the 1990s, global positioning system (GPS) technology emerged and now allows researchers to accurately monitor livestock grazing patterns. Data from collar-mounted GPS units can be collected over weeks or even months at fine temporal intervals to help evaluate livestock movement and thus assist the rancher in improving techniques used to manage grazing distribution. While livestock instinctively seek the best available forage, climate change alters forage distribution, abundance, and quality over time, creating immediate and long-term challenges for cattle nutrition. Understanding grazing patterns and forage selection through tools such as GPS technology enables adaptive management to mitigate these effects by optimizing resource use, preventing overgrazing, improving forage quality, and ensuring the sustainability of grazing systems under climate change challenges. GPS units attached to collars can accurately track cattle movements and grazing patterns [235], providing valuable insights into the time cattle spend foraging, resting, and walking. Using real-time GPS collars, ranchers can accurately monitor cattle behavior and grazing patterns, allowing them to analyze the data within short time ranges and subsequently re-allocate cattle based on their preferences for forage and grazing conditions, helping to mitigate nutritional stress. A longer time spent grazing at a given location indicates a preference for the forage available at that location. For instance, Bello [236] found that real-time GPS collars can be used to identify locations where cattle prefer to forage and thus provide information for targeted interventions, such as enhancing forage availability or creating additional favorable locations, helping to ensure consistent access to nutritious feed during periods of reduced availability.
Current real-time GPS collars also measure air temperatures, offering valuable insights into cattle behavior and habitat preferences. Ranchers can analyze integrated temperature and GPS data to identify specific areas where cattle spend more time due to favorable temperature conditions, allowing for targeted improvements. The data can also help them to understand cattle movement patterns and behavior, enabling farmers to adapt grazing practices to enhance animal welfare and optimize livestock management strategies based on temperature preferences and comfort levels. GPS collars with integrated accelerometers monitor various animal behaviors, including periods when a cow is lying down or grazing. Ramezani [237] successfully used accelerometers to detect periods of lying down, rumination, and activity times in weaned calves during regrouping events, and Versluijs [238] identified common behaviors in free-ranging cattle, such as feeding, resting, walking, scratching, and self-grooming. Hence, an accelerometer provides valuable insights into cattle activity levels and behavior, which are crucial for managing climate change-induced stress. Accelerometers enable timely interventions. Džermeikaitė [239] demonstrated that real-time data from accelerometers helped manage heat stress by allowing farmers to respond promptly to temperature-induced stress, thereby improving animal welfare and productivity. Similarly, Sprinkle [240] highlighted the benefits of integrating accelerometer data with GPS for predictive analysis, assisting in managing grazing areas more effectively and mitigating the impacts of climate variability. However, some studies, such as those by Riaboff [241], have pointed out that accelerometers may sometimes provide inaccurate data under certain environmental conditions or with specific animal behaviors. Despite these challenges, real-time temperature–GPS collars coupled with accelerometers offer a comprehensive approach to managing cattle nutritional stress amidst climate challenges. These advanced technologies enable monitoring of behavioral responses and activity alterations in cattle exposed to heat stress, aiding in assessing the variability of stress experienced. Therefore, using these innovative tools and nutritional interventions, livestock producers can effectively address nutritional stress challenges in the face of climate change.

3. Conclusions

Climate change is linked to increasing levels of greenhouse gases in the atmosphere, leading to rising temperatures, altered precipitation patterns, and more frequent extreme weather events, which impact ecological systems and agricultural practices. Cattle nutrition and productivity are particularly impacted by changes in climate due to reductions in forage quality and productivity. Extreme weather events, such as flooding and droughts, may also reduce water availability and/or accessibility, with negative connotations for cattle nutrition and health. Soil microbial communities, crucial for nutrient cycling, are significantly impacted by temperature and moisture regimes. Extreme weather events such as flooding and droughts intensify these issues by reducing both water access and forage quality, further complicating cattle management. The shifts in climate also affect soil microbial communities, which are crucial for nutrient cycling. While warmer temperatures may increase microbial activity in some ecosystems, excessive heat or dryness can hinder nutrient cycling, disrupting the availability of essential nutrients for cattle. Additionally, changing plant communities, including shifts from C3 to C4 plants and alterations in plant secondary compound concentrations, may further impact forage quality and the nutritional balance of grazing cattle. In response to these challenges, several integrated strategies are needed to strengthen cattle resilience. Genetic selection, breeding for heat tolerance, disease resistance, and epigenetic research can enhance cattle adaptability. Remote sensing and GIS technologies offer valuable tools for monitoring water availability, plant health, and optimal grazing areas. Diversifying plant species, including perennial legumes and shrubs, helps to improve forage quality, enhance cover, and extend grazing seasons while reducing environmental impacts. Moreover, strategic management practices, such as developing water points, using GPS technology, and creating thermoregulatory structures, support cattle resilience in real-time by improving the management of forage, habitat, and environmental stressors. Real-time GPS collars combined with accelerometers provide detailed data on cattle movement and activity, aiding in the management of nutritional and heat stress. These strategies collectively enhance cattle resilience, health, and productivity amidst climate change challenges.

Author Contributions

B.I.M.: Conceptualization, Writing—review and editing original draft. J.J.V.: Review and editing original draft. R.D.R.: Review and editing original draft. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by AFRI Sustainable Agricultural Systems Coordinated Agricultural Project (SAS-CAP) Award Number: 2021-69012-35952 from the USDA National Institute of Food and Agriculture.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Gupta, A.; Anand, A.; Arora, D.; Bhattacharya, A.M.; Chahal, P.; Chauhan, R.; Dhall, A.; Dikshit, V.; Gulati, R.; Haque, A.J.; et al. Conservation, Sustainability, and Environmental Justice in India; Lexington Books: Lanham, MD, USA, 2020. [Google Scholar]
  2. Friedlingstein, P.; Jones, M.W.; O’Sullivan, M.; Andrew, R.M.; Bakker, D.C.E.; Hauck, J.; Le Quéré, C.; Peters, G.P.; Peters, W.; Pongratz, J.; et al. Global Carbon Budget 2021. Earth Syst. Sci. Data 2022, 14, 1917–2005. [Google Scholar] [CrossRef]
  3. Soon, W.; Connolly, R.; Connolly, M.; Akasofu, S.I.; Baliunas, S.; Berglund, J.; Bianchini, A.; Briggs, W.M.; Butler, C.J.; Cionco, R.G.; et al. The Detection and Attribution of Northern Hemisphere Land Surface Warming (1850–2018) in Terms of Human and Natural Factors: Challenges of Inadequate Data. Climate 2023, 11, 179. [Google Scholar] [CrossRef]
  4. Forster, P.M.; Smith, C.J.; Walsh, T.; Lamb, W.F.; Lamboll, R.; Hauser, M.; Ribes, A.; Rosen, D.; Gillett, N.; Palmer, M.D.; et al. Indicators of Global Climate Change 2022: Annual Update of Large-Scale Indicators of the State of the Climate System and Human Influence. Earth Syst. Sci. Data 2023, 15, 2295–2327. [Google Scholar] [CrossRef]
  5. Zhao, S.; Liu, M.; Tao, M.; Zhou, W.; Lu, X.; Xiong, Y.; Li, F.; Wang, Q. The Role of Satellite Remote Sensing in Mitigating and Adapting to Global Climate Change. Sci. Total Environ. 2023, 904, 166820. [Google Scholar] [CrossRef] [PubMed]
  6. Henchion, M.; Moloney, A.P.; Hyland, J.; Zimmermann, J.; McCarthy, S. Review: Trends for Meat, Milk and Egg Consumption for the next Decades and the Role Played by Livestock Systems in the Global Production of Proteins. Animal 2021, 15, 100287. [Google Scholar] [CrossRef]
  7. Cheng, M.; McCarl, B.; Fei, C. Climate Change and Livestock Production: A Literature Review. Atmosphere 2022, 13, 140. [Google Scholar] [CrossRef]
  8. Kelly, A.E.; Goulden, M.L. Rapid Shifts in Plant Distribution with Recent Climate Change. Proc. Natl. Acad. Sci. USA 2008, 105, 11823–11826. [Google Scholar] [CrossRef] [PubMed]
  9. Harrison, S.; Spasojevic, M.J.; Li, D. Climate and Plant Community Diversity in Space and Time. Proc. Natl. Acad. Sci. USA 2020, 117, 4464–4470. [Google Scholar] [CrossRef] [PubMed]
  10. Sprunger, C.D.; Lindsey, A.; Lightcap, A. Above- and Belowground Linkages during Extreme Moisture Excess: Leveraging Knowledge from Natural Ecosystems to Better Understand Implications for Row-Crop Agroecosystems. J. Exp. Bot. 2023, 74, 2845–2859. [Google Scholar] [CrossRef] [PubMed]
  11. 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. [Google Scholar] [CrossRef]
  12. Apgaua, D.M.G.; Tng, D.Y.P.; Forbes, S.J.; Ishida, Y.F.; Vogado, N.O.; Cernusak, L.A.; Laurance, S.G.W. Elevated Temperature and CO2 Cause Differential Growth Stimulation and Drought Survival Responses in Eucalypt Species from Contrasting Habitats. Tree Physiol. 2019, 39, 1806–1820. [Google Scholar] [CrossRef] [PubMed]
  13. Blond, B.; Majkić, M.; Spasojević, J.; Hristov, S.; Radinović, M.; Nikolić, S.; Anđušić, L.; Čukić, A.; Došenović Marinković, M.; Vujanović, B.D.; et al. Influence of Heat Stress on Body Surface Temperature and Blood Metabolic, Endocrine, and Inflammatory Parameters and Their Correlation in Cows. Metabolites 2024, 14, 104. [Google Scholar] [CrossRef]
  14. Crous, K.Y. Plant Responses to Climate Warming: Physiological Adjustments and Implications for Plant Functioning in a Future, Warmer World. Am. J. Bot. 2019, 106, 1049. [Google Scholar] [CrossRef]
  15. Kennedy, P.M.; Milligan, L.P. Effects of Cold Exposure on Digestion, Microbial Synthesis and Nitrogen Transformations in Sheep. Br. J. Nutr. 1978, 39, 105–117. [Google Scholar] [CrossRef]
  16. Bell, M.J.; Eckard, R.J.; Harrison, M.T.; Neal, J.S.; Cullen, B.R. Effect of Warming on the Productivity of Perennial Ryegrass and Kikuyu Pastures in South-Eastern Australia. Crop Pasture Sci. 2013, 64, 61–70. [Google Scholar] [CrossRef]
  17. Havrilla, C.A.; Bradford, J.B.; Yackulic, C.B.; Munson, S.M.; Caroline Havrilla, C.A.; Jarvis, S. Divergent Climate Impacts on C 3 versus C 4 Grasses Imply Widespread 21st Century Shifts in Grassland Functional Composition. Divers. Distrib. 2023, 29, 379–394. [Google Scholar] [CrossRef]
  18. Sanz-Sáez, Á.; Erice, G.; Aguirreolea, J.; Irigoyen, J.J.; Sánchez-Díaz, M. Alfalfa Yield under Elevated CO2 and Temperature Depends on the Sinorhizobium Strain and Growth Season. Environ. Exp. Bot. 2012, 77, 267–273. [Google Scholar] [CrossRef]
  19. Sisco, M.R.; Bosetti, V.; Weber, E.U. When Do Extreme Weather Events Generate Attention to Climate Change? Clim. Chang. 2017, 143, 227–241. [Google Scholar] [CrossRef]
  20. Giridhar, K.; Samireddypalle, A.; Giridhar, K.; Samireddypalle, A. Impact of Climate Change on Forage Availability for Livestock. In Climate Change Impact on Livestock: Adaptation and Mitigation; Springer: New Delhi, India, 2015; pp. 97–112. [Google Scholar] [CrossRef]
  21. Chaplin-Kramer, R.; George, M.R. Effects of Climate Change on Range Forage Production in the San Francisco Bay Area. PLoS ONE 2013, 8, e57723. [Google Scholar] [CrossRef]
  22. Visconti-Moreno, E.F.; Valenzuela-Balcázar, I.G. Pores Distribution Influences the Soil Microorganism’s Response to Changes in Temperature and Moisture. Eurasian J. Soil Sci. 2023, 12, 28–36. [Google Scholar] [CrossRef]
  23. Agom, O.; Gbadebo, A. Effects of Climate Change and Global Warming on Enzymes. Rev. Gestão Soc. Ambient. 2024, 18, e06079. [Google Scholar] [CrossRef]
  24. Choudhary, P.; Bhatt, S.; Chatterjee, S. From Freezing to Functioning: Cellular Strategies of Cold-Adapted Bacteria for Surviving in Extreme Environments. Arch. Microbiol. 2024, 206, 1–13. [Google Scholar] [CrossRef] [PubMed]
  25. Kumar, S.; Najar, I.N.; Sharma, P.; Tamang, S.; Mondal, K.; Das, S.; Sherpa, M.T.; Thakur, N. Temperature—A Critical Abiotic Paradigm That Governs Bacterial Heterogeneity in Natural Ecological System. Environ. Res. 2023, 234, 116547. [Google Scholar] [CrossRef] [PubMed]
  26. Xiao, X.; Ma, Z.; Zhang, J.; Sun, B.; Zhou, J.; Liang, Y. Coupling Temperature-Dependent Spatial Turnover of Microbes and Plants Using the Metabolic Theory of Ecology. New Phytol. 2023, 238, 383–392. [Google Scholar] [CrossRef]
  27. Zhang, J.; Yang, L.; Yu, M.; Chen, X. Response of Extreme Rainfall to Atmospheric Warming and Wetting: Implications for Hydrologic Designs Under a Changing Climate. J. Geophys. Res. Atmos. 2023, 128, e2022JD038430. [Google Scholar] [CrossRef]
  28. Li, W.; Su, T.; Shen, Y.; Ma, H.; Zhou, Y.; Lu, Q.; Wang, G.; Liu, Z.; Li, J. Effects of Warming Seasonal Rotational Grazing on Plant Communities’ Structure and Diversity in Desert Steppe. Ecol. Evol. 2023, 13, e9748. [Google Scholar] [CrossRef] [PubMed]
  29. Rojas, Y.P.; Ghezzehei, T. Determining Soil Microbial Activity Based on Soil Moisture and Average Functions. In Proceedings of the EGU General Assembly 2023, Vienna, Austria, 24–28 April 2023; p. EGU23-899. [Google Scholar] [CrossRef]
  30. Chatterjee, D.; Saha, S. Response of Soil Properties and Soil Microbial Communities to the Projected Climate Change. In Advances in Crop Environment Interaction; Springer: Singapore, 2018; pp. 87–136. [Google Scholar]
  31. Li, D.; Wu, C.; Wu, J. Soil Fungal Community Has Higher Network Stability than Bacterial Community in Response to Warming and Nitrogen Addition in a Subtropical Primary Forest. Appl. Environ. Microbiol. 2024, 90, e00001-24. [Google Scholar] [CrossRef]
  32. Qiu, Y.; Zhang, K.; Zhao, Y.; Zhao, Y.; Wang, B.; Wang, Y.; He, T.; Xu, X.; Bai, T.; Zhang, Y.; et al. Climate Warming Suppresses Abundant Soil Fungal Taxa and Reduces Soil Carbon Efflux in a Semi-Arid Grassland. mLife 2023, 2, 389–400. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, X.; Zhang, Q.; Zhang, Z.; Li, W.; Liu, W.; Xiao, N.; Liu, H.; Wang, L.; Li, Z.; Ma, J.; et al. Decreased Soil Multifunctionality Is Associated with Altered Microbial Network Properties under Precipitation Reduction in a Semiarid Grassland. iMeta 2023, 2, e106. [Google Scholar] [CrossRef]
  34. Huntley, B.J. Soil, Water and Nutrients. In Ecology Angola; Springer International Publishing: Cham, Switzerland, 2023; pp. 127–147. [Google Scholar] [CrossRef]
  35. Naylor, D.; McClure, R.; Jansson, J. Trends in Microbial Community Composition and Function by Soil Depth. Microorganisms 2022, 10, 540. [Google Scholar] [CrossRef] [PubMed]
  36. Bainard, L.D.; Evans, B.; Malis, E.; Yang, T.; Bainard, J.D. Influence of Annual Plant Diversity on Forage Productivity and Nutrition, Soil Chemistry, and Soil Microbial Communities. Front. Sustain. Food Syst. 2020, 4, 560479. [Google Scholar] [CrossRef]
  37. Danar Dewa, D.; Buchori, I. Impacts of Rapid Urbanization on Spatial Dynamics of Land Use-Based Carbon Emission and Surface Temperature Changes in the Semarang Metropolitan Region, Indonesia Dimas Danar Dewa · Imam Buchori. Environ. Monit. Assess. 2023, 195, 259. [Google Scholar] [CrossRef] [PubMed]
  38. Fischer, M.; Günther, M.; Pischinger, S.; Kramer, U.; Nederlof, C.; van Almsick, T. Influence of Desulfurization Strategies for Methane Gaseous Direct Injection Engine on Carbon Dioxide Emissions. J. Nat. Gas. Sci. Eng. 2022, 108, 104822. [Google Scholar] [CrossRef]
  39. Koutsoyiannis, D.; Kundzewicz, Z.W. Atmospheric Temperature and CO2: Hen-Or-Egg Causality? Sci 2020, 2, 83. [Google Scholar] [CrossRef]
  40. Lacis, A.A.; Schmidt, G.A.; Rind, D.; Ruedy, R.A. Atmospheric CO2: Principal Control Knob Governing Earth’s Temperature. Science 2010, 330, 356–359. [Google Scholar] [CrossRef]
  41. Sage, R.F.; Kubien, D.S. The Temperature Response of C3 and C4 Photosynthesis. Plant Cell Environ. 2007, 30, 1086–1106. [Google Scholar] [CrossRef]
  42. Jobe, T.O.; Rahimzadeh Karvansara, P.; Zenzen, I.; Kopriva, S. Ensuring Nutritious Food Under Elevated CO2 Conditions: A Case for Improved C4 Crops. Front. Plant Sci. 2020, 11, 563726. [Google Scholar] [CrossRef] [PubMed]
  43. Wijewardene, I.; Shen, G.; Zhang, H. Enhancing Crop Yield by Using Rubisco Activase to Improve Photosynthesis under Elevated Temperatures. Stress Biol. 2021, 1, 2. [Google Scholar] [CrossRef] [PubMed]
  44. McDowell, N.G. Mechanisms Linking Drought, Hydraulics, Carbon Metabolism, and Vegetation Mortality. Plant Physiol. 2011, 155, 1051–1059. [Google Scholar] [CrossRef]
  45. Buanafina, M.M.d.O.; Morris, P. The Impact of Cell Wall Feruloylation on Plant Growth, Responses to Environmental Stress, Plant Pathogens and Cell Wall Degradability. Agronomy 2022, 12, 1847. [Google Scholar] [CrossRef]
  46. Ashworth, A.J.; Kharel, T.; Sauer, T.; Adams, T.C.; Philipp, D.; Thomas, A.L.; Owens, P.R. Spatial Monitoring Technologies for Coupling the Soil Plant Water Animal Nexus. Sci. Rep. 2022, 12, 3508. [Google Scholar] [CrossRef] [PubMed]
  47. More, S.J.; Ravi, V.; Raju, S. Tropical Tuber Crops. In Vegetables for Nutrition and Entrepreneurship; Springer: Singapore, 2019; Volume 1, pp. 719–758. [Google Scholar] [CrossRef]
  48. Sershen; Perumal, A.; Varghese, B.; Govender, P.; Ramdhani, S.; Berjak, P. Effects of Elevated Temperatures on Germination and Subsequent Seedling Vigour in Recalcitrant Trichilia Emetica Seeds. S. Afr. J. Bot. 2014, 90, 153–162. [Google Scholar] [CrossRef]
  49. Bhattacharya, A. Physiological Processes in Plants Under Low Temperature Stress; Springer: Singapore, 2022. [Google Scholar]
  50. Abobatta, W.F. The Influence of Climate Change on Interactions between Environmental Stresses and Plants. In Plant Stress Mitigators: Types, Techniques and Functions; Academic Press: Cambridge, MA, USA, 2023; pp. 425–434. [Google Scholar] [CrossRef]
  51. Anjum, S.A.; Ashraf, U.; Zohaib, A.; Tanveer, M.; Naeem, M.; Ali, I.; Tabassum, T.; Nazir, U. Žemės Ūkio Augalų Reakcija į Sausros Sukurtą Stresą: Apžvalga. Zemdirbyste 2017, 104, 267–276. [Google Scholar] [CrossRef]
  52. Yadav, S.K. Cold Stress Tolerance Mechanisms in Plants. A Review. Agron. Sustain. Dev. 2010, 30, 515–527. [Google Scholar] [CrossRef]
  53. Bhandari, K.; Nayyar, H. Low Temperature Stress in Plants: An Overview of Roles of Cryoprotectants in Defense. In Physiological Mechanisms and Adaptation Strategies in Plants Under Changing Environment; Springer: New York, NY, USA, 2014; Volume 1, pp. 193–265. ISBN 9781461485919. [Google Scholar]
  54. Ojo, T.A.; Kirkman, K.; Tedder, M. Effects of Warming and Rainfall Variation on Grass Phenology and Regenerative Responses in Mesic Grassland. S. Afr. J. Bot. 2024, 174, 107–115. [Google Scholar] [CrossRef]
  55. Plos, C.; Hensen, I.; Korell, L.; Auge, H.; Römermann, C. Plant Species Phenology Differs between Climate and Land-Use Scenarios and Relates to Plant Functional Traits. Ecol. Evol. 2024, 14, e11441. [Google Scholar] [CrossRef] [PubMed]
  56. Kangombe, F.N.; Throop, H.; Sala, O.; Vivoni, E.; Pigg, K.; Hultine, K.; Kwembeya, E. Environmental Drivers of Vegetative and Flowering Phenology in Drylands. Doctoral Dissertation, Arizona State University, Tempe, AZ, USA, 2023. [Google Scholar]
  57. Fang, H.; Sha, M.; Xie, Y.; Lin, W.; Qiu, D.; Tu, J.; Tan, X.; Li, X.; Sha, Z. Shifted Global Vegetation Phenology in Response to Climate Changes and Its Feedback on Vegetation Carbon Uptake. Remote Sens. 2023, 15, 2288. [Google Scholar] [CrossRef]
  58. Schüle, M.; Heinken, T.; Fartmann, T. Long-Term Effects of Environmental Alterations in Protected Grasslands—Land-Use History Determines Changes in Plant Species Composition. Ecol. Eng. 2023, 188, 106878. [Google Scholar] [CrossRef]
  59. Liu, H.; Mi, Z.; Lin, L.; Wang, Y.; Zhang, Z.; Zhang, F.; Wang, H.; Liu, L.; Zhu, B.; Cao, G.; et al. Shifting Plant Species Composition in Response to Climate Change Stabilizes Grassland Primary Production. Proc. Natl. Acad. Sci. USA 2018, 115, 4051–4056. [Google Scholar] [CrossRef]
  60. French, K.E. Species Composition Determines Forage Quality and Medicinal Value of High Diversity Grasslands in Lowland England. Agric. Ecosyst. Environ. 2017, 241, 193–204. [Google Scholar] [CrossRef]
  61. Zhang, G.; Dai, E.; Dawaqiongda; Luobu; Fu, G. Effects of Climate Change and Fencing on Forage Nutrition Quality of Alpine Grasslands in the Northern Tibet. Plants 2023, 12, 3182. [Google Scholar] [CrossRef] [PubMed]
  62. Yoshihara, Y.; Aoki, R.; Kinugasa, T.; Sasaki, T. Predicted Effects of Simulated Ambient Warming and Moisture on Forage Nutrient Quality and Community Composition in Mongolian an Arid Grassland. Rangel. J. 2022, 44, 159–164. [Google Scholar] [CrossRef]
  63. Petit Bon, M.; Bråthen, K.A.; Ravolainen, V.T.; Ottaviani, G.; Böhner, H.; Jónsdóttir, I.S. Herbivory and Warming Have Opposing Short-Term Effects on Plant-Community Nutrient Levels across High-Arctic Tundra Habitats. J. Ecol. 2023, 111, 1514–1530. [Google Scholar] [CrossRef]
  64. Sebastià, M.T.; Banagar, F.; Palero, N.; Ibáñez, M.; Plaixats, J. Quality Production of Sainfoin Swards Challenged by Global Change in Mountain Areas in the Western Mediterranean. Agronomy 2023, 14, 6. [Google Scholar] [CrossRef]
  65. Habermann, E.; Dias de Oliveira, E.A.; Contin, D.R.; San Martin, J.A.B.; Curtarelli, L.; Gonzalez-Meler, M.A.; Martinez, C.A. Stomatal Development and Conductance of a Tropical Forage Legume Are Regulated by Elevated [CO2] under Moderate Warming. Front. Plant Sci. 2019, 10, 609. [Google Scholar] [CrossRef] [PubMed]
  66. Bykova, O.; Chuine, I.; Morin, X. Highlighting the Importance of Water Availability in Reproductive Processes to Understand Climate Change Impacts on Plant Biodiversity. Perspect. Plant Ecol. Evol. Syst. 2019, 37, 20–25. [Google Scholar] [CrossRef]
  67. Neal, J.S.; Fulkerson, W.J.; Hacker, R.B. Differences in Water Use Efficiency among Annual Forages Used by the Dairy Industry under Optimum and Deficit Irrigation. Agric. Water Manag. 2011, 98, 759–774. [Google Scholar] [CrossRef]
  68. Carter, P.R.; Sheaffer, C.C. Alfalfa Response to Soil Water Deficits. I. Growth, Forage Quality, Yield, Water Use, and Water-Use Efficiency1. Crop Sci. 1983, 23, 669–675. [Google Scholar] [CrossRef]
  69. Peterson, P.R.; Sheaffer, C.C.; Hall, M.H. Drought Effects on Perennial Forage Legume Yield and Quality. Agron. J. 1992, 84, 774–779. [Google Scholar] [CrossRef]
  70. Dumont, B.; Andueza, D.; Niderkorn, V.; Lüscher, A.; Porqueddu, C.; Picon-Cochard, C. A Meta-Analysis of Climate Change Effects on Forage Quality in Grasslands: Specificities of Mountain and Mediterranean Areas. Grass Forage Sci. 2015, 70, 239–254. [Google Scholar] [CrossRef]
  71. Küchenmeister, K.; Küchenmeister, F.; Kayser, M.; Wrage-Mönnig, N.; Isselstein, J. Influence of Drought Stress on Nutritive Value of Perennial Forage Legumes. Int. J. Plant Prod. 2013, 7, 693–710. [Google Scholar]
  72. Wilson, J.R. Effects of Water Stress on in Vitro Dry Matter Digestibility and Chemical Composition of Herbage of Tropical Pasture Species. Aust. J. Agric. Res. 1983, 34, 377–390. [Google Scholar] [CrossRef]
  73. Buxton, D.R. Quality-Related Characteristics of Forages as Influenced by Plant Environment and Agronomic Factors. Anim. Feed. Sci. Technol. 1996, 59, 37–49. [Google Scholar] [CrossRef]
  74. Reddy, G.S.N.; Matsumoto, G.I.; Shivaji, S. Sporosarcina Macmurdoensis Sp. Nov., from a Cyanobacterial Mat Sample from a Pond in the McMurdo Dry Valleys, Antarctica. Int. J. Syst. Evol. Microbiol. 2003, 53, 1363–1367. [Google Scholar] [CrossRef] [PubMed]
  75. Durand, J.-L.; Gonzalez-Dugo, V.; Gastal, F. How Much Do Water Deficits Alter the Nitrogen Nutrition Status of Forage Crops? Nutr. Cycl. Agroecosystems 2021, 88, 231–243. [Google Scholar] [CrossRef]
  76. Liu, Y.; Wu, Q.; Ge, G.; Han, G.; Jia, Y. Influence of Drought Stress on Afalfa Yields and Nutritional Composition. BMC Plant Biol. 2018, 18, 13. [Google Scholar] [CrossRef] [PubMed]
  77. Halim, R.A.; Buxton, D.R.; Hattendorf, M.J.; Carlson, R.E. Water-Stress Effects on Alfalfa Forage Quality After Adjustment for Maturity Differences. Agron. J. 1989, 81, 189–194. [Google Scholar] [CrossRef]
  78. Kalu, B.A.; Fick, G.W. Quantifying Morphological Development of Alfalfa for Studies of Herbage Quality1. Crop Sci. 1981, 21, 267–271. [Google Scholar] [CrossRef]
  79. Van Soest, P.J. Acknowledgments. In Nutritional Ecology of the Ruminant; Cornell University Press: Ithaca, NY, USA, 1994; pp. ix–xii. [Google Scholar]
  80. Tariq, A.; Ahmed, A. Plant Phenolics Production: A Strategy for Biotic Stress Management. In Plant Phenolics in Biotic Stress Management; Springer Nature: Singapore, 2024; pp. 441–454. [Google Scholar] [CrossRef]
  81. Villalba, J.J.; Ramsey, R.D.; Athanasiadou, S. Review: Herbivory and the Power of Phytochemical Diversity on Animal Health. Animal 2024, 101287. [Google Scholar] [CrossRef]
  82. Rahman, A.; Albadrani, G.M.; Waraich, E.A.; Awan, T.H.; Yavaş, İ.; Hussain, S.; Rahman, A.; Albadrani, G.M.; Waraich, E.A.; Awan, T.H.; et al. Plant Secondary Metabolites and Abiotic Stress Tolerance: Overview and Implications. In Plant Abiotic Stress Responses and Tolerance Mechanisms; IntechOpen: London, UK, 2023. [Google Scholar]
  83. La, V.H.; Kim, T.H.; Calderon-Urrea, A.; Janda, T. Editorial: Plant Signaling in Response to Environmental Stresses. Front. Plant Sci. 2023, 14, 1282465. [Google Scholar] [CrossRef] [PubMed]
  84. Ahmad, N.; Xu, Y.; Zang, F.; Li, D.; Liu, Z. The Evolutionary Trajectories of Specialized Metabolites towards Antiviral Defense System in Plants. Mol. Hortic. 2024, 4, 1–11. [Google Scholar] [CrossRef]
  85. Ray, R. Using Natural Variation in Nicotiana Attenuata to Elucidate Its Defense Response Against Herbivory. Doctoral Dissertation, Friedrich Schiller University Jena, Jena, Germany, 1990. [Google Scholar]
  86. Montesinos-Navarro, A.; López-Climent, M.F.; Pérez-Clemente, R.M.; Arenas-Sánchez, C.; Sánchez-Martín, R.; Gómez-Cadenas, A.; Verdú, M. Plant Metabolic Response to Stress in an Arid Ecosystem Is Mediated by the Presence of Neighbors. Ecology 2024, 105, e4247. [Google Scholar] [CrossRef] [PubMed]
  87. Waghorn, G. Beneficial and Detrimental Effects of Dietary Condensed Tannins for Sustainable Sheep and Goat Production—Progress and Challenges. Anim. Feed. Sci. Technol. 2008, 147, 116–139. [Google Scholar] [CrossRef]
  88. Lorenz, K.; Lal, R. The Depth Distribution of Soil Organic Carbon in Relation to Land Use and Management and the Potential of Carbon Sequestration in Subsoil Horizons. Adv. Agron. 2005, 88, 35–66. [Google Scholar] [CrossRef]
  89. Ortiz-López, B.; Mariezcurrena-Berasain, M.D.; Barajas-Cruz, R.; Velázquez-Garduño, G.; Pliego, A.B.; Adegbeye, M.J.; Salem, A.Z.M.; Mariezcurrena-Berasain, M.A. Sustainable Evaluation of Tannin Extract Biomass as a Feed Product Additive: Effects on Growth Performance, Meat Fatty Acid Profile, and Lipid Oxidation in Bullocks. Biomass Convers. Biorefinery 2024, 14, 5101–5107. [Google Scholar] [CrossRef]
  90. Sun, Y.; Alseekh, S.; Fernie, A.R. Plant Secondary Metabolic Responses to Global Climate Change: A Meta-Analysis in Medicinal and Aromatic Plants. Glob. Change Biol. 2023, 29, 477–504. [Google Scholar] [CrossRef]
  91. Laftouhi, A.; Cordero, M.A.W.; Mahraz, M.A.; Zerkani, H.; Hmamou, A.; Idrissi, A.M.; Imane, T.; Eloutassi, N.; Rais, Z.; Taleb, A.; et al. Study of the Physiological Behavior of Some Plants in Response to Climate Change Conditions. Pol. J. Environ. Stud. 2024, 33, 3733–3745. [Google Scholar] [CrossRef]
  92. González-Teuber, M.; Palma-Onetto, V.; Aguirre, C.; Ibáñez, A.J.; Mithöfer, A. Climate Change-Related Warming-Induced Shifts in Leaf Chemical Traits Favor Nutrition of the Specialist Herbivore Battus Polydamas Archidamas. Front. Ecol. Evol. 2023, 11, 1152489. [Google Scholar] [CrossRef]
  93. Beale, P.K.; Foley, W.J.; Moore, B.D.; Marsh, K.J. Warmer Ambient Temperatures Reduce Protein Intake by a Mammalian Folivore. Philos. Trans. R. Soc. B 2023, 378, 20220543. [Google Scholar] [CrossRef] [PubMed]
  94. Zhou, C.; Lv, C.; Miao, T.; Ma, X.; Xia, C. Interactive Effects of Rising Temperature, Elevated CO2 and Herbivory on the Growth and Stoichiometry of a Submerged Macrophyte Vallisneria Natans. Sustainability 2023, 15, 1200. [Google Scholar] [CrossRef]
  95. Balluffi-Fry, J.; Leroux, S.J.; Wiersma, Y.F.; Richmond, I.C.; Heckford, T.R.; Rizzuto, M.; Kennah, J.L.; Vander Wal, E. Integrating Plant Stoichiometry and Feeding Experiments: State-Dependent Forage Choice and Its Implications on Body Mass. Oecologia 2022, 198, 579–591. [Google Scholar] [CrossRef]
  96. Acamovic, T.; Brooker, J.D. Biochemistry of plant secondary metabolites and their effects in animals. Proc. Nutr. Soc. 2005, 64, 403–412. [Google Scholar] [CrossRef] [PubMed]
  97. Torres-Fajardo, R.A.; González-Pech, P.G.; Sandoval-Castro, C.A.; Torres-Acosta, J.F.d.J. Small Ruminant Production Based on Rangelands to Optimize Animal Nutrition and Health: Building an Interdisciplinary Approach to Evaluate Nutraceutical Plants. Animals 2020, 10, 1799. [Google Scholar] [CrossRef]
  98. Reddy, P.R.K.; Elghandour, M.M.M.Y.; Salem, A.Z.M.; Yasaswini, D.; Reddy, P.P.R.; Reddy, A.N.; Hyder, I. Plant Secondary Metabolites as Feed Additives in Calves for Antimicrobial Stewardship. Anim. Feed. Sci. Technol. 2020, 264, 114469. [Google Scholar] [CrossRef]
  99. Beck, M.R.; Gregorini, P. How Dietary Diversity Enhances Hedonic and Eudaimonic Well-Being in Grazing Ruminants. Front. Vet. Sci. 2020, 7, 486377. [Google Scholar] [CrossRef] [PubMed]
  100. Villalba, J.J.; Manteca, X. A Case for Eustress in Grazing Animals. Front. Vet. Sci. 2019, 6, 303. [Google Scholar] [CrossRef] [PubMed]
  101. Reddy, A.R.; Chaitanya, K.V.; Vivekanandan, M. Drought-Induced Responses of Photosynthesis and Antioxidant Metabolism in Higher Plants. J. Plant Physiol. 2004, 161, 1189–1202. [Google Scholar] [CrossRef] [PubMed]
  102. Gourlay, G.; Constabel, C.P. Condensed Tannins Are Inducible Antioxidants and Protect Hybrid Poplar against Oxidative Stress. Tree Physiol. 2019, 39, 345–355. [Google Scholar] [CrossRef] [PubMed]
  103. Popović, B.M.; Štajner, D.; Ždero-Pavlović, R.; Tumbas-Šaponjac, V.; Čanadanović-Brunet, J.; Orlović, S. Water Stress Induces Changes in Polyphenol Profile and Antioxidant Capacity in Poplar Plants (Populus Spp.). Plant Physiol. Biochem. 2016, 105, 242–250. [Google Scholar] [CrossRef]
  104. Bryant, J.P.; Stuart Chapin, F.; Klein Bryant, D.R.; Bryant, J.P.; Chapin, F.S. Herbivore-Plant Interactions at Northern Latitudes. Oikos 1983, 40, 357–368. [Google Scholar] [CrossRef]
  105. Horner, J.D. Nonlinear Effects of Water Deficits on Foliar Tannin Concentration. Biochermcal Syst. Ecol. 1990, 18, 211–213. [Google Scholar] [CrossRef]
  106. Matyssek, R.; Schnyder, H.; Oßwald, W.; Ernst, D.; Munch, J.C.; Pretzsch, H. Ecological Studies 220 Growth and Defence in Plants Resource Allocation at Multiple Scales; Springer: Berlin/Heidelberg, Germany, 2012. [Google Scholar]
  107. AbdElgawad, H.; Peshev, D.; Zinta, G.; Van Den Ende, W.; Janssens, I.A.; Asard, H. Climate Extreme Effects on the Chemical Composition of Temperate Grassland Species under Ambient and Elevated CO2: A Comparison of Fructan and Non-Fructan Accumulators. PLoS ONE 2014, 9, e92044. [Google Scholar] [CrossRef]
  108. Anuraga, M.; Duarsa, P.; Hill, M.; Lovett, J. Soil Moisture and Temperature Affect Condensed Tannin Concentrations and Growth in Lotus Corniculatus and Lotus Pedunculatus. Aust. J. Agric. Res. 1993, 44, 1667–1681. [Google Scholar] [CrossRef]
  109. Windley, H.R.; Shimada, T. Cold Temperature Improves Tannin Tolerance in a Granivorous Rodent. J. Anim. Ecol. 2020, 89, 471–481. [Google Scholar] [CrossRef] [PubMed]
  110. Beale, P.K.; Connors, P.K.; Dearing, M.D.; Moore, B.D.; Krockenberger, A.K.; Foley, W.J.; Marsh, K.J. Warmer Ambient Temperatures Depress Detoxification and Food Intake by Marsupial Folivores. Front. Ecol. Evol. 2022, 10, 888550. [Google Scholar] [CrossRef]
  111. Baumgard, L.H.; Rhoads, R.P. Effects of Heat Stress on Postabsorptive Metabolism and Energetics. Annu. Rev. Anim. Biosci. 2013, 1, 311–337. [Google Scholar] [CrossRef] [PubMed]
  112. 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. [Google Scholar] [CrossRef] [PubMed]
  113. Chen, L.; Thorup, V.M.; Kudahl, A.B.; Østergaard, S. Effects of Heat Stress on Feed Intake, Milk Yield, Milk Composition, and Feed Efficiency in Dairy Cows: A Meta-Analysis. J. Dairy. Sci. 2024, 107, 3207–3218. [Google Scholar] [CrossRef] [PubMed]
  114. Souza, V.C.; Moraes, L.E.; Santos, J.E.P.; Baumgard, L.H.; Mueller, N.D.; Kebreab, E. Modeling the Relationship between Heat Stress, Feed Intake, and Day Relative to Calving in Nonlactating Dairy Cows. J. Dairy. Sci. 2023, 106, 8942–8952. [Google Scholar] [CrossRef]
  115. Sammad, A.; Wang, Y.J.; Umer, S.; Lirong, H.; Khan, I.; Khan, A.; Ahmad, B.; Wang, Y. Nutritional Physiology and Biochemistry of Dairy Cattle under the Influence of Heat Stress: Consequences and Opportunities. Animals 2020, 10, 793. [Google Scholar] [CrossRef] [PubMed]
  116. Safa, S.; Kargar, S.; Moghaddam, G.A.; Ciliberti, M.G.; Caroprese, M. Heat Stress Abatement during the Postpartum Period: Effects on Whole Lactation Milk Yield, Indicators of Metabolic Status, Inflammatory Cytokines, and Biomarkers of the Oxidative Stress. J. Anim. Sci. 2019, 97, 122–132. [Google Scholar] [CrossRef]
  117. Antanaitis, R.; Džermeikaitė, K.; Krištolaitytė, J.; Ribelytė, I.; Bespalovaitė, A.; Bulvičiūtė, D.; Palubinskas, G.; Anskienė, L. The Impacts of Heat Stress on Rumination, Drinking, and Locomotory Behavior, as Registered by Innovative Technologies, and Acid–Base Balance in Fresh Multiparous Dairy Cows. Animals 2024, 14, 1169. [Google Scholar] [CrossRef]
  118. Giannone, C.; Bovo, M.; Ceccarelli, M.; Torreggiani, D.; Tassinari, P. Review of the Heat Stress-Induced Responses in Dairy Cattle. Animals 2023, 13, 3451. [Google Scholar] [CrossRef]
  119. Ross, A.D.; Bryden, W.L.; Bakau, W.; Burgess, L.W. Induction of Heat Stress in Beef Cattle by Feeding the Ergots of Claviceps Purpurea. Aust. Vet. J. 1989, 66, 247–249. [Google Scholar] [CrossRef] [PubMed]
  120. Idris, M.; Uddin, J.; Sullivan, M.; McNeill, D.M.; Phillips, C.J.C. Non-Invasive Physiological Indicators of Heat Stress in Cattle. Animals 2021, 11, 71. [Google Scholar] [CrossRef] [PubMed]
  121. Dash, S.; Chakravarty, A.K.; Singh, A.; Upadhyay, A.; Singh, M.; Yousuf, S. Effect of Heat Stress on Reproductive Performances of Dairy Cattle and Buffaloes: A Review. Vet. World 2016, 9, 235. [Google Scholar] [CrossRef] [PubMed]
  122. Fernandez-Novo, A.; Pérez-Garnelo, S.S.; Villagrá, A.; Pérez-Villalobos, N.; Astiz, S. The Effect of Stress on Reproduction and Reproductive Technologies in Beef Cattle—A Review. Animals 2020, 10, 2096. [Google Scholar] [CrossRef]
  123. Hemantaranjan, A. Heat Stress Responses and Thermotolerance. Adv. Plants Agric. Res. 2014, 1, 1–10. [Google Scholar] [CrossRef]
  124. Alemu, T.W.; Pandey, H.O.; Salilew Wondim, D.; Gebremedhn, S.; Neuhof, C.; Tholen, E.; Holker, M.; Schellander, K.; Tesfaye, D. Oxidative and Endoplasmic Reticulum Stress Defense Mechanisms of Bovine Granulosa Cells Exposed to Heat Stress. Theriogenology 2018, 110, 130–141. [Google Scholar] [CrossRef]
  125. Patra, A.K.; Saxena, J. A New Perspective on the Use of Plant Secondary Metabolites to Inhibit Methanogenesis in the Rumen. Phytochemistry 2010, 71, 1198–1222. [Google Scholar] [CrossRef]
  126. Chen, H.; Fan, W.; Zhang, H.; Yue, P.; Wang, R.; Zhang, W.; Mai, K. Effects of Dietary Methionine on Growth and Body Composition, Indicators of Digestion, Protein Metabolism and Immunity, and Resistance to Heat Stress of Abalone Haliotis Discus Hannai. Aquaculture 2023, 563, 738978. [Google Scholar] [CrossRef]
  127. Czech, B.; Wang, Y.; Szyda, J. Genome-Wide Association Study of Heat Stress Response in Bos taurus. bioRxiv 2023. [Google Scholar] [CrossRef]
  128. Lemal, P.; Tran, M.-N.; Schroyen, M.; Gengler, N. 143 Validation Strategies of Sensor Data in the Context of Heat Stress Detection. J. Anim. Sci. 2023, 101, 36–37. [Google Scholar] [CrossRef]
  129. Nickles, K.R.; Relling, A.E.; Garcia-Guerra, A.; Fluharty, F.L.; Parker, A.J. Environmental Stress during the Last Trimester of Gestation in Pregnant Cows and Its Effect on Offspring Growth Performance and Response to Glucose and Adrenocorticotropic Hormone. J. Anim. Sci. 2023, 101, skac332. [Google Scholar] [CrossRef] [PubMed]
  130. Chase, L.E. Cold Stress: Effects on Nutritional Requirements, Health and Performance. In Reference Module in Food Science; Elsevier: Amsterdam, The Netherlands, 2016. [Google Scholar] [CrossRef]
  131. Kim, W.-S.; Nejad, J.G.; Lee, H.-G. Impact of Cold Stress on Physiological, Endocrinological, Immunological, Metabolic, and Behavioral Changes of Beef Cattle at Different Stages of Growth. Animals 2023, 13, 1073. [Google Scholar] [CrossRef]
  132. Delfino, J.G.; Mathison, G.W. Effects of Cold Environment and Intake Level on the Energetic Efficiency of Feedlot Steers. J. Anim. Sci. 1991, 69, 4577–4587. [Google Scholar] [CrossRef] [PubMed]
  133. Wang, S.; Li, Q.; Peng, J.; Niu, H. Effects of Long-Term Cold Stress on Growth Performance, Behavior, Physiological Parameters, and Energy Metabolism in Growing Beef Cattle. Animals 2023, 13, 1619. [Google Scholar] [CrossRef] [PubMed]
  134. 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. [Google Scholar] [CrossRef]
  135. Coloma-García, W.; Mehaba, N.; Such, X.; Caja, G.; Salama, A.A.K. Effects of Cold Exposure on Some Physiological, Productive, and Metabolic Variables in Lactating Dairy Goats. Animals 2020, 10, 2383. [Google Scholar] [CrossRef] [PubMed]
  136. Berian, S. Effect of Cold Stress on Milk Yield, Physiological and Hemato-Biochemical Profile of Cross Bred Dairy Cattle. J. Anim. Res. 2019, 9, 335–338. [Google Scholar] [CrossRef]
  137. Sahib, Q.S.; Aafaq, I.; Ahmed, H.A.; Sheikh, G.G.; Ganai, I.A. Mitigating Cold Stress in Livestock by Nutritional Interventions: A Comprehensive Review. Indian. J. Anim. Res. 2024, 58, 353–363. [Google Scholar] [CrossRef]
  138. Wagner, D.G. Effects of Cold Stress on Cattle Performance and Management Factors to Reduce Cold Stress and Improve Performance. Bov. Pract. 1988, 23, 88–93. [Google Scholar] [CrossRef]
  139. Cantalapiedra-Hijar, G.; Abo-Ismail, M.; Carstens, G.E.; Guan, L.L.; Hegarty, R.; Kenny, D.A.; Mcgee, M.; Plastow, G.; Relling, A.; Ortigues-Marty, I. Review: Biological Determinants of between-Animal Variation in Feed Efficiency of Growing Beef Cattle. Animal 2018, 12, S321–S335. [Google Scholar] [CrossRef] [PubMed]
  140. Ames, D.R. Adjusting Rations for Climate. Vet. Clin. N. Am. Food Anim. Pract. 1988, 4, 543–550. [Google Scholar] [CrossRef] [PubMed]
  141. Young, B.A. Cold Stress as It Affects Animal Production. J. Anim. Sci. 1981, 52, 154–163. [Google Scholar] [CrossRef]
  142. 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. [Google Scholar] [CrossRef]
  143. Colditz, I.G.; Hine, B.C. Resilience in Farm Animals: Biology, Management, Breeding and Implications for Animal Welfare. Anim. Prod. Sci. 2016, 56, 1961–1983. [Google Scholar] [CrossRef]
  144. Paz, S. Climate Change Impacts on Vector-Borne Diseases in Europe: Risks, Predictions and Actions. Lancet Reg. Health—Eur. 2021, 1, 100017. [Google Scholar] [CrossRef] [PubMed]
  145. Chikezie, F.M.; Opara, K.N.; Ubulom, P.M.E. Impacts of Changing Climate on Arthropod Vectors and Diseases Transmission. Niger. J. Entomol. 2024, 40, 179–192. [Google Scholar] [CrossRef]
  146. Bryer, K.E. Effects of Temperature and Humidity on the Mortality of the Tick Dermacentor Reticulatus in the UK. Doctoral Dissertation, University of Bristol, Bristol, UK, 2020. [Google Scholar]
  147. Deshpande, G.; Beetch, J.E.; Heller, J.G.; Naqvi, O.H.; Kuhn, K.G. Assessing the Influence of Climate Change and Environmental Factors on the Top Tick-Borne Diseases in the United States: A Systematic Review. Microorganisms 2023, 12, 50. [Google Scholar] [CrossRef]
  148. Domatskiy, V.N.; Sivkova, E.I. The Influence of Climatogeographic Conditions on the Expansion of the Range of Ixodes Ticks. Entomol. Appl. Sci. Lett. 2023, 10, 1–9. [Google Scholar] [CrossRef]
  149. Hoch, T.; Madouasse, A.; Jacquot, M.; Wongnak, P.; Beugnet, F.; Bournez, L.; Cosson, J.F.; Huard, F.; Moutailler, S.; Plantard, O.; et al. Seasonality of Host-Seeking Ixodes Ricinus Nymph Abundance in Relation to Climate. Peer Community J. 2024, 4, e2. [Google Scholar] [CrossRef]
  150. Brunner, J.L.; Killilea, M.; Ostfeld, R.S. Overwintering Survival of Nymphal Ixodes Scapularis (Acari: Ixodidae) Under Natural Conditions. J. Med. Entomol. 2012, 49, 981–987. [Google Scholar] [CrossRef] [PubMed]
  151. Lemery, J.; Knowlton, K.; Sorensen, C. (Eds.) Global Climate Change and Human Health: From Science to Practice; John Wiley & Sons: Hoboken, NJ, USA, 2021; p. 636. [Google Scholar]
  152. Ogden, N.H.; Ben Beard, C.; Ginsberg, H.S.; Tsao, J.I. Possible Effects of Climate Change on Ixodid Ticks and the Pathogens They Transmit: Predictions and Observations. J. Med. Entomol. 2021, 58, 1536–1545. [Google Scholar] [CrossRef] [PubMed]
  153. Alexandre, J.; Dionizio, R.; Augusto, J.; Afonso, B.; Silmara, G.; Soares, L.; Pajeú E Silva, B.; Filipe, J.; Cajueiro, P.; Teles Coutinho, L.; et al. Occurrence of Foot Diseases in Cattle Attended at the Clínica de Bovinos de Garanhuns: Epidemiological, Clinical, Therapeutic and Economic Aspects. Ciência Anim. Bras. 2022, 23, e-72731. [Google Scholar] [CrossRef]
  154. Urban-Chmiel, R.; Mudroň, P.; Abramowicz, B.; Kurek, Ł.; Stachura, R. Lameness in Cattle—Etiopathogenesis, Prevention and Treatment. Animals 2024, 14, 1836. [Google Scholar] [CrossRef] [PubMed]
  155. Chepkwony, R.; Castagna, C.; Heitkönig, I.; Van Bommel, S.; Van Langevelde, F. Associations between Monthly Rainfall and Mortality in Cattle Due to East Coast Fever, Anaplasmosis and Babesiosis. Parasitology 2020, 147, 1743–1751. [Google Scholar] [CrossRef] [PubMed]
  156. Correia, L.; Loureiro, A.P.; Lilenbaum, W. Effects of Rainfall on Incidental and Host-Maintained Leptospiral Infections in Cattle in a Tropical Region. Vet. J. 2017, 220, 63–64. [Google Scholar] [CrossRef] [PubMed]
  157. Alizadeh, O. Changes in the Mean and Variability of Temperature and Precipitation over Global Land Areas. Environ. Res. Clim. 2023, 2, 035006. [Google Scholar] [CrossRef]
  158. Luppichini, M.; Bini, M.; Giannecchini, R.; Zanchetta, G.; Luppichini, M.; Bini, M.; Giannecchini, R.; Zanchetta, G. The Effects of the Temperature Increase on the Rainfall Regimes in North-Central Italy. In Proceedings of the EGUGA, Vienna, Austria, 23–28 April 2023; p. EGU-5430. [Google Scholar] [CrossRef]
  159. Lall, U.; Josset, L.; Russo, T. A Snapshot of the World’s Groundwater Challenges. Annu. Rev. Environ. Resour. 2020, 45, 171–194. [Google Scholar] [CrossRef]
  160. Godde, C.M.; Mason-D’Croz, D.; Mayberry, D.E.; Thornton, P.K.; Herrero, M. Impacts of Climate Change on the Livestock Food Supply Chain; a Review of the Evidence. Glob. Food Sec. 2021, 28, 100488. [Google Scholar] [CrossRef] [PubMed]
  161. Maurya, V.P.; Sejian, V.; Kumar, K.; Singh, G.; Naqvi, S.M.K. Walking Stress Influence on Livestock Production. In Environmental Stress and Amelioration in Livestock Production; Springer: Berlin/Heidelberg, Germany, 2012; pp. 75–95. ISBN 9783642292057. [Google Scholar] [CrossRef]
  162. Sejian, V.; Maurya, V.P.; Naqvi, S.M.K. Effect of Walking Stress on Growth, Physiological Adaptability and Endocrine Responses in Malpura Ewes in a Semi-Arid Tropical Environment. Int. J. Biometeorol. 2012, 56, 243–252. [Google Scholar] [CrossRef] [PubMed]
  163. Muzzo, B.I.; Maleko, D.D.; Thacker, E.; Provenza, F.D. Review: Rangeland Management in Tanzania: Opportunities, Challenges, and Prospects for Sustainability. Int. J. Trop. Drylands 2023, 7, 2. [Google Scholar] [CrossRef]
  164. He, T.; Yi, G.; Wang, X.; Sun, Y.; Li, J.; Wu, Z.; Guo, Y.; Sun, F.; Chen, Z. Effects of Heated Drinking Water during the Cold Season on Serum Biochemistry, Ruminal Fermentation, Bacterial Community, and Metabolome of Beef Cattle. Metabolites 2023, 13, 995. [Google Scholar] [CrossRef]
  165. Golher, D.M.; Patel, B.H.M.; Bhoite, S.H.; Syed, M.I.; Panchbhai, G.J.; Thirumurugan, P. Factors Influencing Water Intake in Dairy Cows: A Review. Int. J. Biometeorol. 2021, 65, 617–625. [Google Scholar] [CrossRef]
  166. Bewley, J.M.; Grott, M.W.; Einstein, M.E.; Schutz, M.M. Impact of Intake Water Temperatures on Reticular Temperatures of Lactating Dairy Cows. J. Dairy Sci. 2008, 91, 3880–3887. [Google Scholar] [CrossRef]
  167. Singh, A.K.; Bhakat, C.; Singh, P. A Review on Water Intake in Dairy Cattle: Associated Factors, Management Practices, and Corresponding Effects. Trop. Anim. Health Prod. 2022, 54, 1–13. [Google Scholar] [CrossRef]
  168. Grossi, S.; Rossi, L.; Dell’anno, M.; Biffani, S.; Sgoifo Rossi, C.A. Effects of Heated Drinking Water on the Growth Performance and Rumen Functionality of Fattening Charolaise Beef Cattle in Winter. Animals 2021, 11, 2218. [Google Scholar] [CrossRef] [PubMed]
  169. Valeria González Pereyra, A.; Maldonado May, V.; Catracchia, C.G.; Alejandra Herrero, M.; Flores, M.C.; Mazzini, M. Influence of water temperature and heat stress on drinking water intake in dairy cows. J. Agric. Res 2010, 70, 328–336. [Google Scholar] [CrossRef]
  170. Coffey, R.; Paul, M.J.; Stamp, J.; Hamilton, A.; Johnson, T. A Review of Water Quality Responses to Air Temperature and Precipitation Changes 2: Nutrients, Algal Blooms, Sediment, Pathogens. JAWRA J. Am. Water Resour. Assoc. 2019, 55, 844–868. [Google Scholar] [CrossRef] [PubMed]
  171. Giri, A.; Bharti, V.K.; Kalia, S.; Arora, A.; Balaje, S.S.; Chaurasia, O.P. A Review on Water Quality and Dairy Cattle Health: A Special Emphasis on High-Altitude Region. Appl. Water Sci. 2020, 10, 1–16. [Google Scholar] [CrossRef]
  172. Wagner, J.J.; Engle, T.E. Invited Review: Water Consumption, and Drinking Behavior of Beef Cattle, and Effects of Water Quality. Appl. Anim. Sci. 2021, 37, 418–435. [Google Scholar] [CrossRef]
  173. Arya, A.; Sharma, P.; Trivedi, M.M.; Modi, R.J.; Patel, Y.G. Article No.JSRR.116804 Review Article Arya et Al. J. Sci. Res. Rep. 2024, 30, 427–436. [Google Scholar] [CrossRef]
  174. Rexroad, C.; Vallet, J.; Matukumalli, L.K.; Reecy, J.; Bickhart, D.; Blackburn, H.; Boggess, M.; Cheng, H.; Clutter, A.; Cockett, N.; et al. Genome to Phenome: Improving Animal Health, Production, and Well-Being—A New USDA Blueprint for Animal Genome Research 2018–2027. Front. Genet. 2019, 10, 431609. [Google Scholar] [CrossRef]
  175. Nayak, S.S.; Panigrahi, M.; Rajawat, D.; Ghildiyal, K.; Sharma, A.; Jain, K.; Bhushan, B.; Dutt, T. Deciphering Climate Resilience in Indian Cattle Breeds by Selection Signature Analyses. Trop. Anim. Health Prod. 2024, 56, 1–14. [Google Scholar] [CrossRef] [PubMed]
  176. Tian, R.; Asadollahpour Nanaie, H.; Wang, X.; Dalai, B.; Zhao, M.; Wang, F.; Li, H.; Yang, D.; Zhang, H.; Li, Y.; et al. Genomic Adaptation to Extreme Climate Conditions in Beef Cattle as a Consequence of Cross-Breeding Program. BMC Genom. 2023, 24, 186. [Google Scholar] [CrossRef] [PubMed]
  177. Gutierrez-Reinoso, M.A.; Aponte, P.M.; Garcia-Herreros, M. Genomic Analysis, Progress and Future Perspectives in Dairy Cattle Selection: A Review. Animals 2021, 11, 599. [Google Scholar] [CrossRef]
  178. de Brito, A. Effects of Dietary Forage and Protein Supplements on Production, Nitrogen Utilization and Microbial Protein Synthesis in Lactating Dairy Cows; The University of Wisconsin-Madison: Madison, WI, USA, 2004. [Google Scholar]
  179. Niyonzima, Y.B.; Strandberg, E.; Hirwa, C.D.; Manzi, M.; Ntawubizi, M.; Rydhmer, L. The Effect of High Temperature and Humidity on Milk Yield in Ankole and Crossbred Cows. Trop. Anim. Health Prod. 2022, 54, 1–11. [Google Scholar] [CrossRef] [PubMed]
  180. Castaño-Sánchez, J.P.; Rotz, C.A.; McIntosh, M.M.; Tolle, C.; Gifford, C.A.; Duff, G.C.; Spiegal, S.A. Grass Finishing of Criollo Cattle Can Provide an Environmentally Preferred and Cost Effective Meat Supply Chain from United States Drylands. Agric. Syst. 2023, 210, 103694. [Google Scholar] [CrossRef]
  181. Holliday, R. Epigenetics: A Historical Overview. Epigenetics 2006, 1, 76–80. [Google Scholar] [CrossRef] [PubMed]
  182. Sevane, N.; Martínez, R.; Bruford, M.W. Genome-Wide Differential DNA Methylation in Tropically Adapted Creole Cattle and Their Iberian Ancestors. Anim. Genet. 2019, 50, 15–26. [Google Scholar] [CrossRef]
  183. Verma, S.; Taube, F.; Malisch, C.S. Examining the Variables Leading to Apparent Incongruity between Antimethanogenic Potential of Tannins and Their Observed Effects in Ruminants—A Review. Sustainability 2021, 13, 2743. [Google Scholar] [CrossRef]
  184. Kowalski, K.; Senf, C.; Okujeni, A.; Hostert, P. Large-Scale Remote Sensing Analysis Reveals an Increasing Coupling of Grassland Vitality to Atmospheric Water Demand. Glob. Chang. Biol. 2024, 30, e17315. [Google Scholar] [CrossRef] [PubMed]
  185. Priya, M.V.; Kalpana, R.; Pazhanivelan, S.; Kumaraperumal, R.; Ragunath, K.P.; Vanitha, G.; Nihar, A.; Prajesh, P.J.; Vasumathi, V. Monitoring Vegetation Dynamics Using Multi-Temporal Normalized Difference Vegetation Index (NDVI) and Enhanced Vegetation Index (EVI) Images of Tamil Nadu. J. Appl. Nat. Sci. 2023, 15, 1170–1177. [Google Scholar] [CrossRef]
  186. Mallick, K.; Verfaillie, J.; Wang, T.; Ortiz, A.A.; Szutu, D.; Yi, K.; Kang, Y.; Shortt, R.; Hu, T.; Sulis, M.; et al. Net Fluxes of Broadband Shortwave and Photosynthetically Active Radiation Complement NDVI and near Infrared Reflectance of Vegetation to Explain Gross Photosynthesis Variability across Ecosystems and Climate. Remote Sens. Environ. 2024, 307, 114123. [Google Scholar] [CrossRef]
  187. Bhaga, T.D.; Dube, T.; Shekede, M.D.; Shoko, C. Impacts of Climate Variability and Drought on Surface Water Resources in Sub-Saharan Africa Using Remote Sensing: A Review. Remote Sens. 2020, 12, 4184. [Google Scholar] [CrossRef]
  188. Bellini, E. Assessing Grassland Development by Means of Modelling and Remote Sensing Approaches: Current Situation and Future Projections; University of Florence: Florence, Italy, 2023. [Google Scholar]
  189. König, S.; Thonfeld, F.; Förster, M.; Dubovyk, O.; Heurich, M. Assessing Combinations of Landsat, Sentinel-2 and Sentinel-1 Time Series for Detecting Bark Beetle Infestations. GISci. Remote Sens. 2023, 60, 2226515. [Google Scholar] [CrossRef]
  190. Poitras, T.B.; Villarreal, M.L.; Waller, E.K.; Nauman, T.W.; Miller, M.E.; Duniway, M.C. Identifying Optimal Remotely-Sensed Variables for Ecosystem Monitoring in Colorado Plateau Drylands. J. Arid. Environ. 2018, 153, 76–87. [Google Scholar] [CrossRef]
  191. Regos, A.; Gonçalves, J.; Arenas-Castro, S.; Alcaraz-Segura, D.; Guisan, A.; Honrado, J.P. Mainstreaming Remotely Sensed Ecosystem Functioning in Ecological Niche Models. Remote Sens. Ecol. Conserv. 2022, 8, 431–447. [Google Scholar] [CrossRef]
  192. Bibi, Z.; Maqsood, M.J.; Idrees, A.; Rafique, H.; Butt, A.A.; Ali, R.; Arif, Z.; Nabi, M.U. Retracted: Exploring the Role of Phenotypic Plasticity in Plant Adaptation to Changing Climate: A Review. Asian J. Res. Crop Sci. 2024, 9, 1–9. [Google Scholar] [CrossRef]
  193. Temperton, V.M.; Mwangi, P.N.; Scherer-Lorenzen, M.; Schmid, B.; Buchmann, N. Positive Interactions between Nitrogen-Fixing Legumes and Four Different Neighbouring Species in a Biodiversity Experiment. Oecologia 2007, 151, 190–205. [Google Scholar] [CrossRef]
  194. Pirhofer-Walzl, K.; Eriksen, J.; Rasmussen, J.; Høgh-Jensen, H.; Søegaard, K.; Rasmussen, J. Effect of Four Plant Species on Soil 15N-Access and Herbage Yield in Temporary Agricultural Grasslands. Plant Soil 2013, 371, 313–325. [Google Scholar] [CrossRef]
  195. Moorby, J.M.; Fraser, M.D. Review: New Feeds and New Feeding Systems in Intensive and Semi-Intensive Forage-Fed Ruminant Livestock Systems. Animal 2021, 15, 100297. [Google Scholar] [CrossRef]
  196. Lee, M.A. A Global Comparison of the Nutritive Values of Forage Plants Grown in Contrasting Environments. J. Plant Res. 2018, 131, 641–654. [Google Scholar] [CrossRef]
  197. Provenza, F.D.; Villalba, J.J.; Dziba, L.E.; Atwood, S.B.; Banner, R.E. Linking Herbivore Experience, Varied Diets, and Plant Biochemical Diversity. Small Rumin. Res. 2003, 49, 257–274. [Google Scholar] [CrossRef]
  198. Lagrange, S.; Villalba, J.J. Tannin-Containing Legumes and Forage Diversity Influence Foraging Behavior, Diet Digestibility, and Nitrogen Excretion by Lambs 1,2 3995 Tannin-Containing Legumes and Forage Intake. J. Anim. Sci. 2019, 97, 3994–4009. [Google Scholar] [CrossRef] [PubMed]
  199. Lagrange, S.; Beauchemin, K.A.; MacAdam, J.; Villalba, J.J. Grazing Diverse Combinations of Tanniferous and Non-Tanniferous Legumes: Implications for Beef Cattle Performance and Environmental Impact. Sci. Total Environ. 2020, 746, 140788. [Google Scholar] [CrossRef] [PubMed]
  200. Villalba, J.J.; Provenza, F.D.; K Clemensen, A.; Larsen, R.; Juhnke, J. Preference for Diverse Pastures by Sheep in Response to Intraruminal Administrations of Tannins, Saponins and Alkaloids. Grass Forage Sci. 2011, 66, 224–236. [Google Scholar] [CrossRef]
  201. Villalba, J.J.; Provenza, F.D.; Gibson, N.; López-Ortíz, S. Veterinary Medicine: The Value of Plant Secondary Compounds and Diversity in Balancing Consumer and Ecological Health. In Sustainable Food Production Includes Human and Environmental Health; Springer: Berlin/Heidelberg, Germany, 2014; pp. 165–190. [Google Scholar] [CrossRef]
  202. Garrett, K.; Beck, M.R.; Marshall, C.J.; Maxwell, T.M.R.; Logan, C.M.; Greer, A.W.; Gregorini, P. Varied Diets: Implications for Lamb Performance, Rumen Characteristics, Total Antioxidant Status, and Welfare. J. Anim. Sci. 2021, 99, skab334. [Google Scholar] [CrossRef]
  203. Jensen, T.L.; Provenza, F.D.; Villalba, J.J. Influence of Diet Sequence on Intake of Foods Containing Ergotamine d Tartrate, Tannins and Saponins by Sheep. Appl. Anim. Behav. Sci. 2013, 144, 57–62. [Google Scholar] [CrossRef]
  204. Kamler, J.F.; Ballard, W.B.; Gilliland, R.L.; Mote, K. Coyote (Canis Latrans) Movements Relative to Cattle (Bos Taurus) Carcass Areas. West. N. Am. Nat. 2004, 64, 53–58. [Google Scholar]
  205. Villalba, J.J.; Provenza, F.D.; Manteca, X. Links between Ruminants’ Food Preference and Their Welfare. Animal 2010, 4, 1240–1247. [Google Scholar] [CrossRef]
  206. Tilman, D. Plant Strategies and the Dynamics and Structure of Plant Communities; Princeton University Press: Princeton, NJ, USA, 1988; Volume 26. [Google Scholar]
  207. Gregorini, P.; Villalba, J.J.; Chilibroste, P.; Provenza, F.D. Grazing Management: Setting the Table, Designing the Menu and Influencing the Diner. Anim. Prod. Sci. 2017, 57, 1248–1268. [Google Scholar] [CrossRef]
  208. Distel, R.A.; Arroquy, J.I.; Lagrange, S.; Villalba, J.J. Designing Diverse Agricultural Pastures for Improving Ruminant Production Systems. Front. Sustain. Food Syst. 2020, 4, 596869. [Google Scholar] [CrossRef]
  209. Jordán, M.J.; Martínez-Conesa, C.; Bañón, S.; Otal, J.; Quílez, M.; García-Aledo, I.; Romero-Espinar, P.; Sánchez-Gómez, P. The Combined Effect of Mediterranean Shrubland Pasture and the Dietary Administration of Sage By-Products on the Antioxidant Status of Segureña Ewes and Lambs. Antioxidants 2020, 9, 938. [Google Scholar] [CrossRef]
  210. Athanasiadou, S.; Tzamaloukas, O.; Kyriazakis, I.; Jackson, F.; Coop, R.L. Testing for Direct Anthelmintic Effects of Bioactive Forages against Trichostrongylus Colubriformis in Grazing Sheep. Vet. Parasitol. 2005, 127, 233–243. [Google Scholar] [CrossRef] [PubMed]
  211. Saed-Moucheshi, A.; Shekoofa, A.; Pessarakli, M. Reactive Oxygen Species (ROS) Generation and Detoxifying in Plants. J. Plant Nutr. 2014, 37, 1573–1585. [Google Scholar] [CrossRef]
  212. Monjardino, M.; Revell, D.; Pannell, D.J. The Potential Contribution of Forage Shrubs to Economic Returns and Environmental Management in Australian Dryland Agricultural Systems. Agric. Syst. 2010, 103, 187–197. [Google Scholar] [CrossRef]
  213. Revell, C.K.; Ewing, M.A.; Nutt, B.J. Breeding and Farming System Opportunities for Pasture Legumes Facing Increasing Climate Variability in the South-West of Western Australia. Crop Pasture Sci. 2012, 63, 840–847. [Google Scholar] [CrossRef]
  214. Revell, D.K.; Norman, H.C.; Vercoe, P.E.; Phillips, N.; Toovey, A.; Bickell, S.; Hulm, E.; Hughes, S.; Emms, J. Australian Perennial Shrub Species Add Value to the Feed Base of Grazing Livestock in Low- to Medium-Rainfall Zones. Anim. Prod. Sci. 2013, 53, 1221–1230. [Google Scholar] [CrossRef]
  215. Robertson, M.; Revell, C. Perennial Pastures in Cropping Systems of Southern Australia: An Overview of Present and Future Research. Crop Pasture Sci. 2014, 65, 1084–1090. [Google Scholar] [CrossRef]
  216. Landeen, M.; Jones, C.; Jensen, S.; Whittaker, A.; Summers, D.D.; Eggett, D.; Petersen, S.L. Establishing Seed Islands for Native Forb Species on Rangelands Using N-Sulate Ground Cover Fabric. Nativ. Plants J. 2021, 22, 51–63. [Google Scholar] [CrossRef]
  217. Santos-Gally, R.; Boege, K. Biodiversity Islands: The Role of Native Tree Islands Within Silvopastoral Systems in a Neotropical Region. In Biodiversity Islands: Strategies for Conservation in Human-Dominated Environments; Springer International Publishing: Cham, Switzerland, 2022; pp. 117–138. [Google Scholar] [CrossRef]
  218. Schrader, J.; Wright, I.J.; Kreft, H.; Westoby, M. A Roadmap to Plant Functional Island Biogeography. Biol. Rev. 2021, 96, 2851–2870. [Google Scholar] [CrossRef]
  219. Kim, N.H.; Cho, T.J.; Rhee, M.S. Current Interventions for Controlling Pathogenic Escherichia coli. Adv. Appl. Microbiol. 2017, 100, 1–47. [Google Scholar] [CrossRef]
  220. Charles, A.S.; Baskaran, A.; Murcott, C.; Schreiber, D.; Hoagland, T.; Venkitanarayanan, K. Reduction of Escherichia coli O157:H7 in Cattle Drinking-Water by Trans-Cinnamaldehyde. Foodborne Pathog. Dis. 2008, 5, 763–771. [Google Scholar] [CrossRef]
  221. Zhao, T.; Zhao, P.; West, J.W.; Bernard, J.K.; Cross, H.G.; Doyle, M.P. Inactivation of Enterohemorrhagic Escherichia Coli in Rumen Content- or Feces-Contaminated Drinking Water for Cattle. Appl. Environ. Microbiol. 2006, 72, 3268–3273. [Google Scholar] [CrossRef] [PubMed]
  222. Richards, S.; Rao, L.; Connelly, S.; Raj, A.; Raveendran, L.; Shirin, S.; Jamwal, P.; Helliwell, R. Sustainable Water Resources through Harvesting Rainwater and the Effectiveness of a Low-Cost Water Treatment. J. Environ. Manag. 2021, 286, 112223. [Google Scholar] [CrossRef]
  223. Sobsey, M.D.; Khatib, L.A.; Hill, V.R.; Alocilja, E.; Pillai, S. Pathogens in Animal Wastes and the Impacts of Waste Management Practices on Their Survival, Transport and Fate; ASABE: St. Joseph, MI, USA, 2006; pp. 609–666. [Google Scholar] [CrossRef]
  224. Hunt, L.P.; Mcivor, J.G.; Grice, A.C.; Bray, S.G. Principles and Guidelines for Managing Cattle Grazing in the Grazing Lands of Northern Australia: Stocking Rates, Pasture Resting, Prescribed Fire, Paddock Size and Water Points—A Review. Rangel. J. 2014, 36, 105–119. [Google Scholar] [CrossRef]
  225. Muzzo, B.I.; Provenza, F.D. Review of strategies for overcoming challenges of beef production in Tanzania. Livestock Res. Rural Development 2018, 30, 199. [Google Scholar]
  226. Masters, D.G.; Blache, D.; Lockwood, A.L.; Maloney, S.K.; Norman, H.C.; Refshauge, G.; Hancock, S.N. Shelter and Shade for Grazing Sheep: Implications for Animal Welfare and Production and for Landscape Health. Anim. Prod. Sci. 2023, 63, 623–644. [Google Scholar] [CrossRef]
  227. Valtorta, S.E.; Leva, P.E.; Gallardo, M.R. Evaluation of Different Shades to Improve Dairy Cattle Well-Being in Argentina. Int. J. Biometeorol. 1997, 41, 65–67. [Google Scholar] [CrossRef] [PubMed]
  228. Titto, C.G.; Titto, E.A.L.; Titto, R.M.; Mourão, G.B. Heat Tolerance and the Effects of Shade on the Behavior of Simmental Bulls on Pasture. Anim. Sci. J. 2011, 82, 591–600. [Google Scholar] [CrossRef] [PubMed]
  229. Van Laer, E.; Moons, C.P.H.; Ampe, B.; Sonck, B.; Vandaele, L.; De Campeneere, S.; Tuyttens, F.A.M. Effect of Summer Conditions and Shade on Behavioural Indicators of Thermal Discomfort in Holstein Dairy and Belgian Blue Beef Cattle on Pasture. Animal 2015, 9, 1536–1546. [Google Scholar] [CrossRef]
  230. Davison, T.M.; Silver, B.A.; Lisle, A.T.; Orr, W.N. The Influence of Shade on Milk Production of Holstein-Friesian Cows in a Tropical Upland Environment. Aust. J. Exp. Agric. 1988, 28, 149–154. [Google Scholar] [CrossRef]
  231. McIlvain, E.H.; Shoop, M.C. Shade for Improving Cattle Gains and Rangeland Use (El Uso de Sombreadores Para Mejorar Las Ganancias de Novillas y Pastoreo de Animales). J. Range Manag. 1971, 24, 181. [Google Scholar] [CrossRef]
  232. Moreno García, C.A.; Maxwell, T.M.R.; Hickford, J.; Gregorini, P. On the Search for Grazing Personalities: From Individual to Collective Behaviors. Front. Vet. Sci. 2020, 7, 502292. [Google Scholar] [CrossRef]
  233. Gregorini, P.; Tamminga, S.; Gunter, S.A. Behavior and Daily Grazing Patterns of Cattle. Prof. Anim. Sci. 2006, 22, 201–209. [Google Scholar] [CrossRef]
  234. Lyons, R.K.; Machen, R.V. Livestock Grazing Distribution: Considerations and Management; Texas AgriLife Extension: College Station, TX, USA, 2023. [Google Scholar]
  235. Heins, B.J.; Pereira, G.M.; Sharpe, K.T. Precision Technologies to Improve Dairy Grazing Systems. JDS Commun. 2023, 4, 318–323. [Google Scholar] [CrossRef]
  236. Bello, R.-W.; Talib, A.Z.H.; Mohamed, A.S.A. A Framework for Real-Time Cattle Monitoring Using Multimedia Networks. Int. J. Recent Technol. Eng. (IJRTE) 2020, 8, 974–979. [Google Scholar] [CrossRef]
  237. Ramezani Gardaloud, N.; Guse, C.; Lidauer, L.; Steininger, A.; Kickinger, F.; Öhlschuster, M.; Auer, W.; Iwersen, M.; Drillich, M.; Klein-Jöbstl, D. Short Communications: An Ear-Attached Accelerometer Detects Effects of Regrouping on Lying, Rumination, and Activity Times in Calves. Vet. Res. Commun. 2023, 47, 2333–2337. [Google Scholar] [CrossRef] [PubMed]
  238. Versluijs, E.; Niccolai, L.J.; Spedener, M.; Zimmermann, B.; Hessle, A.; Tofastrud, M.; Devineau, O.; Evans, A.L. Classification of Behaviors of Free-Ranging Cattle Using Accelerometry Signatures Collected by Virtual Fence Collars. Front. Anim. Sci. 2023, 4, 1083272. [Google Scholar] [CrossRef]
  239. Džermeikaitė, K.; Bačėninaitė, D.; Antanaitis, R. Innovations in Cattle Farming: Application of Innovative Technologies and Sensors in the Diagnosis of Diseases. Animals 2023, 13, 780. [Google Scholar] [CrossRef]
  240. Sprinkle, J.E.; Sagers, J.K.; Hall, J.B.; Ellison, M.J.; Yelich, J.V.; Brennan, J.R.; Taylor, J.B.; Lamb, J.B. Predicting Cattle Grazing Behavior on Rangeland Using Accelerometers. Rangel. Ecol. Manag. 2021, 76, 157–170. [Google Scholar] [CrossRef]
  241. Riaboff, L.; Shalloo, L.; Smeaton, A.F.; Couvreur, S.; Madouasse, A.; Keane, M.T. Predicting Livestock Behaviour Using Accelerometers: A Systematic Review of Processing Techniques for Ruminant Behaviour Prediction from Raw Accelerometer Data. Comput. Electron. Agric. 2022, 192, 106610. [Google Scholar] [CrossRef]
Figure 1. A framework of mitigation strategies to improve cattle resilience under climate change stress. Seven key management areas are highlighted: (1) Breeding and Epigenetics: Developing heat-tolerant and disease-resistant cattle with climate-adaptive traits passed down across generations. (2) Remote Sensing and GIS Mapping: Tracking optimal grazing zones and cattle movements as vegetation changes, enabling adaptive land use. (3) Forage Diversity: Enhancing forage species variety to ensure a stable food supply under changing environmental conditions. (4) Chemical Diversity in Forage: Incorporating tannin-rich plants and other chemically diverse forages to improve nutrient utilization and reduce methane emissions, thus contributing to environmental sustainability. (5) Water Treatment and Distribution: Establishing reliable water access systems, particularly important in drought-prone regions, to maintain cattle hydration and health. (6) Vegetation Cover Management: Promoting shade and ground cover to create cooler grazing areas, reduce soil erosion, and retain moisture. (7) GPS Collars and Accelerometers: Employing wearable technologies to monitor cattle movement, behavior, and physiological responses, allowing for precise, data-driven adaptive management.
Figure 1. A framework of mitigation strategies to improve cattle resilience under climate change stress. Seven key management areas are highlighted: (1) Breeding and Epigenetics: Developing heat-tolerant and disease-resistant cattle with climate-adaptive traits passed down across generations. (2) Remote Sensing and GIS Mapping: Tracking optimal grazing zones and cattle movements as vegetation changes, enabling adaptive land use. (3) Forage Diversity: Enhancing forage species variety to ensure a stable food supply under changing environmental conditions. (4) Chemical Diversity in Forage: Incorporating tannin-rich plants and other chemically diverse forages to improve nutrient utilization and reduce methane emissions, thus contributing to environmental sustainability. (5) Water Treatment and Distribution: Establishing reliable water access systems, particularly important in drought-prone regions, to maintain cattle hydration and health. (6) Vegetation Cover Management: Promoting shade and ground cover to create cooler grazing areas, reduce soil erosion, and retain moisture. (7) GPS Collars and Accelerometers: Employing wearable technologies to monitor cattle movement, behavior, and physiological responses, allowing for precise, data-driven adaptive management.
Climate 13 00001 g001
Figure 2. Map identifying areas within a 55-acre grass monoculture pasture—partitioned into 6 paddocks of approximately 9 acres in size, showing the frequency of pixels with above-average greenness across a 7-year period. Colors refer to the number of years a given pixel is 1 standard deviation above the NDVI mean for that year and is considered a surrogate for moisture availability. This frequency layer was derived from NDVI values from Sentinel-2 imagery processed using the IDLT tool. Light blue boxes identify locations with above-average moisture availability that would provide the best chance to successfully establish “feed patches”. Gray areas represent below-average moisture zones that would be unsuitable for forage patch establishment.
Figure 2. Map identifying areas within a 55-acre grass monoculture pasture—partitioned into 6 paddocks of approximately 9 acres in size, showing the frequency of pixels with above-average greenness across a 7-year period. Colors refer to the number of years a given pixel is 1 standard deviation above the NDVI mean for that year and is considered a surrogate for moisture availability. This frequency layer was derived from NDVI values from Sentinel-2 imagery processed using the IDLT tool. Light blue boxes identify locations with above-average moisture availability that would provide the best chance to successfully establish “feed patches”. Gray areas represent below-average moisture zones that would be unsuitable for forage patch establishment.
Climate 13 00001 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Muzzo, B.I.; Ramsey, R.D.; Villalba, J.J. Changes in Climate and Their Implications for Cattle Nutrition and Management. Climate 2025, 13, 1. https://doi.org/10.3390/cli13010001

AMA Style

Muzzo BI, Ramsey RD, Villalba JJ. Changes in Climate and Their Implications for Cattle Nutrition and Management. Climate. 2025; 13(1):1. https://doi.org/10.3390/cli13010001

Chicago/Turabian Style

Muzzo, Bashiri Iddy, R. Douglas Ramsey, and Juan J. Villalba. 2025. "Changes in Climate and Their Implications for Cattle Nutrition and Management" Climate 13, no. 1: 1. https://doi.org/10.3390/cli13010001

APA Style

Muzzo, B. I., Ramsey, R. D., & Villalba, J. J. (2025). Changes in Climate and Their Implications for Cattle Nutrition and Management. Climate, 13(1), 1. https://doi.org/10.3390/cli13010001

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop