Nitrate and/or oils supplementation to diets with different roughage: concentrate ratios on in vitro some rumen parameters and protozoa population

DOI: https://doi.org/10.21203/rs.3.rs-1554028/v1

Abstract

This study aimed to determine the effect of nitrate and/or oils supplementation alone or in combination to diets with different roughage: concentrate ratios on in vitro methane (CH4) production, volatile fatty acid (VFA) and ammonia (NH3) concentrations, pH, and protozoa population. The experimental treatments comprised iso-nitrogenous total mixed rations based on forage with roughage to concentrate ratio of 40R:60C (High concentrate) or 60R:40C (High roughage) supplemented with 2 sources of nitrogen (sodium nitrate (NO3-: 45.94 g/kg DM) or urea (16.45 g/kg DM for control group) and 3 oil sources ((feed-derived oil (FDO: with no oil addition), hazelnut oil (HO: 36.58 g/kg DM) or soybean oil (SO: 36.58 g/kg DM)). For this purpose, 2 x 2 x 3 factorial design in 12 groups were used. The treatments were as follows: 40R:60C + Urea (Control) + FDO, 40R:60C + NO3- + FDO, 40R:60C + Urea (Control) + HO, 40R:60C + Urea (Control) + SO, 40R:60C + NO3- + HO, 40R:60C + NO3- + SO, 60R:40C + Urea (Control) + FDO, 60R:40C + NO3- + FDO, 60R:40C + Urea (Control) + HO, 60R:40C + Urea (Control) + SO, 60R:40C + NO3- + HO, 60R:60C + NO3- + SO. Then, the effect of nitrate, oils (O), roughage:concentrate ratio, and the combined effect of nitrate and oil associated with roughage:concentrate on in vitro methane production, VFA and NH3 concentrations, pH, and protozoa population were evaluated. In this study, while NO3-, O, 40R:60C x NO3-, 40R:60C x O,  40R:60C x NO3- x O, 60R:40C x NO3- 60R:40C x O, 60R:40C x NO3- x O (p<0.01) decreased in vitro CH4, protozoa population, ammonia (NH3) concentration, acetic acid, total VFA, acetic acid: propionic acid ratio and pH, they increased butyric acid and propionic acid concentrations.

Furthermore, in vitro CH4 production (12.44 vs 9.09 ml), NH3 (8.23 vs 7.37 mmol/l), propionic acid (19.14 vs 17.93 mmol/l), butyric acid (15.50 vs 14.50 mmol/l), total VFA (86.46 vs 85.66 mmol/l), protozoa population (32.16 vs 26.96 x104 ml.) were high in the 40R:60C treatment (high concentrate). Although, 40R:60C x NO3- x SO, 60R:40C x NO3- x SO, and SO decreased more acetic acid concentration, protozoa population, and thus in vitro CH4 production than other groups.

1. Introduction

Methane (CH4), a product of ruminal microbial fermentation, is a major contributor to global warming (IPCC, 2014). One of the main actors in CH4 production are ruminants. The amount of CH4 produced in the rumen is an indicator for estimating environmental impacts and energy costs in the animal production sector (Auffret et al., 2018). Archaea are responsible for microbial fermentation in anaerobic rumen environment (Yang et al., 2017). For this reason, the decrease of the population of archaea is related to CH4 mitigation in the rumen.

Using nitrate as a feed additive in ruminant nutrition, is one of best strategies to reduce enteric CH4 emissions (Hristov et al., 2013). In the rumen, nitrate uses free hydrogen for the production of ammonia to the detriment of CH4 production. Thus, CH4 production decreases (van Zijderveld et al., 2011). Some studies show a decrease of enteric methanogenesis, microbial growth (Ungerfeld and Kohn, 2006), inhibition of total gas volume and CH4 emission (Guyader et al., 2016), an increase of ammonia (NH3) concentration in the rumen (Sharifi et al., 2018) due to the use of nitrate in the ration. Another CH4 mitigation way is the use of lipids in the diet. In recent years, the use of lipids as a feed additive has been adopted as an alternative to mitigate CH4 in the rumen (Boadi et al., 2004; Martin et al., 2010). Some researchers have reported that the alfalfa plant (because of its high nutritional value) positively affects the levels of digestion and absorption of nutrients, leading to increased productivity levels in ruminants (Paterson et al., 1982; Hunt et al., 1985; Brandt and Klopfenstein, 1986a; b; Leng 1990; Ørskov et al., 1999). On the other hand, some studies reported a low fiber content for alfalfa (28–30%) (Muir et al., 2003; Koukoura et al., 2009; Kuchenmeister et al., 2013; KanthaRaju et al., 2018). It has been found that the rations used in the present study have low fiber content. In addition, nutritional contents of rations used in both R:C ratios in the present study are similar to some studies (Sharifi et al., 2018; Alvarez- Hess et al., 2019; Villar et al., 2019; Olomonchi et al., 2019). Lipids plays an important role due to their effect on the protozoa population. In fact, protozoa stimulate hydrogen production and so methanogenesis (Guyader et al., 2015). Lloyd et al. (1989), reported rumen protozoa to have a high oxygen-scavenging ability. Thanks to these ability protozoa lead to decrease in production of H2 and CH4 in rumen. Here, oils rich in unsaturated fatty acids (monounsaturated fatty acids (MUFA) or polyunsaturated fatty acids (PUFA)) reduce CH4 emissions (McGinn et al., 2004; Beauchemin et al., 2007). Some studies have shown the complementary mitigation effect of lipids and nitrate on CH4 production in dry cows (Guyader et al., 2015) and dairy cows (Guyader et al., 2016). The aim of our study is to determine the effect of nitrate and/or oils supplementation alone or in combination to diets with different roughage: concentrate ratios on some rumen parameters and protozoa population.

2. Materials And Methods

2.1. Rations and experimental design

This study was conducted in the Laboratory of Animal Nutrition, Department of Animal Science, Faculty of Agriculture, Ondokuz Mayıs University, Samsun, Turkey. Our study was carried out according to 2 x 2 x 3 factorial design in 12 groups. The experimental treatments comprised iso-nitrogenous total mixed rations based on forage with roughage to concentrate ratio of 40R:60C (High concentrate) or 60R:40C (High roughage) supplemented with 2 sources of nitrogen (sodium nitrate (NO3: 45.94 g/kg DM) or urea (16.45 g/kg DM for control group) and 3 oil sources ((feed-derived oil (FDO: with no oil addition), hazelnut oil (HO: 36.58 g/kg DM) or soybean oil (SO: 36.58 g/kg DM)). For this purpose, 2 x 2 x 3 factorial design in 12 groups were used. The treatments were as follows: 40R:60C + Urea (Control) + FDO, 40R:60C + NO3 + FDO, 40R:60C + Urea (Control) + HO, 40R:60C + Urea (Control) + SO, 40R:60C + NO3 + HO, 40R:60C + NO3 + SO, 60R:40C + Urea (Control) + FDO, 60R:40C + NO3 + FDO, 60R:40C + Urea (Control) + HO, 60R:40C + Urea (Control) + SO, 60R:40C + NO3 + HO, 60R:60C + NO3 + SO. Treatment groups and rations content were shown in Table 1. N sources (sodium nitrate and urea) and oils (hazelnut oil and soybean oil) were purchased from market. Medicago sativa was obtained from research farm of Ondokuz Mayis University in Bafra district. Rumen fluid used in this study was taken from a private slaughterhouse operating in Atakum district of Samsun. Chemical composition of rations used are given in Table 2.

 
Table 1

Treatment groups and rations content (g/kg DM)

Ingredients

Treatment (g.kg− 1 DM)

40R:60C

60R:40C

Control

NO3

Control

NO3

FDO

HO

SO

FDO

HO

SO

FDO

HO

SO

FDO

HO

SO

 

Roughage (Medicago sativa)

400

400

400

400

400

400

600

600

600

600

600

600

 

Wheat bran

140.14

140.14

140.14

140.14

140.14

140.14

128

128

128

128

128

128

 

Sunflower Seed Meal (SSM) (28%)

-

-

-

-

-

-

62.08

62.08

62.08

62.08

62.08

62.08

 

D.D.G.S(Corn)

78.6

78.6

78.6

78.6

78.6

78.6

48

48

48

48

48

48

 

SSM (%36)

185.29

185.29

185.29

185.29

185.29

185.29

43.06

43.06

43.06

43.06

43.06

43.06

 

Corn extract

60

60

60

60

60

60

30.45

30.45

30.45

30.45

30.45

30.45

 

Cracked wheat

42

42

42

42

42

42

28

28

28

28

28

28

 

SSM sieved wastes

18

24

18

24

18

24

16

16

16

16

16

16

 

Molasses

23.4

23.4

23.4

23.4

23.4

23.4

15.6

15.6

15.6

15.6

15.6

15.6

 

Corn

13.2

13.2

13.2

13.2

13.2

13.2

-

-

-

-

-

-

 

Sesame sieved wastes

9

9

9

9

9

9

8

8

8

8

8

8

 

Sesame bran

9

9

9

9

9

9

8

8

8

8

8

8

 

Potassium

1.06

1.06

1.06

1.06

1.06

1.06

0.41

0.41

0.41

0.41

0.41

0.41

 

Methionine

0.43

0.43

0.43

0.43

0.43

0.43

1.03

1.03

1.03

1.03

1.03

1.03

 

Lysine

0.74

0.74

0.74

0.74

0.74

0.74

0.37

0.37

0.37

0.37

0.37

0.37

 

Calcium

1.04

1.04

1.04

1.04

1.04

1.04

1.01

1.01

1.01

1.01

1.01

1.01

 

Phosphorus

0.80

0.80

0.80

0.80

0.80

0.80

0.83

0.83

0.83

0.83

0.83

0.83

 

Sugar

5.6

5.6

5.6

5.6

5.6

5.6

5.02

5.02

5.02

5.02

5.02

5.02

 

Starch

17

17

17

17

17

17

16.03

16.03

16.03

16.03

16.03

16.03

 

Bypass starch

7.09

7.09

7.09

7.09

7.09

7.09

8.05

8.05

8.05

8.05

8.05

8.05

 

Halzenut oil

-

36.58

-

-

36.58

-

-

36.58

-

-

36.58

-

 

Soybean oil

-

-

36.58

-

-

36.58

-

-

36.58

-

-

36.58

 

Sodium nitrate

-

-

-

45.94

45.94

45.94

-

-

-

45.65

45.65

45.65

 

Urea

16.45

16.45

16.45

-

-

-

16.34

16.34

16.34

-

-

-

UFL

83.54

83.54

83.54

83.54

83.54

83.54

81.02

81.02

81.02

81.02

81.02

81.02

UFV

79

79

79

79

79

79

76.12

76.12

76.12

76.12

76.12

76.12

PDIN (g/kg DM)

147.35

147.35

147.35

147.35

147.35

147.35

132.7

132.7

132.7

132.7

132.7

132.7

PDIE (g/kg DM)

120.96

120.96

120.96

120.96

120.96

120.96

101.25

101.2

101.25

101.25

101.2

101.2

ME (kcal. Kg− 1 DM)

2550

2550

2550

2550

2550

2550

2500

2500

2500

2500

2500

2500

FDO: Feed-derived oil (with no oil addition), Vitamin bovine A+ (0.6; 0.4 g/kg DM), marble powder (12; 8 g/kg DM), Salt (3; 2 g/kg DM), Niacin 200 (0.6; 0.4 g/kg DM), Novatan (0.6; 0 g/kg DM), Magnesium oxide (0.6; 0.4 g/kg DM), Yeast (0.6; 0.4 g/kg DM), ME: Metabolic energy, PDIE: true protein absorbable in the small intestine, PDIN = true protein absorbable in the small intestine when degradable N is limiting microbial, UFL: Net energy form ilk, UFV: Net energy for meat.


 
Table 2

Nutrient content of rations with different roughage:concentrate ratios supplemented with NO3, O (HO and SO), and NO3 + O.

Nutrients

(g/kg DM)

40R:60C

60R:40C

Control

NO3

Control

NO3

FDO

SO

HO

FDO

SO

HO

FDO

SO

HO

FDO

SO

HO

Ash

6.21

6.33

6.62

6.20

6.85

6.90

8.81

9.05

9.65

8.91

9.54

9.75

NDF

33.27

33.80

33.27

33.90

33.99

33.90

34.25

34.74

34.88

34.52

34.88

34.78

ADF

20.18

20.94

20.33

20.35

20.57

20.77

21.00

21.75

21.29

21.70

21.81

21.60

EE

3.70

8.27

8.85

3.60

9.41

9.35

3.47

9.26

9.52

3.68

9.23

9.45

CP

23.55

24.20

24.05

23.50

24.32

24.03

23.75

24.10

24.01

23.67

24.46

24.54

CF

24.85

23.12

23.00

24.60

23.12

23.63

24.32

22.38

22.46

24.10

21.86

21.33

ADL

7.63

7.90

8.10

7.55

7.79

7.74

8.95

8.25

8.50

9.02

8.60

8.75

HCEL

13.09

12.86

12.94

13.55

13.42

13.13

13.25

12.99

13.59

12.82

13.07

13.18

CEL

12.56

13.04

12.23

12.80

12.78

13.03

12.05

13.50

12.79

12.68

13.21

12.85

OM

84.23

84.00

83.77

84.50

83.80

83.56

83.96

83.90

83.36

83.05

82.96

82.68

NFE

32.13

28.41

27.90

32.80

26.95

26.55

32.42

28.06

27.37

31.60

27.41

27.30

NFC

46.36

40.26

40.70

46.35

38.85

38.95

42.97

35.84

35.53

42.04

34.96

34.66

CT (g/kg DM)

3.6

3.6

3.6

3.6

3.6

3.6

5.4

5.4

5.4

5.4

5.4

5.4

Saponin (g/kg DM)

4.28

4.28

4.28

4.28

4.28

4.28

6.42

6.42

6.42

6.42

6.42

6.42

FDO: Feed-derived oil (with no oil addition), NDF: Neutral Detergent Fiber, ADF: Acid Detergent Fiber, EE: Extract Ether, CP: Crude Protein, CF: Crude Fat, ADL: Acid Detergent Lignin, HCEL: Hemicellulose, CEL: Cellulose, OM: Organic Matter, NFE: Nitrogen Free Extract, NFC: Non-Fiber Carbohydrate, CT: Condensed Tannins.


2.2. Determination of in vitro CH4 production

Infrared CH4 analyzer (Sensor Europe GmbH, Erkrath, Germany model) was used to determine the in vitro CH4 production in the rations used in present study (Goel et al., 2008). After 24 hours, the gas accumulated in the injectors was taken to the CH4 analyzer by means of a special tube (using plastic injectors) and CH4 production (ml) was determined as a percentage of total gas.

CH4 production (ml) = Total gas production (ml) x % CH4

2.3. Determination of NH3 concentration in the rumen fluid.

For the determination of NH3 concentration, 5 ml of rumen fluid was taken from the syringes after 48 hours. As in the protein analyses, a distillation was performed. Then the titration was done and the volume of HCl (0.1) was noted. Due to following formula, the amount of NH3 was determined.

NH3(mg/dl rumen fluid) = 0.1 x 14 x 1.22 (A-B) x 20

A: Volume of HCl titration solution spent in titration for the sample (ml).

B: Volume of HCl titration solution spent in titration for the witness (ml).

0.1: Normality of HCl titration solution.

14: Molar masses of nitrogen.

2.4. pH and VFA analysis in rumen fluid

In our study, 5 ml of rumen fluid was added to 2 wheaton flasks before incubation and 4 drops of H2SO4 were added to determine the content of VFA in the rumen fluid to be used in the study. The rumen fluids thus prepared were kept at room temperature until the analyzes were performed. The rumen fluid taken at 48th hour was subjected to the same treatment. The pH of the rumen liquid used in the experiment was determined by digital pH meter (HANNA INSTRUMENTS 1332 model pH meter) as soon as it was brought to the laboratory. Rumen fluid taken at 48th hour was subjected to the same treatment. VFA content of the ruminal fluid (Obtain after 48th hour of incubation) was made using the procedure described by Wiedmeier et al. (1987) and using gas chromatography (Agilent Tech. 6890N GC, Stabilwax-DA, 30 m, 0.25 mm ID, 0.25 µm df. Max. Sıcaklık: 260°C. Cat. 11023) at the University of Uludağ, Faculty of Agriculture, Department of Animal Sciences. Four drops of sulfuric acid were added to about 5 ml of ruminal liquid fluid and the mixture was maintained at -20°C and centrifuged at 10000 rpm at + 4°C.

2.5. Determination of Protozoa Population

After the mixing 0.6 g methyl green, 8 g sodium chloride (NaCl) and 100 ml 37% formaldehyde solution for staining the protozoa, the volume was increased to 1000 ml with distilled water. One mililiter of rumen inoculum from the fermenter was mixed with 1 ml of methyl-green-formalin solution (MFS). The protozoa number was carried out with the object slide of a light microscope and Fuchs-Rosenthal counting chamber (depth:0.2 mm, small square area: 0.0625 mm2) (Ranilla et al., 1997). This mixture (Rumen fluid + MFS) was kept at -20°C until analysis time. Subsequently, the samples taken from this mixture and shaken were placed on a Fuchs-Rosenthal slide (16 x 16 squared. 0.0625 mm ~ area). Calculation is made with the formula given below:

$${\mathbf{N}\mathbf{u}\mathbf{m}\mathbf{b}\mathbf{e}\mathbf{r} \mathbf{o}\mathbf{f} \mathbf{c}\mathbf{e}\mathbf{l}\mathbf{l}\mathbf{s} \mathbf{i}\mathbf{n} \mathbf{c}\mathbf{m}}^{3 }\left(\text{m}\text{l}\right)=1000 \mathbf{x} \frac{\mathbf{N}\mathbf{u}\mathbf{m}\mathbf{b}\mathbf{e}\mathbf{r} \mathbf{o}\mathbf{f} \mathbf{c}\mathbf{e}\mathbf{l}\mathbf{l}\mathbf{s} \mathbf{c}\mathbf{o}\mathbf{u}\mathbf{n}\mathbf{t}\mathbf{e}\mathbf{d}}{\mathbf{T}\mathbf{o}\mathbf{t}\mathbf{a}\mathbf{l} \mathbf{f}\mathbf{r}\mathbf{a}\mathbf{m}\mathbf{e}\mathbf{s} \mathbf{c}\mathbf{o}\mathbf{u}\mathbf{n}\mathbf{t}\mathbf{e}\mathbf{d} \mathbf{x} \mathbf{D}\mathbf{i}\mathbf{l}\mathbf{u}\mathbf{t}\mathbf{i}\mathbf{o}\mathbf{n} \mathbf{x} \mathbf{V}\mathbf{o}\mathbf{l}\mathbf{u}\mathbf{m}\mathbf{e}}$$


2.6. Statistical analysis

Data obtained as a result of the research (in vitro CH4 production, NH3 concentration, volatile fatty acids (VFA), pH, and protozoa population) were checked for the necessary assumptions (such as normality and homogeneity of variances) and then analyzed in a randomized plot according to factorial experiment. The following model was used in the study.

Yijkl = µ +αi + βj + λk + (αβ)ij + (βλ)jk + (αλ)ik + (αβλ)ijk + eijkl

Where Yijkl: ith application subject to jth feed variety (CH4 production, etc.) kth observation value of the sample (gas production, etc.).

µ: mean population,

αi: Effect of ith ration

βj: Effect of the jth additive

λ: Effect of the kth oil addition

(αβ)ij: Interaction of ith ration and jth additive effect

(βλ)jk: Effect of interaction between jthinci additive with kth oil type

(αλ)ik: Effect of interaction between ith ration with kth vegetable oil type

(αβλ)ijk: Effect of interaction between ith ration, jth additive with kth oil type

eijk: shows a random error.

Duncan Multiple Comparison test was used to compare the means if the differences between the applications or feed types were statistically significant. SPSS 22.0 statistical package program licensed by Ondokuz Mayıs University was used for statistical analysis.

3. Results

3.1. In vitro methane (CH4) production

In current study, R:C ratio (p = 0.002), NO3 (p < 0.001), oils addition (O) (p < 0.001), R:C ratio x NO3 interaction (p < 0.001), NO3 x O interaction (p = 0.007), R:C ratio x O interaction (p < 0.005), and R:C ratio x NO3 x O interaction (p = 0.004) were found to affect in vitro CH4 production (Table 3). After 24 hours of fermentation, the high-concentrate (40R:60C) content rations led to higher in vitro CH4 production. 40R:60C x NO3, 60R:40C x NO3, 40R:60C x NO3 x O, 60R:40C x NO3 x O treatment groups decreased in vitro CH4 production. But, 40R:60C x SO and 60R:40C x SO treatment groups led to lower in vitro CH4 production (compared to 40R:60C x HO and 60R:40C x HO treatment groups).

Table 3

Effects of NO3, O (HO and SO), and NO3 + O supplemented to different roughage:concentrate ratios on rumen fermentation properties.

Roughage: Concentrate

Feed additive

Oils

CH4

NH3

VFA

AA:PA

PP

pH

 

AA

PA

BA

TVFA

   

40R:60C

Control

FDO

13.29a

9.20a

49.57b

17.47d

15.51bc

89.11a

2.84b

36.19a

6.12bc

 

SO

11.05c

8.33b

46.04d

19.19b

14.53def

85.65de

2.40fg

29.25e

5.95ef

 

HO

12.34b

8.98a

48.44bc

18.58c

14.68de

87.63bc

2.61c

32.69c

5.99def

 

NO3

FDO

11.94b

8.35b

47.29cd

19.21b

16.89a

87.14bcd

2.46def

35.06b

6.08c

 

SO

7.18g

6.60f

43.24e

20.84a

15.62bc

83.37f

2.07i

27.69g

5.94f

 

HO

8.62e

7.93c

46.55d

19.57b

15.74b

85.70de

2.38g

32.06d

5.98de

 

60R:40C

Control

FDO

12.27b

8.34b

52.49a

16.77e

13.95fg

88.51ab

3.13a

28.31f

6.25a

 

SO

7.29g

7.28d

46.86d

18.55c

13.82g

84.39ef

2.53cd

26.81h

6.00d

 

HO

9.15e

7.96c

49.36b

17.31de

14.07fg

86.35cd

2.85b

27.19g

6.08c

 

NO3

FDO

9.82d

7.37d

51.19ab

16.80e

15.70b

89.93a

3.04ab

28.06f

6.16b

 

SO

7.35f

6.37f

43.85e

19.66b

14.41efg

81.78g

2.23h

25.36j

5.99de

 

HO

8.68c

6.87e

45.97d

18.47c

15.04cd

83.01fg

2.49de

26.06i

6.08c

 

Roughage:Concentrate (R:C)

40R:60C

12.44

8.23

46.86

19.14

15.50

86.43

2.46

32.16

6.01

 

60R:40C

9.09

7.37

48.45

17.93

14.50

85.66

2.72

26.96

6.09

 

Feed additives (FA)

Control

9.57

8.35

48.79

17.98

15.28

86.94

2.72

30.57

6.07

 

NO3

8.62

7.25

46.52

19.09

14.71

85.15

2.46

29.97

6.04

 

Oils

FDO

11.83

8.32

50.39

17.56

15.52

88.67

2.88

31.91

6.15

 

SO

8.22

7.15

45.00

19.56

14.60

83.79

2.31

27.28

5.97

 

HO

9.70

7.94

47.58

18.48

14.88

85.67

2.58

29.50

6.04

 

S.E.M.

0.306

0.149

0.481

0.207

0.155

0.429

0.54

71.559

0.015

 

Means effects

R:C

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

 

NO3

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

0.006

 

Oils

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

 

F:C × NO3

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

0.031

0.013

 

F:C × Oils

< 0.001

0.008

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

0.033

 

NO3 × Oils

0.015

0.004

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

0.004

0.008

 

F:C × NO3 × Oils

< 0.001

0.001

< 0.001

< 0.001

< 0.001

< 0.001

< 0.001

0.017

0.015

 
a. b. c… The averages shown with different letters in the same column are different from each other.
CH4: Methane (ml) NH3: Ammonia (mg/dl), AA: Acetic acid (mmol/l), PA: Propionic acid (mmol/l), BA: Butiric acid (mmol/l), PP: Protozoa Population (x104 ml), NO3: nitrate, HO: hazelnut oil, SO: soybean oil, S.E.M: Standard error of means, FDO: Feed-derived oil (with no oil addition).

3.2. NH3 concentration in the rumen fluid.

After 48 hours of incubation, R:C ratio (p < 0.001), NO3 (p < 0.001), O (p < 0.001), R:C x NO3 (p < 0.05), R:C ratio x O (p < 0.05), NO3 x O (p < 0.05) and R:C ratio x NO3 x O (p < 0.05) affected NH3 concentration. In the present study, it was found that NH3 concentration (p < 0.001) increased as concentrate level increased. But, 40R:60C x NO3, 60R:40C x NO3, 40R:60C x NO3 x oil, 60R:40C x NO3 x oil treatment groups led to the lowest NH3 concentration. Especially 40R:60C x NO3 x SO and 60R:40C x NO3 x SO treatment groups were found to have the lowest NH3 concentration. By the way, 40R:60C x urea x O and 60R:40C x urea x O treatment groups (control groups) had higher NH3 concentration.

3.3. pH, VFA, and AA: PA ratio

The pH values, concentration of total fatty acids (TVFA), AA, PA, BA, and AA:PA ratio were affected by R:C ratio, NO3, O, and R:C ratio x FA, R:C ratio x O, FA x O, R:C ratio x FA x O interactions (Table 3). In this study, 40R:60C (high concentrate) increased PA, BA, TVFA concentrations. However 60R:40C (high roughage) increased AA, AA:PA ratio and pH. By the way, 40R:60C x NO3, 60R:40C x NO3, 40R:60C x NO3 x oil, 60R:40C x NO3 x oil treatment groups (p < 0.001) increased PA and BA concentration, but decreased AA concentration, TVFA, AA:PA ratio, and pH (p < 0.001).

3.4. Protozoa population

Protozoa number was affected by R:C ratio (p < 0.001), NO3 (p < 0.001), O (p < 0.001), R:C ratio x NO3 (p < 0.05), R:C ratio x O (p < 0.001), NO3 x O (p < 0.05) and R:C ratio x NO3 x O interaction (p < 0.05). 40R:60C (high concentrate) increased protozoa number. But compared to 60R:40C (high roughage) decreased protozoa number. At the end of 48 hours of fermentation, 40R:60C x NO3, 60R:40C x NO3, 40R:60C x NO3 x oil, 60R:40C x NO3 x oil treatment groups decreased protozoa population more than other groups (p < 0.05). The decreasing effect of NO3 and Oils on protozoa population was more evident with 40R:60C x NO3 x SO and 60R: 40C x NO3 x SO treatment groups.

4. Discussion

4.1. CH4 production

A low in vitro CH4 production recorded in 60R:40C (high roughage) may be associated with the presence of high levels of secondary metabolites (saponin and condensed tannin) in Medicago sativa used as a roughage in this study (Kozlowska et al., 2020). Previously, Castro-Montoya et al. (2012) and recently Kozlowska et al. (2020) found that the plantes rich in saponin (quillaja plant, alfalfa or Medicago sativa) used in high level in diet (high roughage for example) decreased CH4 production compared to high concentrate. Likewise, in some studies, saponin has been found to reduce CH4 production (Morgavi et al., 2012; Jayanegara et al., 2014; Chen et al., 2019). A low in vitro CH4 production found in high roughage and high roughage can be explained by low NDF and high CP of all of diets used in this study (Table 2). However a lowest in vitro CH4 production found in a high roughage (60R:40C) compare to the high concentrate (40R:60C) is due to the presence of saponin and condensed tannins (in high level, Table 2) and the combined effect of secondary metabolites (Saponin and Condensed tannins), NO3 and oils. While some researchers found that a low condensed tannin (CT < 0.001%) content in alfalfa hay decreased CH4 production, some researchers reported a decrease in CH4 production due to the low NDF and a high crude protein content and a presence of secondary metabolites in alfalfa (Cheok et al., 2014; Rira et al., 2015; Moate et al., 2017; Szumacher -Strabel et al., 2019). These reports are consistent with our findings.

In current study, the effect of NO3 supplementation on in vitro CH4 production in the high concentrate and high roughage was different. In this study the increase of roughage level (60R:40C) led to a decrease of in vitro CH4 production. This can be explained by the presence of secondary metabolites (Saponin and Condensed tannin) and the effect of nitrate which act to reduce in vitro CH4 production. Previously an interaction was found between ration type (roughage/concentrate ratio) and CH4 reducing agents (such as nitrate) in cattle (Alvarez-Hess et al., 2019). This result is consistent with our findings.

In vitro CH4 production was high when a concentrate ratio was high in the diet and was gradually decrease with the high roughage ration in the diet. This could be due to that roughage (Medicago sativa) content a high secondary metabolites (Saponin and condensed tannin) which could explain the lower in vitro CH4 production when roughage fraction was maximun (60%) in the diets. A high in vitro CH4 production found in a high concentrate level, is consistent with some studies (Hristov et al., 2015; 2017; Moate et al., 2017; 2019; Olomonchi et al., 2022). In this study, a relationship between in vitro CH4 production, dietary starch rate and digestion can be established. This is in line with previous findings (Herrera-Saldana et al., 1990; McAllister et al., 1996; Alvarez-Hess et al., 2019).

In our study, it was found that NO3 + O added rations decreased in vitro CH4 production in the high concentrate (40R:60C) and high roughage (60R:40C) (p < 0.001). Some researchers reported that a combined effect of oils (rich MUFA or PUFA) and NO3 is an effective method to reduce CH4 production in rumen (Leng and Preston, 2010; Yang et al., 2016). It has been determined that a lower effect of NO3 on in vitro CH4 production is associated to NO3 and nitrite reducing microorganisms (Guo et al., 2009). Nitrate acts as a hydrogen acceptor. There are studies showing that nitrate has a significant inhibitory effect on CH4 production (El-Zaiat et al., 2014; Olijhoek et al., 2016). In the current study, the reduction of nitrate to nitrite and then to NH3 reduces H ions concentration required for the conversion of the CO2 to CH4 compound in the rumen and thus in vitro CH4 production decreases. Previously some studies reported the same result (Zhou et al., 2012; Liu et al., 2017). In addition, it has been determined that nitrite has a toxic effect on methanogens (Božic et al., 2009; Zhou et al., 2011). However, in this study the reducing effect of NO3 on CH4 is more evident in rations with a high roughage which is rich in condensed tannins or saponins. Similar results have been reported from an experiment by Pal et al. (2014).

In our study, while a high in vitro CH4 production was observed in 40R:60C (12.44 ml), a lower CH4 production (9.09 ml) was recorded in low a high roughage level (Table 3). This result is related to the increase in saponin and condensed tannins level and their effects on CH4 production in a high roughage level.

In our study, oils used were rich in PUFA (SO) and MUFA (HO). As expected, SO with a high PUFA content decreased in vitro CH4 production at a higher level than HO. This finding is consistent with studies reporting that the mitigation effect of oils on CH4 production is related to degree of unsaturation (Rodrigues et al., 2017; Vargas et al., 2017).

In high concentrate and high roughage, the higher negative effect of SO (rich in PUFA) in vitro CH4 production (compared to the effect of HO) is associated with a high presence of α-linolenic acid (C18:3 cis-9, cis-12, cis-15) and linoleic acid (C18:2 cis-9, cis-12) in SO. Previously, the effect of oils such as flaxseed and rapeseed rich in PUFA on CH4 mitigation was found by some researchers (Chung et al., 2011; Benchaar et al., 2015; Veneman et al., 2015). As a matter of fact, a lowering effect of oils rich in MUFA (oleic acid (C18:1)) on CH4 mitigation was found (Dong et al., 1997). Although, in the high concentrate and high roughage, HO decreased in vitro CH4 production. Likewise, in some studies canola oil (22% linoleic acid, 11% linolenic acid, and 54% oleic acid) caused a reduction of CH4 production (Dohme et al., 2000; Beauchemin and McGinn, 2015)). It was found that oil (rich in MUFA or PUFA) reduced the cellulolytic bacteria population, methanogenic bacteria, and then CH4 production (Freitas et al., 2018; Nur Atikah et al., 2018).

In present study, the decrease in CH4 production due to oils addition can be associated with the decrease in protozoa population. In the high concentrate and high roughage, it was determined that NO3 + O supplementation caused a higher decrease in CH4 production compared to the use of O and NO3 separately. This result is consistent with previous studies (Duthie et al., 2017; Villar et al., 2019). Likewise, Guyader et al. (2015) and Veneman et al. (2015) reported that CH4 production decreases when nitrate and flaxseed oil (high in MUFA) are added to ration.

4.2. NH3 concentration

In the current study, rumen NH3 values determined for rations used in high concentrate and high roughage are above the recommended minimum NH3 concentration (4.39 to 7.32 mmol/l) (Satter and Slyter, 1974), which is considered sufficient for maximum microbial growth rates. The high NH3 concentration found in the high concentrate (40R:60C) might be related to the high number of proteolytic bacteria in rumen. Because, proteolytic bacteria increase ruminal NH3 concentration by accelerating protein degradation in rumen. While this finding is in agreement with some studies (Kljak et al., 2017; Liu et al., 2019), it disagreed with other studies (Jadhav et al., 2017; Liu et al., 2018).

A high roughage (60R:40C) decreased NH3 concentration in present study. This finding can be associated with a high level of alfalfa (rich in saponins), which increased saponin level in ration. Saponin decreased or inhibited NH3 production. Our results were consistent with some previous studies (Belanche et al., 2016; Jadhav et al., 2018).

In our study, 40R:60C x NO3 and 60R:40C x NO3 treatment groups decreased NH3 (p < 0.001). NO3 is converted to nitrite, which has a toxic effect on rumen bacteria, and therefore NO3 addition reduces NH3 concentration at a high level compared to urea addition (control group). Nitrate alters the fermentation profile and decreases the NH3 production. However, the conversion rate of NO3 to NH3 in rumen is slower than urea to NH3. This can explained the high NH3 concentrattions found in control groups (urea supplementation) compared to other treatment groups.

Various studies investigated the effect of O (rich in MUFA or PUFA) supplementation on NH3 concentration. While in some studies oil supplementation had no effect (Jalc et al., 2005) on NH3 concentration, in some studies oil supplementation increased (Jalc et al., 2002) or decreased (Szumacher-Strabel et al., 2009; Doreau et al., 2017) NH3 concentration.

In this study, oils rich in PUFA (SO) or in MUFA (HO) associated with the R:C ratios (40R:60C, and 60R:40C) decreased NH3 concentration (p < 0.001). This can be explained by the presence of linolenic acid (SO) and oleic acid (HO). But the effect of SO (rich in PUFA) was more evident. In fact, the biohydrogenation of linoleic acid consumes more hydrogen (compare to biohydrogenation of oleic acid). Thus, in our study, the lack of hydrogen causes the decrease in NH3 production. Previously, while Bayat et al. (2017), and Kubelkova et al. (2018) found that flaxseed oil (rich in PUFA) compared to rapeseed oil (rich in MUFA) decreased the rumen pH and NH3 concentration at a high level, some researchers reported that Moringa oleifera oil rich in MUFA (oleic acid (74.99%), stearic acid (2.09%), linolenic acid (1.75%), and linoleic acid (1.27%)) increased rumen protected (by-pass) protein and decreased NH3 concentration (Gassenschmidt et al., 1995; Belewu et al., 2014).

In our study, a combined effect of NO3 + O supplementation link to the high concentrate or high rougahge decreased NH3 concentration (p < 0.05). However, combined effect of NO3 + SO (compared to NO3 + HO) was more evident on NH3 concentration in high concentrate or high roughage. In the same time, the biohydrogenation (due to O supplementation) and hydrogen sink reaction (due to NO3 supplementation) were happened to use the free hydrogen in rumen. Like that, NH3 production decreased because of lack of hydrogen. Previously, combined effect of NO3 + O supplementation was reported in some studies (Veneman et al., 2015 (NO3 + linseed oil supplementation); Villar et al., 2019 (NO3 + canola oil supplementation)).

4.3. pH, VFA, and AA: PA ratio

In the present study, pH values of rations used, are determined from the fluids remaining in the injectors after 48 hours of incubation. The pH values vary between 5.99 and 6.25 (Table 3). The pH difference in this study is due to R:C ratio. In this study while a high concentrate decreased pH, a high roughage increased pH. Firstly, this could be to that a high roughage (60R:40C) contained more NDF, ADF and cellulose than a high concentrate (40R:60C). Secondly, a decreased in pH due to the high concentrate can be associated to a high starch content which creates an environment to inhibit nitrate and nitrite metabolism. This means that a high concentrate provided sufficient energy for the microorganisms to convert nitrate to nitrite and then nitrite to NH3. For this reason, NH3 concentration was high in the high concentrate (Table 3).

Although, it was found that NO3 addition link to R:C ratios decreased pH values (p < 0.05). This finding is in agreement with some studies (Li et al., 2012; Villar et al., 2019). Likewise, rumen pH values found in our study are consistent with the value reported by Latham et al. (2016). A decline in pH observed due to NO3 supplementation indicates that microorganisms were not accustomed to digesting nitrate. It suggested that NO3 supplementation caused a dramatic change in rumen conditions.

In the present study, oil addition associated to R:C ratios decreased pH values. A decrease in ruminal pH, AA concentration, and CH4 production observed due to oil addition in our study can be associated with the degree of unsaturation of oils used (SO and HO). Some researchers have reported that oils addition reduces ruminal pH, AA concentration, and CH4 production with oils (rich in MUFA or in PUFA) supplementation (Wu et al., 2016; Majewska et al., 2017; Alvarez-Hess et al., 2019). However, it was found that SO (compared to HO) decreased significantly pH, AA concentration, and CH4 production (Compared to HO). This can be associated to the high level of linolenic acid in SO. This result is consistent with some studies (Russell and Wilson, 1996; Mertens, 1997). By the way, NO3 + O supplementation combine with R:C ratios decreased more pH, AA concentration, and CH4 production. This is associated to the biohydrogenation of unsaturated fatty acids (PUFA, and MUFA) provided by oil (SO, and HO), and hydrogen sink reaction (due to NO3 supplementation) which occurred in the same time.

In this experiment, TVFA, individual concentration of VFA (AA, PA, BA, and AA: PA ratio) were affected by R:C ratios, NO3, O, R:C ratios x NO3, R:C ratios x O, NO3 x O.

It was determined that a high concentrate decreased AA concentration, and AA: PP ratio (p < 0.001). An increase in PA, BA, TVFA concentration found in the high concentrate can be explained by the lowering pH due to the increase in lactic acid content derived from the high easily fermentable carbohydrates content of rations used, and an increase in carbohydrate fermentation. An increase in BA concentration can be also associated with the increase in ammonia concentration which inhibited bacterial growth and promotes a fermentation for BA production in this study. As it is known, VFA are produced as a result of microbial fermentation of carbohydrates in the rumen. However, the increase in AA, TVFA, and AA:PA ratio found in the high roughage, is associated with the increase in fiber content (in this case NDF and ADF). Depending on an increase in fibrous content of ration, ruminal hydrogen concentration used in the production of AA and in vitro CH4 increased. Our results are consistent with some studies (Kljak et al., 2017; Moate et al., 2017; Alende et al., 2019).

The increase in PA concentration due to NO3 addition can be explained by the competition between the mechanism of PA production and nitrate (for ammonia production). In other words, propionic acid producing bacteria population (Selenomonas ruminantium, Propionibacterium and Tessaracoccus) increased and they used free H ions present in the rumen to produce propionic acid. For this reason, hydrogen required for nitrate reduction (nitrite then ammonia) decreased. Consequently, PA concentration increased and NH3 concentration, AA and AA: PA ratio decreased in the rumen. But, a decrease in BA (due to NO3 supplementation) was caused by the rapid reduction of NO3 (to nitrite then ammonia) which use up the electrons needed for the production of BA. However, a high roughage level decreased BA concentration. This was due to the combined effect of tannin and NO3. Our results were consistent with some studies (van Zijderveld et al., 2011; Adejoro and Hassen, 2017; Wang et al., 2018).

In this study, the decrease in AA due to NO3 supplementation can be explained by the use of free hydrogen for production of NH3 and PA. Like this, hydrogen concentration required for the production of AA decreased. One of the possible reasons for the reduction in the concentration of AA due to the combined effect of NO3 and the two types of oil (MUFA or PUFA) is the use of free hydrogens for the production of PA and BA.

In our study, the effect of NO3 + O on VFA and AA: PA ratio changed according to the source of fatty acids (MUFA and PUFA). For that, NO3 + SO (rich in PUFA) associated with R:C ratios decreased AA concentration but it increased PA and BA concentrations. This result can be explained by the simultaneous effect of NO3 (hydrogen sinks) and the biohydrogenation of PUFA which consume more free hydrogen than the biohydrogenation of MUFA. A high concentrate level increased more the combined effect of NO3 + SO on AA, CH4, NH3, TVFA and AA:PA ratio. However, while NO3 stimulated the population of propionic acid-producing bacteria, the unsaturated fatty acids (PUFA and MUFA) in SO and HO used free hydrogens for biohydrogenation. Thus, the production of AA, CH4, NH3, TVFA and AA:PA ratio decreased. Our findings are in conformity with those found by Popova et al. (2017) and Villar et al. (2019). Our results showed that AA and CH4 were more decreased due to the combined effect of NO3 + SO which can be explained by biohydrogenation of PUFA and NO3 mechanism (transformation of NO3 to nitrite then to NH3) for obtaining PA having different and associative mechanism for using available hydrogen. Use of NO3 and O in the same time in the ration led to reduction in ruminal hydrogen concentration.

4.4. Protozoa population (PP)

In the current study, a high concentrate level increased the number of PP compared to the high roughage (p < 0.001). Previously, it was shown that a high concentrate can increase (Franzolin and Dehority, 1996; Lengowski et al., 2016) or decrease (Gozho et al., 2005; Khafipour et al., 2009; Hook et al., 2011) protozoa population.

However, the increase in roughage (rich in secondary metabolites: saponins and tannins) content of ration link to NO3 addition caused a decrease in the protozoa number. This can be explained by the combine effect of NO3 and saponins which acted negatively on protozoa population. Our findings are consistent with those of Lin et al. (2013).

In the present study, NO3 added rations associated with R:C ratios decreased PP. Otherwise, nitrite which come from a transformation of nitrate, inhibits rumen protozoa population and thus in vitro CH4 production. This is consistent with findings of Iwamoto et al. (2001).

Furthermore, in our study, there is a parallelism between NH3 concentration and PP in a high roughage level, and this finding was reported by some studies (Hu et al., 2005; Liu et al., 2018).

In our study, use of NO3 alone or in combination with HO (rich in MUFA: oleic acid) and SO (rich in linolenic acid and linoleic acid) reduced protozoa population. However, the combined effect of NO3 + SO decreased protozoa population more than individual use of NO3 and O. This can be explained by the simultaneous mitigation effect of NO3 and PUFA (linolenic acid and linoleic acid) on PP. Previously it was demonstrated that NO3 alone (Sar et al., 2005; Asanuma et al., 2015) or in combination with linseed oil (Veneman et al., 2015) or canola oil (Villar et al., 2019) decreased PP. While some authors notified a toxic effect of NO3 and lipids on protozoa PP (Morgavi et al., 2010), other researchers reported no significant effect on PP (Guyader et al., 2016).

Conclusion

In this study, a high roughage (because of a presence of saponins and condansed tannins in high level) decreased CH4, AA, NH3, TVFA, PP, and AA:PA ratio. While the 60R:40C associated to NO3-O alone or in combination decreased CH4, AA, TVFA, PP, and AA:PA ratio, it increased PA, BA. This study shows that NO3- and O (HO and SO) affect in vitro CH4 production, protozoa population, NH3 and VFA concentrations. The combination of NO3- and O (HO and SO) reduced acetic acid, protozoa population (thus in vitro CH4), and increased propionic acid and butyric acid more than individual use of nitrate and oils. Our study showed that a combined effect of nitrate and oils can be considered more avantageous to reduce methane production and protozoa population without a negatif effect on PA and BA concentrattions. 

The fermentation properties of rations supplemented with nitrate or oils have a potential to improve rumen fermentation. It has been found that the degree of unsaturated fat alone or in combination with NO3- decreases CH4 production and increases VFA. 

Declarations

Funding

This study was supported by Ondokuz Mayıs University as PYO. ZRT.1904.19.010 Scientific Research Project.

Conflit of interests

The authors declare that they have no conflict of interest.

Ethics Approval

Not applicable

Consent to participate

Not applicable

Consent for publication

Not applicable

Availability of data and material (data transparency)

Data on the parameters that were the subject of this study are available from the corresponding author on reasonable request.

Code availability (Software application or custom code)

Not applicable

Author’s contributions

Euloge O.A. OLOMONCHI and Ali V. GARIPOGLU conceived and designed the present study. Euloge O.A. OLOMONCHI conducted the literature search, analysed, interpreted data, and drafted the manuscript. The study was supervised by Euloge O.A. OLOMONCHI and Ali V. GARIPOGLU. All authors read and approved the final manuscript. 

References

  1. Adejoro, F. A., and Hassen, A., 2017. Effect of supplementing or treating Eragrostis curvula  hay with urea or nitrate on its digestibility and in vitro fermentation. South African Journal of Animal Science, 47(2), 168-177.
  2. Alende, M., Lascano, G. J., Jenkins, T. C., Andrae, J. G., 2019. Contrasting levels of fructose and urea adaded to an annual ryegrass based diet: effects on microbial protein synthesis, nutrient digestibility and fermentation parameters in continuous culture fermenters. Semiárida, 29(1), 33-41.
  3. Alvarez-Hess, P. S., Moate, P. J., Williams, S. R. O., Jacobs, J. L., Beauchemin, K. A.,  Hannah, M. C., ... Eckard, R. J., 2019. Effect of combining wheat grain with nitrate, fat or 3-nitrooxypropanol on in vitro methane production. Animal Feed Science and Technology, 256(11), 42-37.
  4. Asanuma, N., Yokoyama, S., Hino, T., 2015. Effects of nitrate addition to a diet on  fermentation and microbial populations in the rumen of goats, with special reference to S elenomonas ruminantium having the ability to reduce nitrate and nitrite. Animal Science Journal, 86(4), 378-384.
  5. Auffret, M. D., Stewart, R., Dewhurst, R. J., Duthie, C. A., Rooke, J.A., Wallace, R. J.,  Freeman, T. C., Snelling, T. J., Watson, M., Roehe, R., 2018. Identification, comparison, and validation of robust rumen microbial biomarkers for methane emissions using diverse Bos Taurus breeds and basal diets. Front. Microb, 8(2), 36-42. https://doi.org/10.3389/fmicb.2017.02642.
  6. Bayat, A. R., Tapio, I., Vilkki, J., Shingfield, K. J., Leskinen, H., 2017. Plant oilupplements  reduce methane emissions and improve milk fatty acid composition in dairy cows fed grass silage-based diets without affecting milk yield. J. Dairy Sci. 101 : 1-16. 
  7. Beauchemin, K. A., McGinn, S. M., Martinez, T. F., McAllister, T. A., 2007. Use of  condensed tannin extract from quebracho trees to reduce methane emissions from cattle. Journal of Animal Science, 85(8), 1990-1996.
  8. Beauchemin, K. A., McGinn, S. M. (2015). Methane emissions from feedlot cattle fed barley  or corn diets, Journal of Animal Science, 83, 653–661.
  9. Benchaar, C., Hassanat, F., Martineau, R., Gervais, R., 2015. Linseed oil supplementation to  dairy cows fed diets based on red clover silage or corn silage: Effects on methane production, rumen fermentation, nutrient digestibility, N balance, and milk production, Journal of Dairy Science, 98, 7993–8008.
  10. Belanche, A., Ramos‐Morales, E., and Newbold, C. J., 2016. In vitro screening of natural  feed additives from crustaceans, diatoms, seaweeds and plant extracts to manipulate rumen fermentation. Journal of the Science of Food and Agriculture, 96(9), 3069-3078.
  11. Belewu, M. A., Ahmed, M. A., Badmos, A. H. A., Esan, T. O., Abdulsalam, K. O., Odebisi  M. B., and Arise, A. K., 2014. Effect of different levels of moringa oleifera oil on performance characteristics of pregnant goat. Nigerian J. of Agric., Food and Environ., 10(2):29-33.
  12. Boadi, D., Benchaar, C., Chiquette, J., Masse, D., 2004. Mitigation strategies to reduce  enteric methane emissions from dairy cows: update review, Canadian Journal of Animal Science, 84, 319-335.
  13. Bougouin, A., Ferlay, A., Doreau, M., Martin, C., 2018. Effects of carbohydrate type or  bicarbonate addition to grass silage-based diets on enteric methane emissions and milk fatty acid composition in dairy cows. Journal of dairy science, 101(7), 6085-6097.
  14. Božic, A. K., Anderson, R. C., Carstens, G. E., Ricke, S. C., Callaway, T. R., Yokoyama, M.  T., ... & Nisbet, D. J., 2009. Effects of the methane-inhibitors nitrate, nitroethane, lauric acid, Lauricidin and the Hawaiian marine algae Chaetoceros on ruminal fermentation in vitro. Bioresource technology, 100(17), 4017-4025.
  15. Brandt, R. T., Klopfenstein, T. J., 1986a. Evaluation of alfalfa-corn cob associative action. I.  Interaction between alfalfa hay and ruminal escape protein on growth of lambs and steers. Journal of Animal Science 63, 894-901.
  16. Brandt, R. T., Klopfenstein, T. J., 1986b. Evaluation of alfalfa-corn cob associative action. II.  Comparative tests of alfalfa hay as a source of ruminal degradable protein. Journal of Animal Science 63, 902-910.
  17. Castro-Montoya, J., De Campeneere, S., Van Ranst, G., & Fievez, V., 2012. Interactions  between methane mitigation additives and basal substrates on in vitro methane and VFA production. Animal Feed Science and Technology, 176(1-4), 47-60.
  18. Chen, L., Dong, Z., Li, J., & Shao, T., 2019. Ensiling characteristics, in vitro rumen  fermentation, microbial communities and aerobic stability of lowdry matter silages produced with sweet sorghum and alfalfa mixtures. Journal of the Science of Food and Agriculture, 99(5), 2140-2151.
  19. Cheok, C. Y., Salman, H. A. K., Sulaiman, R., 2014. Extraction and quantification of  saponins: a review. Food Res. Int. 59, 16-40.
  20. Chung, Y. H., He, M. L., McGinn, S. M., McAllister, T. A., Beauchemin, K. A., 2011.  Linseed suppresses enteric methane emissions from cattle fed barley silage, but not from those fed grasshay, Animal Feed Science and Technology, 166-167, 321-329.
  21. Dong, Y., Bae, H. D., McAllister, T. A., Mathison, G. W., Cheng, K. J., 1997. Lipid-induced  depression of methane production and digestibility in the artificial rumen system (RUSITEC), Canadian Journal of Animal Science, 77(2), 269-278.
  22. Doreau, M., Arturo‐Schaan, M. and Laverroux, S., 2017. Garlic oil reduces ruminal fatty acid  biohydrogenation in vitro. European Journal of Lipid Science and Technology, 119(4) 1500388.
  23. Duthie, C. A., Haskell, M. J., Hyslop, J. J., Waterhouse, A., Wallace, R. J., Roehe, R., &  Rooke, J. A., 2017. The impact of divergent breed types and diets on methane emissions, rumen characteristics and performance of finishing beef cattle. Animal, 11(10), 1762-1771.
  24. El-Zaiat, H. M., Araujo, R. C., Soltan, Y. A., Morsy, A. S., Louvandini, H., Pires, A. V., ... &  Abdalla, A. L., 2014. Encapsulated nitrate and cashew nut shell liquid on blood and rumen constituents, methane emission, and growth performance of lambs. Journal of animal science, 92(5), 2214-2224.
  25. Franzolin, R., Dehority, B. A., 1996. Effect of prolonged high-concentrate feeding on  ruminal protozoa concentrations. Journal of animal science, 74(11), 2803-2809.
  26. Freitas, D. S., Terry, S. A., Ribeiro, R. S., Pereira, L. G. R., Tomich, T. R., Machado, F. S.,  Campos, M. M., Corrêa, P. S., Abdalla, A. L., Maurício, R. M., Chaves, A. V., 2018. Unconventional vegetable oils for a reduction of methanogenesis and modulation of ruminal fermentation. Front. Vet. Sci. 5:201. doi:10.3389/fvets.2018.00201.
  27. Gassenschmidt, U., Jany, K. D., Tauscher, B., and Niebergall, H., 1995. Isolation and  characterization of a flocculating protein from Moringa oleifera Lam. Biochimica Biophysica Acta, 1243: 477 – 481.
  28. Goel, G., Makkar, H. P. S., Becker, K., 2008. Changes in microbial community structure,  methanogenesis and rumen fermentation in response to saponinrich fractions from different plant materials. Journal of Applied Microbiology, 105(3), 770-777.
  29. Gozho, G. N., Plaizier, J. C., Krause, D. O., Kennedy, A. D., & Wittenberg, K. M., 2005.  Subacute ruminal acidosis induces ruminal lipopolysaccharide endotoxin release and triggers an inflammatory response. Journal of Dairy Science, 88(4), 1399–1403. https://doi.org/10.3168/jds.S0022-0302(05)72807-1.
  30. Guo, W. S., Schaefer, D. M., Guo, X. X., Ren, L. P., Meng, Q. X., 2009. Use of nitrate- nitrogen as a sole dietary nitrogen source to inhibit ruminal methanogenesis and to improve microbial nitrogen synthesis in vitro, Asian-Australasian Journal of Animal Sciences, 22(4), 542-549.
  31. Guyader, J., Eugène, M., Meunier, B., Doreau, M., Morgavi, D. P., Silberberg, M., ... & 
  32. Martin, C., 2015. Additive methane-mitigating effect between linseed oil and nitrate fed to cattle, Journal of Animal Science, 93(7), 3564-3577.
  33. Guyader, J., Janzen, H. H., Kroebel, R., Beauchemin, K. A., 2016. Forage use to improve  environmental sustainability of ruminant production, Journal of animal science, 94(8), 3147-3158.
  34. Herrera-Saldana, R. E., Huber, J. T., Poore, M. H., 1990. Dry matter, crude protein, and starch degradability of five cereal grains1. J Dairy Sci. 73: 2386–93.  doi:10.3168/jds.S0022-0302(90)78922-9
  35. Hook, S. E., Steele, M. A., Northwood, K. S., Dijkstra, J., France, J., Wright, A. D., & 
  36. McBride, B. W., 2011. Impact of subacute ruminal acidosis (SARA) adaptation and recovery on the density and diversity of bacteria in the rumen of dairy cows. FEMS Microbiology Ecology, 78(2), 275–284. https://doi.org/10.1111/j.1574-6941.2011.01154.x
  37. Hristov, A. N., Oh, J., Firkins, J. L., Dijkstra, J., Kebreab, E., Waghorn, G., ... & Gerber, P. J.,  2013. Special topics Mitigation of methane and nitrous oxide emissions from animal operations: I. A review of enteric methane mitigation options. Journal of animal science, 91(11), 5045-5069.
  38. Hristov, A. N., Oh, J., Giallongo, F., Frederick, T. W., Harper, M. T., Weeks, H. L., ... & 
  39. Kindermann, M., 2015. An inhibitor persistently decreased enteric methane emission from dairy cows with no negative effect on milk production. Proceedings of the National Academy of Sciences, 112(34), 10663-10668.
  40. Hristov, A. N., Harper, M., Meinen, R., Day, R., Lopes, J., Ott, T., ... & Randles, C. A., 2017.  Discrepancies and uncertainties in bottom-up gridded inventories of livestock methane emissions for the contiguous United States. Environmental science & technology, 51(23), 13668-13677.
  41. Hu, W., Liu, J., Ye, J., Wu, Y., Guo, Y., 2005. Effect of tea saponin on rumen fermentation  in vitro. Anim. Feed Sci. Technol. 120, 333-339.
  42. Hunt, C.W., Paterson, J. A., Williams, J. E., 1985. Intake and digestibility of alfalfa-tall  fescue combination diets fed to lambs. Journal of Animal Science 60, 301-312.
  43. IPCC, Field, C. B. (Ed.)., 2014. Climate change 2014–Impacts, adaptation and vulnerability:  Regional aspects. Cambridge University Press.
  44. Iwamoto, M., Asanuma, N., Hino, T., 2001. Effects of pH and electron donors on nitrate  and nitrite reduction in ruminal microbiota. Nihon Chikusan Gakkaiho, 72(2), 117-125.
  45. Jadhav, R. V., Kannan, A., Bhar, R., Sharma, O. P., Bhat, T. K., Gulati, A., ... & Sharma, V.  K., 2017. Effect of tea (Camellia sinensis) seed saponin supplementation on growth performance, nutrient utilization, microbial protein synthesis and hemato-biochemical attributes of Gaddi Goats. Animal Nutrition and Feed Technology, 17(2), 255-268.
  46. Jadhav, R. V., Kannan, A., Bhar, R., Sharma, O. P., Gulati, A., Rajkumar, K., ... & Verma, M.  R., 2018. Effect of tea (Camellia sinensis) seed saponins on in vitro rumen fermentation, methane production and true digestibility at different forage to concentrate ratios. Journal of Applied Animal Research, 46(1), 118-124.
  47. Jalč, D., Kišidayová, S., & Nerud, F., 2002. Effect of plant oils and organic acids on rumen  fermentation in vitro. Folia microbiologica, 47(2), 171-177.
  48. Jalč, D., & Čertík, M., 2005. Effect of microbial oil, monensin and fumarate on rumen  fermentation in artificial rumen. Czech J. Anim. Sci, 50(10), 467-472.
  49. Jayanegara, A., Wina, E., & Takahashi, J., 2014. Meta-analysis on methane mitigating  properties of saponin-rich sources in the rumen: influence of addition levels and plant sources. Asian-Australasian journal of animal sciences, 27(10), 1426.
  50. KanthaRaju, M., Jagadeeswary, V., Satyanarayan, K., Veeranna, K., Rajeshwari, Y., Nagaraj,  C., Shilpa Shree, J., 2018. Intensive Cultivation of Medicago sativa for Sustainable Milk Production an Action Oriented Approach. International Journal of Livestock Research, 8(4), 101-108.
  51. Khafipour, E., Krause, D. O., & Plaizier, J. C., 2009. Alfalfa pellet-induced subacute ruminal  acidosis in dairy cows increases bacterial endotoxin in the rumen without causing inflammation. Journal of Dairy Science, 92(4), 1712–1724. 
  52. Kljak, K., Pino, F., Heinrichs, A. J., 2017. Effect of forage to concentrate ratio with sorghum  silage as a source of forage on rumen fermentation, N balance, and pürine derivative excretion in limit-fed dairy heifers. J. Dairy Sci. 100: 213-223.
  53. Koukoura, Z., Kykriazopoulos, A. P., & Parissi, Z. M., 2009. Growth characteristics and  nutrient content of some herbaceous species under shade and fertilization. Spanish Journal of Agricultural Research, (2), 431-438.
  54. Kozlowska, M., Cieślak, A., Jóźwik, A., El‐Sherbiny, M., Stochmal, A., Oleszek, W., ... &  Szumacher-Strabel, M., 2020. The effect of total and individual alfalfa saponins on rumen methane production. Journal of the Science of Food and Agriculture, 100(5), 1922-1930.
  55. Kuchenmeister, K., Kuchenmeister, F., Kayser, M., WRAGE, M. N., & Isselstein, J. (2013).  Influence of drought stress on nutritive value of perennial forage legumes. Int. J., Plant Prod., 7(4): 693-710.
  56. Kubelková, P., Jalč, D., Jančík, F., Homolka, P., 2018. In vitro ruminal fermentation and fatty  acid production by various oil seeds. South African Journal of Animal Science, 48(3), 526. doi:10.4314/sajas. v48i3.13. 
  57. Kubkomawa, I. H., Ogundu, M. A., Okoli, I. C., & Udedibie, A. B. I., 2017. Biochemical  profiling of the values of dry season feed resources in pastoralcattle environment in Nigeria. International Journal of Research in Agriculture and Forestry, 4(4), 23-36.
  58. Latham, E. A., Anderson, R. C., Pinchak, W. E., & Nisbet, D. J., 2016. Insights on  alterations to the rumen ecosystem by nitrate and nitrocompounds. Frontiers in microbiology, 7, 228.
  59. Leng, R. A., 1990. Factors affecting the utilization of 'poor-quality' forages by ruminants  particularly under tropical conditions. Nutrition Research Reviews 3, 277-303. 
  60. Leng, R. A., & Preston, T. R., 2010. Further considerations of the potential of nitrate as a  high affinity electron acceptor to lower enteric methane production in ruminants. Livestock Research for Rural Development, 22 (221).
  61. Lengowski, M. B., Zuber, K. H., Witzig, M., Mohring, J., Boguhn, J., & Rodehutscord, M.,  2016. Changes in rumen microbial community composition during adaption to an in vitro system and the im pact of different forages. PLoS One, 11(2), e0150115. 
  62. Li, L., Davis, J., Nolan, J., & Hegarty, R., 2012. An initial investigation on rumen  fermentation pattern and methane emission of sheep offered diets containing urea or nitrate as the nitrogen source. Animal Production Science, 52(7), 653-658
  63. Lin, M., Schaefer, D. M., Zhao, G. Q., & Meng, Q. X., 2013. Effects of nitrate adaptation by  rumen inocula donors and substrate fiber proportion on in vitro nitrate disappearance, methanogenesis, and rumen fermentation acid. Animal: an international journal of animal bioscience, 7(7), 1099.
  64. Liu, K. H., Niu, Y., Konishi, M., Wu, Y., Du, H., Chung, H. S., ... Ishida, T., 2017. Discovery  of nitrate CPK NLP signalling in central nutrient–growth networks, Nature, 545(7654), 311.
  65. Liu, C., Qu, Y. H., Guo, P. T., Xu, C. C., Ma, Y., Luo, H. L., 2018. Effects of dietary  supplementation with alfalfa (Medicago sativa L.) on lamb growth performance, nutrient digestibility, and plasma parameters. Animal Feed Science and Technology 236 (2018) 98–106.
  66. Liu, X., Xu, Q., Wang, D., Wu, Y., Yang, Q., Liu, Y., ... & Yang, G., 2019. Unveiling the  mechanisms of how cationic polyacrylamide affects short-chain fatty acids accumulation during long-term anaerobic fermentation of waste activated sludge. Water research, 55, 142-151.
  67. Lloyd, Hillman, K., Yarlett, A., Williams, A.G. Hydrogen Production by Rumen Holotrich Protozoa: Effects of Oxygen and Implications for Metabolic Control by In Situ Conditions. The Journal of Protozoology,36, 2, 205-213.
  68. Majewska, M.P., Miltko, R., Bełżecki, G., Skomiał, J., Kowalik, B., 2017. Supplementation  of rapeseed and linseed oils to sheep rations: effects on ruminal fermentation characteristics and protozoal populations. Czech J. Anim. Sci., 62, 527-538.
  69. Martin, C., Morgavi, D., & Doreau, M., 2010. Methane mitigation in ruminants: from  microbe to the farm scale. Animal, 4(3), 351-365.
  70. Mertens, D. R., 1997. Creating a system for meeting the fiber requirements of dairy cows.  Journal of dairy science, 80(7), 1463-1481.
  71. McAllister, T. A., Cheng, K. J., Okine, E. K., & Mathison, G. W., 1996. Dietary,  environmental and microbiological aspects of methane production in ruminants. Canadian Journal of Animal Science, 76(2), 231-243.
  72. McGinn, S. M., Bodas, K. A., Coates, T., Colombatto, D., 2004. Methane emissions from  beef cattle: Effects of monensin, sunflower oil, enzymes, yeast, and fumaric acid, Journal of animal science, 82(11), 3346-3356.
  73. Moate, P. J., Williams, S. R. O., Jacobs, J. L., Hannah, M. C., Beauchemin, K. A., Eckard, R.  J., & Wales, W. J., 2017. Wheat is more potent than corn or barley for dietary mitigation of enteric methane emissions from dairy cows. Journal of Dairy Science, 100(9), 7139-7153.
  74. Moate, P. J., Williams, S. R. O., Deighton, M. H., Hannah, M. C., Ribaux, B. E., Morris, G.  L., ... & Wales, W. J., 2019. Effects of feeding wheat or corn and of rumen fistulation on milk production and methane emissions of dairy cows. Animal Production Science, 59(5), 891-905.
  75. Morgavi, D. P., Martin, C., Jouany, J. P., and Ranilla, M. J., 2012. Rumen protozoa and  methanogenesis: not a simple cause effect relationship. The British Journal of Nutrition 107, 388–397.
  76. Muir, J. P., Ocumpaugh, W. R, Read, J. C., 2003. Spring forage yield and nutritive value of Texas black medic accessions. Agron. J., 95 (4): 908-912.
  77. Nur Atikah, I., Alimon, A. R., Yaakub, H., Abdullah, N., Jahromi, M. F., Ivan, M., Samsudin,  A. A., 2018. Profiling of rumen fermentation, microbial population and digestibility in goats fed with dietary oils containing different fatty acids. BMC Veterinary Research, 14:344. 
  78. Olijhoek, D. W., Hellwing, A. L. F., Brask, M., Weisbjerg, M. R., Højberg, O., Larsen, M. K.,  Lund, P., 2016. Effect of dietary nitrate level on enteric methane production, hydrogen emission, rumen fermentation, and nutrient digestibility in dairy cows. Journal of Dairy Science, 99(8), 6191-6205.
  79. Olomonchi, O. A. E., Kiliç, Ü., Garipoğlu, A. V., Erişek, A., 2019. Comparison of forage quality and in vitro digestibilities of fodder beet (Beta vulgaris var. rapa) fresh material and silage. Journal of Scientific and Engineering Research, 6 (11), 172-178.
  80. Olomonchi, O. A. E., Garipoğlu, A. V., Nuh, O., Adem, K., 2022. Nutritional values and in vitro fermentation parameters of some fodder species found in two rangeland areas in the Republic of Benin. Turk J Vet Anim Sci (2022) 46: 88-94. doi:10.3906/vet-2101-69
  81. Ørskov, E. R., Liu, J. X., Chen, X. B., 1999. Optimization of steam treatment as a method for  upgrading rice straw as feeds. Animal Feed Science and Technology 76, 345-357.
  82. Pal, K., Patra, A. K., Sahoo, A., & Mandal, G. P., 2014. Effect of nitrate and fumarate in  Prosopis cineraria and Ailanthus excelsa leaves-based diets on methane production and rumen fermentation. Small Ruminant Research, 121(2-3), 168-174.
  83. Paterson, J. A., Klopfenstein, T. J., Britton, R. A., 1982. Digestibility of sodium hydroxide- treated crop residues when fed with alfalfa hay. Journal of Animal Science 54, 1056-1066.
  84. Popova, M., Guyader, J., Silberberg, M., Seradj, A. R., Saro, C., Bernard, A., ... & Morgavi,  D. P. (2019). Changes in the rumen microbiota of cows in response to dietary supplementation with nitrate, linseed, and saponin alone or in combination. Applied and environmental microbiology, 85(4).
  85. Ranilla, M. J., Carro, M. D., Valdés, C., Giráldez, F. J., & López, S., 1997. A comparative  study of ruminal activity in Churra and Merino sheep offered alfalfa hay. Animal Science, 65(1), 121-128.
  86. Rira, M., Chentli, A., Boufenera, S., Bousseboua, H., 2015. Effects of plants containing  secondary metabolites on ruminal methanogenesis of Sheep in vitro. Energy Procedia 74, 15-24.
  87. Rodrigues, J. P. P., Ramin, M., Huhtanen, P., Aru, F., Detmann, E., Marcondes, M. I., 2017.  Effect of soyabean oil supplementation and forage type on methane production and fibre digestibility using the in vitro gas production system, Grass Forage Sci, 00: 1-13.
  88. Russell, J. B., & Wilson, D. B., 1996. Why are ruminal cellulolytic bacteria unable to digest  cellulose at low pH? Journal of dairy science, 79(8), 1503-1509.
  89. Sar, C., Mwenya, B., Pen, B., Takaura, K., Morikawa, R., Tsujimoto, A., Kuwaki, K., Isogai,  N.,;Shinzato, I., Asakura, Y., Toride, Y., Takahashi, J., 2005. Effect of ruminal administration of Escherichia coli wild type or a genetically modifi ed strain with enhanced high nitrite reductase activity on methane emission and nitrate toxicity in nitrate-infused sheep, Br. J. Nutr. 94:691-697. 
  90. Satter, L. D., Slyter, L. L., 1974. Effect of ammonia concentration on rumen microbial proteinproduction in vitro, Brit. J. Nutr, 32:199-208.
  91. Sharifi, M., Taghizadeh, A., Khadem, A. A., Hosseinkhani, A., Mohammadzadeh, H., 2018. Effects of nitrate supplementation and forage level on gas production, nitrogen balance and dry-matter degradation in sheep, Animal Production Science. Animal Production Science, 59(3), 515-523.
  92. Szumacher-Strabel, M., Cieslak, A., Nowakowska, A., 2009. Effect of oils rich in linoleic  acid on in vitro rumen fermentation parameters of sheep, goats and dairy cows, J. Anim. Feed Sci, 18(3), 440-452
  93. Szumacher-Strabel, M., Stochmal, A., Cieslak, A., Kozłowska, M., Kuznicki, D., Kowalczyk,  M., & Oleszek, W., 2019. Structural and quantitative changes of saponins in fresh alfalfa compared to alfalfa silage. Journal of the Science of Food and Agriculture, 99(5), 2243-2250. Ungerfeld, E. M., & Kohn, R. A., 2006. The role of thermodynamics in the control of ruminal  fermentation. Ruminant physiology: digestion, metabolism and impact of nutrition on gene expression, immunology and stress, 55-85. Van Zijderveld, S. M., Gerrits, W. J. J., Dijkstra, J., Newbold, J. R., Hulshof, R. B. A., & 
  94. Perdok, H. B., 2011. Persistency of methane mitigation by dietary nitrate supplementation in dairy cows. Journal of Dairy Science, 94(8), 4028-4038.
  95. Vargas, J. E., Andrés, S., Snelling, T. J., López-Ferreras, L., Yáñez-Ruíz, D. R., García-
  96. Estrada, C., & López, S., 2017. Effect of sunflower and marine oils on ruminal microbiota, in vitro fermentation and digesta fatty acid profile. Frontiers in microbiology, 8, 1124.
  97. Veneman, J. B., Muetzel, S., Hart, K. J., Faulkner, C. L., Moorby, J. M., Perdok, H. B., &  Newbold, C. J., 2015. Does dietary mitigation of enteric methane production affect rumen function and animal productivity in dairy cows? PloS one, 10(10), e0140282.
  98. Villar, M. L., Hegarty, R. S., Nolan, J. V., Godwin, I. R., & McPhee, M., 2019. The effect of  dietary nitrate and canola oil alone or in combination on fermentation, digesta kinetics and methane emissions from cattle. Animal Feed Science and Technology, 259, 114294.
  99. Wang, R., Wang, M., Ungerfeld, E. M., Zhang, X. M., Long, D. L., Mao, H. X., ... & Tan, Z.  L., 2018. Nitrate improves ammonia incorporation into rumen microbial protein in lactating dairy cows fed a low-protein diet. Journal of dairy science, 101(11), 9789-9799.
  100. Wu, D., Xu, L., Tang, S., Guan, L., He, Z., Guan, Y., ... & Wang, M., 2016. Influence of oleic acid on rumen fermentation and fatty acid formation in vitro. PLoS One, 11(6), e0156835.
  101. Yang, C., Rooke, J. A., Cabeza, I., and Wallace, R. J., 2016. Nitrate and inhibition of ruminal  methanogenesis: Microbial ecology, obstacles, and opportunities for lowering methane emissions from ruminant livestock. Front. Microbiol. 7:132. doi: 10.3389/fmicb.2016.00132
  102. Yang, K., Wei, C., Zhao, G. Y., Xu, Z. W., & Lin, S. X., 2017. Effects of dietary  supplementing tannic acid in the ration of beef cattle on rumen fermentation, methane emission, microbial flora and nutrient digestibility. Journal of animal physiology and animal nutrition, 101(2), 302-310.
  103. Zhou, Z., Meng, Q., Yu, Z., 2011. Effects of methanogenic inhibitors on methane production  and abundance of methanogen and cellulolytic bacteria in in-vitro ruminal cultures. Appl. Environ. Microbiol. 77, 2634–2639.
  104. Zhou, Z., Yu, Z., & Meng, Q., 2012. Effects of nitrate on methane production, fermentation,  and microbial populations in in vitro ruminal cultures. Bioresource Technology, 103(1), 173-179.