Next Article in Journal
Improving Energy Efficiency of Barley Production Using Joint Data Envelopment Analysis (DEA) and Life Cycle Assessment (LCA): Evaluation of Greenhouse Gas Emissions and Optimization Approach
Next Article in Special Issue
Mitigating Greenhouse Gas Emissions from Beef Cattle Production in Brazil through Animal Management
Previous Article in Journal
Research on Supplier Selection of Prefabricated Building Elements from the Perspective of Sustainable Development
Previous Article in Special Issue
Effect of Dietary Inclusion of Alpha-Monolaurin on the Growth Performance, Lipid Peroxidation, and Immunity Response in Broilers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Towards Sustainable Livestock Production: Estimation of Methane Emissions and Dietary Interventions for Mitigation

by
Pragna Prathap
1,
Surinder Singh Chauhan
1,
Brian Joseph Leury
1,
Jeremy James Cottrell
1 and
Frank Rowland Dunshea
1,2,*
1
Faculty of Veterinary and Agricultural Sciences, The University of Melbourne, Parkville, VIC 3010, Australia
2
Faculty of Biological Sciences, The University of Leeds, Leeds LS2 9JT, UK
*
Author to whom correspondence should be addressed.
Sustainability 2021, 13(11), 6081; https://doi.org/10.3390/su13116081
Submission received: 8 March 2021 / Revised: 30 April 2021 / Accepted: 24 May 2021 / Published: 28 May 2021
(This article belongs to the Special Issue Sustainable Livestock Production)

Abstract

:
The increasing need for sustainable livestock production demands more research in the field of greenhouse gas (GHG), particularly methane (CH4), measurement and mitigation. Dietary interventions, management, and biotechnological strategies to reduce the environmental impacts and economic implications of enteric CH4 emissions are needed. While the use of biotechnological interventions and management strategies can be challenging on a routine basis, feed additive supplementation appears to be the most researched, developed, and ready to use strategy to mitigate enteric CH4 emissions. This paper discusses various recently developed feeding strategies to reduce enteric CH4 emissions in livestock. Additionally, the manuscript reviews various technologies developed for CH4 estimation since the accurate and reliable estimation of CH4 emissions can be a limiting step in the development and adoption of any mitigation strategy.

Graphical Abstract

1. Introduction

The contribution of the agriculture sector to the climate crisis is typically underestimated due to numerous overlooked emission sources. According to the environmental protection agency, the agricultural sector alone accounts for 10–12% of total global anthropogenic greenhouse gas (GHG) emissions [1]. Of the net global emissions, this accounts for 13% of carbon dioxide (CO2), 44% of methane (CH4), and 82% of nitrous oxide emissions through anthropogenic activities [2]. Among the total agriculture GHG emissions, the ruminant supply chains alone release around 5.7 gigatonnes CO2-equivalent GHGs in a year, contributing 80% of the total emission from the entire livestock sector [3,4]. The contribution of livestock towards the existing GHG pool is mainly in the form of enteric CH4 (around 63%), followed by 25% from the use of manure as a fertiliser to the plants and pastures and 12% emissions from dung and urine management [5]. Therefore, the scientific communities are concentrating on research activities to reduce enteric CH4 emission from livestock and substantial progress has been achieved during the last decade [6,7,8]. For example, Australia’s agricultural GHG emissions have declined by 15.77% since 2005, now reaching 72.04 MtCO2e GHG emission from the total agricultural sector [9]. However, the drastic increase in animal population could offset the efforts to reduce CH4 emission. Pertinently, there is an increased focus on sustainable livestock production along with suitable CH4 mitigation measures.
Decreasing the production of CH4 from ruminant animals is desirable both as a strategy to reduce global GHG emissions and as a way of improving feed conversion efficiency. Among the various approaches available to reduce CH4 emission, feed manipulation is the most widely used strategy to target enteric CH4 reduction in livestock since they are near market-ready products [10]. To help meet this goal, more reliable and repeatable CH4 measurement strategies are required to ensure the effectiveness and reliability of these feed additives and other CH4 abatement strategies. Currently, there are a range of techniques being used around the world for the measurement of CH4 from the ruminants. Lately, researchers and manufacturers are considering this as a vital area of research to develop new technologies and to modify existing instrumentation [11]. There are several methods available for the quantification of CH4 although these vary in cost, application, repeatability, precision, and reliability [12]. Therefore, it is the premise of this manuscript to review the recent advances in technologies for the measurement of enteric CH4 emission from ruminants and discuss the potential nutritional interventions for CH4 mitigation. Recently, a few authors have reviewed the effects of different nutritional supplements for their CH4 mitigation potential [4,13,14,15,16] and the readers are redirected to those reviews for regional or species-specific CH4 mitigation strategies or mitigation using a particular group of feed additives.

2. Global Climate Change and the Role of Methane

The Earth’s climate has varied tremendously in the past few decades due to enhanced anthropogenic activity [17]. The effects of climate change are unequivocal, as is now evident from the rise in global temperature and sea level, the shift in rainfall patterns, glacial retreat, increased frequency of extreme events, and prolonged periods of dry spells and frost [17]. The global mean surface temperature has increased by around 0.9 °C from the late 19th century, a change driven mainly by augmented CH4, CO2, and other anthropogenic emissions to the global gas pool [18]. Likewise, since the last century, sea levels have risen by approximately 20 cm, but the rate of rise in sea levels has doubled alarmingly in the past two decades, as compared to the figures of the last century [19]. Similarly, the widespread availability and pattern of rainfall has changed across the globe. The regional climate patterns are also in a changing phase, with an increase in the frequency and duration of extreme weather events [18]. Over the period from 1971 to 2008, the Earth has witnessed greater numbers of heatwaves and hot days, with the hottest days during these heatwaves becoming even hotter. The increased frequency of heatwaves may have devastating effects on livestock production and other agricultural sectors [20]. Further, the chemical composition of the global atmosphere has changed greatly since the 1700s because of anthropogenic activities, including livestock production and farming [21]. Among the various GHGs emitted from the agriculture sector, CH4 is the second most abundant and is a very potent GHG since it has 25 times more global warming potential than CO2, making it a current target for action [22,23,24]. Further, due to the short atmospheric half-life of CH4, efforts to mitigate CH4 would achieve substantial and swift effects on the global warming potential [24]. The CH4 emissions from the agriculture sector have doubled since the pre-industrial time. Change in atmospheric CH4 can substantially increase water vapor concentration and it can affect stratospheric and tropospheric chemistry. Ultimately, an increased concentration of CH4 in the atmosphere can increase the magnitude of the GHG effect and the Earth’s temperature [21,25].

3. Mechanism of Methane Production in the Rumen

Methane is produced by ruminants as a product of the fermentation of ingested feed [26]. Ruminant animals emit CH4 mainly through two pathways, via midgut fermentation and hindgut fermentation. Midgut fermentation or enteric fermentation solely accounts for 89% of total CH4 emission from the animal. Apart from the deleterious effects on global warming, CH4 is also a dietary energy loss and ruminants can lose between 2 and 12% of ingested energy in the form of CH4 [27]. The ruminant forestomach or rumen is host to a large group of diverse microorganisms [28]. These microbes ferment feed materials consumed by the animal through the process of enteric fermentation [29]. The products derived from the enteric fermentation of plant materials provide the nutrients required for the animal’s survival. Microorganisms present in the rumen, such as bacteria, fungi, archaea, and protozoa, hydrolyse the dietary polysaccharides present in the feed materials into simple sugars through their enzymatic activity and finally yield volatile fatty acids (VFA), primarily acetate, propionate, and butyrate [30]. In unison, varying amounts of hydrogen (H2), formic acid, and CO2 are produced as end-products of fermentation [26]. Most of the methanogenic archaea and some bacteria in the rumen use H2 ions to reduce CO2 to produce CH4 since this process is thermodynamically favourable to microbes. This process keeps the partial pressure of H2 low, which directs fermentation towards the production of less reduced end-products including acetate [31]. The abundance of H2 ions is determined by the proportion of end-products from the ruminal fermentation of ingested feed. Formate, which is abundant in most of the ruminant archaea, is also considered to be a part of this hydrogenotrophic pathway. Methanobrevibacter ruminantium and Methanobrevibacter gottschalkii are the major hydrogenotrophic archaea that alone encompass 74% of the methanogenic archaeal community in the ruminant stomach [32]. Likewise, methyl groups present in the methanol and methylamines may serve as another category of substrates that favour methanogenesis [33]. The formation of acetate results in an increase in hydrogen ions, while the process that yields propionate consumes hydrogen ions [34]. Therefore, the greater the production of acetate, the more CH4 can be expected, whilst an increase in propionate production is associated with lower production of CH4 [13].

4. Estimates of Enteric Methane Production: How Much Does Each Species Contribute?

Livestock production is an integral component of global agriculture and also a significant contributor to anthropogenic GHG emissions. The enteric CH4 emission is responsible for 44% of the GHG emission from the total livestock group and 55% from the ruminants [35]. Altogether, ruminants contribute 2,098,787.77 CO2-eq of enteric CH4 to the global GHG pool, of which 54.7% is from non-dairy cattle, 18.9% is from dairy cattle, 10.5% is from buffaloes, 7% is from sheep, 4.4% is from goats, 1.3% is from camels, 1.1 is from horses, and 0.5% is from donkeys [36]. As per FAOSTAT [36], on a yearly average, dairy cattle, non-dairy cattle, buffaloes, sheep, and goat emit 1419.1, 926.5, 1155, 117.2, 105 kg CO2-eq of CH4, respectively. There is a strong link between the quantity of enteric CH4 emission and species differences in animals. Variations in the quantity of feed intake/body mass and the quality of the feed are the major reasons behind this drastic difference in enteric CH4 emission between each species [3]. Within the species, there can be breed-wise variation in enteric CH4 emission. Differences in the genetic potential and feed digestion efficiency among breeds better explain this variation [3,37]. Figure 1 shows the percentage of enteric CH4 from total CH4 emission from each species. Figure 2 represents species-wise contributions to the global enteric methane pool.

5. Methods Used to Quantify Methane Production

Knowledge about different CH4 measuring techniques is increasingly in demand due to the crucial role of CH4 in global warming and several commitments to reduce global CH4 emissions. There are several existing methods for estimating CH4 production that include in vitro estimation and on-farm measuring techniques [40,41].

5.1. In Vitro Estimation

There are several methods for quantifying gas production through in vitro fermentation, with variations in complexity and sophistication. This is a relatively cheap method that is suitable for analyzing CH4 emissions from a vast variety of feed additives and plant extracts without the error of animal to animal variation [42]. It is particularly useful for ranking different dietary interventions. The basic principle of every in vitro fermentation technique relies on the incubation of feed samples along with the rumen microbial inoculum and buffer solution in an anaerobic environment [43]. The anaerobic fermentation of feed samples can yield various gases in the container and the cumulative volume can be later recorded [44]. The typical gas compositions and CH4 concentrations can be estimated using the gas samples harvested from the headspace of the container [43].
The ANKOM rumen fluid gas production system is one of the easiest to use and reliable gas production systems that is commercially available. The system is equipped with sample bottles and pressure sensor modules [45]. The gases produced in the headspace of bottles, filled with buffered rumen fluid and feed samples, can be aspirated through the vents using gas-tight syringes [46]. Rumen motility and temperature are simulated by placing the bottles in a shaking water bath maintained at 39 °C [47,48]. The in vitro gas production technique (IVGPT) is another routine method for evaluating feed samples and gas production. This system uses glass syringes, to incubate feed samples and rumen liquor, instead of bottles [49]. Rusitech is another type of artificial rumen apparatus that is currently available. The main advantage of this system is that it can simulate the rumen for several days and can maintain protozoal numbers [50].

5.2. Respiration Chamber

Respiratory chambers are one of the main in vivo CH4 measuring technologies that have been used for more than 125 years, with varying degrees of complexity [11,51]. This equipment provides the user with an opportunity to measure enteric CH4 and other gases emitted from the mouth, nostrils, and rectum [41]. Among the two models of respiration chambers available, open-circuit chambers are more widely used over closed-circuit chambers. The basic working principle of this technique is based on the first law of thermodynamics and measures the concentrations of CH4 leaving the chamber. Gas samples are collected from the inflow and exhaust ducts during this period by creating a negative pressure inside the chamber [52]. Internal ventilation fans fitted inside the chamber ensure proper mixing of the incoming air and exhaled gas. The CH4 produced by the animal is calculated by multiplying the air circulation through the system by the concentration difference between incoming and outgoing air. Animals can be fed and watered inside the chamber for 1–7 sequential days [11,53]; this gives an opportunity to measure CH4 emission in terms of dry matter intake [54]. Further, open-circuit respiratory chambers are often referred to as the ‘gold standard’ for measuring accurate CH4 emission measurements; hence, they account for the losses from rectum and rumen fistulas in addition to the losses through regurgitation [12,55]. However, CH4 can still be lost if the chamber is imperfectly sealed [56].

5.3. Ventilated Hood Systems

Ventilated hood systems or head hood systems are a simplified version of respiratory chambers that only enclose the head of the animal [57]. The mobile head hood system used by Fernández Martínez et al. [58] consists of a headbox, rotameter, flow meter, air volume totaliser, adjustable and precise membrane pump, gas cooler, and gas analyser unit. The gas analyser associated with the instrument analyses the composition of gases drawn through the head hood. The head hood has blowers that can move air from the inlet to the exhaust. Further, the head hood, made up of clear polycarbonate material, allows a full range of vision to the animal during the sampling time [59]. The sufficiently large headbox allows the animal to move its head effortlessly to access feed and water and to lay down [11,60]. The ventilated hood systems can be used consecutively over 24 h or for longer periods to measure CH4 emission. This can be useful for researchers trying to establish a link between CH4 emission, feed consumption, and energy metabolism [60]. In addition to the CH4 measurement, this equipment can measure ethanol, methanol, water vapor, nitrous oxide, acetic acid, CO2, and oxygen (O2) from the animal on a real-time basis [61]. As compared to respiration chambers, ventilated hood systems are less expensive and they require much less space [62]. The main demerit with this system is that it can only account for the CH4 produced from midgut fermentation [63]. Furthermore, a sufficient amount of time and training is required to make the animals accustomed to the head hood apparatus. Like the respiratory chamber, this equipment also cannot be used in grazing conditions [64].

5.4. Sulfur Hexafluoride Tracer Gas

Another CH4-measuring technique that can be used for grazing animals, as well as penned animals, is the sulphur hexafluoride (SF6) tracer technique [65]. In this basic system, animals are orally administered with a permeation tube that releases a known amount of SF6 into the reticulorumen junction of the animal [66]. The exhaled air from the animal is collected from a point near the nostrils and mouth by means of a tube with in-line flow restrictors connected to an evacuated canister connected to a halter and attached with a capillary tube around the neck or held in a harness on the back of the animal [67,68]. At the end of each day, gas accumulated in the canisters will be collected and subjected to gas chromatography. This method assumes that the rate of tracer gas emission is the same as the CH4 emitted from the animal [59,64]. The enteric CH4 production is determined by multiplying the CH4 to SF6 ratio by the release rate of the permeation tube, corrected for the actual length of sample collection and the background CH4 level [69]. As compared to the respiration chambers and ventilated hood systems, the SF6 tracer method is less costly and can be concurrently used in a greater number of animals. However, the release rate of the tracer can affect the CH4 release and it can account for 6 to 13% variation in the CH4 accountability [70]. Furthermore, among- and within-animal variation in CH4 emission are greater when using the SF6 technique as compared to the respiratory chamber [71].

5.5. Open-Path Lasers

An open-path laser or tunable diode laser is another possible CH4-measuring instrument that uses beams of infrared light to quantify CH4 from a grazing herd [72]. The sensor associated with the instrument captures the reflected light and analyses the intensity of the received light as an indicator of CH4 levels along the path [73]. The laser system used by Laubach and Kelliher [74] consists of a main unit that contains an infrared laser source and a reference cell, remote heads that contain a photodiode that converts the reflected light to an electric signal, and four retroreflector units. The efficiency of an open-path laser is highly dependent on the weather elements and the location of the animals. Sometimes, insufficient laser lighting and wind variation can create disruptions in the CH4 measurements [75]. The CH4 emission rates are usually calculated using a backward-Lagrangian stochastic model [76]. Usually, the laser path will be located at 0.5 m height and 1 to 1.5 m outside the perimeter of the pens. The instrument considers cattle herd as a surface source and the individual animals fitted with collars as a point source [77]. The main advantage of this technique is that it can cover a large area and large herd and it does not affect the normal grazing behaviour of the animals. The accuracy of the measurements depends heavily on the positioning of the animals. For example, in larger paddocks, animals may congregate around water tanks and feeders or away from the beam of infrared path of the instrument, resulting in inconsistent measurements [78].

5.6. GreenFeed Emission Monitoring System

GreenFeed is a short-term CH4-measuring system that offers a small amount of pelleted bait to attract animals to the measuring unit [79]. The gas collection head of the instrument measures CH4 in the exhaled air every time the animal approaches the equipment. This instrument is highly compatible with in-house conditions as well as in an extensive grazing condition [80]. The user has the freedom to change the type, frequency, and amount of the pellet. The user can adjust the pellet flow remotely to make the animal spend more time in the semi-enclosed hood in order to capture more eructations. The GreenFeed system automatically monitors the positioning of the muzzle in the hood and omits incorrect data due to incorrect head positioning [81]. The extractor fan inside the hood samples the eructed and exhaled air for analysis of various gases. Each animal has to be tagged using unique radio frequency identification (RFID) tags so that the instrument can identify each animal [82]. The instrument can restrict the excessive visit of animals for accessing bait with the aid of RFID tags. The instrument will not dispense the bait feed to the animals that visit the instrument more frequently than the interval set by the user [83]. The data generated from the GreenFeed equipment can be accessed on a real-time basis using a web-based data management system [84]. The main advantage of the GreenFeed system is that it does not require extensive labour or any other laboratory equipment and more animals can be monitored over a short span of time [79]. However, the main drawback associated with this instrument is the supply of an attractant that can modify the VFA concentration and overall digestibility of the diet. Furthermore, in the grazing paddocks, some animals might be reluctant to approach the instrument [82]. Animals must be trained thoroughly to use GreenFeed equipment before the experiment [85].

5.7. Portable Accumulation Chambers

Portable accumulation chambers are another short-term CH4-measuring technique that shows resemblance to the respiratory chamber [86].The accumulation chambers are essentially a portable, airtight polycarbonate box that contains CH4, CO2, and O2 analysers mounted on it [64,87]. The portable accumulation chamber captures all the exhaled and erected air from the animal during the sampling period and analyses it at the end [11]. The CH4 emission is calculated as the airflow inside the chamber multiplied by the level of CH4 inside the chamber that is corrected for the CH4 concentration of the incoming air, pressure, and temperature in the chamber [12,88]. This method is suitable for measuring CH4 from a large number of animals and to classify the animals based on their genetic potential to produce CH4 gas [88]. Portable accumulation chambers can also be used to assess the impact of different types of feeds and feeding regimes on CH4 production [89]. Moreover, this method is relatively inexpensive compared to many other pieces of commercially available CH4 measuring equipment [90]. However, as compared to the respiration chambers, results from portable accumulation chambers seem to be less repeatable [91].
The recent advances in CH4 estimation techniques and technologies have played an important role in the accurate quantification and mitigation of CH4 emissions and in preparing the inventories. Furthermore, different types of measuring devices have helped researchers and producers to cover emissions from heterogenous farming systems to develop national inventories. Accurate quantification of CH4 is not only critical to track our industry emissions but is equally important for the assessment of mitigation technologies that are highly needed to reduce global methane emissions. There are various mitigation technologies in addition to nutritional interventions, which are one of the main focuses of this manuscript and are reviewed in the next section. Each technology has certain merits and demerits and proper field configurations, which are outside the scope of this review but have been previously reviewed, and readers are directed to the review by Pragna et al. [13].

6. Nutritional Interventions as One of the Important Methane Mitigation Options

Among various CH4 mitigation strategies, nutritional intervention or dietary manipulation is the most effective and increasingly used strategy to mitigate enteric CH4 emission in ruminant livestock [92,93,94]. Table 1 summarises the effect of various feed additives on the CH4 and other rumen fermentation characteristics.

6.1. Concentrate Supplementation

It is obvious that the use of concentrate feed can reduce enteric CH4 production in ruminants. This is achieved mainly through shifting the fibre-based fermentation to starch fermentation [93,95]. The fermentation of starch creates an alternative hydrogen sink in the rumen by lowering the ruminal pH and inhibiting the growth of methanogens, thereby promoting more propionate production [96]. Nampoothiri et al. [97] investigated the effects of different levels of concentrate supplementation (20, 40, and 60%) on the CH4 emission from Murrah buffalo calves housed in a well-ventilated shed and reported a reduction in daily CH4 emission and yield while using the SF6 technique to measure CH4. Jiao et al. [98] fed perennial ryegrass grazing Holstein Friesian dairy cows with different ranges of concentrate feeding levels (2 kg, 4 kg, 6 kg and 8 kg as-fed basis) and reported a decline in CH4 emission (using SF6) with the increase in the level of concentrate supplementation when expressed in terms of emission per unit of feed intake and energy-corrected milk. Further, individually housed Charolais cross heifers showed a decline in enteric CH4 production when they were supplemented with 80 and 90% concentrate, although the effect was not significant at 35 or 60% concentrate inclusions [99]. A recent study conducted in Alpine Grey and Brown Swiss cattle fed on different levels of concentrate diets (low and high) showed a decrease in emission of CH4 biogenic compared to low concentrate when estimated using a life cycle assessment model [100]. Van Wyngaard et al. [101] tested three different levels of concentrate intake (0, 4, and 8 kg) in lactating Jersey cattle reared under medium-quality summer pasture and observed a decrease in the CH4 yield and intensity with increasing concentrate level, though the CH4 production peaked with the increase in concentrate supplementation. Holstein cows, tied in a modified respiratory chamber, were fed different forage to concentrate levels, 47:53, 54:46, 61:39, and 68:32, and showed 25.9, 28.2, 29.1, and 31.9 g/kg of DMI CH4 production, respectively [102]. Moreover, a very low CH4 production of around 2–3% of gross energy ingested was reported in cows supplemented with 90% concentrates [27]. In contrast, Muñoz et al. [103] observed an increase in CH4 emission (measured using the SF6 technique) per unit of milk yield with an increase in the level of concentrate supplementation. This was plausibly due to the high digestibility of perennial ryegrass pasture ingested by the Holstein animals. However, concentrate feeding beyond a certain limit is not recommended as it can cause severe damage to the animal itself and its production performance because of acute or sub-acute acidosis. Furthermore, grains that may be used for concentrates are more valuable for human feeds in arid and semi-arid regions, where much of the global ruminant production is located.

6.2. Lipid Supplementation

The use of lipid compounds offers another possible strategy to decrease enteric CH4 emission from ruminants. Addition of lipid compounds inhibits the methanogenic and ciliate protozoan population in the rumen [104,105]. Lipid addition also decreases organic matter and fibre degradability and reduces fermentable substrate in order to reduce CH4 production [106]. Machmüller and Kreuzer [107] suggested coconut oil as an efficient natural additive to reduce CH4 production without causing detrimental effects on the nutrient utilisation of the animals. On average, they observed 28 and 73% reductions in daily CH4 emission/animal when the Swiss Brown Hill wethers housed in respiratory chambers are fed with a ration containing 3.5 and 7% coconut oil, respectively. The reduction in CH4 release could be due to the suppressive effect of coconut oil on methanogens and ciliate protozoa populations. Further, Hereford × Friesian cross steers, fitted with SF6 breath sample collection canisters, were reared on canola-oil-sprayed (Oil-spray, 12 L/strip) ryegrass pasture and showed reduced CH4 production by 18% in terms of g per day [108]. Using soybean oil, Mao et al. [109] demonstrated around a 13.9% decrease in CH4 production in Huzhou lambs when measured using a simple, open-circuit respiratory chamber. Similarly, Chuntrakort et al. [110] investigated the effect of different feeding oil plant diets on CH4 emission using a headbox respiration chamber system from Thai native Brahman crossbred cattle and observed a reduction in CH4 production with oil supplementation. Among the oil-plant-supplemented diets, the coconut kernel diet was most effective in mitigating enteric CH4 emission, followed by the sunflower seed and cottonseed diets. Using open-circuit respiratory chambers, Machmüller et al. [111] reported decreased CH4 production in lambs fed different types of lipids along with total mixed rations. Within a short span of 3 weeks, they observed a 26, 27, and 10% reduction in CH4 production per kg LW when the lambs were supplemented with coconut oil, sunflower oil, and linseed oil, respectively. Conversely, Cosgrove et al. [112] reported no significant change in CH4 production (measured using a SF6 marker) from penned ryegrass pasture fed sheep supplemented with different concentrations of linseed and sunflower oil mixture (0, 1.2, 2.5, 3.7, 5.0, and 6.2% of DMI). While using lipid supplementation, one caution has to be observed that fat supplementation should not exceed over 6–7% to prevent a possible decline in dry matter intake by animals due to the inconvenient odour [93].

6.3. Ionophore Supplementation

Ionophores are generally used in livestock feed to improve feed efficiency and to increase body weight [113]. Commonly available forms of ionophores include lasalocid, monensin, laidlomycin propionate, tetronasin, salinomycin, narasin, and lysocellin [113,114]. Ionophores act as a CH4-inhibiting factor by shifting the fermentation acids from acetic acid and butyric acid to propionic acid by promoting the growth and proportion of Gram-positive bacteria in the rumen. Stall-fed Holstein cows treated with 18 mg/kg of dry matter monensin showed a 24.3% decline in CH4 production when expressed in g/day. According to the authors, this reduction could be due to the positive effect of monensin on the Gram-negative bacteria that produce propionate and due to the negative effect of monensin on the acetate and hydrogen-producing bacteria such as Eubacterium, Lactobacillus, and Streptococcus [115]. Likewise, feedlot-type penned Angus steers fed with 33 mg/kg monensin showed a 30% reduction in enteric CH4 production (measured using the SF6 technique) along with a numeric reduction in ciliated protozoan populations [116]. The CH4-reducing effect of monensin is mainly due to changes in the production of ruminal volatile fatty acids. Li et al. [117] reported a 20.3 L/day reduction in CH4 in goats supplemented with monensin and housed in a closed portable static environmental chamber. Additionally, stall-fed Murrah buffalo heifers supplemented with sodium monensin showed 8–9% reduced energy loss in the form of CH4 when estimated using the SF6 tracer technique [118]. However, the excessive supplementation of ionophores can lead to toxicity in ruminants, and they also need to be screened for their residual level in the animal products [119,120].

6.4. Anti-Methanogenic Compounds

Anti-methanogenic compounds are another important nutritional intervention in the enteric CH4 mitigation studies, though the usage of some chemical anti-methanogenic compounds is not allowed in some countries because of their anti-nutritional effects. Bromochloromethane (BCM) is one of the widely researched anti-methanogenic compounds that has the potential to reduce a considerable amount of CH4 from the ruminants. When incorporated into the diets of stall-fed steers, BCM-Cyclodextrin (1 g/100 kg BW/day) has been reported to reduce CH4 production by around 95% by hindering the cobamide-dependent methyltransferase step in the process of methanogenesis through its reaction with vitamin B12 [121]. Furthermore, Lalu et al. [122] reported a 90% reduction in CH4 emission from penned rams supplemented with BCM. In addition, 3-nitrooxypropanol (3-NOP) is another possible anti-methanogenic compound and Romero-Perez et al. [123] conducted a series of experiments in barn-tied beef cattle using 3-NOP. In the first set of experiments, the authors compared different dosages of 3-NOP (0, 0.75, 2.25, and 4.50 mg/kg BW) and observed a linear decline in CH4 production with the level of dosage, with 33% CH4 reduction from the highest dosage. The authors found a shift in VFA production from acetate to more propionate without hampering body weight gain or feed digestibility. In the next long-duration experiment (112 days) using 3-NOP at 2 g/day level, the author had observed a 60% reduction in the enteric CH4 production without the microbial adaptation to 3-NOP when measured using closed-circuit respiratory chambers [123]. Likewise, Lopes et al. [124] also reported a 31% decline in CH4 production (CH4 estimated using GreenFeed system) from lactating Holstein cows fed 3-NOP at the rate of 60 mg/kg of feed dry matter. In another study using ethyl-3-nitrooxy propionate (E3NP) and 3-NOP, Martínez-Fernández et al. [125] demonstrated 14 and 25% decreases in enteric CH4 production, respectively, from E3NP- and 3-NOP-fed Segureña sheep using respiration chambers.
Apart from the synthetic or artificial anti-methanogenic derivatives, some of the naturally occurring red algae, seaweeds, fungus, and lichens can produce haloforms, dihalomethanes, and some other organobromine compounds that have an anti-methanogenic effect [126,127]. Li et al. [128] demonstrated an 80% reduction in enteric CH4 emission from penned Merino-cross wethers when they were supplemented with 3% Asparagopsis taxiformis organic matter; here, the authors used open-circuit respiration chambers for the measurement of CH4. Kinley et al. [129] reported the anti-methanogenic effect of red macroalgae, Asparagopsis taxiformis, in vitro when fermented with a high-quality Rhodes grass. This work demonstrated a significant enteric CH4 reduction with the supplementation of 1% Asparagopsis. Further, Roque et al. [130] reported a 26.4% reduction in CH4 production (estimated using the GreenFeed Large Animal System) without compromising feed intake or milk yield when the freestall barn-housed cattle were supplemented with 0.5% level Asparagopsis armata (organic matter basis). However, when they increased the inclusion level to 1%, it resulted in a 67.2% enteric CH4 reduction but with negative effects on feed intake and milk yield. Martínez et al. [131] found an anti-methanogenic effect of garlic-derived compound propyl propane thiosulfinate. In order to test the anti-methanogenic potential of allyl disulphide and lovastatin, Klevenhusen et al. [132] conducted a study with caged swiss Black-Brown Mountain sheep. Briefly, the sheep were randomly allocated to a diet supplemented with 4 g diallyl disulphide and a diet supplemented with 80 mg lovastatin per kg of total dietary dry matter for 23 days and the animals of the experiment were kept inside the open-circuit respiratory chambers for 4 days for measuring CH4 emission. In summary, they could not find any significant influence of dietary supplements on daily CH4 production. However, diallyl disulphide showed a reduction in CH4 production when expressed in per kg NDF digested. Therefore, this study revealed the potential of diallyl disulphide, a garlic oil derivative, to improve fibre digestion and to limit energy loss in the form of CH4.

6.5. Probiotic Feeding

Probiotics are potential feed additives that have many beneficial properties, including immunity stimulation, stabilisation of the microbes in the digestive tract, production of anti-microbial substances, prevention of feed-related allergies, improved dry matter intake and fibre digestibility, and CH4 mitigation [133,134]. Some of the direct-fed ruminant specific probiotics include Saccharomyces cerevisiae, Enterococcus, Bifidobacterium, Lactobacillus, Propionibacterium, Prevotellabryantii, Bacillus, and Megasphaeraelsdenii [8,135]. Recently, Hassan et al. [134] studied the effect of a Ruminococcus flavefaciens-based probiotic supplement on CH4 production in Barki lambs kept in caged conditions. Their results showed a significant reduction in CH4 production as compared to controls; this change could be attributed to variation in the rumen microflora. Likewise, in another CH4 estimation study using head hood systems conducted in Bacillus licheniformis-supplemented Dorper × thin-tailed Han wethers, Deng et al. [136] reported a 6% reduction in daily CH4 production. Additionally, Latham et al. [137] found the possibility of using Paenibacillus 79R4 as a probiotic supplement in order to reduce nitrate toxification and CH4 production in nitrate-treated steers grazing on Bermuda grass pasture. Suryani et al. [138] found that Saccharomyces cerevisiae and combination of Saccharomyces cerevisiae with Bacillus amyloliquefaciens could reduce the CH4 production from Bali cattle kept in individual pens by stimulating the acetogens in the rumen to compete with methanogenic bacteria. Recently, Chen et al. [139] conducted an in vitro experiment using a cluster of different propionic acid bacterial strains and reported the ability of Propionibacterium jensenii LMGT2826 and Propionibacterium thoenii LMGT2827 and Propionibacterium thoenii T159 bacterial strains to mitigate CH4 emission by 18, 8, and 20%, respectively, compared to the control. However, the application of Propionibacterium acidipropionici as a feed additive did not affect the CH4 production (measured using open-circuit respiration chambers) from Merino wethers [140].

6.6. Essential Oils

Generally, essential oils are volatile aromatic substances extracted from herbs and spices [141]. Essential oils contain a variety of chemical substances, such as isoprenes, terpenes, diterpenes, triterpenes, hemiterpenes, sesquiterpenes, and tetraterpenes, etc. Essential oils possess antimicrobial properties against ruminal inhabitants such as bacteria, fungi, and protozoa [142]. Additionally, essential oils have shown promising potential in improving the production potential and in mitigating enteric CH4 emission [141]. The use of essential oils as a CH4 mitigation strategy has been greatly tested by several authors over the last decade and extracts from citrus, oregano, garlic, thyme, and cinnamon have given consistent results [141,143]. For example, Wu et al. [144] suggested intermittent feeding of citrus essential oil as a potential CH4 reduction strategy in Hu sheep housed in individual cages by reducing microbial adaptation to additives. Further, Hart et al. [145] demonstrated 6% less CH4 production per day in cows kept in freestall barns with the supplementation of a commercial essential oil blend (Agolin Ruminant Liquid Formulation); this was measured using a GreenFeed large animal monitor. Cows fed with a feed additive rich in thyme essential oil have also shown a significant reduction in CH4 produced (CH4 was measured using an indirect calorimetry facemask system) [146]. Soltan et al. [147] conducted an essential oil feeding experiment in Santa Inês sheep. Briefly, sheep were fed with a microencapsulated blend consisting of cinnamaldehyde, carvacrol, capsicum oleoresin, and eugenol. Sheep were kept inside the respiratory chamber for measuring CH4 emission. The sheep fed with the essential oil bled had significantly lower CH4 production, without any antagonistic effect on nutrient digestibility. In another experiment, Sallama et al. [148] supplemented sheep kept in open-circuit respiration chambers with 10 mL and 20 mL/day eucalyptus essential oil and reported 31 and 22% reductions in CH4 production, respectively.

6.7. Organic Acid Supplementation

Predominantly, organic acids fed to the animals are of natural origin, with low potential for toxicity, as they naturally occur in the cell metabolism. Organic acid supplementation helps the animals to prevent their ruminal pH from falling; at the same time, it also helps to reduce the methanogenesis in the rumen [149]. Among the various organics acids, aspartate, malate, and fumarate are known for their ability to act as an alternative hydrogen sink to promote more propionate production [150]. Dietary supplementation of fumaric acid (2% of the diet dry matter) with a roughage-based diet has been reported to decrease CH4 production (measured using the head hood system) by 23% in stall-fed Holstein steers, changes that were also accompanied by increased total VFA production and propionic acid production. However, the potential of organic acids to lower CH4 may depend on the level of organic acid supplementation and the dietary condition [151]. To omit acidity-related issues due to organic acid supplementation, Wallace RJ [152] encapsulated fumaric acid with a shell of hydrogenated vegetable oil and observed a greater reduction (75%) in CH4 production from Welsh Mule Cross lambs. CH4 was estimated with the help of a polythene tunnel system. Further, Dorper × Thin-tailed Han crossbred ewes showed a decrease in the daily enteric CH4 output from 66.1 L/kg digestible organic matter (DOP) to 61.01 L/kg/DOP when supplemented with allicin and this was measured using a headbox system [153]. Additionally, Charolais cross heifers supplemented with DL-malic acid also showed a 16% reduction in daily total CH4 emission (measured using the SF6 technique) [154]. When tested using in vitro batch fermentation with a mixed diet of meadow hay, barley, and sugar beet molasses, sodium aspartate gained a 21.56% reduction in CH4 production. Furthermore, supplementation of aspartate increased the production of propionate without reducing acetate production [155]. However, the high cost of organic acid makes its commercial usage a somewhat economically unviable option [150].

6.8. Exogenous Enzymes

Exogenous enzymes are widely used to remove the anti-nutritional factors in livestock feed and to improve digestibility [156]. The enzymes are generally sourced from bacteria such as Lactobacillus acidophilus, 5 Streptococcus faecium, spp. L. plantarum, and Bacillus subtilis, and fungi like Trichoderma reesei, Aspergillus oryzae, and 6 Saccharomyces cerevisiae. The studies linking CH4 production and exogenous enzymes are very limited and equivocal. Some studies showed that enzyme addition decreased CH4 production by ruminant animals but others did not [157]. Arriola et al. [158] tested the effect of a fibrolytic enzyme on CH4 production from two groups of Holstein cows fed low- and high-concentrate diets, respectively, and they observed a reduction in CH4 production when the animals were supplemented with the fibrolytic enzyme; these animals were housed in a freestall, open-sided barn. Further, these effects were more prominent in the high-concentrate-based diet. In a review of the nutritional management for enteric CH4 abatement, evidence was presented to support a role for exogenous enzymes in the mitigation of enteric CH4 produced from ruminants [93]. Zhao et al. [159] demonstrated a reduction in CH4 production from feed substrates supplemented with cellulose and xylanase enzymes and tested in vitro. Contrastingly, negative effects of exogenous enzyme supplementation have also been reported in cattle [160,161] and goats [162].

6.9. Plant Secondary Metabolites

Plant secondary metabolites are the secondary group of molecules that help the plants to adapt to their micro and macro environment. Protease inhibitors, lectins, alkaloids, nonprotein amino acids, cyanogenic glycosides, terpenes, saponins, and tannins are some of the key plant secondary metabolites [163,164]. Secondary metabolites such as condensed tannins and saponins have an anti-methanogenic and anti-protozoal effect [165]. There are a plethora of studies citing the ability of secondary metabolites to mitigate CH4 production from the ruminant animals [166,167]. SF6 canister-fitted Santa Inês lambs fed with Leucaena leucocephala showed a 25.7% enteric CH4 reduction [168]. Similarly, hydrolysed tannins from Castanea sativa wood have also shown enteric CH4 depressing activity (20% reduction) with significant anti-protozoal effects when tested with Swiss White Hill lambs kept in respiratory chambers [169]. Further, Baruah et al. [167] reported a 19–21% reduction in enteric CH4 emission in penned, SF6-equipped Mandya lambs when they were supplemented with Syzygium cumini and Machilus bomycina leaves containing phyto-sources. The presence of condensed tannins in Tamarindus indica seed husk was found to be inhibitory to methanogenic activity in the rumen. Condensed tannins found in the Tamarindus indica seed husk could affect enzymatic activity, cell membrane composition, and metallic ion exchange in methanogens. Additionally, most of the tanniferous compounds present in the plants can increase the duodenal protein flow by reducing the rate of protein breakdown in the rumen when supplemented at a moderate dosage to ruminants [164,170]. Malik et al. [171] reported a 10–50% reduction in methanogenic activity following Tamarindus indica seed husk supplementation to penned crossbred cattle that were equipped with SF6 canisters. Furthermore, penned Thai native beef cattle fed Bamboo-Cass that contains 2.8% and 1.3% condensed tannins and crude saponins, respectively, showed a reduction in CH4 production by suppressing protozoal populations [172]. Dietary supplementation of chestnut tannins was shown to reduce CH4 production in Rideau Arcott sheep, without any negative effect on their growth performance. However, the sheep showed a reduction in methanogen and protozoa populations, and the authors used respiratory chambers to measure CH4 emission [173]. Albores-Moreno et al. [174] supplemented caged Pelibuey × Katahdin lambs with saponin-rich ground pods of Enterolobium cyclocarpum and observed a 36% reduction in CH4 production. Moreover, Váradyová et al. [175] reported that asphonins help the animal nutrients to bypass the rumen, thereby lowering the methanogenesis as a result of reductive acetogenesis. Plant secondary metabolites have been observed to have differing effects on the rumen methanogenesis depending on the plant sources and dosages; however, their mode of action depends on their direct or indirect effect on the microbes responsible for CH4 production. It is important to consider the risk associated with anti-nutritional factors while feeding plant secondary metabolites, which might cause detrimental effects on animal health and feed palatability; in particular, plant saponins could cause haemolysis in animals [92].
Table 1. Effect of various feed additives on CH4 and other rumen fermentation characteristics—a summary.
Table 1. Effect of various feed additives on CH4 and other rumen fermentation characteristics—a summary.
Feed AdditivesDosageSpeciesImpact on CH4Current FeasibilityReference
Concentrate feed containing vitamins and mineral supplement6 kg/dayCow↑ CH4 (g/day)feasible[176]
Ground corn3.2 kg/dayCow↓ CH4 yield by 7.3 g/kg DMIfeasible[177]
Concentrate feed containing maize, rapeseed meal, soybean, Molaferm, and Megalac, etc.6 kg/dayCowNo effect (g/kgDMI) but ↓ CH4 (g/kg Energy corrected milk)feasible[178]
Concentrate feed containing barley, beet pulp, soybean meal, maize meal, molasses, vitamins, and minerals0.5 kg/dayLambNo effect (CH4 g/kg DMI)feasible[179]
Concentrate feed mixture18.1% of DMCow↓ CH4 (g/day and g/kg DMI)feasible[180]
Concentrate feed containing maize, deoiled mustard cake, soybean meal, wheat bran, rice bran, mineral mixture, and salt etc.15% high ME content (2.82 Mcal/kg)buffalo↓ CH4 g/day, g/kg DM intake)feasible[181]
Coconut oil4%Goat34% ↓ CH4 emissionSomewhat feasible[182]
Coconut oil2%GoatMore than 50% ↓ CH4 emissionSomewhat feasible[183]
Soybean oil4%Goat32% ↓ CH4 emissionSomewhat feasible[182]
Soybean
oil
50 g/kg DMSheep 35.8% ↓ CH4 emissionSomewhat feasible[184]
Corn oil30 g/kg DM Goat15.1% ↓ CH4 emission (g/kg DMI)feasible[185]
Corn oil5%Cattle~30% ↓ CH4 emissionfeasible[186]
Lasalocid200 mg/hd/dCattle30.91% ↓ CH4 production (g/kg DMI)feasible[187]
Monensin30 mg/kgSteer16.67% ↓ CH4 (MJ/100 MJ GE intake)Currently not feasible[188]
Monensin22 mg/kgGoat28% ↓ CH4 emissionCurrently not feasible[182]
Monensin0.6 mg/kg of body weightBuffalo8–9%↓ CH4 emissionCurrently not feasible[118]
Nitrate11 g/kg DMCow↓ CH4 by 8%Not permitted in some countries[101]
Nitrate23 g/kg DMCow↓ CH4 by 15%Not permitted in some countries[101]
Ethyl-3-NOP50 and 500 mg/animal per daySheep↓ CH4 by 29% (L/kg of DMI)Not permitted in some countries[125]
3NOP60 mg of 3NOP/kg DMCow↓ CH4 by 31%Not permitted in some countries[124]
Bacillus licheniformis2.5 × 108 colony forming units (CFU)Sheep↓ CH4 by 6%Currently not economically feasible[136]
Bacillus licheniformis2.5 × 109 CFUsSheep↓ CH4 by 12%Currently not economically feasible[136]
Saccharomyces cerevisiae(1.2–2.3) × 107 CFU/gSheep↓ CH4 by 10% (L/day)Currently not economically feasible[189]
Leuconostoc mesenteroides(1.5–1.8) × 109 CFU/gSheepNo effectCurrently not economically feasible[189]
Orange leavesTMRGoat↓ CH4 by 32% (g/day)feasible[190]
Citrus essential oil0. 0.8 and 1.6 mL/Lsheep↓ CH4feasible[144]
Commercial essential oil blend1 g/day Cow↓ CH4 by 6% (g/day)feasible[145]
Encapsulated fumaric acid117 g EFA/kgLamb↓ CH4 by 76% (L/day)feasible[191]
Fumaric acid100 g FA and 17 g partially hydrogenated vegetable oil/kgLamb↓ CH4 by 62% (L/day)Currently not economically feasible[191]
Dl-malic acid7.5% on a DM basisBeef cattle↓ CH4 by 9% (L/day)Currently not economically feasible[154]
Cellulase10,000 IU/gGoatNo effectCurrently not economically possible[192]
Cellulose/xylanase7000 IU/g of cellulase and 5000 IU/g of xylanaseGoatNo effectCurrently not economically feasible[192]
Leucaena leucocephala350 g/kg DMsheep↓ CH4 by 14.1% g/kg DMI [193]
White grape marc5.0 kg DMCow↓ CH4 by 15% g/kg DMIfeasible[194]
Red grape marc5.0 kg DMCow↓ CH4 by 15% g/kg DMIfeasible[194]
Willow fodder (Salix spp.)12 g CT kg/DMISheep↓ CH4 by 19% (g/kg BW0.75/day)feasible[195]

7. Conclusions and Future Perspectives

Most of the CH4 emitted from the livestock production systems is mainly in the form of enteric CH4. With the changing climate and global warming, it is very important to develop strategies to reduce or mitigate CH4 emissions from livestock production systems. However, it is equally important to develop methods and technologies to measure CH4 emissions efficiently and accurately. There are several methods and equipment available for the estimation of CH4 emission from ruminants. However, most of these techniques have certain advantages and disadvantages and therefore a careful selection of methods is needed for specific production systems. For example, SF6 is more suitable for grazing studies while respiration chambers and hood systems are only useful for indoor studies. Likewise, a plethora of feed supplements for CH4 mitigation from ruminants have been developed but some of these feed additives may not be feasible for farm usage because of their toxic levels, accessibility, and cost. However, some of the strategies, such as adjusting the roughage to concentrate ratio and using feeding additives such as lipids, essential oils, and plant secondary metabolites, can be used on the farm level to achieve CH4 mitigation.
Mitigation strategies that do not hamper production and are able to reduce CH4 emissions in ruminants have better acceptance among farmers and the industry. In practice, farmers are less likely to adopt any of the mitigation technologies that do not attain a minimum sustainable production level or are not economically viable, while reducing methane emissions. Therefore, during the initial phase of transitioning and adoption, the provision of rewards or some incentives might encourage farmers to adopt these mitigation strategies. Furthermore, most of the nutritional interventions have been developed and assessed under the intensive system or in the in-house conditions; therefore, further research is necessary to evaluate the long-term effectiveness of these feeding strategies in grazing farm systems, which are contributing a high proportion of livestock methane emissions.

Author Contributions

Conceptualisation, P.P., F.R.D. and S.S.C.; methodology, P.P., F.R.D. and S.S.C.; writing—original draft preparation, P.P.; writing—review and editing, P.P., F.R.D., S.S.C., J.J.C. and B.J.L.; supervision, F.R.D., S.S.C., J.J.C. and B.J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This project received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We wish to thank the University of Melbourne for providing the Melbourne Research Scholarship to Pragna Prathap in order to support this work at the University of Melbourne.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. UN-EPA. Global Anthropogenic Non-CO2 Greenhouse Gas Emissions: 1990–2030; Climate Change Division U.S. Environmental Protection Agency 1200 Pennsylvania Avenue NW: Washington, DC, USA, 2012.
  2. IPCC. Climate Change and Land, an Ipcc Special Report on Climate Change, Desertification, Land Degradation, Sustainable Land Management, Food Security, and Greenhouse Gas Fluxes in Terrestrial Ecosystems; Inter Governmental Panel on Climate Change: Geneva, Switzerland, 2019; p. 41.
  3. Opio, C.; Gerber, P.; Mottet, A.; Falcucci, A.; Tempio, G.; MacLeod, M.; Vellinga, T.; Henderson, B.; Steinfeld, H. Greenhouse Gas Emissions from Ruminant Supply Chains–A Global Life Cycle Assessment; Food and Agriculture Organization of the United Nations: Rome, Italy, 2013. [Google Scholar]
  4. Davison, T.M.; Black, J.L.; Moss, J.F. Red meat—An essential partner to reduce global greenhouse gas emissions. Anim. Front. 2020, 10, 14–21. [Google Scholar] [CrossRef] [PubMed]
  5. Ammar, H.; Abidi, S.; Ayed, M.; Moujahed, N.; de Haro Martí, M.E.; Chahine, M.; Bouraoui, R.; López, S.; Cheikh M’hamed, H.; Hechlef, H. Estimation of tunisian greenhouse gas emissions from different livestock species. Agriculture 2020, 10, 562. [Google Scholar] [CrossRef]
  6. Giraldo, P.; Marín, A.; Giraldo, L.A. Effects of different feed additives on methane emissions from beef cattle. In Impacts on and Adaptation of Livestock Production Systems to Climate Change; Universidad Nacional de Colombia: Cundinamarca, Colombia, 2013; pp. 1386–1387. [Google Scholar]
  7. Stewart, E.K.; Beauchemin, K.A.; Dai, X.; MacAdam, J.W.; Christensen, R.G.; Villalba, J.J. Effect of tannin-containing hays on enteric methane emissions and nitrogen partitioning in beef cattle1. J. Anim. Sci. 2019, 97, 3286–3299. [Google Scholar] [CrossRef] [PubMed]
  8. Kim, S.-H.; Lee, C.; Pechtl, H.A.; Hettick, J.M.; Campler, M.R.; Pairis-Garcia, M.D.; Beauchemin, K.A.; Celi, P.; Duval, S.M. Effects of 3-nitrooxypropanol on enteric methane production, rumen fermentation, and feeding behavior in beef cattle fed a high-forage or high-grain diet. J. Anim. Sci. 2019, 97, 2687–2699. [Google Scholar] [CrossRef]
  9. DISER-Department of Industry; Energy and Resources. National greenhouse gas inventory quarterly update: September 2020. In Quarterly; Commonwealth of Australia: Sydney, Australia, 2021. [Google Scholar]
  10. Bhatta, R.; Malik, P.K.; Sejian, V. Enteric methane emission and reduction strategies in sheep. In Sheep Production Adapting to Climate Change; Sejian, V., Bhatta, R., Gaughan, J., Malik, P.K., Naqvi, S.M.K., Lal, R., Eds.; Springer: Singapore, 2017; pp. 291–305. [Google Scholar]
  11. Hammond, K.J.; Crompton, L.A.; Bannink, A.; Dijkstra, J.; Yáñez-Ruiz, D.R.; O’Kiely, P.; Kebreab, E.; Eugène, M.; Yu, Z.; Shingfield, K.J. Review of current in vivo measurement techniques for quantifying enteric methane emission from ruminants. Anim. Feed Sci. Technol. 2016, 219, 13–30. [Google Scholar] [CrossRef] [Green Version]
  12. Patra, A.K. Recent advances in measurement and dietary mitigation of enteric methane emissions in ruminants. Front. Vet. Sci. 2016, 3, 39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Pragna, P.; Chauhan, S.S.; Sejian, V.; Leury, B.J.; Dunshea, F.R. Climate change and goat production: Enteric methane emission and its mitigation. Animals 2018, 8, 235. [Google Scholar] [CrossRef] [Green Version]
  14. Min, B.R.; Solaiman, S.; Waldrip, H.M.; Parker, D.; Todd, R.W.; Brauer, D. Dietary mitigation of enteric methane emissions from ruminants: A review of plant tannins mitigation options☆. Anim. Nutr. 2020, 6, 231–246. [Google Scholar] [CrossRef]
  15. McCauley, J.I.; Labeeuw, L.; Jaramillo-Madrid, A.C.; Nguyen, L.N.; Nghiem, L.D.; Chaves, A.V.; Ralph, P.J. Management of enteric methanogenesis in ruminants by algal-derived feed additives. Curr. Pollut. Rep. 2020, 6, 1–18. [Google Scholar] [CrossRef]
  16. Baca-González, V.; Asensio-Calavia, P.; González-Acosta, S.; Pérez de la Lastra, J.M.; Morales de la Nuez, A. Are vaccines the solution for methane emissions from ruminants? A systematic review. Vaccines 2020, 8, 460. [Google Scholar] [CrossRef]
  17. IPCC. Global Warming of 1.5 °C-An Ipcc Special Report on the Impacts of Global Warming of 1.5 °C above Pre-Industrial Levels and Related Global Greenhouse Gas Emission Pathways, in the Context of Strengthening the Global Response to the Threat of Climate Change, Sustainable Development, and Efforts to Eradicate Poverty; IPCC: Geneva, Switzerland, 2018. [Google Scholar]
  18. IPCC. Synthesis report: Summary for policy makers. In Fifth Assessment Report (ar5); Pachauri, R.K., Riahi, K., Eds.; Intergovernmental Panel on Climate Change: Geneva, Switzerland, 2014; p. 31. [Google Scholar]
  19. IPCC. Summary for policymakers ipcc special report on the ocean and cryosphere in a changing climate. In Ipcc Special Report on the Ocean and Cryosphere in a Changing Climate; Pörtner, H.O., Masson-Delmotte, V., Zhai, P., Tignor, M., Poloczanska, E., Mintenbeck, K., Nicolai, M., Okem, A., Petzold, J., Rama, B., et al., Eds.; IPCC: Geneva, Switzerland, 2019; in press. [Google Scholar]
  20. Steffen, W.; Hughes, L.; Perkins, S. Heatwaves: Hotter, Longer, More Often; Climate Council of Australia: Sydney, Australia, 2014. [Google Scholar]
  21. Wuebbles, D.J.; Hayhoe, K. Atmospheric methane and global change. Earth-Sci. Rev. 2002, 57, 177–210. [Google Scholar] [CrossRef]
  22. Malyan, S.K.; Bhatia, A.; Kumar, A.; Gupta, D.K.; Singh, R.; Kumar, S.S.; Tomer, R.; Kumar, O.; Jain, N. Methane production, oxidation and mitigation: A mechanistic understanding and comprehensive evaluation of influencing factors. Sci. Total Environ. 2016, 572, 874–896. [Google Scholar] [CrossRef]
  23. Malyan, S.K.; Bhatia, A.; Kumar, S.S.; Fagodiya, R.K.; Pugazhendhi, A.; Duc, P.A. Mitigation of greenhouse gas intensity by supplementing with azolla and moderating the dose of nitrogen fertilizer. Biocatal. Agric. Biotechnol. 2019, 20, 101266. [Google Scholar] [CrossRef]
  24. EPA. Importance of Methane. In Global Methane Initiative; United States Environmental Protection Agency: Washington, DC, USA, 2016. Available online: https://www.epa.gov/gmi/importance-methane#:~:text=Methane%20is%20also%20a%20greenhouse,%2Dinfluenced)%20and%20natural%20sources.&text=Methane%20is%20more%20than%2025,trapping%20heat%20in%20the%20atmosphere (accessed on 15 July 2020).
  25. Brasseur, G.P.; Schultz, M.; Granier, C.; Saunois, M.; Diehl, T.; Botzet, M.; Roeckner, E.; Walters, S. Impact of climate change on the future chemical composition of the global troposphere. J. Clim. 2006, 19, 3932–3951. [Google Scholar] [CrossRef] [Green Version]
  26. Hook, S.E.; Wright, A.D.; McBride, B.W. Methanogens: Methane producers of the rumen and mitigation strategies. Archaea 2010, 2010, 945785. [Google Scholar] [CrossRef] [Green Version]
  27. Johnson, D.E.; Ward, G.M. Estimates of animal methane emissions. Environ. Monit. Assess. 1996, 42, 133–141. [Google Scholar] [CrossRef] [PubMed]
  28. Huws, S.A.; Creevey, C.J.; Oyama, L.B.; Mizrahi, I.; Denman, S.E.; Popova, M.; Muñoz-Tamayo, R.; Forano, E.; Waters, S.M.; Hess, M. Addressing global ruminant agricultural challenges through understanding the rumen microbiome: Past, present, and future. Front. Microbiol. 2018, 9, 2161. [Google Scholar] [CrossRef] [PubMed]
  29. Matthews, C.; Crispie, F.; Lewis, E.; Reid, M.; O’Toole, P.W.; Cotter, P.D. The rumen microbiome: A crucial consideration when optimising milk and meat production and nitrogen utilisation efficiency. Gut Microbes 2019, 10, 115–132. [Google Scholar] [CrossRef] [PubMed]
  30. Kumari, S.; Fagodiya, R.K.; Hiloidhari, M.; Dahiya, R.P.; Kumar, A. Methane production and estimation from livestock husbandry: A mechanistic understanding and emerging mitigation options. Sci. Total Environ. 2020, 709, 136135. [Google Scholar] [CrossRef] [PubMed]
  31. Moss, A.R.; Jouany, J.-P.; Newbold, J. Methane Production by Ruminants: Its Contribution to Global Warming; Annales de zootechnie, 2000; EDP Science: London, UK, 2000; pp. 231–253. [Google Scholar]
  32. Henderson, G.; Cox, F.; Ganesh, S.; Jonker, A.; Young, W.; Janssen, P.H. Rumen microbial community composition varies with diet and host, but a core microbiome is found across a wide geographical range. Sci. Rep. 2015, 5, 1–15. [Google Scholar] [CrossRef]
  33. Tapio, I.; Snelling, T.J.; Strozzi, F.; Wallace, R.J. The ruminal microbiome associated with methane emissions from ruminant livestock. J. Anim. Sci. Biotechnol. 2017, 8, 1–11. [Google Scholar] [CrossRef] [PubMed]
  34. Van Lingen, H.J.; Plugge, C.M.; Fadel, J.G.; Kebreab, E.; Bannink, A.; Dijkstra, J. Thermodynamic driving force of hydrogen on rumen microbial metabolism: A theoretical investigation. PLoS ONE 2016, 11, e0161362. [Google Scholar]
  35. FAO. Five Practical Actions towards Low-Carbon Livestock; Animal Production and Health Division: Rome, Italy, 2019. [Google Scholar]
  36. FAOSTAT. Fao Statistical Data Base; FAOSTAT: Rome, Italy, 2018. [Google Scholar]
  37. Gerber, P.J.; Steinfeld, H.; Henderson, B.; Mottet, A.; Opio, C.; Dijkman, J.; Falcucci, A.; Tempio, G. Tackling Climate Change through Livestock: A Global Assessment of Emissions and Mitigation Opportunities; Food and Agriculture Organization of the United Nations (FAO): Rome, Italy, 2013. [Google Scholar]
  38. Grossi, G.; Goglio, P.; Vitali, A.; Williams, A.G. Livestock and climate change: Impact of livestock on climate and mitigation strategies. Anim. Front. 2019, 9, 69–76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. FAO. Global Livestock Environmental Assessment Model; Food and Agriculture Organization of the United Nations: Rome, Italy, 2018; Volume 2, p. 109. [Google Scholar]
  40. Goopy, J.P.; Chang, C.; Tomkins, N. A comparison of methodologies for measuring methane emissions from ruminants. In Methods for Measuring Greenhouse Gas Balances and Evaluating Mitigation Options in Smallholder Agriculture; Rosenstock, T.S., Rufino, M.C., Butterbach-Bahl, K., Wollenberg, L., Richards, M., Eds.; Springer International Publishing: Cham, Switzerland, 2016; pp. 97–117. [Google Scholar]
  41. Brouček, J. Methods of methane measurement in ruminants. Slovak J. Anim. Sci. 2014, 47, 51–60. [Google Scholar]
  42. Alvarez Hess, P.S.; Eckard, R.J.; Jacobs, J.L.; Hannah, M.C.; Moate, P.J. Comparison of five methods for the estimation of methane production from vented in vitro systems. J. Sci. Food Agric. 2019, 99, 109–116. [Google Scholar] [CrossRef]
  43. Russo, V.; Jacobs, J.; Hannah, M.; Moate, P.; Dunshea, F.; Leury, B. In vitro evaluation of the methane mitigation potential of a range of grape marc products. Anim. Prod. Sci. 2017, 57, 1437–1444. [Google Scholar] [CrossRef]
  44. Gonzalez-Rivas, P.; DiGiacomo, K.; Russo, V.; Leury, B.; Cottrell, J.; Dunshea, F. Feeding slowly fermentable grains has the potential to ameliorate heat stress in grain-fed wethers. J. Anim. Sci. 2016, 94, 2981–2991. [Google Scholar] [CrossRef]
  45. Hess, P.A.; Giraldo, P.; Williams, R.; Moate, P.; Beauchemin, K.; Eckard, R. A novel method for collecting gas produced from the in vitro ankom gas production system. J. Anim. Sci. 2016, 94, 570. [Google Scholar] [CrossRef]
  46. Ramin, M.; Huhtanen, P. Development of an in vitro method for determination of methane production kinetics using a fully automated in vitro gas system—A modelling approach. Anim. Feed Sci. Technol. 2012, 174, 190–200. [Google Scholar] [CrossRef]
  47. Dubois, B.; Tomkins, N.W.; Kinley, R.D.; Bai, M.; Seymour, S.; Paul, N.A.; de Nys, R. Effect of tropical algae as additives on rumen in vitro gas production and fermentation characteristics. Am. J. Plant Sci. 2013, 4, 34–43. [Google Scholar] [CrossRef] [Green Version]
  48. Ramin, M.; Krizsan, S.; Jančík, F.; Huhtanen, P. Measurements of methane emissions from feed samples in filter bags or dispersed in the medium in an in vitro gas production system. J. Dairy Sci. 2013, 96, 4643–4646. [Google Scholar] [CrossRef]
  49. Bhatta, R.; Tajima, K.; Takusari, N.; Higuchi, K.; Enishi, O.; Kurihara, M. Comparison of Sulfur Hexafluoride Tracer Technique, Rumen Simulation Technique and In Vitro Gas Production Techniques for Methane Production from Ruminant Feeds; International Congress Series; Elsevier: Amsterdam, The Netherlands, 2006; pp. 58–61. [Google Scholar]
  50. Czerkawski, J.; Breckenridge, G. Design and development of a long-term rumen simulation technique (rusitec). Br. J. Nutr. 1977, 38, 371–384. [Google Scholar] [CrossRef] [Green Version]
  51. Kellner, O.J.; Goodwin, W. The Scientific Feeding of Animals; Duckworth: Richmond-upon-Thames, UK, 1913. [Google Scholar]
  52. Muñoz, C.; Yan, T.; Wills, D.; Murray, S.; Gordon, A. Comparison of the sulfur hexafluoride tracer and respiration chamber techniques for estimating methane emissions and correction for rectum methane output from dairy cows. J. Dairy Sci. 2012, 95, 3139–3148. [Google Scholar] [CrossRef] [PubMed]
  53. Schwarm, A.; Schweigel-Röntgen, M.; Kreuzer, M.; Ortmann, S.; Gill, F.; Kuhla, B.; Meyer, U.; Lohölter, M.; Derno, M. Methane emission, digestive characteristics and faecal archaeol in heifers fed diets based on silage from brown midrib maize as compared to conventional maize. Arch. Anim. Nutr. 2015, 69, 159–176. [Google Scholar] [CrossRef] [Green Version]
  54. Hammond, K.; Hoskin, S.; Burke, J.; Waghorn, G.; Koolaard, J.; Muetzel, S. Effects of feeding fresh white clover (trifolium repens) or perennial ryegrass (lolium perenne) on enteric methane emissions from sheep. Anim. Feed Sci. Technol. 2011, 166, 398–404. [Google Scholar] [CrossRef]
  55. Vanlierde, A.; Soyeurt, H.; Gengler, N.; Colinet, F.G.; Froidmont, E.; Kreuzer, M.; Grandl, F.; Bell, M.; Lund, P.; Olijhoek, D.W.; et al. Short communication: Development of an equation for estimating methane emissions of dairy cows from milk fourier transform mid-infrared spectra by using reference data obtained exclusively from respiration chambers. J. Dairy Sci. 2018, 101, 7618–7624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Li, J.; Green-Miller, A.R.; Shike, D.W. Integrity assessment of open-circuit respiration chambers for ruminant animal indirect calorimetry. Trans. Am. Soc. Agric. Biol. Eng. 2019, 62, 1185–1193. [Google Scholar] [CrossRef]
  57. Takahashi, J.; Chaudhry, A.; Beneke, R.; Young, B. An open-circuit hood system for gaseous exchange measurements in small ruminants. Small Rumin. Res. 1999, 32, 31–36. [Google Scholar] [CrossRef]
  58. Fernández Martínez, C.J.; López Luján, M.; Lachica, M. Low-cost mobile open-circuit hood system for measuring gas exchange in small ruminants: From manual to automatic recording. J. Agric. Sci. 2015, 153, 1302–1309. [Google Scholar] [CrossRef]
  59. Place, S.E.; Pan, Y.; Zhao, Y.; Mitloehner, F.M. Construction and operation of a ventilated hood system for measuring greenhouse gas and volatile organic compound emissions from cattle. Animals 2011, 1, 433–446. [Google Scholar] [CrossRef] [Green Version]
  60. Fernández, C.; Gomis-Tena, J.; Hernández, A.; Saiz, J. An open-circuit indirect calorimetry head hood system for measuring methane emission and energy metabolism in small ruminants. Animals 2019, 9, 380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Troy, S.; Rooke, J.; Duthie, C.; Ross, D.; Hyslop, J.; Roehe, R.; Waterhouse, T. Measurement of methane from finishing cattle fed either a forage-based or high concentrate diet from both feeder-mounted samplers and respiration chambers. In Proceedings of the Greenhouse Gases and Animal Agriculture Conference, Dublin, Ireland, 23–26 June 2013; Volume 4, p. 551. [Google Scholar]
  62. Hill, J.; McSweeney, C.; Wright, A.-D.G.; Bishop-Hurley, G.; Kalantar-Zadeh, K. Measuring methane production from ruminants. Trends Biotechnol. 2016, 34, 26–35. [Google Scholar] [CrossRef] [PubMed]
  63. Suzuki, T.; McCrabb, G.; Nishida, T.; Indramanee, S.; Kurihara, M. Construction and operation of ventilated hood-type respiration calorimeters for in vivo measurement of methane production and energy partition in ruminants. In Measuring Methane Production from Ruminants; Springer: Berlin/Heidelberg, Germany, 2007; pp. 125–135. [Google Scholar]
  64. Kebreab, E.; Clark, K.; Wagner-Riddle, C.; France, J. Methane and nitrous oxide emissions from canadian animal agriculture: A review. Can. J. Anim. Sci. 2006, 86, 135–157. [Google Scholar] [CrossRef]
  65. Boadi, D.; Wittenberg, K.; Kennedy, A. Validation of the sulphur hexafluoride (sf6) tracer gas technique for measurement of methane and carbon dioxide production by cattle. Can. J. Anim. Sci. 2002, 82, 125–131. [Google Scholar] [CrossRef]
  66. Deighton, M.H.; Williams, S.R.O.; Hannah, M.C.; Eckard, R.J.; Boland, T.M.; Wales, W.J.; Moate, P.J. A modified sulphur hexafluoride tracer technique enables accurate determination of enteric methane emissions from ruminants. Anim. Feed Sci. Technol. 2014, 197, 47–63. [Google Scholar] [CrossRef] [Green Version]
  67. Wright, A.; Kennedy, P.; O’Neill, C.; Toovey, A.; Popovski, S.; Rea, S.; Pimm, C.; Klein, L. Reducing methane emissions in sheep by immunization against rumen methanogens. Vaccine 2004, 22, 3976–3985. [Google Scholar] [CrossRef] [PubMed]
  68. Grainger, C.; Clarke, T.; McGinn, S.; Auldist, M.; Beauchemin, K.; Hannah, M.; Waghorn, G.; Clark, H.; Eckard, R. Methane emissions from dairy cows measured using the sulfur hexafluoride (sf6) tracer and chamber techniques. J. Dairy Sci. 2007, 90, 2755–2766. [Google Scholar] [CrossRef]
  69. Lassey, K.; Pinares-Patiño, C.; Martin, R.; Molano, G.; McMillan, A. Enteric methane emission rates determined by the sf6 tracer technique: Temporal patterns and averaging periods. Anim. Feed Sci. Technol. 2011, 166, 183–191. [Google Scholar] [CrossRef]
  70. Pinares-Patiño, C.; Clark, H. Reliability of the sulfur hexafluoride tracer technique for methane emission measurement from individual animals: An overview. Aust. J. Exp. Agric. 2008, 48, 223–229. [Google Scholar] [CrossRef]
  71. Pinares-Patiño, C.; Lassey, K.; Martin, R.; Molano, G.; Fernandez, M.; MacLean, S.; Sandoval, E.; Luo, D.; Clark, H. Assessment of the sulphur hexafluoride (sf6) tracer technique using respiration chambers for estimation of methane emissions from sheep. Anim. Feed Sci. Technol. 2011, 166, 201–209. [Google Scholar] [CrossRef]
  72. Van Well, B.; Murray, S.; Hodgkinson, J.; Pride, R.; Strzoda, R.; Gibson, G.; Padgett, M. An open-path, hand-held laser system for the detection of methane gas. J. Opt. A Pure Appl. Opt. 2005, 7, S420. [Google Scholar] [CrossRef] [Green Version]
  73. Detto, M.; Verfaillie, J.; Anderson, F.; Xu, L.; Baldocchi, D. Comparing laser-based open-and closed-path gas analyzers to measure methane fluxes using the eddy covariance method. Agric. For. Meteorol. 2011, 151, 1312–1324. [Google Scholar] [CrossRef]
  74. Laubach, J.; Kelliher, F.M. Methane emissions from dairy cows: Comparing open-path laser measurements to profile-based techniques. Agric. For. Meteorol. 2005, 135, 340–345. [Google Scholar] [CrossRef]
  75. Desjardins, R.; Denmead, O.; Harper, L.; McBain, M.; Massé, D.; Kaharabata, S. Evaluation of a micrometeorological mass balance method employing an open-path laser for measuring methane emissions. Atmos. Environ. 2004, 38, 6855–6866. [Google Scholar] [CrossRef]
  76. Gao, Z.; Desjardins, R.L.; Flesch, T.K. Assessment of the uncertainty of using an inverse-dispersion technique to measure methane emissions from animals in a barn and in a small pen. Atmos. Environ. 2010, 44, 3128–3134. [Google Scholar] [CrossRef]
  77. McGinn, S.; Turner, D.; Tomkins, N.; Charmley, E.; Bishop-Hurley, G.; Chen, D. Methane emissions from grazing cattle using point-source dispersion. J. Environ. Qual. 2011, 40, 22–27. [Google Scholar] [CrossRef]
  78. Charmley, E.; McSweeney, C.; Eady, S. Strategies for measuring and reducing methane emissions from beef cattle in northern australia. In Proceedings of the Northern Beef Research Update Conference, Darwin, Austrilia, 2 August 2011; pp. 73–80. [Google Scholar]
  79. Hammond, K.; Waghorn, G.; Hegarty, R. The greenfeed system for measurement of enteric methane emission from cattle. Anim. Prod. Sci. 2016, 56, 181–189. [Google Scholar] [CrossRef]
  80. C-Lock. Greenfeed Large Animals. Available online: https://www.c-lockinc.com/researchers/products/greenfeed-pasture-system (accessed on 23 May 2020).
  81. Huhtanen, P.; Ramin, M.; Hristov, A.N. Enteric methane emission can be reliably measured by the greenfeed monitoring unit. Livest. Sci. 2019, 222, 31–40. [Google Scholar] [CrossRef]
  82. Hristov, A.N.; Oh, J.; Giallongo, F.; Frederick, T.; Weeks, H.; Zimmerman, P.R.; Harper, M.T.; Hristova, R.A.; Zimmerman, R.S.; Branco, A.F. The use of an automated system (greenfeed) to monitor enteric methane and carbon dioxide emissions from ruminant animals. JoVE 2015, 103, e52904. [Google Scholar] [CrossRef] [Green Version]
  83. Waghorn, G.; Jonker, A.; Macdonald, K. Measuring methane from grazing dairy cows using greenfeed. Anim. Prod. Sci. 2016, 56, 252–257. [Google Scholar] [CrossRef]
  84. Hammond, K.; Humphries, D.; Crompton, L.; Green, C.; Reynolds, C.J.A.F.S. Methane emissions from cattle: Estimates from short-term measurements using a greenfeed system compared with measurements obtained using respiration chambers or sulphur hexafluoride tracer. Anim. Feed Sci. Technol. 2015, 203, 41–52. [Google Scholar] [CrossRef]
  85. Lind, V. Ruminants and Methane 2: 4 Methods and Techniques for Measuring Ghg Emissions from Ruminants; 2464-1170; Norsk Institutt for Bioøkonomi: Høgskoleveien, Norway, 2020; pp. 1–4. [Google Scholar]
  86. Goopy, J.; Robinson, D.L.; Woodgate, R.; Donaldson, A.; Oddy, V.; Vercoe, P.; Hegarty, R. Estimates of repeatability and heritability of methane production in sheep using portable accumulation chambers. Anim. Prod. Sci. 2016, 56, 116–122. [Google Scholar] [CrossRef]
  87. Robinson, D.L.; Goopy, J.P.; Hegarty, R.; Oddy, V. Comparison of repeated measurements of methane production in sheep over 5 years and a range of measurement protocols. J. Anim. Sci. 2015, 93, 4637–4650. [Google Scholar] [CrossRef] [PubMed]
  88. Goopy, J.P.; Woodgate, R.; Donaldson, A.; Robinson, D.L.; Hegarty, R. Validation of a short-term methane measurement using portable static chambers to estimate daily methane production in sheep. Anim. Feed Sci. Technol. 2011, 166, 219–226. [Google Scholar] [CrossRef]
  89. Robinson, D.; Dominik, S.; Donaldson, A.; Oddy, V. Repeatabilities, heritabilities and correlations of methane and feed intake of sheep in respiration and portable chambers. Anim. Prod. Sci. 2020, 60, 880–892. [Google Scholar] [CrossRef]
  90. Robinson, D.L.; Goopy, J.P.; Donaldson, A.; Woodgate, R.; Oddy, V.; Hegarty, R. Sire and liveweight affect feed intake and methane emissions of sheep confined in respiration chambers. Animal 2014, 8, 1935–1944. [Google Scholar] [CrossRef] [Green Version]
  91. Bell, M.J. Measuring enteric methane emissions from individual ruminant animals in their natural environment. In Greenhouse Gas Emissions; Springer: Berlin/Heidelberg, Germany, 2019; pp. 111–126. [Google Scholar]
  92. Jayanegara, A. Reducing Methane Emissions from Livestock: Nutritional Approaches. In Proceedings of the Indonesian Students Scientific Meeting (ISSM), Manchester, UK, 25–26 August 2011; pp. 18–21. [Google Scholar]
  93. Beauchemin, K.A.; Kreuzer, M.; O’Mara, F.; McAllister, T.A. Nutritional management for enteric methane abatement: A review. Aust. J. Exp. Agric. 2008, 48, 21–27. [Google Scholar] [CrossRef]
  94. Sejian, V.; Lal, R.; Lakritz, J.; Ezeji, T. Measurement and prediction of enteric methane emission. Int. J. Biometeorol. 2011, 55, 1–16. [Google Scholar] [CrossRef]
  95. Pesta, A.C. Dietary strategies for mitigation of methane production by growing and finishing cattle. Ph.D. Thesis, University of Nebraska, Lincoln, NE, USA, 2015. [Google Scholar]
  96. Haque, M. Dietary manipulation: A sustainable way to mitigate methane emissions from ruminants. J. Anim. Sci. Technol. 2018, 60, 15. [Google Scholar] [CrossRef] [Green Version]
  97. Nampoothiri, V.M.; Mohini, M.; Malla, B.A.; Mondal, G.; Pandita, S. Growth performance, and enteric and manure greenhouse gas emissions from murrah calves fed diets with different forage to concentrate ratios. Anim. Nutr. 2018, 4, 215–221. [Google Scholar] [CrossRef]
  98. Jiao, H.P.; Dale, A.J.; Carson, A.F.; Murray, S.; Gordon, A.W.; Ferris, C.P. Effect of concentrate feed level on methane emissions from grazing dairy cows. J. Dairy Sci. 2014, 97, 7043–7053. [Google Scholar] [CrossRef] [PubMed]
  99. Lovett, D.; Lovell, S.; Stack, L.; Callan, J.; Finlay, M.; Conolly, J.; O’Mara, F. Effect of forage/concentrate ratio and dietary coconut oil level on methane output and performance of finishing beef heifers. Livest. Prod. Sci. 2003, 84, 135–146. [Google Scholar] [CrossRef]
  100. Sabia, E.; Kühl, S.; Flach, L.; Lambertz, C.; Gauly, M. Effect of feed concentrate intake on the environmental impact of dairy cows in an alpine mountain region including soil carbon sequestration and effect on biodiversity. Sustainability 2020, 12, 2128. [Google Scholar] [CrossRef] [Green Version]
  101. Van Wyngaard, J.D.V.; Meeske, R.; Erasmus, L.J. Effect of concentrate level on enteric methane emissions, production performance, and rumen fermentation of jersey cows grazing kikuyu-dominant pasture during summer. J. Dairy Sci. 2018, 101, 9954–9966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Aguerre, M.J.; Wattiaux, M.A.; Powell, J.; Broderick, G.A.; Arndt, C. Effect of forage-to-concentrate ratio in dairy cow diets on emission of methane, carbon dioxide, and ammonia, lactation performance, and manure excretion. J. Dairy Sci. 2011, 94, 3081–3093. [Google Scholar] [CrossRef]
  103. Muñoz, C.; Hube, S.; Morales, J.M.; Yan, T.; Ungerfeld, E.M. Effects of concentrate supplementation on enteric methane emissions and milk production of grazing dairy cows. Livest. Sci. 2015, 175, 37–46. [Google Scholar] [CrossRef]
  104. Dohme, F.; Machmüller, A.; Wasserfallen, A.; Kreuzer, M. Comparative efficiency of various fats rich in medium-chain fatty acids to suppress ruminal methanogenesis as measured with rusitec. Can. J. Anim. Sci. 2000, 80, 473–484. [Google Scholar] [CrossRef]
  105. Grainger, C.; Beauchemin, K. Can enteric methane emissions from ruminants be lowered without lowering their production? Anim. Feed Sci. Technol. 2011, 166, 308–320. [Google Scholar] [CrossRef]
  106. Knapp, J.R.; Laur, G.L.; Vadas, P.A.; Weiss, W.P.; Tricarico, J.M. Invited review: Enteric methane in dairy cattle production: Quantifying the opportunities and impact of reducing emissions. J. Dairy Sci. 2014, 97, 3231–3261. [Google Scholar] [CrossRef] [Green Version]
  107. Machmüller, A.; Kreuzer, M. Methane suppression by coconut oil and associated effects on nutrient and energy balance in sheep. Can. J. Anim. Sci. 1999, 79, 65–72. [Google Scholar] [CrossRef] [Green Version]
  108. Pinares-Patiño, C.S.; Franco, F.E.; Molano, G.; Kjestrup, H.; Sandoval, E.; MacLean, S.; Battistotti, M.; Koolaard, J.; Laubach, J. Feed intake and methane emissions from cattle grazing pasture sprayed with canola oil. Livest. Sci. 2016, 184, 7–12. [Google Scholar] [CrossRef]
  109. Mao, H.-L.; Wang, J.-K.; Zhou, Y.-Y.; Liu, J.-X. Effects of addition of tea saponins and soybean oil on methane production, fermentation and microbial population in the rumen of growing lambs. Livest. Sci. 2010, 129, 56–62. [Google Scholar] [CrossRef]
  110. Chuntrakort, P.; Otsuka, M.; Hayashi, K.; Takenaka, A.; Udchachon, S.; Sommart, K. The effect of dietary coconut kernels, whole cottonseeds and sunflower seeds on the intake, digestibility and enteric methane emissions of zebu beef cattle fed rice straw based diets. Livest. Sci. 2014, 161, 80–89. [Google Scholar] [CrossRef]
  111. Machmüller, A.; Ossowski, D.; Kreuzer, M. Comparative evaluation of the effects of coconut oil, oilseeds and crystalline fat on methane release, digestion and energy balance in lambs. Anim. Feed Sci. Technol. 2000, 85, 41–60. [Google Scholar] [CrossRef]
  112. Cosgrove, G.; Waghorn, G.; Anderson, C.; Peters, J.; Smith, A.; Molano, G.; Deighton, M. The effect of oils fed to sheep on methane production and digestion of ryegrass pasture. Aust. J. Exp. Agric. 2008, 48, 189–192. [Google Scholar] [CrossRef]
  113. Hersom, M.; Thrift, T. Application of Ionophores in Cattle Diets; University of Florida: Gainesville, FL, USA, 2013; Volume 14. [Google Scholar]
  114. Valli, C. Mitigating enteric methane emission from livestock through farmer-friendly practices. In Global Climate Change and Environmental Policy; Springer: Berlin/Heidelberg, Germany, 2020; pp. 257–273. [Google Scholar]
  115. Perna Junior, F.; Vásquez, D.C.Z.; Gardinal, R.; Meyer, P.M.; Berndt, A.; Friguetto, R.T.S.; Demarchi, J.A.; Rodrigues, P.H.M. Short-term use of monensin and tannins as feed additives on digestibility and methanogenesis in cattle. Rev. Bras. Zootec. 2020, 49. [Google Scholar] [CrossRef]
  116. Guan, H.; Wittenberg, K.; Ominski, K.; Krause, D. Efficacy of ionophores in cattle diets for mitigation of enteric methane. J. Anim. Sci. Technol. 2006, 84, 1896–1906. [Google Scholar]
  117. Li, Z.; Ren, H.; Liu, S.; Cai, C.; Han, J.; Li, F.; Yao, J. Dynamics of methanogenesis, ruminal fermentation, and alfalfa degradation during adaptation to monensin supplementation in goats. J. Dairy Sci. 2018, 101, 1048–1059. [Google Scholar] [CrossRef] [PubMed]
  118. Gupta, S.; Mohini, M.; Malla, B.A.; Mondal, G.; Pandita, S. Effects of monensin feeding on performance, nutrient utilisation and enteric methane production in growing buffalo heifers. Trop. Anim. Health Prod. 2019, 51, 859–866. [Google Scholar] [CrossRef] [PubMed]
  119. Kunkle, W.; Johns, J.; Poore, M.; Herd, D. Designing supplementation programs for beef cattle fed forage-based diets. J. Anim. Sci. 2000, 77, 1–12. [Google Scholar] [CrossRef] [Green Version]
  120. Lascano, C.E.; Cárdenas, E. Alternatives for methane emission mitigation in livestock systems. Rev. Bras. Zootecn 2010, 39, 175–182. [Google Scholar] [CrossRef] [Green Version]
  121. Matsui, H.; Tajima, K.; Itabashi, H. Diversity of prokaryotes in the rumen of steers fed a diet supplemented with or without bromochloromethane, an anti-methanogenic compound. Jpn. Agric. Res. Q. JARQ 2020, 54, 179–183. [Google Scholar] [CrossRef]
  122. Lalu, K.; Bhar, R.; Das, A.; Mandal, A. Effect of bromochloromethane supplementation and dietary energy restriction on methane production and efficiency of energy utilization in rams. Indian J. Anim. Nutr. 2009, 26, 97–102. [Google Scholar]
  123. Romero-Perez, A.; Okine, E.K.; McGinn, S.M.; Guan, L.L.; Oba, M.; Duval, S.M.; Kindermann, M.; Beauchemin, K.A. Sustained reduction in methane production from long-term addition of 3-nitrooxypropanol to a beef cattle diet. J. Anim. Sci. 2015, 93, 1780–1791. [Google Scholar] [CrossRef] [PubMed]
  124. Lopes, J.; de Matos, L.; Harper, M.; Giallongo, F.; Oh, J.; Gruen, D.; Ono, S.; Kindermann, M.; Duval, S.; Hristov, A.N. Effect of 3-nitrooxypropanol on methane and hydrogen emissions, methane isotopic signature, and ruminal fermentation in dairy cows. J. Dairy Sci. 2016, 99, 5335–5344. [Google Scholar] [CrossRef] [Green Version]
  125. Martínez-Fernández, G.; Abecia, L.; Arco, A.; Cantalapiedra-Hijar, G.; Martín-García, A.I.; Molina-Alcaide, E.; Kindermann, M.; Duval, S.; Yáñez-Ruiz, D.R. Effects of ethyl-3-nitrooxy propionate and 3-nitrooxypropanol on ruminal fermentation, microbial abundance, and methane emissions in sheep. J. Dairy Sci. 2014, 97, 3790–3799. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Machado, L.; Magnusson, M.; Paul, N.A.; Kinley, R.; de Nys, R.; Tomkins, N. Identification of bioactives from the red seaweed asparagopsis taxiformis that promote antimethanogenic activity in vitro. J. Appl. Phycol. 2016, 28, 3117–3126. [Google Scholar] [CrossRef]
  127. Paul, N.A.; de Nys, R.; Steinberg, P. Chemical defence against bacteria in the red alga asparagopsis armata: Linking structure with function. Mar. Ecol. Prog. Ser. 2006, 306, 87–101. [Google Scholar] [CrossRef] [Green Version]
  128. Li, X.; Norman, H.C.; Kinley, R.D.; Laurence, M.; Wilmot, M.; Bender, H.; de Nys, R.; Tomkins, N. Asparagopsis taxiformis decreases enteric methane production from sheep. Anim. Prod. Sci. 2018, 58, 681–688. [Google Scholar] [CrossRef]
  129. Kinley, R.D.; de Nys, R.; Vucko, M.J.; Machado, L.; Tomkins, N. The red macroalgae asparagopsis taxiformis is a potent natural antimethanogenic that reduces methane production during in vitro fermentation with rumen fluid. Anim. Prod. Sci. 2016, 56, 282–289. [Google Scholar] [CrossRef]
  130. Roque, B.M.; Salwen, J.K.; Kinley, R.; Kebreab, E. Inclusion of asparagopsis armata in lactating dairy cows’ diet reduces enteric methane emission by over 50 percent. J. Clean. Prod. 2019, 234, 132–138. [Google Scholar] [CrossRef]
  131. Martínez, G.; Abecia, L.; Martín-García, A.; Ramos-Morales, E.; Molina-Alcaide, E.; Yáñez-Ruiz, D. Effect of antimethanogenic garlic-derived compounds on amylolytic and xylanolytic activities in the rumen. Options Méditerranéennes. Série A Séminaires Méditerranéens 2013, 107, 277–282. [Google Scholar]
  132. Klevenhusen, F.; Duval, S.; Zeitz, J.O.; Kreuzer, M.; Soliva, C.R. Diallyl disulphide and lovastatin: Effects on energy and protein utilisation in, as well as methane emission from, sheep. Arch. Anim. Nutr. 2011, 65, 255–266. [Google Scholar] [CrossRef] [PubMed]
  133. Bidarkar, V.K.; Swain, P.S.; Ray, S.; Dominic, G. Probiotics: Potential alternative to antibiotics in ruminant feeding. Trends Vet. Anim. Sci. 2014, 1, 1–4. [Google Scholar]
  134. Hassan, A.; Gado, H.; Anele, U.Y.; Berasain, M.A.; Salem, A.Z. Influence of dietary probiotic inclusion on growth performance, nutrient utilization, ruminal fermentation activities and methane production in growing lambs. Anim. Biotechnol. 2019, 31, 1–8. [Google Scholar] [CrossRef]
  135. Arowolo, M.A.; He, J. Use of probiotics and botanical extracts to improve ruminant production in the tropics: A review. Anim. Nutr. 2018, 4, 241–249. [Google Scholar] [CrossRef]
  136. Deng, K.D.; Xiao, Y.; Ma, T.; Tu, Y.; Diao, Q.Y.; Chen, Y.H.; Jiang, J.J. Ruminal fermentation, nutrient metabolism, and methane emissions of sheep in response to dietary supplementation with bacillus licheniformis. Anim. Feed Sci. Technol. 2018, 241, 38–44. [Google Scholar] [CrossRef]
  137. Latham, E.A.; Pinchak, W.E.; Trachsel, J.; Allen, H.K.; Callaway, T.R.; Nisbet, D.J.; Anderson, R.C. Paenibacillus 79r4, a potential rumen probiotic to enhance nitrite detoxification and methane mitigation in nitrate-treated ruminants. Sci. Total Environ. 2019, 671, 324–328. [Google Scholar] [CrossRef]
  138. Suryani, H.; Zain, M.; Ningrat, R.; Jamarun, N. Effect of dietary supplementation based on an ammoniated palm frond with direct fed microbials and virgin coconut oil on the growth performance and methane production of bali cattle. Pak. J. Nutr. 2017, 16, 599–604. [Google Scholar] [CrossRef] [Green Version]
  139. Chen, J.; Harstad, O.M.; McAllister, T.; Dörsch, P.; Holo, H. Propionic acid bacteria enhance ruminal feed degradation and reduce methane production in vitro. Acta Agric. Scand. Sect. A Anim. Sci. 2020, 69, 1–7. [Google Scholar] [CrossRef] [Green Version]
  140. De Raphelis-Soissan, V.; Li, L.; Godwin, I.; Barnett, M.; Perdok, H.; Hegarty, R. Use of nitrate and propionibacterium acidipropionici to reduce methane emissions and increase wool growth of merino sheep. Anim. Prod. Sci. 2014, 54, 1860–1866. [Google Scholar] [CrossRef]
  141. Cobellis, G.; Trabalza-Marinucci, M.; Yu, Z. Critical evaluation of essential oils as rumen modifiers in ruminant nutrition: A review. Sci. Total Environ. 2016, 545, 556–568. [Google Scholar] [CrossRef] [PubMed]
  142. Benchaar, C.; Chaves, A.; Fraser, G.; Beauchemin, K.; McAllister, T. Effects of essential oils and their components on in vitro rumen microbial fermentation. Can. J. Anim. Sci. 2007, 87, 413–419. [Google Scholar] [CrossRef]
  143. Benchaar, C.; Greathead, H. Essential oils and opportunities to mitigate enteric methane emissions from ruminants. Anim. Feed Sci. Technol. 2011, 166–167, 338–355. [Google Scholar] [CrossRef]
  144. Wu, P.; Liu, Z.B.; He, W.F.; Yu, S.B.; Gao, G.; Wang, J.K. Intermittent feeding of citrus essential oils as a potential strategy to decrease methane production by reducing microbial adaptation. J. Clean. Prod. 2018, 194, 704–713. [Google Scholar] [CrossRef]
  145. Hart, K.J.; Jones, H.G.; Waddams, K.E.; Worgan, H.J.; Zweifel, B.; Newbold, C.J. An essential oil blend decreases methane emissions and increases milk yield in dairy cows. Open J. Anim. Sci. 2019, 9, 259. [Google Scholar] [CrossRef] [Green Version]
  146. Laabouri, F.; Guerouali, A.; Alali, S.; Remmal, A.; Ajbilou, M. Effect of a natural food additive rich in thyme essential oil on methane emissions in dairy cows. Rev. Maroc. Des Sci. Agron. Et Vétérinaires 2017, 5, 287–292. [Google Scholar]
  147. Soltan, Y.A.; Natel, A.S.; Araujo, R.; Morsy, A.S.; Abdalla, A.L. Progressive adaptation of sheep to a microencapsulated blend of essential oils: Ruminal fermentation, methane emission, nutrient digestibility, and microbial protein synthesis. Anim. Feed Sci. Technol. 2018, 237, 8–18. [Google Scholar] [CrossRef]
  148. Sallama, S.; Nassera, M.; Araujoc, R.; Abdallab, A. Methane production by sheep consuming diets with different levels of eucalyptus essential oil. In Proceedings of the FAO/IAEA International Symposium on Sustainable Improvement of Animal Production and Health, Vienna, Austria, 8–11 June 2009; Available online: https://inis.iaea.org/search/search.aspx?orig_q=RN:41037857 (accessed on 18 June 2020).
  149. Sahoo, A.; Jena, B. Organic acids as rumen modifiers. Int. J. Sci. Res. 2014, 3, 2262–2266. [Google Scholar]
  150. Newbold, C.J.; Rode, L.M. Dietary additives to control methanogenesis in the rumen. Int. Congr. Ser. 2006, 1293, 138–147. [Google Scholar] [CrossRef]
  151. Bayaru, E.; Kanda, S.; Kamada, T.; Itabashi, H.; Andoh, S.; Nishida, T.; Ishida, M.; Itoh, T.; Nagara, K.; Isobe, Y. Effect of fumaric acid on methane production, rumen fermentation and digestibility of cattle fed roughage alone. Anim. Sci. J. 2001, 72, 139–146. [Google Scholar] [CrossRef] [Green Version]
  152. Wallace, R.J.; Wood, T.A.; Rowe, A.; Price, J.; Yanez, D.R.; Williams, S.P.; Newbold, C.J. Encapsulated fumaric acid as a means of decreasing ruminal methane emissions. Int. Congr. Ser. 2006, 1293, 148–151. [Google Scholar] [CrossRef]
  153. Ma, T.; Chen, D.; Tu, Y.; Zhang, N.; Si, B.; Deng, K.; Diao, Q. Effect of supplementation of allicin on methanogenesis and ruminal microbial flora in dorper crossbred ewes. J. Anim. Sci. Biotechnol. 2016, 7, 1. [Google Scholar] [CrossRef] [Green Version]
  154. Foley, P.A.; Kenny, D.A.; Callan, J.J.; Boland, T.M.; O’Mara, F.P. Effect of dl-malic acid supplementation on feed intake, methane emission, and rumen fermentation in beef cattle. J. Anim. Sci. 2009, 87, 1048–1057. [Google Scholar] [CrossRef]
  155. Jalč, D.; Čerešňáková, Z. Effect of plant oils and aspartate on rumen fermentation in vitro. J. Anim. Physiol. Anim. Nutr. 2001, 85, 378–384. [Google Scholar] [CrossRef]
  156. McAllister, T.; Hristov, A.; Beauchemin, K.; Rode, L.; Cheng, K. Enzymes in ruminant diets. In Enzymes in Farm Animal Nutrition; Bedford, M.R., Partridge, G.G., Eds.; CABI Publishing: Wallingford, UK, 2001; pp. 273–298. [Google Scholar]
  157. Zhou, M.; Chung, Y.H.; Beauchemin, K.; Holtshausen, L.; Oba, M.; McAllister, T.; Guan, L. Relationship between rumen methanogens and methane production in dairy cows fed diets supplemented with a feed enzyme additive. J. Appl. Microbiol. 2011, 111, 1148–1158. [Google Scholar] [CrossRef] [PubMed]
  158. Arriola, K.G.; Kim, S.C.; Staples, C.R.; Adesogan, A.T. Effect of fibrolytic enzyme application to low- and high-concentrate diets on the performance of lactating dairy cattle. J. Dairy Sci. 2011, 94, 832–841. [Google Scholar] [CrossRef] [PubMed]
  159. Zhao, L.; Peng, Y.; Wang, J.; Liu, J. Effects of exogenous fibrolytic enzyme on in vitro ruminal fiber digestion and methane production of corn stover and corn stover based mixed diets. Life Sci. J. 2015, 12, 1–9. [Google Scholar] [CrossRef]
  160. McGinn, S.M.; Beauchemin, K.A.; Coates, T.; Colombatto, D. Methane emissions from beef cattle: Effects of monensin, sunflower oil, enzymes, yeast, and fumaric acid. J. Anim. Sci. 2004, 82, 3346–3356. [Google Scholar] [CrossRef] [PubMed]
  161. Oh, J.; Harper, M.; Melgar, A.; Compart, D.M.P.; Hristov, A.N. Effects of saccharomyces cerevisiae-based direct-fed microbial and exogenous enzyme products on enteric methane emission and productivity in lactating dairy cows. J. Dairy Sci. 2019, 102, 6065–6075. [Google Scholar] [CrossRef] [PubMed]
  162. Lu, Q.; Jiao, J.; Tang, S.; He, Z.; Zhou, C.; Han, X.; Wang, M.; Kang, J.; Odongo, N.; Tan, Z. Effects of dietary cellulase and xylanase addition on digestion, rumen fermentation and methane emission in growing goats. Arch. Anim. Nutr. 2015, 69, 251–266. [Google Scholar] [CrossRef] [PubMed]
  163. Makkar, H.; Siddhuraju, P.; Becker, K. Preface. In Plant Secondary Metabolites; Humana Press: Totowa, NJ, USA, 2007; pp. i–xi. [Google Scholar]
  164. Rochfort, S.; Parker, A.J.; Dunshea, F.R. Plant bioactives for ruminant health and productivity. Phytochemistry 2008, 69, 299–322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Bodas, R.; Prieto, N.; García-González, R.; Andrés, S.; Giráldez, F.J.; López, S. Manipulation of rumen fermentation and methane production with plant secondary metabolites. Anim. Feed Sci. Technol. 2012, 176, 78–93. [Google Scholar] [CrossRef]
  166. Malik, P.; Uyeno, Y.; Kolte, A.; Kumar, R.; Trivedi, S.; Bhatta, R. Screening of phyto-sources from foothill of himalayan mountain for livestock methane reduction. Appl. Sci. 2019, 1, 232. [Google Scholar] [CrossRef]
  167. Baruah, L.; Malik, P.K.; Kolte, A.P.; Goyal, P.; Dhali, A.; Bhatta, R. Rumen methane amelioration in sheep using two selected tanniferous phyto-leaves. Carbon Manag. 2019, 10, 299–308. [Google Scholar] [CrossRef]
  168. Moreira, G.D.; Lima, P.d.M.T.; Borges, B.O.; Primavesi, O.; Longo, C.; McManus, C.; Abdalla, A.; Louvandini, H. Tropical tanniniferous legumes used as an option to mitigate sheep enteric methane emission. Trop. Anim. Health Prod. 2013, 45, 879–882. [Google Scholar] [CrossRef]
  169. Śliwiński, B.; Kreuzer, M.; Wettstein, H.-R.; Machmüller, A. Rumen fermentation and nitrogen balance of lambs fed diets containing plant extracts rich in tannins and saponins, and associated emissions of nitrogen and methane. Arch. Anim. Nutr. 2002, 56, 379–392. [Google Scholar] [CrossRef] [PubMed]
  170. Carulla, J.; Kreuzer, M.; Machmüller, A.; Hess, H. Supplementation of acacia mearnsii tannins decreases methanogenesis and urinary nitrogen in forage-fed sheep. Aust. J. Agric. Res. 2005, 56, 961–970. [Google Scholar] [CrossRef]
  171. Malik, P.K.; Kolte, A.P.; Bakshi, B.; Baruah, L.; Dhali, A.; Bhatta, R. Effect of tamarind seed husk supplementation on ruminal methanogenesis, methanogen diversity and fermentation characteristics. Carbon Manag. 2017, 8, 319–329. [Google Scholar] [CrossRef]
  172. Wann, C.; Wanapat, M.; Mapato, C.; Ampapon, T.; Huang, B.-z. Effect of bamboo grass (tiliacora triandra, diels) pellet supplementation on rumen fermentation characteristics and methane production in thai native beef cattle. Asian Australas. J. Anim. Sci. 2019, 32, 1153. [Google Scholar] [CrossRef]
  173. Liu, H.; Vaddella, V.; Zhou, D. Effects of chestnut tannins and coconut oil on growth performance, methane emission, ruminal fermentation, and microbial populations in sheep. J. Dairy Sci. 2011, 94, 6069–6077. [Google Scholar] [CrossRef] [PubMed]
  174. Albores-Moreno, S.; Alayón-Gamboa, J.; Ayala-Burgos, A.; Solorio-Sánchez, F.; Aguilar-Pérez, C.; Olivera-Castillo, L.; Ku-Vera, J. Effects of feeding ground pods of enterolobium cyclocarpum jacq. Griseb on dry matter intake, rumen fermentation, and enteric methane production by pelibuey sheep fed tropical grass. Trop. Anim. Health Prod. 2017, 49, 857–866. [Google Scholar] [CrossRef]
  175. Váradyová, Z.; Zeleňák, I.; Siroka, P. In vitro study of the rumen and hindgut fermentation of fibrous materials (meadow hay, beech sawdust, wheat straw) in sheep. Anim. Feed Sci. Technol. 2000, 83, 127–138. [Google Scholar] [CrossRef]
  176. Lovett, D.K.; Stack, L.J.; Lovell, S.; Callan, J.; Flynn, B.; Hawkins, M.; O’Mara, F.P. Manipulating enteric methane emissions and animal performance of late-lactation dairy cows through concentrate supplementation at pasture. J. Dairy Sci. 2005, 88, 2836–2842. [Google Scholar] [CrossRef]
  177. Dall-Orsoletta, A.C.; Oziemblowski, M.M.; Berndt, A.; Ribeiro-Filho, H.M.N. Enteric methane emission from grazing dairy cows receiving corn silage or ground corn supplementation. Anim. Feed Sci. Technol. 2019, 253, 65–73. [Google Scholar] [CrossRef]
  178. Ferris, C.P.; Jiao, H.; Murray, S.; Gordon, A.; Laidlaw, S. Effect of dairy cow genotype and concentrate feed level on cow performance and enteric methane emissions during grazing. Agric. Food Sci. 2020, 29, 130–138. [Google Scholar] [CrossRef]
  179. Wang, C.; Zhao, Y.; Aubry, A.; Arnott, G.; Hou, F.; Yan, T. Effects of concentrate input on nutrient utilization and methane emissions of two breeds of ewe lambs fed fresh ryegrass. Transl. Anim. Sci. 2019, 3, 485–492. [Google Scholar] [CrossRef] [Green Version]
  180. Hynes, D.N.; Stergiadis, S.; Gordon, A.; Yan, T. Effects of concentrate crude protein content on nutrient digestibility, energy utilization, and methane emissions in lactating dairy cows fed fresh-cut perennial grass. J. Dairy Sci. 2016, 99, 8858–8866. [Google Scholar] [CrossRef] [Green Version]
  181. Talukdar, P.; Kundu, S.S.; Mondal, G. Quantification of methane emissions from murrah buffaloes fed different energy diets during various temperature humidity index periods in a tropical environment. Anim. Prod. Sci. 2019, 59, 169–176. [Google Scholar] [CrossRef]
  182. Puchala, R.; LeShure, S.; Gipson, T.A.; Tesfai, K.; Flythe, M.D.; Goetsch, A.L. Effects of different levels of lespedeza and supplementation with monensin, coconut oil, or soybean oil on ruminal methane emission by mature boer goat wethers after different lengths of feeding. J. Appl. Anim. Res. 2018, 46, 1127–1136. [Google Scholar] [CrossRef] [Green Version]
  183. Dong, N.T.K.; Van Thu, N. Dietary supplementation of coconut oil markedly suppressed enteric methane production without compromising growth performance in bach thao goats. In Strengthening Development of Dairy Goat Production Adapting to Climate Change, Proceedings of the 4th Asian-Australasian Dairy Goat Conference, Tra Vinh, Viet Nam, 17–19 October 2018; Van Thu, N., Liang, J.B., Eds.; Abstract Number 54; Available online: https://www.researchgate.net/profile/Aliah-Mohsin/publication/333773806_4thAADC_Proceedings_TVU_2018/links/5d03284e92851c874c65066b/4thAADC-Proceedings-TVU-2018.pdf#page=404 (accessed on 22 March 2020).
  184. Lima, P.R.; Apdini, T.; Freire, A.S.; Santana, A.S.; Moura, L.M.L.; Nascimento, J.C.S.; Rodrigues, R.T.S.; Dijkstra, J.; Garcez Neto, A.F.; Queiroz, M.A.Á.; et al. Dietary supplementation with tannin and soybean oil on intake, digestibility, feeding behavior, ruminal protozoa and methane emission in sheep. Anim. Feed Sci. Technol. 2019, 249, 10–17. [Google Scholar] [CrossRef]
  185. Zhang, X.M.; Medrano, R.F.; Wang, M.; Beauchemin, K.A.; Ma, Z.Y.; Wang, R.; Wen, J.N.; Lukuyu, B.A.; Tan, Z.L.; He, J.H. Corn oil supplementation enhances hydrogen use for biohydrogenation, inhibits methanogenesis, and alters fermentation pathways and the microbial community in the rumen of goats. J. Anim. Sci. 2019, 97, 4999–5008. [Google Scholar] [CrossRef] [PubMed]
  186. Martin, C.; Coppa, M.; Fougère, H.; Bougouin, A.; Baumont, R.; Eugène, M.; Bernard, L. Diets supplemented with corn oil and wheat starch, marine algae, or hydrogenated palm oil modulate methane emissions similarly in dairy goats and cows, but not feeding behavior. Anim. Feed Sci. Technol. 2021, 272, 114783. [Google Scholar] [CrossRef]
  187. Pickett, A.T. Effects Of Lasalocid and Energy Supplementation on Forage Intake, Energy Metabolism, and Performance of Cattle Grazing Wheat Pasture; University of Arkansas: Fayetteville, AR, USA, 2020. [Google Scholar]
  188. Mwenya, B.; Sar, C.; Santoso, B.; Kobayashi, T.; Morikawa, R.; Takaura, K.; Umetsu, K.; Kogawa, S.; Kimura, K.; Mizukoshi, H.; et al. Comparing the effects of β1-4 galacto-oligosaccharides and l-cysteine to monensin on energy and nitrogen utilization in steers fed a very high concentrate diet. Anim. Feed Sci. Technol. 2005, 118, 19–30. [Google Scholar] [CrossRef]
  189. Mwenya, B.; Santoso, B.; Sar, C.; Gamo, Y.; Kobayashi, T.; Arai, I.; Takahashi, J. Effects of including β1–4 galacto-oligosaccharides, lactic acid bacteria or yeast culture on methanogenesis as well as energy and nitrogen metabolism in sheep. Anim. Feed Sci. Technol. 2004, 115, 313–326. [Google Scholar] [CrossRef]
  190. Fernández, C.; Pérez-Baena, I.; Marti, J.V.; Palomares, J.L.; Jorro-Ripoll, J.; Segarra, J.V. Use of orange leaves as a replacement for alfalfa in energy and nitrogen partitioning, methane emissions and milk performance of murciano-granadina goats. Anim. Feed Sci. Technol. 2019, 247, 103–111. [Google Scholar] [CrossRef]
  191. Wood, T.A.; Wallace, R.J.; Rowe, A.; Price, J.; Yáñez-Ruiz, D.R.; Murray, P.; Newbold, C.J. Encapsulated fumaric acid as a feed ingredient to decrease ruminal methane emissions. Anim. Feed Sci. Technol. 2009, 152, 62–71. [Google Scholar] [CrossRef]
  192. Wang, L.; Xue, B. Effects of cellulase supplementation on nutrient digestibility, energy utilization and methane emission by boer crossbred goats. Asian Australas. J. Anim. Sci. 2016, 29, 204–210. [Google Scholar] [CrossRef] [Green Version]
  193. Soltan, Y.A.; Morsy, A.S.; Sallam, S.M.; Lucas, R.C.; Louvandini, H.; Kreuzer, M.; Abdalla, A.L. Contribution of condensed tannins and mimosine to the methane mitigation caused by feeding leucaena leucocephala. Arch. Anim. Nutr. 2013, 67, 169–184. [Google Scholar] [CrossRef]
  194. Moate, P.J.; Jacobs, J.L.; Hixson, J.L.; Deighton, M.H.; Hannah, M.C.; Morris, G.L.; Ribaux, B.E.; Wales, W.J.; Williams, S.R.O. Effects of feeding either red or white grape marc on milk production and methane emissions from early-lactation dairy cows. Animals 2020, 10, 976. [Google Scholar] [CrossRef]
  195. Ramírez-Restrepo, C.A.; Barry, T.N.; Marriner, A.; López-Villalobos, N.; McWilliam, E.L.; Lassey, K.R.; Clark, H. Effects of grazing willow fodder blocks upon methane production and blood composition in young sheep. Anim. Feed Sci. Technol. 2010, 155, 33–43. [Google Scholar] [CrossRef]
Figure 1. The proportion of enteric methane in the total methane emissions arising from different species; data adapted from Grossi et al. [38], FAO [39].
Figure 1. The proportion of enteric methane in the total methane emissions arising from different species; data adapted from Grossi et al. [38], FAO [39].
Sustainability 13 06081 g001
Figure 2. Distribution of enteric global methane emissions by species (CO2 equivalents); data adapted from FAOSTAT [36].
Figure 2. Distribution of enteric global methane emissions by species (CO2 equivalents); data adapted from FAOSTAT [36].
Sustainability 13 06081 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Prathap, P.; Chauhan, S.S.; Leury, B.J.; Cottrell, J.J.; Dunshea, F.R. Towards Sustainable Livestock Production: Estimation of Methane Emissions and Dietary Interventions for Mitigation. Sustainability 2021, 13, 6081. https://doi.org/10.3390/su13116081

AMA Style

Prathap P, Chauhan SS, Leury BJ, Cottrell JJ, Dunshea FR. Towards Sustainable Livestock Production: Estimation of Methane Emissions and Dietary Interventions for Mitigation. Sustainability. 2021; 13(11):6081. https://doi.org/10.3390/su13116081

Chicago/Turabian Style

Prathap, Pragna, Surinder Singh Chauhan, Brian Joseph Leury, Jeremy James Cottrell, and Frank Rowland Dunshea. 2021. "Towards Sustainable Livestock Production: Estimation of Methane Emissions and Dietary Interventions for Mitigation" Sustainability 13, no. 11: 6081. https://doi.org/10.3390/su13116081

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