Next Article in Journal
Social Acceptance of a Thermal Architectural Implementation Proposal
Next Article in Special Issue
Effects of Different Materials on Biogas Production during Anaerobic Digestion of Food Waste
Previous Article in Journal
Melting the Psychological Boundary: How Interactive and Sensory Affordance Influence Users’ Adoption of Digital Heritage Service
Previous Article in Special Issue
Upcycled Animal Feed: Sustainable Solution to Orange Peels Waste
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Sustainable Strategies for Greenhouse Gas Emission Reduction in Small Ruminants Farming

by
Elisavet Giamouri
1,*,
Foivos Zisis
1,
Christina Mitsiopoulou
1,
Christos Christodoulou
1,
Athanasios C. Pappas
1,
Panagiotis E. Simitzis
2,
Charalampos Kamilaris
3,
Fenia Galliou
4,
Thrassyvoulos Manios
4,
Alexandros Mavrommatis
1 and
Eleni Tsiplakou
1,*
1
Laboratory of Nutritional Physiology and Feeding, Department of Animal Science, Agricultural University of Athens, Iera Odos 75, 11855 Athens, Greece
2
Laboratory of Animal Breeding & Husbandry, Department of Animal Science, Agricultural University of Athens, Iera Odos 75, 11855 Athens, Greece
3
Centre for Innovation Excellence in Livestock, Innovation Centre, York Science Park, Innovation Way, Heslington, York YO10 5DG, UK
4
Department of Agriculture, Hellenic Mediterranean University, 71410 Heraklion, Greece
*
Authors to whom correspondence should be addressed.
Sustainability 2023, 15(5), 4118; https://doi.org/10.3390/su15054118
Submission received: 26 December 2022 / Revised: 15 February 2023 / Accepted: 16 February 2023 / Published: 24 February 2023
(This article belongs to the Special Issue RETASTE: Rethink Food Waste)

Abstract

:
During the previous decades, the growing demand for animal origin products has gained considerable attention. As a result, livestock breeding has faced a rapid intensification in order to fulfil market expectations. This increase in livestock production has led to a large scale of manure that is associated with many environmental impacts, such as climate change, to an increase of greenhouse gases (GHG) emissions. Livestock production is considered to generate significant amounts of GHG, mainly carbon dioxide (CO2), methane (CH4) and nitrous oxide (N2O). Methane and nitrous oxide are the main emissions from livestock systems. Ruminants contribute highly to total livestock emissions. In the present study, the contribution of livestock and especially of the small ruminants in GHG emissions is reviewed. Additionally, useful sustainable strategies for farming and feeding of small ruminants are highlighted. Some of the practices discussed include but are not limited to efficient manure management, the replacement of mineral fertilizers by farm manure, the improvement of feed efficiency and provision of feed supplements. Moreover, the use of food waste or agro-industrial by-products is discussed as a sustainable strategy.

1. Introduction

In 2022 around 8 billion people are estimated to live on Earth, while in 2050 this number is expected to be 10.4 billion [1]. Nutritional demands are going to rise in order to satisfy human needs. Meanwhile, the increasing human population in accordance with urbanization will alter the dietary habits to more processed rather than natural origin products. As a result, a large quantity of animal origin products will be necessary, putting pressure on the food market [2]. Until 2050, around two times the amount of milk and meat that is produced nowadays will be necessary to cover human needs; for instance, 1077 million tons of milk should be produced compared to the 580 tons that are produced today [3]. In order to meet this demand, livestock farming grows constantly. However, there are many concerns about the environmental impact of this intensified animal production [4]. The generation of greenhouse gas emissions (GHG) is the most important environmental impact as far as livestock farming is concerned. As the emissions from livestock increase, many changes concerning the atmosphere, the land and the oceans are occurring that lead to climate change.
Agriculture, and therefore livestock, is concurrently one of the main contributors to GHG emissions and consequently to climate change [5]. The significant amounts of GHG emissions are produced either through their physiological processes or during the production of animal origin products. The livestock sector is responsible for 14.5% of total anthropogenic GHG emissions, or 7.1 Gt CO2e [6]. Of this total, the share of feed production and processing is approximately 45% or 3.2 Gt CO2e, the share of enteric fermentation is approximately 39% or 2.8 Gt CO2e, and the share of manure management is approximately 10% or 0.71 Gt. The processing and transportation of animal products account for the remaining 6% or 0.42 Gt CO2e. [6]. On the other hand, by accounting only for the direct CH4 and N2O emissions from enteric fermentation, manure management and distribution, the contribution is estimated to be 5.4 Gt CO2e [6,7]. In EU, a report by Perez Dominguez et al. [8] indicates that agriculture accounts for 10.3% of total GHG emissions, with CH4 from enteric fermentation accounting for 32% and manure management contributing another 16%. Cattle production is the largest contributor to global livestock emissions, contributing 4.6Gt CO2e or 61%, while other species contribute much less: pigs 0.7 Gt CO2e (9%), poultry 0.7 Gt CO2e (8%) and small ruminants 0.474 Gt CO2e (6%) [6,7].
An adequate level of reduction in GHG emissions by 2050 can be achieved, with selected strategies including but not limited to policies, infrastructures and technology in order to change human habits and behavior [9]. Agriculture could mitigate related emissions. The era of climate denial is over, human being is the main contributor of climate change and livestock farming could help reduce its footprint. Mitigation strategies need to be adopted in the livestock sector to reduce the environmental impact.
This review aimed to demonstrate the contribution of livestock and especially small ruminants in the climate change through its GHG emissions. It should be mentioned that for research purposes cattle literature will be used when appropriate in the following parts, since GHG mitigation strategies for small ruminants are sparse. Moreover, mitigation techniques that would provide a sustainable approach for the emissions’ reduction concerning feed, energy and manure are explained. Additionally, the inclusion of food waste and agro-industrial by-products in ruminant diets are discussed as a circular bioeconomy leverage which can improve livestock sustainability.

2. Contribution of Small Ruminants to GHG

Small ruminants’ contribution to GHG emissions are about 475 million tons CO2e, which represents about 6.5% of the agriculture sector global emissions [7]. Sheep and goat world meat and milk production accounts for around 254 and 175 million tons CO2e, respectively [7]. Small ruminants’ milk production contributes about 12% of the total GHG emissions that come from CH4 from enteric fermentation and manure and 19% from the N2O of manure management [10]. The global average of GHG emission intensity of milk is lower for goats than for sheep with 5.2 and 8.4 kg CO2e/kg product, respectively, mainly because goats have higher milk yields on average at the global level [7]. The GHG emission intensity of meat is very similar between the two species at about 23 kg CO2e/kg meat [7].
Different geographical regions and production systems have different carbon footprints [7]. The term “carbon footprint” (CF) refers to the sum of greenhouse gas emissions and removals in a product system related to the climate change and associated with any human activity (e.g., livestock, domestic energy combustion, etc.) [11]. In extensive systems, the CF of goat milk is high while in semi-intensive and intensive systems, emissions per kg of milk are low [12,13,14] because of the high productivity and the high amount of compound feed used in the rations [13]. In the extensive, semi-intensive and intensive farms, methane from milk production accounts for 75%, 65% and 52% of total emissions, respectively [13]. More specifically, emission intensities for sheep and goat milk production ranged from 1.6 kg to 14.2 kg of CO2e (in humid grassland areas of Western Europe and arid grassland areas of North Africa, respectively), while emission intensities for lamb and goat meat production ranged from 7.4 kg to 57.5 kg of CO2e (in mixed areas of Western Europe and temperate grassland areas of North Africa, respectively) [7]. Nevertheless, to accurately develop mitigation options for the small ruminant farms, it is important to detect all potential sources of GHG emissions. The three ‘greenhouse gases’ (GHG) which contribute most to global warming are carbon dioxide (CO2), methane (CH4) and nitrous oxide (N2O) [15]. The contribution of small ruminants is presented in Figure 1.

2.1. Sources of Methane (CH4) Emissions

Methane is the most important greenhouse gas generated from the animal sector [17]. In ruminants, methane (CH4) is mainly produced by enteric fermentation and manure storage [18]. More than 55 percent of emissions from small ruminant meat and milk production originate from enteric fermentation and the vast majority is formed in the rumen by methanogenic microbes. Therefore, any mitigation strategy that reduces methanogen populations and consequently methane emissions should not be limited to it but also include an alternative pathway for H2 neutralization since its accumulation can impair rumen function [19]. In ruminants, methane emissions represent a loss of about 2 to 15% of dietary gross energy [20]. Sheep and goats produce 10 to 16 kg CH4/year and cattle 60 to 160 kg/year, depending on their size and dry matter intake (DMI) [21] and, especially in Greece, CH4 represents the main GHG from agriculture, with a contribution ranging from 48% to 58% [22].

2.2. Sources of Nitrous oxide (N2O) Emissions

Slightly more than 35 percent of N2O emissions derive from feed production, whereas manure emissions are lower because manure is mainly deposited onto pasture [6]. Crop activities in small ruminants’ farms involve forage and grain production for livestock feeding but also crop production for sale. Direct N2O emissions from manure storage are lower compared with CH4 emissions. Direct N2O emissions occur from when manure is treated aerobically; at that time, organic N is converted to NO3- and NO2- through nitrification and then manure is treated aerobically where N2O and NO are produced through denitrification [23]. The N2O emissions from manure management account for 20%, 25%, and 34% of total emissions for the extensive, semi-intensive, and intensive dairy goat farming systems, respectively [13].

2.3. Sources of Carbon Dioxide (CO2) Emissions

Carbon dioxide emissions are mainly generated from anaerobic degradation of manure, producing biogas [24] and combustion of fuel [25]. Carbon dioxide is linked to the energy combustion of mixed crop and livestock farms. This energy use is related to fuel use (mainly diesel) and electricity [26], which is attributed to the milking machinery, the transportation, the ventilation equipment used in the animal houses, and the whole crop production use if the feed is produced on site. The emissions from this combustion can be accounted for by estimating the requirements for energy use, fuel, electricity, and type of machinery used for every farm operation, multiplied by the appropriate emission factors.

3. Sustainable Strategies for Mitigation of GHG Emissions

3.1. Feed Related Strategies

Methane represents a significant loss of dietary energy; thus, the aim is to reduce its production favoring both environment and feed conservation rate. Therefore, the key role on the mitigation practices for small ruminants is the manipulation of the methanogenesis route. Although there are several options for reducing methane, it has to be balanced so that an economically sustainable way for producers, together with the ability to improve the production efficiency, are supported simultaneously.

3.1.1. Forage Quality, Growth Stage, and Legumes Grazing

By improving pasture quality, the animal productivity may be enhanced and lower CH4 will be emitted, with less dietary energy loss [27]. A mitigation option is the harvesting of forages at an earlier growth stage, in order to increase the soluble carbohydrates and decrease the lignification of cell walls (less NDF) [28]. This also would increase its digestibility, leading to a reduction in GHG emissions. A methane reduction is reported with the replacement of fibrous forage with unlignified immature grasses and early cut forage at a level of 15%, while for processed forages, the reduction was 21%, while for grass forage and good quality silage the reduction was 28% and 20%, respectively [29]. The overall objective is to combine the maximum yield of dry matter per hectare with the highest digestibility. For example, by shifting from higher cell wall content (C4) to lower (C3) grasses or even grazing on less mature pastures, mitigation of CH4 production can be achieved [27]. Archimède et al. [30] recommended the use of C3 legumes in warm climates as a mitigation option, as animals fed warm climate legumes produced 20% less CH4 than animals fed C4 grasses. Consumption of several levels of legumes by ruminants results in lower CH4 yield than that associated with the consumption of grasses [27], due to the tannins content of some legumes [31]. In sheep fed diets with legumes such as white clover, Lotus pedunculatus, the outcome was 16% lower CH4 production compared to sheep fed ryegrass [32]. At the same time, ryegrass cut at 21 d revealed ameliorated nutrient intake than that cut at 35 d, as well as increased microbial crude protein formation with silage feeding than with hay fed lambs [33]. Moreover, as reported by Gonzalez-Ronquillo et al. [33], the ensiling of ryegrass at 21 d had no negative effects in the nutrient intake, N balance, and sustainability of locally produced feed which could be an overall aim. Nevertheless, there are important benefits of legumes such as the replacement of mineral N fertiliser [34]. Further, the supplementation of 300 g/kg DM of rapeseed forage harvested at pod stage, presented assuring results as functioning as an alternative protein source for alfalfa hay, without affecting nutrient intake and digestibility [35]. Overall, the inclusion of less lignified forages into ruminants’ diets appears to be an efficient strategy to increase organic matter digestibility and feed efficiency, therefore reducing the overall carbon footprint of meat and milk production while the optimum rumen conditions are preserved. However, further holistic studies should evaluate if the collection of forages at an earlier growth stage could negatively affect other socio-economic balances such as feed-food or inter-species feed competition.

3.1.2. Elevated Proportion of Concentrate Feed in Diets

The inclusion of concentrate feed at elevated rates of 35% to 40% of DMI in ruminants’ diets may conceivably decrease the enteric CH4, although it unambiguously depends on other factors such as the type of grain, etc. [19]. By including grain, the starch is raised while the NDF (Neutral Detergent Fiber) intake is lessened, a fact that subsequently reduces rumen pH and advances propionate over acetate production in the rumen [36]. Lovett et al. [37] indicated that the emissions of CH4, N2O, and CO2 were reduced by 9.5%, 16%, and 5%, respectively, by increasing the feeding of concentrates in dairy cows. The inclusion of concentrate feeds will enhance animal productivity and a shrink in GHG emissions intensity will occur, but the absolute GHG emissions may not always decrease as a result of intense feed production [7]. Additionally, special attention should be paid to forage to concentrate ratio since the overload of rapidly fermentable carbohydrate reaching the rumen can impair its degradative potential and therefore animal performance.

3.1.3. Inclusion of Feed Supplements in Diets Manipulating Rumen Methanogenesis

Dietary Lipids

Dietary lipids’ supplementation reduces CH4 emissions [38] by interfering in rumen function. The effectiveness depends on the influence on animals’ productivity and the cost of oil products. Researchers in Australia observed a 12% CH4 reduction in dairy cattle [39] and a 42% CH4 reduction in dairy cows [40] following the inclusion of cottonseed meal. For adding 1% lipids (limited to 6% supplementation level), the decrease of CH4 was 3.5–5.4%, as measured in cattle and sheep [27,41,42]. Specifically, Grainger and Beauchemin [43] demonstrated that addition 10 g/kg of dietary fat in the sheep diet would decrease CH4 emissions for meat production by 2.6 g/kg. Dietary lipids can be found in a plethora of feeds such as coconut oil and whole crushed oil seeds (rapeseed, sunflower seed, etc.) [44]. Moreover, Vargas et al. [45] added 6% of olive and linseed oil in rumen’ diets and managed to reduce methane production by 21 to 28%. In recent studies, supplementing long chain fatty acids in the goat diet decreased methanogens population on both liquid and solid fraction in the rumen [46,47] unveiling promising scenarios for methane suppression. On the contrary, Cosgrove et al. [48] reported no mitigation result of lipid supplementation (in 0, 1.2, 2.5, 3.8, and 5%) to the sheep diet. The quantity of inclusion for avoiding negative effects in ruminants was kept between 4% and 8% of DMI [49]. However, it has been stated that it is the fatty acid profile of the fat source, rather than the fat level per se, that impairs rumen methanogenesis [50]. In this light, a plethora of scientific questions remain open such as, is it the absolute number of methanogens that correlated with methane formation or is it the microbiome structure [51]. Additionally, it must be highlighted if specific fatty acids affect specific rumen “residents” per se or if the disruption of their biofilm is compromising their structure or their metabolism. Understanding the regulatory mechanisms of specific fatty acids in rumen methane formation will shape the key manipulation strategy for ruminants GHG mitigation. On the other hand, it has been observed that fat mode of action in the rumen microbiome is not methanogenic-specific; thus, other microorganisms involved in feed degradation and volatile fatty acid production could be harmed as well. The deeper exploring of the inhibitory specificity of certain fatty acids against methanogenic species or the synergetic action of fatty acids and other feed additives and bioactive compounds aiming to target manipulate rumen populations can facilitate the on-farm scale implementation of such strategies.

Electron Receptors

In rumen, archaea produce CH4 during fermentation, mainly by using H2 and CO2 as substances [49]. Nitrates are used as electron acceptors, replacing CO2, so that NH3 is produced instead of CH4 [36], which also negatively affect soil and water. Further, nitrates have been shown to reduce CH4 production in sheep [52], with the actual level of reduction being 23% in KNO3-supplemented sheep [53], 16% with nitrate supplementation in dairy cows [54], and 17% after supplementing forage-based diets in finishing beef cattle [55]. However, nitrate must be supplemented with caution as it can be toxic above certain doses [21]. Other electron receptors such as fumaric or myristic acid have been tested and reviewed, showing reductions of CH4 [36]. In cattle, inclusion of 2% fumaric acid in silage reduced CH4 by 23% [56]. Wood et al. [57] provided 100 g kg−1 fumaric acid in a free form to lambs and reported a 62% reduction in CH4 output while in an encapsulated form found a 76% reduction. Moreover, in sheep, myristic acid (50 mg kg−1 DM) reduced CH4 by 22% when fed in a forage-based diet and 58% in a concentrate-based diet [58]. However, due to the high doses required, dicarboxylic acids are too expensive to be used widely [59].

Microorganisms

This category contains mainly suggestions based on the activity of yeast and some bacteria, widely known as acetogens. Previous research conducted by McGinn et al. [60] using two strains of yeast (Levucell SC and Procreatin-7) in beef cattle demonstrated no effects on CH4 production. Conversely, yeast products, based on Saccharomyces cerevisiae, can decrease CH4 by promoting rumen fermentation [61]. Similarly, Grainger and Beauchemin [44] reduced CH4 emissions by 7% but negative effects on rumen pH followed. Moreover, another effective way could be the utilization of bacteriophages [62], based on their ability to specifically target methanogenic archaea and surpass host’s cells, which has already shown promising results for methane mitigation [63]. However, more emphasis on the metagenomics part of the phages’ identification is needed.

Antibodies and Vaccines

Due to the rising concern over the use of antibiotics, new areas of interest have been researched. Subharat et al. [64] examined the prospect of immunization and production of antibodies over methanogens. A vaccine targeting the cell surface or membrane proteins of methanogens may establish vaccination as an effective CH4 mitigation practice [36,65], by diminishing the methanogen population or its activity in rumen. In the same context, Wright et al. [66] managed to decrease CH4 production in Australian sheep using a vaccine over specific methanogens, but did not have the expected results with different microbial species or rearing conditions [36]. Nevertheless, this approach is based on the possible cost neutrality of the action but needs further research regarding the complexity of the target species.

Enzymes

In a meta-analysis on the effects of exogenous fibrolytic enzymes in ruminant diets, Tirado-González et al. [67] reported that enzymes may improve the productive performance of dairy cows and beef cattle, but the response depends upon the proper ratio of cellulases to xylanases according to the diet’s composition. The addition of cellulases and hemicellulases in ruminants’ diets improved fiber digestion and productivity [27,68] and, thus, reduced in vivo CH4 production by 28% and 9%, respectively [27].

Plant Bioactive Compounds

This category includes a variety of plant secondary compounds that are used as feed supplements, such as tannins, saponins, essential oils, plant extracts and their ingredients, that have exert anti-protozoal effects and decrease the enteric emissions [69]. Tannins have a great variety of type, concentration, and protein-binding capacity as well as technique accounting for the plants’ concentration [70], so the effects differ among studies. For instance, the dietary supplementation of chestnut tannins, examined in vitro, portrayed beneficial results for methane and ammonia emissions without affecting acetate production and thus inducing a promising solution for reducing them [71]. Moreover, low CH4 yield has been shown on pastures rich in condensed tannins [32]. Furthermore, Carulla et al. [72] managed to decrease CH4 emission by 13% with a partial replacement of tannins from Acacia mearnsii included in the sheep diet. Similarly, Tiemann et al. [73] indicated that 25 g/kg DM of tannins of Calliandra calothyrsus and Flemingia macrophylla reduced CH4 yield by 13%. Further, for Angora goats fed with Lespedeza cuneata containing condensed tannins, CH4 production was reduced compared to goats fed a combination of Digitaria ischaemum and Festuca arundinacea [74]. Likewise, the same anti-protozoal properties can be ascertained for plant saponins [27], flaunting some vital potential, but still more and extended studies are required before they could be recommended for use [21]. The same potential was presented by in vitro research of green tea ethanolic extracts to lessen methanogenic processes by using four doses (250, 500, 750, and 1000 μg/mL) which concluded in supplementing the lowest dose as better adapted [75]. However, considering methane mitigation through essential oils, Cobellis et al. [76] pointed out by summarizing the effect of essential oils as rumen modifiers that the in vivo reduction of methane yield seems to be fictitious since there is ample evidence that indicates an overall degradative suppression in the rumen due to their action.
Although essential oils and other bioactive compound derivatives of plant materials have been extensively investigated as rumen methane mitigation agents, their on-farm scale implementation, accompanied by crucial limitations such as the amount that the livestock sector needed, the financial sustainability of its inclusion in animal diets, the food-feed and nutraceutical industry competition, etc.

Chemical Inhibitors

Halogenated analogues, such as bromochloromethane (BCM), 2-bromo-ethane sulphonate (BES), and chloroform are potent inhibitors of CH4 formation in ruminants [49]; thus, considerable reduction can be achieved by utilizing them [21,77]. Although some compounds such as BCM are banned and therefore cannot be recommended, other compounds with a similar mode of action could be acquired [21]. Such compounds can be found in macroalgae, such as Asparagopsis taxiformis, which is acknowledged to potentially reduce enteric CH4 production, by inhibiting cobamide-dependent methanogenesis [78]. The same researchers reported 50 to 80% CH4 reduction over a 72-day feeding period by adding 3% of Asparagopsis taxiformis to the sheep diet, in comparison with lower inclusion levels. Roque et al. [79] mentioned a 95% reduction in CH4 by adding 5% of feed organic matter of the same macroalgae species to the diet of dairy cows. Similar results were shown by Kinley et al. [80] when steers received 0.10% and 0.20% of Asparagopsis taxiformis and CH4 decreased up to 40% and 98%, respectively. However, Patra [69] described the risk of toxicity, even up to death, after a long period of consumption. Recent research has identified alternative approaches capable of inhibiting methanogenesis, such as 3-nitrooxypropanol (3NP) [81], which achieved 24% reduction in CH4 emissions in sheep [77] and up to 70% in cattle [82,83], while nontoxic effects were observed [21,84].

Amino Acid

Furthermore, more practices should be adapted in order to decrease the environmental burden and maximize the exploitation of nutrients. Amongst these, amino acids operate a fundamental role not only in fulfilling the protein needs but also in the bioavailability of nutrients, such as nitrogen efficiency [85], thus reducing the cost and increasing the environmental effectiveness [86]. Furthermore, amino acids are vastly involved in animals’ metabolic processes, improving growth and reproductive performance plus strengthening the immune system [87]. In spite of this, the combination of amino acids with trace elements, under the format of chelated amino acid, could also be a favorable practice not only for the N use efficiency but to the total feed efficiency and performance burden [88]. A summary of the practices mentioned below is presented in Table 1.

3.2. Energy-Related Strategies

Energy management is an important factor for sustainable livestock production and mitigation strategies. Primarily, measures should be taken to improve the heating requirements of livestock buildings. Furthermore, using, selecting, and maintaining efficient lighting systems, exhaust fans or milking machines, and improving building insulation appear to be of paramount importance. In addition, an immediate reduction in GHG emissions can be achieved by replacing fossil fuels with renewable fuels such as wind, solar, biomass, energy crops, and energy from CH4 produced from manure [89]. In addition, farm buildings should be heated, ventilated, cooled, and lit using renewable energy sources, such as roof-mounted solar panels [90]. In addition, LED technology is more reliable, energy efficient, longer lasting, and greener. Nonetheless, it is crucial to install and maintain a simple control sensor system, for example, temperature control sensors to preserve the appropriate temperature.

3.3. Manure-Related Strategies

Animal manure is generally classified into liquid, slurry, and solid. Its management can lessen the hazards for GHG emissions as well as environmental pollution and public health. Manure management includes collecting, storing, treating, transporting, and disposing of manure. An efficient system needs to keep manure and its hazardous constituents off the environment, and concurrently be profitable. Furthermore, by applying it properly on the field for fertilization and feed production, the reduced use of nitrogen fertilizer and subsequently nitrogen loss can be achieved.

3.3.1. Storage

For proper management, animal houses must allow the easy collection to prevent losses. The floor of the house should be covered from rain to prevent loss of nutrients due to evaporation, runoff, and leaching. Frequent removal from animal housing is required for the prevention of manure fermentation and GHG emissions, as a storage period extension can raise CH4 emissions [21,83]. The area in which the manure is to be stored should be in compliance with the appropriate time limits so that there is proper time for digestion. Additionally, the duration of storage is highly dependent on the climatic conditions and, as a result, may vary depending on the conditions, from 3 months in drier countries including Greece to even 10 months in Finland [91].

Covering

Storage of liquids should be assembled in tanks that can be covered or opened, and sealed with specific water permeability indicators [21,92,93,94]. Covering liquid manure maintains aerobic conditions during storage that can help reduce N2O and CH4 emissions and regulate the odors from NH3 emissions. Less CH4 is produced by solid manure or when deposited on pasture or rangelands [23]. The average CH4 emissions as reported by Chianese et al. [25] were 6.5, 5.4, and 2.3 kg per m2 per year from covered, uncovered slurry, and manure in stacks, respectively. N2O emissions from liquid manure are negligible during storage unless a surface crust is present, while a crust may create N2O emissions while drying and aerobic conditions are developed [92,95]. The use of straw as a cover for dairy cattle slurry, instead of a solid cover may produce higher CH4 and N2O emissions [96]. Covering semipermeable manure may reduce NH3, CH4, and odor emissions, but possibly increase N2O emissions [93,94].

Cooling

An applicable practice, not expensive yet simple and efficient, is the cooling of manure [97]. Nevertheless, the proper temperature conditions for the storage of liquid manure should be lower than 15 oC, as this has been reported to decrease the methane emissions [98]. In the same line, the combination of frequent emptying and slurry cooling can be used in cold or mild climates under conditions of significant temperature difference [99].

3.3.2. Process

There are various methods available for treating manure from simple to extremely complex. Firstly, the separation into liquids and solids. The addition of solid materials to the liquid manure is appropriate. Külling et al. [100] showed that farmyard manure and deep litter manure handling systems tend to produce greater N2O emissions than slurry-based systems.

Anaerobic Digestion

The process of anaerobic digestion refers to the deprivation of microorganisms occurring under anaerobic conditions. The outcome is the generation of CH4, CO2, and other gases as by-products. The aftermath is energy production, mainly biogas in the first place. However, the biogas can produce electricity, and electricity is on the last level of the process, thus signifying a very promising approach for lessening the burden of the GHG emissions from manure and energy [101]. Therefore, biogas can replace and decrease farms’ fuel combustion, implying it as one of the most promising practices. The production of biogas by materials such as biomass, manure, sewage, municipal waste, green waste, or plant material and energy crops [102,103,104] makes it a relatively simple practice that can be efficient at a large or small scale (village and household level) [105].

Composting

Composting is a natural exothermic process of aerobic decomposition, where microorganisms transform degradable organic matter into CO2 and water [95,106], providing several advantages pertaining to manure handling, odor, pathogen control, organic matter equilibrium, additional farm income, etc. [21] as well as lower density. Composting can be organized in piles or ditches, either on a small or large scale whereas solid or liquid manure can be used. Aeration may reduce CH4 but increase N2O and NH3 emissions [21,107,108] due to the probability of generating aerobic environment, but the benefits on lessened odor and CH4 emissions, compared with anaerobically stored manure, make it a favorable option [21].

Acidification

Another modest recommendation is the use of acidic agents for NH3 emissions [21,109]. For instance, 14 to 100% reduction in NH3 emissions was reported by Ndegwa et al. [110], by applying sulfuric, hydrochloric, or phosphoric acids and 67 to 87% reductions in CH4 by using sulfuric acid (pH of 5.5) [111]. However, the long-term impact of acidified manure applied on soil pH have not yet been reported [21].

Ventilation-Biofiltration

A ventilation system provides the opportunity to filtrate emissions from facilities or absorb them. Singh and Mallick [112] reported that the utilization of ventilation air methane is considered as the most effective strategy to reduce methane emissions but nonetheless a challenging one from the economical perspective. Additionally, some chemical compounds show promising results based on filtration of NH3, NO2, or even CO2. For example, the stimulation of photocatalytic properties the using of TiO2 by UV light, oxidates NH3 and NOx [113,114] by using titanium dioxide (TiO2) paint on the walls [115,116]. Another option is the use of oxides or hydroxides as well as carbonates or bicarbonates which are known to exert high CO2 absorption and have therefore been examined as such [117].

3.3.3. Distribution and Deposition of Manure

Manure is a valuable source of available nutrients for crops, successfully replacing artificial fertilizer [21]. However, there are a plethora of factors such as the type of soil, moisture or the application technique affecting the level of emissions. Measures such as improved timing (e.g., avoiding application before a rain), fitting nutrient application to crop requirements [12,21], as well as avoiding spreading slurry on wet soils [118] can reduce emissions and additional cost for farmers. Another option is the dilution or reduction of degradable carbon by solid separation or pre-treatment of anaerobic degradation [96,119]. Useful also could be the control of the nitrogen amount available for nitrification and denitrification in soil [21].
A summary of the described applications is presented in Table 2.

3.4. Other Strategies

3.4.1. Precision Feeding

Precision feeding is a technique that aims to harmonize nutrient requirements of the different animals with the nutrient supply. A sustainable technique for nutrition is precision feeding in order to maintain a healthy rumen and to maximize feed efficiency as to decrease CH4 production [21]. Some promising results have already been shown, presented as enhanced ewes’ milk production efficiency by 50%, as well as 42% lower environmental impact measured as kg of FPCM milk (fat protein corrected milk) in a comparative study of innovative farms that used precision feeding versus conventional ones [120]. However, since it needs specific feed, equipment and management practices, it is not easily adapted by extensive farmers because there is a lack of data on the nutrient requirements of the animal breeds and on the quality of feed [21].

3.4.2. Genetics and Selection

The concept of this category is the selection of breeds that emit low levels of GHG; so, to attain this, it is of foremost importance to further understand the mechanisms underlying the correlation between host genetics and the rumen microbial consistency to obtain optimal production with low environmental impact [121,122]. Therefore, a suggestion considerably effective yet simple and reliable, is the handling of animals with low emissions through genetics selection, as the trait for low emissions is heritable [123,124]. Moreover, the host’s genome can affect the CH4 generation by triggering the growth of certain microbes that regulate the amount of H2 [125]. This premise is based on the perception that host–microbiome interaction is substantially orchestrated by host’s genes; thus, the methanogens inhabitation and the methane formation could be manipulated through animal selection.

4. Circular Bio-Economy Sustainable Strategies in Ruminant Nutrition

4.1. WASTE Material: Food Waste as an Alternative Feed Source

Every year on a global scale about 32% of the food produced gets lost or wasted [126]. About 800 million people could cover their needs for food more than four times with this portion of food [127]. One third of the total food produced every year are lost or wasted, which accounts for 1.3 billion tons of food that was destined for human consumption. In the food supply chain, food might be lost or wasted in every stage. In the early stages of the food chain, during production, post-harvest and processing the food loss occurs [127,128,129], while in the stages of retail and consumption the food waste occurs [128,129,130].
Inadequate handling of food waste is responsible for negative environmental impacts and many methods are used for disposal. Some widely used methods are anaerobic digestion, composting or discarding in landfills. An alternative option of recycling food waste might be its use as feedstuffs [131]. Food waste is a valuable source of energy and nutrients and can be used in animal diets. The utilization of food waste is an economic and sustainable method that also reduces the cost of animal feeding [132]. However, there are many obstacles that makes difficult the reutilization of food waste as feedstuffs. The high content of moisture [132], the variability of nutrients [19] or the potential content of pathogenic microorganisms are some of them. Moreover, factors such as the source of collection [133], the period of the year [134] and others [135,136,137,138] could determine the composition of food waste.
The mixing of food waste with other feedstuffs is a practical method to overcome these nutritional obstacles and create a balanced diet for animals [139]. Food waste can be fed in different forms, such as wet, dried or ensiled. In the angle of safety, many safety concerns emerge for the direct utilization of food waste in animal diets. This is the reason that many prohibitions are enforced in the European Union about its use in animal diets [140] and many regulations are applied in the United States [141], Japan, South Korea and Taiwan [142].

4.2. Use of Agro-Industrial By-Products in the Diets

Agro-industrial by-products are considered as a valuable and cheap feed ingredient for animal nutrition. A sustainable utilization of these products is supported within the food chain, developing a circular economy strategy. Meanwhile, in the last decades consumers seem to have moved towards natural and clean animal products from farms which follow a sustainable approach in order to be safe and healthy [143]. The use of agro-industrial by-products in ruminant diets could support their growth and productivity, resulting in edible food for human consumption. Many assets could emerge, such as a reduced feed cost for farmers, the production of value-added products or an improved health profile for animals [144].
The different types of agro-industrial by-products are obtained during processing methods, such as the production of oil, sugar, products derived from fruits or vegetables, roots or tubes [145]. Production of wine and oil are corresponded for grape and olive pomace, fruit by-products (apple, citrus) are derived from fruits, jelly or jam industries and vegetable by-products from the processing of tomatoes, potatoes or carrots [146]. All of them can be used as raw materials or after processing with several methods, for example drying or more advanced techniques in order to pick specific compounds. However, there are some limitations concerning their use, with the most important of them the large compositional variability depending on the processing method used, as their availability is strictly affected by the season of the year and the area, while undesirable contaminants with or inorganic origin are present [147,148]. Preservation treatments are important for product stabilization, seasonal availability and for the increasing of products’ shelf-life, especially for those rich in moisture and lipids [149,150].

4.2.1. Apple By-Products

The process of apple fruit generates large quantities of solid and liquid by-products. As far as the solid part is concerned, it consists mainly of skin, pulp and seeds that are generated from the production of apple juice, jam and sweets and is known as apple pomace. The disposal of apple waste is connected with the release of greenhouse gases, because as a high biodegradable material it is disposed of in landfills, incinerated or sent for composting [151].
Using apple pomace in animal diets could help to reduce the environmental issues of disposal. Rich in fermentable carbohydrate, pectin, crude fiber and minerals, it is a valuable industrial by-product for ruminant diets [152]. Steyn et al. [153] replaced ground maize with dried apple pomace in the diets of Jersey cows in order to study its effect on milk yield, with four inclusion levels (0, 25, 50, 75%). They found no differences concerning the composition of the protein and fat of the milk. A mix of ensiled apple and tomato pomace in a ratio of 50:50 was used by Abdollahzadeh et al. [154] in order to evaluate the performance of Holstein cows. According to their results, this mix could replace 30% of cows’ diet with a small increase in milk production, and with no other adverse effect.

4.2.2. Citrus By-Products

The remaining residues of citrus fruits from juice extraction are considered as the citrus by-products [155]. It consists mainly of peels at about 60–65%, 30–35% is the pulp and 0–10% from seeds. Citrus pulp is an extraordinary source of energy for ruminants. The utilization in animal diets could limit management problems concerning the disposal of citrus residues. The dried citrus pulp is also rich in fiber and calcium and can be a valuable and cheap feedstuff. Fibers from citrus contains associated bioactive compounds (flavonoids) and this is a remarkable trait compared to the usual dietary fibers.
In the diet of dairy cows a replacement of corn grain (in 0, 15, 30%) with pelleted citrus pulp was studied by Santo et al. [156]. They did not observe any negative effect on the composition and quantity of the milk produced. Moreover, effects in many parameters were studied after the inclusion of citrus pulp and the replacement of cereals in sheep diets. Accordingly, meat oxidative could improve with an inclusion of 24–35% [157,158] of the protein content [159] and an increased concentration of meat PUFAs [160]. Different levels of citrus pulp (10–45%) were evaluated [161,162] and no significant effects were found in lamb growth performance parameters, feed intake and carcass traits.

4.2.3. Grape By-Products

Grape by-products are derived during the destemming, pressing and extracting process of grapes. Annually, large portions of grape waste are produced, posing management concerns [163]. Waste from grape processing are mainly stems and grape pomace such as grape skins, stems and seeds that compose the solid part [164]. From the total amount of waste, grape pomace accounts for 62%, wine lees accounts for 14% and stems 12% [165].
In a study carried out by Moate et al. [166], a reduction of methane emissions of about 20% was observed when dairy cows consumed about 5 kg of grape pomace per day. Moreover, no negative effects were found in milk quality traits and plasma biochemical parameters with the exception of urea in cows fed 3 kg/day grape pomace [167]. In fattening lambs, the inclusion of 10% in their diets did not affect their growth performance [168]. Furthermore, as far as the meat fatty acid concentration is concerned, a significant increase in linolenic acid and an improvement in oxidative stability of the meat was observed [169].

4.2.4. Olive By-Products

The extraction of olive oil generates large quantities of olive by-products that can lead to the degradation of the environment (pulp, skin, olive grains) [170]. An approximate amount of 800 g olive cake are produced by 1 kg olives [171]. The use of olive by-products in livestock diets would alleviate the environmental impacts caused and eliminate the feeding cost of animals [172]. Due to its high fiber content, olive cake could be used in ruminant diets. The composition content of it depends on the cultivation technique, the method of oil extraction, factors that affect the nutritional value and the preservation techniques that were followed [173].
Many studies were carried out in order to investigate any possible effects on the growth performance, after the inclusion of olive cake in ruminant diets, with neutral or positive results. An addition of 15% olive cake had no significant negative results in the daily weight gain, final body weight and dressing percentage of lambs [174]. Farmers could reduce lambs’ feeding cost by 75%, with the use of olive-cake based feed blocks, with no negative effect on their performance [174]. Moreover, Chiofalo et al. [175] used olive cake to substitute cereals in beef cattle diets. In that study, after the inclusion of 7 and 15% olive cake an increase in the body weight, average daily weight gain and carcass traits was found.

4.2.5. Tomato By-Products

The quantity of tomato grown every year on a global scale is estimated at 180 million tons [176]. The production of tomato juice and puree related with the generation of several wastes, mainly peels and seeds. About 4.5% of the fresh tomato weight accounts as residues, 3.5% comes from peels and 1.5% comes from seeds. Tomato pomace is a by-product of the canning industry depending on the following procedure and the source of raw tomatoes. The high moisture content is a limiting factor that decreases the storage life of this by-product. It can be used in animal diets fresh, dried or ensiled as an alternative and economic feed.
Tomato by-products could be used in ruminant diets as a valuable and cheap ingredient, representing a rich source of energy and nutrients. The addition of tomato by-product at a level of 40% could improve the quantity and the quality of milk but did not affect physiological characteristics of goats [177]. The supplementation of fattening ruminants diets with tomato pomace at the level of 180 g/kg did not affect negatively rumen fermentation [178,179].
A summary of the aforementioned strategies is presented in Table 3.

5. Conclusions

Conclusively, there is a plethora of actions for maintaining sustainable and environmentally friendly small ruminant production. As has been described, the nutritional strategies, mainly through the utilization of novel feed ingredients, additives and practices over those previously well established (e.g., insects, agro waste or by-products) have been regarded as potentially promising options. Moreover, the nutrition also affects manure content and thus the emissions generated. Therefore, options for the manure as well as the energy, needed for the farm operations, underline another way of sustainable production. The composition of the proportions and additives affect the form and amount of N manure. The future is promising if available strategies are applied in time.

Author Contributions

Conceptualization, E.G. and A.C.P..; investigation, E.G., F.Z., C.M., C.C., A.C.P., P.E.S., C.K., F.G., T.M., A.M. and E.T.; writing—original draft preparation, E.G., F.Z., C.M., C.C., A.C.P., P.E.S., C.K., F.G., T.M., A.M. and E.T.; writing—review and editing, E.G., F.Z., C.M., C.C., A.C.P., P.E.S., C.K., F.G., T.M., A.M. and E.T.; visualization, E.G.; supervision, E.T. and A.C.P.; project administration, E.T., A.C.P., and C.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. United Nations News. Global Perspective Human Stories. Available online: https://news.un.org/en/story/2022/07/1122272 (accessed on 17 November 2022).
  2. United Nations, Department of Economic and Social Affairs, Population Division. World Population Prospects 2019, Online Edition. Rev. 1; United Nations: New York, NY, USA, 2019. [Google Scholar]
  3. Rojas-Downing, M.M.; Nejadhashemi, A.P.; Harrigan, T.; Woznicki, S.A. Climate change and livestock: Impacts, adaptation, and mitigation. Clim. Risk Manag. 2017, 16, 145–163. [Google Scholar] [CrossRef]
  4. Tubiello, F.N.; Salvatore, M.; Cóndor Golec, R.D.; Ferrara, A.; Rossi, S.; Biancalani, R.; Federici, S.; Jacobs, H.; Flammini, A. Agriculture, Forestry and Other Land Use Emissions by Sources and Removals by Sinks 1990–2011 Analysis; ESS Working Paper No.2; FAO: Rome, Italy, 2014. [Google Scholar]
  5. 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]
  6. 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]
  7. Opio, C.; Gerber, P.; Mottet, A.; Falcucci, A.; Tempio, G.; MacLeod, M.; Vellinga, T.; Henderson, B.; Steinfeld, H. Greenhouse Gas Emission from Ruminant Supply Chains: A Global Life Cycle Assessment; Food and Agriculture Organization of the United Nations Animal Production and Health Division: Rome, Italy, 2013. [Google Scholar]
  8. Perez Dominguez, I.; Fellmann, T.; Weiss, F.; Witzke, H.; Barreiro Hurle, J.; Himics, M.; Jansson, T.; Salputra, G.; Leip, A. An Economic Assessment of GHG Mitigation Policy Options for EU Agriculture (EcAMPA 2); Publications Office of the European Union: Luxembourg, 2016. [Google Scholar]
  9. International Institute for Applied Systems Analysis. Available online: https://iiasa.ac.at (accessed on 20 November 2022).
  10. Zervas, G.; Tsiplakou, E. An assessment of GHG emissions from small ruminants in comparison with GHG emissions from large ruminants and monogastric livestock. Atmos. Environ. 2012, 49, 13–23. [Google Scholar] [CrossRef]
  11. Cabeza, L.F.; Tay, N.S. (Eds.) Environmental Approach. In High-Temperature Thermal Storage Systems Using Phase Change Materials; Academic Press: Cambridge, MA, USA, 2018; pp. 277–295. [Google Scholar]
  12. Lesschen, J.P.; van den Berg, M.; Westhoek, H.J.; Witzke, H.P.; Oenema, O. Greenhouse gas emission profiles of European livestock sectors. Anim. Feed. Sci. Technol. 2011, 166–167, 16–28. [Google Scholar] [CrossRef]
  13. Sintori, A.; Tzouramani, I.; Liontakis, A. Greenhouse gas emissions in dairy goat farming systems: Abatement potential and cost. Animals 2019, 9, 945. [Google Scholar] [CrossRef] [Green Version]
  14. Steinfeld, H.; Gerber, P.; Wassenaar, T.D.; Castel, V.; Rosales, M.; Rosales, M.; de Haan, C. Livestock’s Long Shadow: Environmental issues and Options; Food & Agriculture Org: Rome, Italy, 2006. [Google Scholar]
  15. Naser, H.M.; Nagata, O.; Sultana, S.; Hatano, R. Carbon Sequestration and Contribution of CO2, CH4 and N2O Fluxes to Global Warming Potential from Paddy-Fallow Fields on Mineral Soil Beneath Peat in Central Hokkaido, Japan. Agriculture 2019, 10, 6. [Google Scholar] [CrossRef] [Green Version]
  16. FAO. GLEAM v3.0 Dashboard In: Shiny Apps. 2022. Available online: https://foodandagricultureorganization.shinyapps.io/GLEAMV3_Public/ (accessed on 14 February 2022).
  17. 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, 7. [Google Scholar] [CrossRef] [Green Version]
  18. Wolf, J.; Asrar, G.R.; West, T.O. Revised Methane Emissions Factors and Spatially Distributed Annual Carbon Fluxes for Global Livestock. Carbon Balance Manag. 2017, 12, 16. [Google Scholar] [CrossRef] [Green Version]
  19. Ungerfeld, E.M. Metabolic Hydrogen Flows in Rumen Fermentation: Principles and Possibilities of Interventions. Front. Microbiol. 2020, 11, 589. [Google Scholar] [CrossRef] [Green Version]
  20. Johnson, K.A.; Johnson, D.E. Methane emissions from cattle. J. Anim. Sci. 1995, 73, 2483–2492. [Google Scholar] [CrossRef]
  21. Hristov, A.N.; Oh, J.; Firkins, J.L.; Dijkstra, J.; Kebreab, E.; Waghorn, G.; Makkar, H.P.S.; Adesogan, A.T.; Yang, W.; Lee, C.; et al. SPECIAL TOPICS—Mitigation of Methane and Nitrous Oxide Emissions from Animal Operations: I. A Review of Enteric Methane Mitigation Options. J. Anim. Sci. 2013, 91, 5045–5069. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Ministry of Environment and Energy. Climate Change—National Inventory Report of Greece for Greenhouse and Other Gases for the Years 1990–2019; Ministry of Environment and Energy: Athens, Greece, 2021. Available online: https://ypen.gov.gr/wp-content/uploads/2021/06/2021_NIR_Greece.pdf (accessed on 30 January 2023).
  23. Environmental Protection Agency. Inventory of U.S. Greenhouse Gas Emissions and Sinks: 1990–2018; EPA: Washington, DC, USA, 2020; p. 733.
  24. Holly, M.A.; Larson, R.A.; Powell, J.M.; Ruark, M.D.; Aguirre-Villegas, H. Greenhouse gas and ammonia emissions from digested and separated dairy manure during storage and after land application. Agric. Ecosyst. Environ. 2017, 239, 410–419. [Google Scholar] [CrossRef]
  25. Chianese, D.S.; Rotz, C.A.; Richard, T.L. Whole-Farm Greenhouse Gas Emissions: A Review with Application to a Pennsylvania Dairy Farm. Appl. Eng. Agric. 2009, 25, 431–442. [Google Scholar] [CrossRef]
  26. Olesen, J.E.; Schelde, K.; Weiske, A.; Weisbjerg, M.R.; Asman, W.A.H.; Djurhuus, J. Modelling Greenhouse Gas Emissions from European Conventional and Organic Dairy Farms. Agric. Ecosyst. Environ. 2006, 112, 207–220. [Google Scholar] [CrossRef]
  27. 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]
  28. 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] [PubMed] [Green Version]
  29. Benchaar, C.; Pomar, C.; Chiquette, J. Evaluation of dietary strategies to reduce methane production in ruminants: A modelling approach. Can. J. Anim. Sci. 2001, 81, 563–574. [Google Scholar] [CrossRef]
  30. Archimède, H.; Eugène, M.; Marie Magdeleine, C.; Boval, M.; Martin, C.; Morgavi, D.P.; Lecomte, P.; Doreau, M. Comparison of methane production between C3 and C4 grasses and legumes. Anim. Feed. Sci. Technol. 2011, 166–167, 59–64. [Google Scholar] [CrossRef]
  31. Lee, J.M.; Woodward, S.L.; Waghorn, G.C.; Clark, D.A. Methane emissions by dairy cows fed increasing proportions of white clover (Trifolium repens) in pasture. Proc. New Zealand Grassl. Assoc. 2004, 66, 151–155. [Google Scholar] [CrossRef]
  32. Waghorn, G.C.; Tavendale, M.H.; Woodfield, D.R. Methanogenesis from forages fed to sheep. Proc. New Zealand Grassl. Assoc. 2002, 64, 167–171. [Google Scholar]
  33. Gonzalez-Ronquillo, M.; Robles-Jiménez, L.E.; Romero-Bernal, J.; Ariciaga-Gonzalez, C.; Valdez-Ramírez, B.; Chay-Canul, A.J.; Vargas-Bello-Pérez, E. Effect of ryegrass hay and ryegrass silage, cut at two stages of development, on nutrient digestibility, nitrogen balance, and purine derivative excretion in growing sheep. Ital. J. Anim. Sci. 2021, 20, 2110–2121. [Google Scholar] [CrossRef]
  34. van Dorland, H.A.; Wettstein, H.R.; Leuenberger, H.; Kreuzer, M. Effect of supplementation of fresh and ensiled clovers to ryegrass on nitrogen loss and methane emission of dairy cows. Livest. Sci. 2007, 111, 57–69. [Google Scholar] [CrossRef]
  35. Robles Jimenez, L.E.; Zetina Sánchez, A.; Castelán Ortega, O.A.; Osorio Avalos, J.; Estrada Flores, J.G.; González-Ronquillo, M.; Vargas-Bello-Pérez, E. Effect of different growth stages of rapeseed (brassica rapa L.) on nutrient intake and digestibility, nitrogen balance, and rumen fermentation kinetics in sheep diets. Ital. J. Anim. Sci. 2021, 20, 698–706. [Google Scholar] [CrossRef]
  36. McAllister, T.A.; Newbold, C.J. Redirecting rumen fermentation to reduce methanogenesis. Aust. J. Exp. Agric. 2008, 48, 7–13. [Google Scholar] [CrossRef]
  37. Lovett, D.K.; Shalloo, L.; Dillon, P.; O’Mara, F.P. A systems approach to quantify greenhouse gas fluxes from pastoral dairy production as affected by management regime. Agric. Syst. 2006, 88, 156–179. [Google Scholar] [CrossRef]
  38. Alvarez-Hess, P.S.; Williams, S.R.O.; Jacobs, J.L.; Hannah, M.C.; Beauchemin, K.A.; Eckard, R.J.; Wales, W.J.; Morris, G.L.; Moate, P.J. Effect of dietary fat supplementation on methane emissions from dairy cows fed wheat or corn. J. Dairy Sci. 2019, 102, 2714–2723. [Google Scholar] [CrossRef] [Green Version]
  39. Grainger, C.; Clarke, T.; Beauchemin, K.A.; McGinn, S.M.; Eckard, R.J. Supplementation with whole cottonseed reduces methane emissions and can profitably increase milk production of dairy cows offered a forage and cereal grain diet. Aust. J. Exp. Agric. 2008, 48, 73–76. [Google Scholar] [CrossRef]
  40. Nogueira, R.G.S.; Perna Junior, F.; Pereira, A.S.C.; Cassiano, E.C.O.; Carvalho, R.F.; Rodrigues, P.H.M. Methane Mitigation and Ruminal Fermentation Changes in Cows Fed Cottonseed and Vitamin E. Sci. Agric. 2020, 77, e20180247. [Google Scholar] [CrossRef]
  41. Martin, C.; Morgavi, D.P.; Doreau, M. Methane mitigation in ruminants: From microbe to the farm scale. Animal 2010, 4, 351–365. [Google Scholar] [CrossRef] [Green Version]
  42. Moate, P.J.; Williams, S.R.O.; Grainger, C.; Hannah, M.C.; Ponnampalam, E.N.; Eckard, R.J. Influence of cold-pressed canola, brewers grains and hominy meal as dietary supplements suitable for reducing enteric methane emissions from lactating dairy cows. Anim. Feed. Sci. Technol. 2011, 166, 254–264. [Google Scholar] [CrossRef]
  43. Grainger, C.; Beauchemin, K.A. Can enteric methane emissions from ruminants be lowered without lowering their production? Anim. Feed. Sci. Technol. 2011, 166–167, 308–320. [Google Scholar] [CrossRef]
  44. Wiedmann, T.; Minx, J. A definition of ‘carbon footprint’. Ecol. Econ. Res. Trends 2008, 1, 1–11. [Google Scholar]
  45. Vargas, J.E.; Andrés, S.; López-Ferreras, L.; Snelling, T.J.; Yáñez-Ruíz, D.R.; García-Estrada, C.; López, S. Dietary supplemental plant oils reduce methanogenesis from anaerobic microbial fermentation in the rumen. Sci. Rep. 2020, 10, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Mavrommatis, A.; Skliros, D.; Flemetakis, E.; Tsiplakou, E. Changes in the Rumen Bacteriome Structure and Enzymatic Activities of Goats in Response to Dietary Supplementation with Schizochytrium spp. Microorganisms 2021, 9, 1528. [Google Scholar] [CrossRef] [PubMed]
  47. Mavrommatis, A.; Skliros, D.; Simoni, M.; Righi, F.; Flemetakis, E.; Tsiplakou, E. Alterations in the rumen particle-associated microbiota of goats in response to dietary supplementation levels of Schizochytrium spp. Sustainability 2021, 13, 607. [Google Scholar] [CrossRef]
  48. Cosgrove, G.P.; Waghorn, G.C.; Anderson, C.B.; Peters, J.S.; 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]
  49. Llonch, P.; Haskell, M.J.; Dewhurst, R.J.; Turner, S.P. Current available strategies to mitigate greenhouse gas emissions in livestock systems: An animal welfare perspective. Animal 2017, 11, 274–284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Mavrommatis, A.; Sotirakoglou, K.; Skliros, D.; Flemetakis, E.; Tsiplakou, E. Dose and time response of dietary supplementation with Schizochytrium sp. on the abundances of several microorganisms in the rumen liquid of dairy goats. Livest. Sci. 2021, 247, 104489. [Google Scholar] [CrossRef]
  51. Abbott, D.W.; Aasen, I.M.; Beauchemin, K.A.; Grondahl, F.; Gruninger, R.; Hayes, M.; Huws, S.; Kenny, D.A.; Krixsan, S.J.; Kirwan, S.F.; et al. Seaweed and seaweed bioactives for mitigation of enteric methane: Challenges and opportunities. Animals 2020, 10, 2432. [Google Scholar] [CrossRef]
  52. Sar, C.; Mwenya, B.; Santoso, B.; Takaura, K.; Morikawa, R.; Isogai, N.; Asakura, Y.; Toride, Y.; Takahashi, J. Effect of Escherichia coli wild type or its derivative with high nitrite reductase activity on in vitro ruminal methanogenesis and nitrate/nitrite reduction. J. Anim. Sci. 2005, 83, 644–652. [Google Scholar] [CrossRef]
  53. Nolan, J.V.; Hegarty, R.S.; Hegarty, J.; Godwin, I.R.; Woodgate, R. Effects of dietary nitrate on fermentation, methane production and digesta kinetics in sheep. Anim. Prod. Sci. 2010, 50, 801–806. [Google Scholar] [CrossRef]
  54. Van Zijderveld, S.M.; Gerrits, W.J.J.; Dijkstra, J.; Newbold, J.R.; Hulshof, R.B.A.; Perdok, H.B. Persistency of methane mitigation by dietary nitrate supplementation in dairy cows. J. Dairy Sci. 2011, 94, 4028–4038. [Google Scholar] [CrossRef] [Green Version]
  55. Troy, S.M.; Duthie, C.A.; Hyslop, J.J.; Roehe, R.; Ross, D.W.; Wallace, R.J.; Wterhouse, A.; Rooke, J.A. Effectiveness of nitrate addition and increased oil content as methane mitigation strategies for beef cattle fed two contrasting basal diets. J. Anim. Sci. 2015, 93, 1815–1823. [Google Scholar] [CrossRef] [PubMed]
  56. 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. Nihon Chikusan Gakkaiho 2001, 72, 139–146. [Google Scholar] [CrossRef] [Green Version]
  57. 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]
  58. Machmüller, A.; Soliva, C.R.; Kreuzer, M. Methane-suppressing effect of myristic acid in sheep as affected by dietary calcium and forage proportion. Br. J. Nutr. 2003, 90, 529–540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Eckard, R.J.; Grainger, C.; De Klein, C.A.M. Options for the abatement of methane and nitrous oxide from ruminant production: A review. Livest. Sci. 2010, 130, 47–56. [Google Scholar] [CrossRef]
  60. 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]
  61. Newbold, C.J.; Rode, L.M. Dietary additives to control methanogenesis in the rumen. Int. Congr. Ser. 2006, 1293, 138–147. [Google Scholar] [CrossRef]
  62. Lobo, R.R.; Faciola, A.P. Ruminal Phages–A Review. Front. Microbiol. 2021, 12, 763416. [Google Scholar]
  63. Leahy, S.C.; Kelly, W.J.; Altermann, E.; Ronimus, R.S.; Yeoman, C.J.; Pacheco, D.M.; Li, D.; Kong, Z.; Mctavish, S.; Sang, C.; et al. The genome sequence of the rumen methanogen Methanobrevibacter ruminantium reveals new possibilities for controlling ruminant methane emissions. PloS ONE 2010, 5, e8926. [Google Scholar] [CrossRef] [Green Version]
  64. Subharat, S.; Shu, D.; Zheng, T.; Buddle, B.M.; Janssen, P.H.; Luo, D.; Wedlock, D.N. Vaccination of cattle with a methanogen protein produces specific antibodies in the saliva which are stable in the rumen. Vet. Immunol. Immunopathol. 2015, 164, 201–207. [Google Scholar] [CrossRef] [PubMed]
  65. Buddle, B.M.; Denis, M.; Attwood, G.T.; Altermann, E.; Janssen, P.H.; Ronimus, R.S.; Pinares Patino, C.S.; Muetzel, S.; Wedlock, D.N. Strategies to reduce methane emissions from farmed ruminants grazing on pasture. Vet. J. 2011, 188, 11–17. [Google Scholar] [CrossRef] [PubMed]
  66. Wright, A.D.G.; Kennedy, P.; O’neill, C.J.; Toovey, A.F.; Popovski, S.; Rea, S.M.; Pimm, C.L.; Klein, L. Reducing methane emissions in sheep by immunization against rumen methanogens. Vaccine 2004, 22, 3976–3985. [Google Scholar] [CrossRef]
  67. Tirado-González, D.N.; Miranda-Romero, L.A.; Ruíz-Flores, A.; Medina-Cuéllar, S.E.; Ramírez-Valverde, R.; Tirado-Estrada, G. Meta-analysis: Effects of exogenous fibrolytic enzymes in ruminant diets. J. Appl. Anim. Res. 2018, 46, 771–783. [Google Scholar] [CrossRef] [Green Version]
  68. Eun, J.S.; Beauchemin, K.A. Assessment of the efficacy of varying experimental exogenous fibrolytic enzymes using in vitro fermentation characteristics. Anim. Feed. Sci. Technol. 2007, 132, 298–315. [Google Scholar] [CrossRef]
  69. Patra, A.K. Enteric methane mitigation technologies for ruminant livestock: A synthesis of current research and future directions. Environ. Monit. Assess. 2012, 184, 1929–1952. [Google Scholar] [CrossRef] [PubMed]
  70. Makkar, H.P.S. Effects and fate of tannins in ruminant animals, adaptation to tannins, and strategies to overcome detrimental effects of feeding tannin-rich feeds. Small Rumin. Res. 2003, 49, 241–256. [Google Scholar] [CrossRef]
  71. Cappucci, A.; Mantino, A.; Buccioni, A.; Casarosa, L.; Conte, G.; Serra, A.; Mannelli, F.; Luciano, G.; Foggi, G.; Mele, M. Diets supplemented with condensed and hydrolysable tannins affected rumen fatty acid profile and plasmalogen lipids, ammonia and methane production in an in vitro study. Ital. J. Anim. Sci. 2021, 20, 935–946. [Google Scholar] [CrossRef]
  72. Carulla, J.E.; Kreuzer, M.; Machmüller, A.; Hess, H.D. 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]
  73. Tiemann, T.T.; Lascano, C.E.; Wettstein, H.R.; Mayer, A.C.; Kreuzer, M.; Hess, H.D. Effect of the tropical tannin-rich shrub legumes Calliandra calothyrsus and Flemingia macrophylla on methane emission and nitrogen and energy balance in growing lambs. Animal 2008, 2, 790–799. [Google Scholar] [CrossRef] [Green Version]
  74. Puchala, R.; Min, B.R.; Goetsch, A.L.; Sahlu, T. The effect of a condensed tannin-containing forage on methane emission by goats. J. Anim. Sci. 2005, 83, 182–186. [Google Scholar] [CrossRef] [PubMed]
  75. Sahebi Ala, M.; Pirmohammadi, R.; Khalilvandi-Behroozyar, H.; Anassori, E. Changes in vitro rumen fermentation, methane production and microbial populations in response to green tea extract. Ital. J. Anim. Sci. 2021, 20, 1114–1125. [Google Scholar] [CrossRef]
  76. Cobellis, G.; Trabalza-Marinucci, M.; Yu, Z. Critical evaluation of essential oils as rumen modifiers in ruminant nutrition: A review. Sci. Total Environ. 2016, 545–546, 556–568. [Google Scholar] [CrossRef] [PubMed]
  77. 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] [Green Version]
  78. 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. 2016, 58, 681–688. [Google Scholar] [CrossRef]
  79. Roque, B.M.; Brooke, C.G.; Ladau, J.; Polley, T.; Marsh, L.J.; Najafi, N.; Pandey, P.; Singh, L.; Kinley, R.; Salwen, J.K.; et al. Effect of the macroalgae Asparagopsis taxiformis on methane production and rumen microbiome assemblage. Anim. Microbiome 2019, 1, 1–14. [Google Scholar]
  80. Kinley, R.D.; Martinez-Fernandez, G.; Matthews, M.K.; Nys, R.; Magnusson, M.; Tomkins, N.W. Mitigating the carbon footprint and improving productivity of ruminant livestock agriculture using a red seaweed. J. Clean. Prod. 2020, 259, 120836. [Google Scholar] [CrossRef]
  81. Jayanegara, A.; Sarwono, K.A.; Kondo, M.; Matsui, H.; Ridla, M.; Laconi, E.B. Nahrowi Use of 3-nitrooxypropanol as feed additive for mitigating enteric methane emissions from ruminants: A meta-analysis. Ital. J. Anim. Sci. 2018, 17, 650–656. [Google Scholar] [CrossRef] [Green Version]
  82. Haisan, J.; Sun, Y.; Guan, L.L.; Beauchemin, K.A.; Iwaasa, A.; Duval, S.; Barreda, D.; Oba, M. The effects of feeding 3-nitrooxypropanol on methane emissions and productivity of Holstein cows in mid lactation. J. Dairy Sci. 2014, 97, 3110–3119. [Google Scholar] [CrossRef] [Green Version]
  83. Reynolds, C.K.; Humphries, D.J.; Kirton, P.; Kindermann, M.; Duval, S.; Steinberg, W. Effects of 3-nitrooxypropanol on methane emission, digestion, and energy and nitrogen balance of lactating dairy cows. J. Dairy Sci. 2014, 97, 3777–3789. [Google Scholar] [CrossRef] [Green Version]
  84. Hristov, A.N.; Oh, J.; Giallongo, F.; Frederick, T.W.; Harper, M.T.; Weeks, H.L.; Branco, A.F.; Moate, P.J.; Deighton, M.H.; Williams, S.R.O.; et al. An inhibitor persistently decreased enteric methane emission from dairy cows with no negative effect on milk production. Proc. Natl. Acad. Sci. USA 2015, 112, 10663–10668. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Laroche, J.P.; Gervais, R.; Lapierre, H.; Ouellet, D.R.; Tremblay, G.F.; Halde, C.; Boucher, M.S.; Charbonneau, É. Milk production and efficiency of utilization of nitrogen, metabolizable protein, and amino acids are affected by protein and energy supplies in dairy cows fed alfalfa-based diets. J. Dairy Sci. 2022, 105, 329–346. [Google Scholar] [CrossRef] [PubMed]
  86. Apelo, S.A.; Bell, A.L.; Estes, K.; Ropelewski, J.; De Veth, M.J.; Hanigan, M.D. Effects of reduced dietary protein and supplemental rumen-protected essential amino acids on the nitrogen efficiency of dairy cows. J. Dairy Sci. 2014, 97, 5688–5699. [Google Scholar] [CrossRef] [Green Version]
  87. Abbasi, I.H.R.; Abbasi, F.; El-Hack, A.; Mohamed, E.; Abdel-Latif, M.A.; Soomro, R.N.; Hayat, K.; Mohamed, M.A.E.; Bodinga, B.M.; Yao, J.; et al. Critical analysis of excessive utilization of crude protein in ruminants ration: Impact on environmental ecosystem and opportunities of supplementation of limiting amino acids—A review. Environ. Sci. Pollut. Res. 2018, 25, 181–190. [Google Scholar] [CrossRef] [PubMed]
  88. Pomport, P.H.; Warren, H.E.; Taylor-Pickard, J. Effect of total replacement of inorganic with organic sources of key trace minerals on performance and health of high producing dairy cows. J. Appl. Anim. Nutr. 2021, 9, 23–30. [Google Scholar] [CrossRef]
  89. Pucker, J.; Jungmeier, G.; Siegl, S.; Pötsch, E.M. Anaerobic digestion of agricultural and other substrates—Implications for greenhouse gas emissions. Animal 2013, 7, 283–291. [Google Scholar] [CrossRef] [PubMed]
  90. Cui, Y.; Theo, E.; Gurler, T.; Su, Y.; Saffa, R. A comprehensive review on renewable and sustainable heating systems for poultry farming. Int. J. Low-Carbon Technol. 2020, 15, 121–142. [Google Scholar] [CrossRef]
  91. Burton, C.H.; Turner, C. Manure Management: Treatment Strategies for Sustainable Agriculture, 2nd ed.; Burton, C.H., Turner, C., Eds.; Silsoe Research Institute: Silsoe, UK, 2003. [Google Scholar]
  92. Sommer, S.G.; Petersen, S.O.; Søgaard, H.T. Greenhouse gas emission from stored livestock slurry. J. Environ. Qual. 2000, 29, 744–751. [Google Scholar] [CrossRef]
  93. Guarino, M.; Fabbri, C.; Brambilla, M.; Valli, L.; Navarotto, P. Evaluation of simplified covering systems to reduce gaseous emissions from livestock manure storage. Trans. ASABE 2006, 49, 737–747. [Google Scholar] [CrossRef]
  94. Sonesson, U.; Cederberg, C.; Berglund, M. Greenhouse Gas Emissions in Animal Feed Production; Klimatmärkning för mat, Svenskt Sigill: Stockholm, Sweden, 2009. [Google Scholar]
  95. Petersen, S.O.; Blanchard, M.; Chadwick, D.; Del Prado, A.; Edouard, N.; Mosquera, J.; Sommer, S.G. Manure management for greenhouse gas mitigation. Animal 2013, 7 (Suppl. S2), 266–282. [Google Scholar] [CrossRef] [Green Version]
  96. Amon, B.; Kryvoruchko, V.; Amon, T.; Zechmeister-Boltenstern, S. Methane, nitrous oxide and ammonia emissions during storage and after application of dairy cattle slurry and influence of slurry treatment. Agric. Ecosyst. Environ. 2006, 112, 153–162. [Google Scholar] [CrossRef]
  97. Sejian, V.; Bhatta, R.; Soren, N.M.; Malik, P.K.; Ravindra, J.P.; Prasad, C.S.; Lal, R. Introduction to concepts of climate change impact on livestock and its adaptation and mitigation. In Climate Change Impact on Livestock: Adaptation and Mitigation; Sejian, V., Gaughan, J., Baumgard, L., Prasad, C., Eds.; Springer India: New Delhi, India, 2015; pp. 1–23. ISBN 9788132222644. [Google Scholar]
  98. Rennie, T.J.; Gordon, R.J.; Smith, W.N.; VanderZaag, A.C. Liquid manure storage temperature is affected by storage design and management practices—A modelling assessment. Agric. Ecosyst. Environ. 2018, 260, 47–57. [Google Scholar] [CrossRef]
  99. Sommer, S.G.; Olesen, J.E.; Petersen, S.O.; Weisbjerg, M.R.; Valli, L.; Rodhe, L.; Béline, F. Region-specific assessment of greenhouse gas mitigation with different manure management strategies in four agroecological zones. Glob. Chang. Biol. 2009, 15, 2825–2837. [Google Scholar] [CrossRef]
  100. Külling, D.R.; Menzi, H.; Sutter, F.; Lischer, P.; Kreuzer, M. Ammonia, nitrus oxide and methane emissions from differently stored dairy manure derived from grass-and hay-based rations. Nutr. Cycl. Agroecosyst. 2003, 65, 13–22. [Google Scholar] [CrossRef]
  101. Aguirre-Villegas, H.A.; Larson, R.A. Evaluating greenhouse gas emissions from dairy manure management practices using survey data and lifecycle tools. J. Clean. Prod. 2017, 143, 169–179. [Google Scholar] [CrossRef]
  102. Bhatia, S.C. Biogas. In Advanced Renewable Energy Systems; WPI Publishing: New York, NY, USA, 2015; p. 47. [Google Scholar]
  103. Da Costa Gomez, C. Biogas as an Energy Option: An Overview. In The Biogas Handbook; Elsevier: Amsterdam, The Netherlands, 2013; pp. 1–16. [Google Scholar]
  104. Teng, Z.; Hua, J.; Wang, C.; Lu, X. Design and optimization principles of biogas reactors in large scale applications. In Reactor and Process Design in Sustainable Energy Technology; Elsevier: Amsterdam, The Netherlands, 2014; pp. 99–134. [Google Scholar]
  105. Teenstra, E.; De Buisonjé, F.; Ndambi, A.; Pelster, D. Manure Management in the (Sub-)Tropics: Training Manual for Extension Workers; Livestock Research Report; Wageningen UR (University & Research centre) Livestock Research: Rome/Wageningen, Italy, 2015. [Google Scholar]
  106. Hoeksma, P.; Losada, J.M.; Melse, R.W. Monitoring Methane and Nitrous Oxide Reduction by Manure Treatment (No. 627); Wageningen UR Livestock Research: Wageningen, The Netherlands, 2012. [Google Scholar]
  107. Pattey, E.; Trzcinski, M.K.; Desjardins, R.L. Quantifying the reduction of greenhouse gas emissions as a result of composting dairy and beef cattle manure. Nutr. Cycl. Agroecosyst. 2005, 72, 173–187. [Google Scholar] [CrossRef]
  108. Webb, J.; Sommer, S.G.; Kupper, T.; Groenestein, K.; Hutchings, N.J.; Eurich-Menden, B.; Rodhe, L.; Misselbrook, T.H.; Amon, B. Emissions of ammonia, nitrous oxide and methane during the management of solid manures. In Agroecology and Strategies for Climate Change; Springer: Dordrecht, The Netherlands, 2012; pp. 67–107. [Google Scholar]
  109. Petersen, S.O.; Sommer, S.G. Ammonia and nitrous oxide interactions: Roles of manure organic matter management. Anim. Feed. Sci. Technol. 2011, 166, 503–513. [Google Scholar] [CrossRef]
  110. Ndegwa, P.M.; Hristov, A.N.; Ogejo, J.A. Ammonia emission from animal manure: Mechanisms and mitigation techniques. In Environmental Chemistry of Animal Manure; Nova Science Publishers, Inc.: New York, NY, USA, 2011; pp. 107–151. [Google Scholar]
  111. Petersen, S.O.; Andersen, A.J.; Eriksen, J. Effects of cattle slurry acidification on ammonia and methane evolution during storage. J. Environ. Qual. 2012, 41, 88–94. [Google Scholar] [CrossRef]
  112. Singh, H.; Mallick, J. Utilization of ventilation air methane in Indian coal mines: Prospects and Challenges. Procedia Earth Planet. Sci. 2015, 11, 56–62. [Google Scholar] [CrossRef] [Green Version]
  113. Lee, J.; Park, H.; Choi, W. Selective photocatalytic oxidation of NH3 to N2 on platinized TiO2 in water. Environ. Sci. Technol. 2012, 36, 5462–5468. [Google Scholar] [CrossRef]
  114. Allen, N.S.; Edge, M.; Sandoval, G.; Verran, J.; Stratton, J.; Maltby, J. Photocatalytic Coatings for Environmental Applications. Photochem. Photobiol. 2005, 81, 279–290. [Google Scholar] [CrossRef] [PubMed]
  115. Guarino, M.; Costa, A.; Porro, M. Photocatalytic TiO2 coating—To reduce ammonia and greenhouse gases concentration and emission from animal husbandries. Bioresour. Technol. 2008, 99, 2650–2658. [Google Scholar] [CrossRef] [PubMed]
  116. Costa, A.; Chiarello, G.L.; Selli, E.; Guarino, M. Effects of TiO2 based photocatalytic paint on concentrations and emissions of pollutants and on animal performance in a swine weaning unit. J. Environ. Manag. 2012, 96, 86–90. [Google Scholar] [CrossRef] [PubMed]
  117. Duan, Y.; Luebke, D.; Pennline, H. Efficient theoretical screening of solid sorbents for CO2 capture applications. Int. J. Clean Coal Energy 2012, 1, NETL-PUB-318. [Google Scholar] [CrossRef] [Green Version]
  118. Mkhabela, M.S.; Gordon, R.; Burton, D.; Madani, A.; Hart, W.; Elmi, A. Ammonia and nitrous oxide emissions from two acidic soils of Nova Scotia fertilised with liquid hog manure mixed with or without dicyandiamide. Chemosphere 2006, 65, 1381–1387. [Google Scholar] [CrossRef] [PubMed]
  119. Clemens, J.; Trimborn, M.; Weiland, P.; Amon, B. Mitigation of greenhouse gas emissions by anaerobic digestion of cattle slurry. Agric. Ecosyst. Environ. 2006, 112, 171–177. [Google Scholar] [CrossRef]
  120. Bosco, S.; Volpi, I.; Cappucci, A.; Mantino, A.; Ragaglini, G.; Bonari, E.; Mele, M. Innovating feeding strategies in dairy sheep farming can reduce environmental impact of ewe milk. Ital. J. Anim. Sci. 2021, 20, 2147–2164. [Google Scholar] [CrossRef]
  121. Shi, W.; Moon, C.D.; Leahy, S.C.; Kang, D.; Froula, J.; Kittelmann, S.; Fan, C.; Deutsch, C.; Gagic, D.; Seedorf, H.; et al. Methane yield phenotypes linked to differential gene expression in the sheep rumen microbiome. Genome Res. 2014, 24, 1517–1525. [Google Scholar] [CrossRef] [Green Version]
  122. Difford, G.F.; Plichta, D.R.; Løvendahl, P.; Lassen, J.; Noel, S.J.; Højberg, O.; Wright, A.D.G.; Shu, S.; Kristensen, L.; Nielsen, H.B.; et al. Host genetics and the rumen microbiome jointly associate with methane emissions in dairy cows. PLoS Genet. 2018, 14, e1007580. [Google Scholar] [CrossRef] [Green Version]
  123. Pinares-Patiño, C.S.; Hickey, S.M.; Young, E.A.; Dodds, K.G.; MacLean, S.; Molano, G.; Sandoval, E.; Kjestrup, H.; Harland, R.; Hunt, C.; et al. Heritability estimates of methane emissions from sheep. Animal 2013, 7, 316–321. [Google Scholar] [CrossRef] [Green Version]
  124. Bittante, G.; Cecchinato, A. Heritability estimates of enteric methane emissions predicted from fatty acid profiles, and their relationships with milk composition, cheese-yield and body size and condition. Ital. J. Anim. Sci. 2020, 19, 114–126. [Google Scholar] [CrossRef] [Green Version]
  125. Martínez-Álvaro, M.; Auffret, M.D.; Duthie, C.A.; Dewhurst, R.J.; Cleveland, M.A.; Watson, M.; Roehe, R. Bovine host genome acts on rumen microbiome function linked to methane emissions. Commun. Biol. 2022, 5, 1–16. [Google Scholar] [CrossRef]
  126. Mapping Supply and Demand for Animal-Source Foods to 2030; Robinson, T.P.; Pozzi, F. (Eds.) FAO: Rome, Italy, 2011. [Google Scholar]
  127. FAO. The Future of Food and Agriculture: Trends and Challenges; FAO: Rome, Italy, 2017. [Google Scholar]
  128. Gustavsson, J.; Cederberg, C.; Sonesson, U.; Otterdijk, R.; McYbeck, A. Global Food Losses and Food Waste: Extent, Causes and Prevention; FAO: Rome, Italy, 2011. [Google Scholar]
  129. FAO. The State of Food and Agriculture 2019. Moving Forward on Food Loss and Waste Reduction; FAO: Rome, Italy, 2019. [Google Scholar]
  130. Parfitt, J.; Barthel, M.; MacNaughton, S. Food waste within food supply chains: Quantification and potential for change to 2050. Philos. Trans. R. Soc. B Biol. Sci. 2010, 365, 3065–3081. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. De Boer, I.J.M.; Van Ittersum, M.K. Circularity in Agricultural Production; Wageningen University & Research: Wageningen, The Netherlands, 2018. [Google Scholar]
  132. Westendorf, M.L. Food Waste as Swine Feed. In Food Waste to Animal Feed; Westendorf, M.L., Ed.; Iowa State University Press: Iowa City, IA, USA, 2000; pp. 69–90. [Google Scholar]
  133. Westendorf, M.L. Food Waste as Animal Feed: An Introduction. In Food Waste to Animal Feed; Westendorf, M.L., Ed.; Iowa State University Press: Iowa City, IA, USA, 2000; pp. 3–16. [Google Scholar]
  134. Esteban, M.B.; García, A.J.; Ramos, P.; Márquez, M.C. Evaluation of fruit-vegetable and fish wastes as alternative feedstuffs in pig diets. Waste Manag. 2007, 27, 193–200. [Google Scholar] [CrossRef] [PubMed]
  135. Myer, R.O.; Brendemuhl, J.H.; Johnson, D.D. Evaluation of dehydrated restaurant food waste products as feedstuffs for finishing pigs. J. Anim. Sci. 1999, 77, 685–692. [Google Scholar] [CrossRef]
  136. Castrica, M.; Tedesco, E.A.D.; Panseri, S.; Ferrazzi, G.; Ventura, V.; Frisio, G.D.; Balzaretti, M.C. Pet Food as the most concrete strategy for using food waste as feedstuff within the European context: A feasibility study. Sustainability 2018, 10, 2035. [Google Scholar] [CrossRef] [Green Version]
  137. Garcia, A.J.; Esteban, M.B.; Marquez, M.C.; Ramos, P. Biodegradable municipal solid waste: Characterization and potential use as animal feedstuffs. Waste Manag. 2005, 25, 780–787. [Google Scholar] [CrossRef]
  138. Westendorf, M.L.; Schuler, T.; Zirkle, E.W.; Hays, V.W.; Wilson, L.L. Nutritional quality of recycled food plate waste in diets fed to swine. Profes. Ani. Sci. 1999, 15, 106–111. [Google Scholar] [CrossRef]
  139. Kwak, W.S.; Kang, J.S. Effect of feeding food waste-broiler litter and bakery by-product mixture to pigs. Bioresour. Technol. 2006, 97, 243–249. [Google Scholar] [CrossRef]
  140. EC. Regulation (EC) No 1069/2009 of the European Parliament and of the Council of 21 October 2009 Laying down Health Rules as Regards Animal By-Products and Derived Products Not Intended for Human Consumption and Repealing Regulation (EC) No 1774/2002; EC: Brussels, Belgium, 2009. [Google Scholar]
  141. Leib, E.B.; Balkus, O.; Rice, C.; Maley, M.; Taneja, R.; Cheng, R.; Civita, N.; Alvoid, T. Leftovers for Livestock: A Legal Guide for Using Excess Food as Animal Feed; The Harvard Food Law and Policy Clinic and the Food Recovery Project at the University of Arkansas School of Law: Fayetteville, Carolina, 2016. [Google Scholar]
  142. Cheng, J.Y.; Lo, I. Investigation of the available technologies and their feasibility for the conversion of food waste into fish feed in Hong Kong. Environ. Sci. Pollut. Res. 2016, 23, 7169–7177. [Google Scholar] [CrossRef]
  143. Font i Furnols, M.F.; Realini, C.; Montossi, F.; Sañudo, C.; Campo, M.M.; Oliver, M.A.; Nute, G.R.; Guerrero, L. Consumer’s purchasing intention for lamb meat affected by country of origin, feeding system and meat price: A conjoint study in Spain, France and United Kingdom. Food Qual. Prefer. 2011, 22, 443–451. [Google Scholar] [CrossRef]
  144. Correddu, F.; Lunesu, M.F.; Buffa, G.; Atzori, A.S.; Nudda, A.; Battacone, G.; Pulina, G. Can agro-industrial by-products rich in polyphenols be advantageously used in the feeding and nutrition of dairy small ruminants? Animals 2020, 10, 131. [Google Scholar] [CrossRef] [Green Version]
  145. Pinotti, L.; Manoni, M.; Fumagalli, F.; Rovere, N.; Luciano, A.; Ottoboni, M.; Ferrari, L.; Cheli, F.; Djuragic, O. Reduce, reuse, recycle for food waste: A second life for fresh-cut leafy salad crops in animal diets. Animals 2020, 10, 1082. [Google Scholar] [CrossRef]
  146. Dilucia, F.; Lacivita, V.; Conte, A.; Del Nobile, M.A. Sustainable use of fruit and vegetable by-products to enhance food packaging performance. Foods 2020, 9, 857. [Google Scholar] [CrossRef] [PubMed]
  147. Grasser, L.A.; Fadel, J.G.; Garnett, I.; DePeters, E.J. Quantity and economic importance of nine selected by-products used in California dairy rations. J. Dairy Sci. 1995, 78, 962–971. [Google Scholar] [CrossRef]
  148. Mussatto, S.I.; Dragone, G.; Roberto, I.C. Brewers’ spent grain: Generation, characteristics and potential applications. J. Cereal Sci. 2006, 43, 1–14. [Google Scholar] [CrossRef]
  149. Halmemies-Beauchet-Filleau, A.; Rinne, M.; Lamminen, M.; Mapato, C.; Ampapon, T.; Wanapat, M.; Vanhatalo, A. Alternative and novel feeds for ruminants: Nutritive value, product quality and environmental aspects. Animal 2018, 12, s295–s309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Salami, S.A.; Luciano, G.; O’Grady, M.N.; Biondi, L.; Newbold, C.J.; Kerry, J.P.; Priolo, A. Sustainability of feeding plant by-products: A review of the implications for ruminant meat production. Anim. Feed. Sci. Technol. 2019, 251, 37–55. [Google Scholar] [CrossRef]
  151. Dhillon, G.S.; Kaur, S.; Brar, S.K. Perspective of apple processing wastes as low-cost substrates for bioproduction of high value products: A review. Renew. Sustain. Energy Rev. 2013, 27, 789–805. [Google Scholar] [CrossRef]
  152. Mirzaei-Aghsaghali, A.; Maheri-Sis, N.; Mansouri, H.; Razeghi, M.E.; Shayegh, J.; Aghajanzadeh-Golshani, A. Evaluating nutritional value of apple pomace for ruminants using in vitro gas production technique. Ann. Biol. Res. 2011, 2, 100–106. [Google Scholar]
  153. Steyn, L.; Meeske, R.; Cruywagen, C.W. The effect of replacing maize with dried apple pomace in the concentrate on performance of Jersey cows grazing kikuyu pasture. Anim. Feed. Sci. Technol. 2018, 239, 85–93. [Google Scholar] [CrossRef]
  154. Abdollahzadeh, F.; Pirmohammadi, R.; Fatehi, F.; Bernousi, I. Effect of feeding ensiled mixed tomato and apple pomace on performance of Holstein dairy cows. Slovak J. Anim. Sci. 2010, 43, 31–35. [Google Scholar]
  155. Bampidis, V.A.; Robinson, P.H. Citrus by-products as ruminant feeds: A review. Anim. Feed. Sci. Technol. 2006, 128, 175–217. [Google Scholar] [CrossRef]
  156. Santos, G.T.; Lima, L.S.; Schogor, A.L.B.; Romero, J.V.; De Marchi, F.E.; Grande, P.A.; Santos, N.W.; Santos, F.S.; Kazama, R. Citrus pulp as a dietary source of antioxidants for lactating Holstein cows fed highly polyunsaturated fatty acid diets. Asian-Australas. J. Anim. Sci. 2014, 27, 1104. [Google Scholar] [CrossRef]
  157. Inserra, L.; Priolo, A.; Biondi, L.; Lanza, M.; Bognanno, M.; Gravador, R.; Luciano, G. Dietary citrus pulp reduces lipid oxidation in lamb meat. Meat Sci. 2014, 96, 1489–1493. [Google Scholar] [CrossRef]
  158. Luciano, G.; Roscini, V.; Mattioli, S.; Ruggeri, S.; Gravador, R.S.; Natalello, A.; Lanza, M.; De Angelis, A.; Priolo, A. Vitamin E is the major contributor to the antioxidant capacity in lambs fed whole dried citrus pulp. Animal 2017, 11, 411–417. [Google Scholar] [CrossRef] [PubMed]
  159. Gravador, R.S.; Jongberg, S.; Andersen, M.L.; Luciano, G.; Priolo, A.; Lund, M.N. Dietary citrus pulp improves protein stability in lamb meat stored under aerobic conditions. Meat Sci. 2014, 97, 231–236. [Google Scholar] [CrossRef]
  160. Lanza, M.; Scerra, M.; Bognanno, M.; Buccioni, A.; Cilione, C.; Biondi, L.; Priolo, A.; Luciano, G. Fatty acid metabolism in lambs fed citrus pulp. J. Anim. Sci. 2015, 93, 3179–3188. [Google Scholar] [CrossRef]
  161. Caparra, P.; Foti, F.; Scerra, M.; Sinatra, M.C.; Scerra, V. Solar-dried citrus pulp as an alternative energy source in lamb diets: Effects on growth and carcass and meat quality. Small Rumin. Res. 2007, 68, 303–311. [Google Scholar] [CrossRef]
  162. Gobindram, M.N.N.E.; Bognanno, M.; Luciano, G.; Avondo, M.; Piccione, G.; Biondi, L. The effects of barley replacement by dehydrated citrus pulp on feed intake, performance, feeding behaviour and serum metabolic indicators in lambs. Anim. Prod. Sci. 2017, 57, 133–140. [Google Scholar] [CrossRef]
  163. Skuras, D.; Psaltopoulos, D. A broad overview of the main problems derived from climate change that will affect agricultural production in the Mediterranean area. In Building Resilience for Adaptation to Climate Change in the Agriculture Sector; FAO: Rome, Italy, 2012; pp. 217–260. [Google Scholar]
  164. Brenes, A.; Viveros, A.; Chamorro, S.; Arija, I. Use of polyphenol-rich grape by-products in monogastric nutrition. A review. Anim. Feed Sci. Technol. 2016, 211, 1–17. [Google Scholar] [CrossRef]
  165. Naziri, E.; Nenadis, N.; Mantzouridou, F.T.; Tsimidou, M.Z. Valorization of the major agrifood industrial by-products and waste from Central Macedonia (Greece) for the recovery of compounds for food applications. Food Res. Int. 2014, 65, 350–358. [Google Scholar] [CrossRef]
  166. Moate, P.J.; Williams, S.R.O.; Torok, V.A.; Hannah, M.C.; Ribaux, B.E.; Tavendale, M.H.; Eckard, R.J.; Jacobs, J.L.; Auldist, M.J.; Wales, W.J. Grape marc reduces methane emissions when fed to dairy cows. J. Dairy Sci. 2014, 97, 5073–5087. [Google Scholar] [CrossRef] [Green Version]
  167. Chedea, V.S.; Pelmus, R.S.; Lazar, C.; Pistol, G.C.; Calin, L.G. Toma, S.M.; Dragomir, S.; Taranu, I. Effects of a diet containing dried grape pomace on blood metabolites and milk composition of dairy cows. J. Sci. Food Agric. 2017, 97, 2516–2523. [Google Scholar] [CrossRef]
  168. Bahrami, Y.; Foroozandeh, A.D.; Zamani, F.; Modarresi, M.; Eghbal-Saeid, S.; Chekani-Azar, S. Effect of diet with varying levels of dried grape pomace on dry matter digestibility and growth performance of male lambs. J. Anim. Plant Sci. 2010, 6, 605–610. [Google Scholar]
  169. Ianni, A.; Di Luca, A.; Martino, C.; Bennato, F.; Marone, E.; Grotta, L.; Cichelli, A.; Martino, G. Dietary supplementation of dried grape pomace increases the amount of linoleic acid in beef, reduces the lipid oxidation and modifies the volatile profile. Animals 2019, 9, 578. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Chouchene, A.; Jeguirim, M.; Khiari, B.; Zagrouba, F.; Trouvé, G. Thermal degradation of olive solid waste: Influence of particle size and oxygen concentration. Resour. Conserv. Recycl. 2010, 54, 271–277. [Google Scholar] [CrossRef]
  171. Camposeo, S.; Vivaldi, G.A.; Gattullo, C.E. Ripening indices and harvesting times of different olive cultivars for continuous harvest. Sci. Hortic. 2010, 151, 1–10. [Google Scholar] [CrossRef]
  172. Molina-Alcaide, E.; Yáñez-Ruiz, D.R. Potential use of olive by-products in ruminant feeding: A review. Anim. Feed. Sci. Technol. 2008, 147, 247–264. [Google Scholar] [CrossRef]
  173. Mioc, B.; Pavic, V.; Vnucec, I.; Prpic, Z.; Kostelic, A.; Susic, V. Effect of olive cake on daily gain, carcass characteristics and chemical composition of lamb meat. Czech J. Anim. Sci. 2007, 52, 31. [Google Scholar] [CrossRef] [Green Version]
  174. Salem, H.B.; Znaidi, I.A. Partial replacement of concentrate with tomato pulp and olive cake-based feed blocks as supplements for lambs fed wheat straw. Anim. Feed. Sci. Technol. 2008, 147, 206–222. [Google Scholar] [CrossRef]
  175. Chiofalo, V.; Liotta, L.; Lo Presti, V.; Gresta, F.; Di Rosa, A.R.; Chiofalo, B. Effect of dietary olive cake supplementation on performance, carcass characteristics, and meat quality of beef cattle. Animals 2020, 10, 1176. [Google Scholar] [CrossRef] [PubMed]
  176. FAOSTAT. Crops Production Quantity. 2022. Available online: http://www.fao.org/faostat/en/#data/QC (accessed on 2 February 2022).
  177. Mizael, W.C.; Costa, R.G.; Rodrigo Beltrão Cruz, G.; Ramos de Carvalho, F.F.; Ribeiro, N.L.; Lima, A.; Dominguez, R.; Lorenzo, J.M. Effect of the use of tomato pomace on feeding and performance of lactating goats. Animals 2020, 10, 1574. [Google Scholar] [CrossRef]
  178. Valenti, B.; Luciano, G.; Pauselli, M.; Mattioli, S.; Biondi, L.; Priolo, A.; Natalello, A.; Morbidini, L.; Lanza, M. Dried tomato pomace supplementation to reduce lamb concentrate intake: Effects on growth performance and meat quality. Meat Sci. 2018, 145, 63–70. [Google Scholar] [CrossRef]
  179. Marcos, C.N.; de Evan, T.; Molina-Alcaide, E.; Carro, M.D. Nutritive value of tomato pomace for ruminants and its influence on in vitro methane production. Animals 2019, 9, 343. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Contribution of small ruminants to GHG/year, globally. Blue color letters: general data about livestock’s GHG contribution such as feed production, including rice products and manure management, and excluding manure application on feed crops or pasture, and total contribution of small ruminant sector, were obtained by Gerber et al. [6]. Black color letters: data about small ruminants’ GHG emission related to feed production, manure management and enteric fermentation were accessed by GLEAMv3.0 on 14/02/2023 using the AR6 equation (CH4 = 27, N2O = 273) [16]).
Figure 1. Contribution of small ruminants to GHG/year, globally. Blue color letters: general data about livestock’s GHG contribution such as feed production, including rice products and manure management, and excluding manure application on feed crops or pasture, and total contribution of small ruminant sector, were obtained by Gerber et al. [6]. Black color letters: data about small ruminants’ GHG emission related to feed production, manure management and enteric fermentation were accessed by GLEAMv3.0 on 14/02/2023 using the AR6 equation (CH4 = 27, N2O = 273) [16]).
Sustainability 15 04118 g001
Table 1. Feed-related strategies applied for the mitigation of GHG emissions.
Table 1. Feed-related strategies applied for the mitigation of GHG emissions.
Feed-Related StrategiesNutritional PracticeImpact on GHGs EmissionsAnimalsReference
Replacing fibrous forageUnlignified immature grasses and early cut forage↓ 15% CH4 (g/kg DMI)Ruminants—modelling approach[29]
Processed forage↓ 21% CH4 (g/kg DMI)
Use of grass forage↓ 28% CH4 (g/kg DMI)
Good quality silage↓ 20% CH4 (g/kg DMI)
Replacing C4 grassesUse of C3 legumes in warm climates↓ 20% CH4 (L/kg OM) Ruminants—meta-analysis[30]
Replacing ryegrassLegumes white clover, Lotus pedundulatus (condensed tannins)↓ 16% CH4 (g/kg DMI)Sheep—in vivo[32]
Ryegrass cut at 21 d instead of 35 d nutrient intake, N balanceSheep—in vitro[33]
300 g/kg DM rapeseed forage harvested at pod stage− nutrient intake and digestibilitySheep—in vitro[35]
Inclusion of concentrate feedElevated levels of concentrate feed (35–40%)↓ 9.5% CH4
↓ 16% N2O
↓ 5% CO2
Ruminants
—modelling approach
[37]
Inclusion of dietary lipidsCottonseed meal↓ 12% CH4 (g/cow/day) Dairy cattle
—in vivo
[38]
Cottonseed meal↓ 42% CH4 in dairy cowsDairy cows—in vivo[40]
10 g/kg dietary fat↓ CH4 for meat by 2.6 g/kg DMISheep—meta-analysis[43]
6% oil and linseed oil↓ 21–28% CH4 (mL/d)Sheep—in vitro[45]
Lipid supplementation in 0, 1.2, 2.5, 3.8, and 5% in sheep diet− alteration (g/kg DMI)Sheep—in vivo[48]
Inclusion of electron receptorsNitrate KNO3 in sheep↓ 23% CH4 (L/kg DMI)Sheep—in vivo[53]
Nitrate in dairy cows↓ 16% CH4 (g/kg DMI)Dairy cows—in vivo[54]
Nitrates in forage-based diets in beef cattle↓ 17% CH4 (g/kg DMI)Beef cattle—in vivo[55]
2% fumaric in cattle silage↓ 23% CH4 (L/d)Cattle—in vivo[56]
100 g/kg fumaric acid in free form↓ 62% CH4 (L/d)Lambs—in vivo[57]
100 g/kg fumaric acid in encapsulated form↓ 76% CH4 (L/d)Lambs—in vivo
50 mg/kg DM myristic acid to concentrate based diet↓ 58% CH4Sheep—in vivo[58]
50 mg/kg DM myristic acid to forage-based diet↓ 22% CH4
Inclusion of microorganisms
2
strains of yeast (Levucell SC and Procreatin-7)
− CH4 (g/kg of DMI)Beef cattle—in vivo[60]
EnzymesAddition of cellulases and hemicellulases in ruminants↓ 28% CH4 and 9% respectively (g/g DMI) Dairy cows—in vitro[28]
Plant bioactive compoundsTannins from Acacia mearnsii ↓ 13% CH4 (kJ/MJ energy intake)Sheep—in vivo[72]
25 g/kg DM of Calliandra calothyrsus and Flemingia macrophylla↓ 13% CH4 (kJ/MJ gross energy intake)Lambs—in vivo[73]
Chemical inhibitors3% of Asparagopsis taxiformis in sheep for 72-day feeding period↓ 50–80% CH4 (g/kg DMI)Sheep—in vivo[78]
5% feed organic matter of Asparagopsis taxiformis↓ 95% CH4 (ml/g OM)Dairy cattle—in vitro[79]
0.10–0.20% of Asparagopsis taxiformis ↓ 40–98% CH4 (g/kg DMI)Steers—in vivo[80]
↓ = decrease, = increase, − = no effect, DMI = dry matter intake, OM = organic matter.
Table 2. Sustainable strategies applied in manure for the mitigation of GHG emissions.
Table 2. Sustainable strategies applied in manure for the mitigation of GHG emissions.
Manure Related StrategiesPracticeImpact on GHG EmissionsReferences
Storage of manureCovering dairy cattle slurry with straw instead of solid cover↓ CH4 and N2O[96]
Semipermeable covering↓ NH3, CH4 and odour
↑ N2O
[93,94]
Cooling slurry below 15 ↓ CH4 [98]
Process of manureAnaerobic digestion↓ emissions from manure and energy[101]
AcidificationSulfuric, hydrochloric or phosphoric acids↓ 14–100% NH3[110]
Sulfuric acid (pH = 5.5)↓ 67–87% CH4 (mg/kg/h) [111]
↓ = decrease, = Increase
Table 3. Circular bio-economy sustainable strategies applied in ruminants’ nutrition.
Table 3. Circular bio-economy sustainable strategies applied in ruminants’ nutrition.
StrategiesPracticeEffects in the SustainabilityReference
Precision feedingInnovative vs. conventional farming systems↑ 50% in milk production of ewes and ↓ 42% CF (kg CO2eq/kg FPCM)[120]
Amino acids and trace elements ↑ feed and N efficiency[88]
Waste valorizationFood wasteRecycling food waste[131]
Reduce cost of feeding[132]
Agro-industrial By-Products
Replacement of ground maize0, 25, 50, 75% dried apple pomace in diets of dairy cows− protein and fat in milk[153]
50:50 ensiled apple and tomato pomace milk production[154]
− adverse effect with inclusion of 30%
Replacement of corn grain in dairy cows0, 15, 30% of pelleted citrus pulp in dairy cows’ diet− composition and quality of milk[156]
Replacement of cerealsDifferent levels of citrus pulp in sheep diet 24–35% oxidative stability[157,158]
↑ protein content[159]
meat PUFAs[160]
Different levels 10–45% in lambs− growth performance parameters, feed intake, carcass traits[161,162]
5 kg of grape pomace/day in dairy cows↓ 20% CH4[166]
10, 15, 20% grape pomace in lamb diet– performance on level of 10% [167]
15% olive cake − daily weight gain, final body weight, dressing percentage of lambs[174]
7 and 15% olive cake body weight, average daily weight gain, carcass traits of beef cattle[175]
40% tomato by-product in goat diet milk composition and production[177]
− physiological characteristics of goats
60, 120, 180 g/kg tomato pomace in fattening ruminants− rumen fermentation[179]
↓ = decrease, = increase, − = no effect
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Giamouri, E.; Zisis, F.; Mitsiopoulou, C.; Christodoulou, C.; Pappas, A.C.; Simitzis, P.E.; Kamilaris, C.; Galliou, F.; Manios, T.; Mavrommatis, A.; et al. Sustainable Strategies for Greenhouse Gas Emission Reduction in Small Ruminants Farming. Sustainability 2023, 15, 4118. https://doi.org/10.3390/su15054118

AMA Style

Giamouri E, Zisis F, Mitsiopoulou C, Christodoulou C, Pappas AC, Simitzis PE, Kamilaris C, Galliou F, Manios T, Mavrommatis A, et al. Sustainable Strategies for Greenhouse Gas Emission Reduction in Small Ruminants Farming. Sustainability. 2023; 15(5):4118. https://doi.org/10.3390/su15054118

Chicago/Turabian Style

Giamouri, Elisavet, Foivos Zisis, Christina Mitsiopoulou, Christos Christodoulou, Athanasios C. Pappas, Panagiotis E. Simitzis, Charalampos Kamilaris, Fenia Galliou, Thrassyvoulos Manios, Alexandros Mavrommatis, and et al. 2023. "Sustainable Strategies for Greenhouse Gas Emission Reduction in Small Ruminants Farming" Sustainability 15, no. 5: 4118. https://doi.org/10.3390/su15054118

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