Hostname: page-component-7c8c6479df-ws8qp Total loading time: 0 Render date: 2024-03-28T20:35:03.764Z Has data issue: false hasContentIssue false

Propionate precursors and other metabolic intermediates as possible alternative electron acceptors to methanogenesis in ruminal fermentation in vitro

Published online by Cambridge University Press:  08 March 2007

C. J. Newbold*
Affiliation:
Institute of Rural Sciences, University of Wales, Aberystwyth SY23 3AL, UK
S. López
Affiliation:
Department of Animal Production, University of León E-24071, León, Spain
N. Nelson
Affiliation:
Rowett Research Institute, Bucksburn, Aberdeen AB21 9SB, UK
J. O. Ouda
Affiliation:
ADAS Feed Evaluation and Nutritional Sciences, Alcester Road, Stratford-upon-Avon CV37 9RQ, UK
R. J. Wallace
Affiliation:
Rowett Research Institute, Bucksburn, Aberdeen AB21 9SB, UK
A. R. Moss
Affiliation:
ADAS Feed Evaluation and Nutritional Sciences, Alcester Road, Stratford-upon-Avon CV37 9RQ, UK
*
*Corresponding author: Dr C. J. Newbold, fax +44 (0) 1970 611264, email cjn@aber.ac.uk
Rights & Permissions [Opens in a new window]

Abstract

Core share and HTML view are not available for this content. However, as you have access to this content, a full PDF is available via the ‘Save PDF’ action button.

Fifteen potential precursors of propionate were tested for their ability to decrease CH4 production by ruminal fluid in vitro. Sodium acrylate and sodium fumarate produced the most consistent effects in batch cultures, with 50 % of the added precursors being fermented to propionate and CH4 production decreasing by between 8 and 17 %, respectively. Additives were more effective when added as free acids, but this also decreased the pH and may have inhibited fibre digestion. Changing the dietary substrate from predominantly grass hay to predominantly concentrate had no influence on the effectiveness of acrylate and fumarate. In an in vitro fermentor (the rumen simulating technique, Rusitec) with a grass hay—concentrate (50:50, w/w) diet as substrate, both compounds were again fermented to propionate (33 and 44 % conversion to propionate, respectively). However, fumarate appeared more effective as a H2 sink compound. It was calculated to capture 44 % of the H2 previously used for CH4 formation compared with a 22 % capture of H2 with acrylate. Fumarate also caused a stimulation in fibre digestion. Thus, sodium fumarate was the preferred propionate precursor for use as a feed ingredient to decrease CH4 emissions from ruminants.

Type
Research Article
Copyright
Copyright © The Nutrition Society 2005

References

Asanuma, N & Hino, T (2000) Activity and properties of fumarate reductase in ruminal bacteria. J Gen Appl Microbiol 46, 119125.CrossRefGoogle ScholarPubMed
Asanuma, N, Iwamoto, M & Hino, T (1999) Effect of the addition of fumarate on methane production by ruminal microorganisms in vitro. J Dairy Sci 82, 780787.CrossRefGoogle ScholarPubMed
Baldwin, RL & Kim, WY (1993) Lactation. In Quantitative Aspects of Ruminant Digestion and Metabolism, pp. 433451[Forbes, JM and France, J, editors]. Wallingford, UK: CAB International.Google Scholar
Bayaru, E, Kanda, S & Kamada, T (2001) Effect of fumaric acid on methane production, rumen fermentation and digestibility of cattle fed roughage alone. Anim Sci J 72, 139146.Google Scholar
Bryant, MP (1972) Commentary on the Hungate technique for culture of anaerobic bacteria. Am J Clin Nutr 25, 13241328.CrossRefGoogle ScholarPubMed
Callaway, TR & Martin, SA (1996) Effects of organic acid and monensin treatment on in vitro mixed ruminal micro-organisms fermentation of cracked corn. J Anim Sci 74, 19821989.CrossRefGoogle Scholar
Carro, MD & Ranilla, MJ (2003 a) Effect of the addition of malate on in vitro rumen fermentation of cereal grains. Br J Nutr 89, 181188.CrossRefGoogle ScholarPubMed
Carro, MD & Ranilla, MJ (2003 b) Influence of different concentrations of disodium fumarate on methane production and fermentation of concentrate feeds by rumen micro-organisms in vitro. Br J Nutr 90, 617623.CrossRefGoogle ScholarPubMed
Czerkawski, JW & Breckenridge, G (1977) Design and development of a long term rumen simulation technique (Rusitec). Br J Nutr 38, 371384.CrossRefGoogle ScholarPubMed
Demeyer, D & Fievez, V (2000) Ruminants and environment: methanogenesis. Ann Zootech 49, 95112.CrossRefGoogle Scholar
Demeyer, DI & Henderickx, HK (1967) Competitive inhibition of in vitro methane production by mixed rumen bacteria. Arch Int Physiol Biochim 75, 157159.Google ScholarPubMed
Demeyer, DI, Van Nevel, CJ (1975) Methanogenesis, an integrated part of carbohydrate fermentation and its control. In Digestion and Metabolism in the Ruminant, pp. 366382[McDonald, IW and Warner, ACI, editors]. Armidale, Australia: University of New England Publishing Unit.Google Scholar
García-López, PM, Kung, L & Odom, JM (1996) In vitro inhibition of microbial methane production by 9,10-anthraquinone. J Anim Sci 74, 22762284.CrossRefGoogle ScholarPubMed
Goering, HK & Van Soest, PJ (1970) Forage Fiber Analyses (Apparatus, Reagents, Procedures and Some Applications). USDA Handbook no.375. Washington DC: USDA.Google Scholar
Hobson, PN (1969) Rumen bacteria. Methods Microbiol 3, 133159.CrossRefGoogle Scholar
Hungate, RE (1969) A roll tube method for cultivation of strict anaerobes. Methods Microbiol 3, 117132.CrossRefGoogle Scholar
Jalc, D & Ceresnakova, Z (2002) Effect of plant oils and malate on rumen fermentation in vitro. Czech J Anim Sci 47, 106111.Google ScholarPubMed
Kung, L Jr, Smith, KA, Smagala, AM, Endres, KM, Bessett, CA, Ranjit, NK & Yaissle, J (2003) Effects of 9,10 anthraquinone on ruminal fermentation, total-tract digestion, and blood metabolite concentrations in sheep. J Anim Sci 81, 323328.CrossRefGoogle ScholarPubMed
López, S, McIntosh, FM, Wallace, RJ & Newbold, CJ (1999 a) Effect of adding acetogenic bacteria on methane production by mixed rumen microorganisms. Anim Feed Sci Technol 78, 19.CrossRefGoogle Scholar
López, S, Valdes, C, Newbold, CJ & Wallace, RJ (1999 b) Influence of sodium fumarate on rumen fermentation in vitro. Br J Nutr 81, 5964.CrossRefGoogle ScholarPubMed
McDougall, EI (1948) Studies on ruminal saliva 1. The composition and output of sheep's saliva. Biochem J 43, 99109.CrossRefGoogle ScholarPubMed
Mann, SO (1968) An improved method for determining cellulolytic activity in anaerobic bacteria. J Appl Bacteriol 31, 241244.CrossRefGoogle Scholar
Martin, SA (1998) Manipulation of ruminal fermentation with organic acids: a review. J Anim Sci 76, 31233132.CrossRefGoogle ScholarPubMed
Martin, SA & Park, CM (1996) Effect of extracellular hydrogen on organic acid utilization by the ruminal bacterium Selenomonas ruminantium. Curr Microbiol 32, 327331.CrossRefGoogle ScholarPubMed
Martin, SA, Streeter, MN, Nisbet, DJ, Hill, GM & Williams, SE (1999) Effects of DL-malate on ruminal metabolism and performance of cattle fed a high-concentrate diet. J Anim Sci 77, 10081015.CrossRefGoogle ScholarPubMed
Miller, TL & Wolin, MJ (2001) Inhibition of growth of methane-producing bacteria of the ruminant forestomach by hydroxymethylglutaryl-SCoA reductase inhibitors. J Dairy Sci 84, 14451448.CrossRefGoogle ScholarPubMed
Moss, AR (1993) Methane Global Warming and Production by Animals. Canterbury: Chalcombe Publications.Google Scholar
Moss, AR, Jouany, JP & Newbold, CJ (2000) Methane production by ruminants: its contribution to global warming. Ann Zootech 49, 231253.CrossRefGoogle Scholar
National Research Council (2001) Nutrient Requirements of Dairy Cattle, 7th revised ed. Washington DC: National Academy Press.Google Scholar
Newbold, CJ, Lassalas, B & Jouany, JP (1995) The importance of methanogens associated with ciliate protozoa in ruminal methane production in vitro. Lett Appl Microbiol 21, 230234.CrossRefGoogle ScholarPubMed
Newbold, CJ, Wallace, RJ & McIntosh, FM (1997) Mode of action of the yeast Saccharomyces cerevisiae as a feed additive for ruminants. Br J Nutr 76, 249261.CrossRefGoogle Scholar
Newbold, CJ, Williams, AG & Chamberlain, DG (1987) The in vitro metabolism of D, L-lactic acid by rumen microorganisms. J Sci Food Agric 38, 919.CrossRefGoogle Scholar
Oba, M & Allen, MS (2003) Extent of hypophagia caused by propionate infusion is related to plasma glucose concentration in lactating dairy cows. J Nutr 133, 11051112.CrossRefGoogle ScholarPubMed
Ørskov, ER & Ryle, M (1990) Energy Metabolism in Ruminants. London: Elsevier Science.Google Scholar
Russell, JB (1998) The importance of pH in the regulation of ruminal acetate to propionate ratio and methane production in vitro. J Dairy Sci 81, 32223230.CrossRefGoogle ScholarPubMed
Steel, RGD & Torrie, JH (1980) Principles and Procedures of Statistics. New York: McGraw-Hill.Google Scholar
Stewart, CS & Duncan, SH (1985) The effect of avoparcin on cellulolytic bacteria of the ovine rumen. J Gen Microbiol 131, 427435.Google Scholar
Takahashi, J (2001) Nutritional manipulation of methanogenesis in ruminants. Asian-Australasian J Anim Sci 14, 131135.Google Scholar
Ungerfeld, EM, Rust, SR & Burnett, R (2003 a) Use of some novel alternative electron sinks to inhibit ruminal methanogenesis. Reprod Nutr Dev 43, 189202.CrossRefGoogle ScholarPubMed
Ungerfeld, EM, Rust, SR & Burnett, R (2003 b) Attempts to inhibit ruminal methanogenesis by blocking pyruvate oxidative decarboxylation. Can J Microbiol 49, 650654.CrossRefGoogle ScholarPubMed
Van Nevel, CJ & Demeyer, DI (1996) Control of rumen methanogenesis. Environ Monit Assess 42, 7397.CrossRefGoogle ScholarPubMed
Weatherburn, MW (1967) Phenol-hypochlorite reaction for determination of ammonia. Anal Chem 39, 971974.CrossRefGoogle Scholar
Wolin, MJ, Miller, TL & Stewart, CS (1997) Microbe-microbe interactions. In The Rumen Microbial Ecosystem, 2nd edition, 467491[Hobson, PN and Stewart, CS, editors]. London: Blackie Academic & Professional.CrossRefGoogle Scholar