Вы находитесь на странице: 1из 15

AnilKumarPuniya RameshwarSingh

DevkiNandanKamra Editors

Rumen
Microbiology:
From Evolution to
Revolution

Rumen Microbiology: An Overview


Prasanta Kumar Choudhury,
Abdelfattah Zeidan Mohamed Salem,
Rajashree Jena, Sanjeev Kumar,
Rameshwar Singh, and Anil Kumar Puniya

Abstract

The systematic exploration of microbial ecosystem of the rumen was


commenced by the father of rumen microbiology, Robert Hungate, in
1950s. His contributions toward the development of anaerobic culture techniques have illustrated the ways to explore the complex microbial structures of the rumen and other anaerobic ecosystems. The understanding of
rumen microbiology has strengthened an awareness to improve the feed
utilization and manipulation of microbial compositions. Microbes and their
interactions in interspecies H2 transfers were first studied in the rumen ecosystems and attracted pioneers to investigate the alternate abatement strategies of methane production along with enhanced animal productivity. The
discovery of alternate hydrogenotrophs and industrially important novel
microbes and the management of rumen disorders via microbial manipulations make this community an interesting research platform for different
microbial theories. The discovery of anaerobic fungi as a part of rumen
flora by Orpin during the 1970s disproved their mistaken identity as flagellated protozoa and the concept that all fungi are aerobic organisms.
Upcoming biotechnological strategies and deciphering on microbial com-

P.K. Choudhury R. Jena


Dairy Microbiology Division, ICAR-National Dairy
Research Institute, Karnal 132001, India
A.Z.M. Salem
Facultad de Medicina Veterinaria y Zootecnia,
Universidad Autnoma del Estado de Mxico,
Estado de Mxico, Mexico
S. Kumar
Department of Life Science, Assam University,
Silchar 788011, India

R. Singh
Directorate of Knowledge Management in
Agriculture, Indian Council of Agricultural Research,
KAB-1 (PUSA), New Delhi 110012, India
A.K. Puniya (*)
College of Dairy Science and Technology,
Guru Angad Dev Veterinary and Animal Sciences
University, Ludhiana, Punjab, India
e-mail: akpuniya@gmail.com

Springer India 2015


A.K. Puniya et al. (eds.), Rumen Microbiology:
From Evolution to Revolution, DOI 10.1007/978-81-322-2401-3_1

P.K. Choudhury et al.

munity using molecular tools, novel gene discovery, community-community


interactions, and phylogenetic relationships have opened new avenues of
microbial ecology in rumen ecosystem. The present chapter deals with the
microbial ecosystems of rumen and their interactions.
Keywords

Rumen Bacteria Anaerobic fungi Methanogens Microbial


interactions

1.1

Introduction

Rumen, the four-chambered stomach of grazing


animals, harbors a complex bionetwork, where all
forms of primitive starting from archaea to protozoa exist in close proximity (Choudhury et al.
2012). Some of these microbes interact with each
other in a synergistic relationship to extract energy
while producing highly active lignocellulolytic
enzymes supporting digestion of the host. This
yields volatile fatty acids (VFAs, acetate, butyrate,
propionate), formic acid, H2, CO2, and CH4
(Krause et al. 2003). For many years, nutritionists,
microbiologists, and physiologists have been
studying the rumen with the aim of maximizing
productivity and improving overall host health via
manipulating the rumen and its microbial ecosystem. Without rumen microorganisms, or when the
rumen ecosystem is disturbed, there can be a cascade of detrimental effects on animal health and
productivity. Altering the rumen ecosystem to
reduce methane emissions is a significant scientific challenge that must take the effects on animal
health, farm management, and animal productivity
into account. Understanding of complex microbial
communities function and microbial interactions
within their niches represents a major challenge
for rumen microbiologists even today. Advances
in molecular biology and genomics offer new
opportunities to conduct a holistic examination of
the structure and function of rumen microbial
communities (Sirohi et al. 2012a). Analyzing the
role of individual group of microbes, microbial
structure, composition, and interaction may provide functional dimensions of the rumen network
that will help to achieve a major goal of rumen
microbiology.

1.2

Rumen Microbial Ecosystem

Herbivores feed on plant structural carbohydrates


such as lignin, cellulose, and hemicellulose,
which they are not able to utilize themselves
because of their inability to produce respective
hydrolytic enzymes. Therefore, symbiotic microorganisms are established in their alimentary
tracts that can hydrolyze these compounds to
generate energy for themselves, as well as for the
host animal. Since strict anaerobic conditions
(150 to 350 mv) prevail in the rumen (Clarke
1977), the natural rumen microflora consists of
strict anaerobes. Woeses classification represented all microbes in the rumen ecosystem and
distinguished them into three domains: Bacteria
(bacteria), Archaea (methanogens), and Eucarya
(protozoa and fungi). The rumen is open to the
external environment, and there is a continuous
flow of material into and out of the rumen. It has
a dry matter content of 1013 %, and temperature
is typically regulated between 38 and 41 C. The
physiological pH range is between 5.5 and 6.9,
and it is one of the most variable factors in the
rumen environment. However, the buffering
capacity of the rumen content is quite high
(Hungate 1966; Church 1969; Clarke 1977;
Dehority 2003). The details of different rumen
physical, chemical, and microbiological parameters are discussed in Table 1.1.

1.2.1

Bacteria

Rumen harbors different types of bacteria, which


are most actively involved in the plant fiber degradation, as revealed by the fact that bacteria associ-

Rumen Microbiology: An Overview

Table 1.1 Physical, chemical, and microbiological characteristics of rumen ecosystem


Physical properties
Dry matter (%)
Osmolality
pH
Redox potential
Temperature
Chemical properties
Amino acids and oligopeptides
Ammonia
Dietary (cellulose, hemicelluloses, pectin) component
Endogenous (mucopolysaccharides)
Gas phase (%)
Growth factors
Lignin
Minerals
Nonvolatile acids (mmol L1)
Soluble carbohydrates
Trace elements/vitamins
Volatile fatty acids (mmol L1)

1018
250350 mOsmol/Kg1
5.56.9 (Mean 6.4)
350 to 400 mV
3841 C
<1 mmolL1 present 23 h post feeding
212 mmolL1
Always present
Always present
CO2 65; CH4 27, N2 7; O2 0.6, H2 0.2
Good supply; branched-chain fatty acids, long-chain
fatty acids, purines, pyrimidines, other unknown
Always present
High Na; generally good supply
Lactate <10
<1 mmolL1 present 23 h post feeding
Always present; good supply of B vitamins
Acetate 6090, propionate 1530, butyrate 1025,
branched chain and higher 25

Microbiological properties
Anaerobic fungi
1035 g1 (6 genera)
Bacteria
101011 g1 (>200 species)
Bacteriophage
1079 g1 particles ml1
Ciliate protozoa
1046 g1 (25 genera)
Source: Adapted from Mackie et al. (1999), Castro-Montoya et al. (2011)

ated with feed particles account for nearly


5075 % of the total microbial population (Minato
et al. 1966). The adhesion of microbes to solid
substances is an important factor in successful
competition and survival in the rumen, as well as
in the digestion of solid feed (Minato et al. 1966;
Mitsumori and Minato 1997). In addition, most of
the rumen endoglucanase and xylanase activities
are contributed by bacteria, and therefore, fiberassociated bacteria are pivotal to the rumen digestion. These are divided into four groups on the
basis of their association as free floating in the
liquid phase, attached to the feed particles (firmly/
loose), rumen epithelium, protozoa, and fungi.
Besides, these microbes have also diversified with
respect to their functions. These carry out in degradation of plant ingredients. These have been
classified as fiber (i.e., cellulose, hemicelluloses,
and pectin)-degrading bacteria, lactic acid utilizers, acetogens, starch utilizers, etc. (Table 1.2).

Starch and sugar digesters make up a significant


part of the rumen bacterial population. Generally,
high-producing dairy cows are fed diets containing more than 30 % starch and sugars; therefore,
these bacteria are greatly needed. Even if a cow is
on an all-straw diet, the fiber digesters still never
account for more than 25 % of the rumen bacterial
population. Some microbes utilize the gaseous
end products of the rumen to form methane and
are grouped under archaea, while others are
involved in converting the reducing power to useful end products as acetate, propionate, H2S, and
butyrate. Always, there is an interest for nonmethanogenic sinks to release H2 produced during
fermentation (Joblin 1996, 1999). In the rumen,
CH4 is produced by methanogens using H2 to
reduce CO2. The non-methanogenic hydrogenotrophic bacteria residing in the rumen include acetogens that reduce CO2 to form acetate by the
Wood-Ljungdahl pathway (i.e., reductive aceto-

P.K. Choudhury et al.

6
Table 1.2 Microbes involved in various rumen functions
Microbial types
Bacteria
Acetogens
Acid utilizers

Important genera and species

Acetitomaculum ruminis, Eubacterium limosum


Megasphaera elsdeni, Wolinella succinogenes, Veillonella gazogene, Micrococcus
lactolytica, Oxalobacter formigenes, Desulfovibrio desulfuricans, Desulfotomaculum
ruminis, Succiniclasticum ruminis
Cellulolytic
Fibrobacter succinogenes, Butyrivibrio fibrisolvens, Ruminococcus flavefaciens,
Ruminococcus albus, Clostridium cellobioparum, Clostridium longisporum, Clostridium
lochheadii, Eubacterium cellulosolvens
Hemicellulolytic
Prevotella ruminicola, Eubacterium xylanophilum, Eubacterium uniformis
Lipolytic
Anaerovibrio lipolytica
Pectinolytic
Treponema saccharophilum, Lachnospira multiparus
Proteolytic
Prevotella ruminicola, Ruminobacter amylophilus, Clostridium bifermentans
Amylolytic
Streptococcus bovis, Ruminobacter amylophilus, Prevotella ruminicola
Saccharolytic
Succinivibrio dextrinosolvens, Succnivibrio amylolytica, Selenomonas ruminantium,
Lactobacillus acidophilus, Lactobacillus casei, Lactobacillus fermentum, Lactobacillus
plantarum, Lactobacillus brevis, Lactobacillus helveticus, Bifidobacterium globosum,
Bifidobacterium longum, Bifidobacterium thermophilum, Bifidobacterium ruminale,
Bifidobacterium ruminantium
Tanninolytic
Streptococcus caprinus, Eubacterium oxidoreducens
Ureolytic
Megasphaera elsdenii
Bacteriophages
Methanobacterium phage M1, Methanobacterium phage M10, Methanobacterium
phage M100, Methanothermobacter phage M100, Methanobacterium phage M2
Fungi
Piromyces communis, Piromyces mae, Piromyces minutus, Piromyces dumbonicus,
Piromyces rhizinflatus, Piromyces spiralis, Piromyces citronii, Piromyces polycephalus,
Anaeromyces mucronatus, Anaeromyces elegans, Caecomyces communis, Caecomyces
equi, Caecomyces sympodialis, Cyllamyces aberensis, Cyllamyces icaris, Neocallimastix
frontalis, Neocallimastix patriciarum, Neocallimastix hurleyensis, Neocallimastix
variabilis, Orpinomyces joynii, Orpinomyces intercalaris
Methanogens
Methanobacterium formicicum, Methanobacterium bryantii, Methanobrevibacter
ruminantium, Methanobrevibacter smithii, Methanomicrobium mobile, Methanosarcina
barkeri, Methanoculleus olentangyi
Protozoa
Entodinium bovis, Entodinium bubalum, Entodinium bursa, Entodinium caudatum,
Entodinium chatterjeei, Entodinium parvum, Entodinium longinucleatum, Entodinium
dubardi, Entodinium exiguum, Epidinium caudatum, Isotricha prostoma, Isotricha
intestinalis, Dasytricha ruminantium, Diplodinium dendatum, Diplodinium indicum,
Oligoisotricha bubali, Polyplastron multivesiculatum, Eremoplastron asiaticus,
Eremoplastron bubalus
Source: Pfister et al. (1998), Luo et al. (2001), Kamra (2005), Janssen and Kirs (2008), Wright and Klieve (2011), Sirohi
et al. (2012b), Choudhury et al. (2012), Kumar et al. (2014)

genesis), sulfate-reducing bacteria (SRB) that


reduce sulfate to H2S (Morvan et al. 1996), and
bacteria that use H2 to reduce fumarate to succinate (Asanuma and Hino 2000). Succinate, the
product of fumarate reduction, is decarboxylated
to propionate, a valuable animal nutrient (Wollin
et al. 1997). Dissimilatory sulfate-reducing bacteria produce a toxic end product as H2S from the
reducing power.

The type of feed also controls the bacterial populations inside the rumen. A high-fiber-containing
feed increases the population of Fibrobacter succinogenes, Ruminococcus albus, and Ruminococcus
flavefaciens. The H2 and VFA production are also
enhanced by high-fiber diet. Minato et al. (1989)
clarified the effect of feeding ammonia-treated
straw on rumen bacterial flora, showing that it stimulates the growth of Eubacterium ruminantium.
Orpin et al. (1985) described seasonal variations of

Rumen Microbiology: An Overview

rumen bacteria in wild reindeer by demonstrating


an increased cellulolytic Butyrivibrio fibrisolvens
under severe dietary conditions in winter.
Interestingly, non-cellulolytic bacteria were also
detected as members of the fiber-associating consortia in the rumen (Koike et al. 2003; Larue et al.
2005). Tannin-degrading bacteria have also been
isolated (Goel et al. 2005a, b, 2011) from ruminants of wild origin identified as Selenomonas
ruminantium (Odenyo and Osuji 1998) as these
consume tanniferous plants. Streptococcus bovis is
present only when large amounts of starch or
sugars are fed and pH is low. It produces lactic acid,
a stronger acid than many of the other VFAs produced in the rumen. When conditions are favorable
for Streptococcus bovis, it will grow explosively.
This type of growth causes rumen acidosis.
Megasphaera elsdenii requires lactic acid to grow
that helps to clean up the rumen a bit and raise
rumen pH, sustaining the growth of the acid-intolerant fiber digesters in the rumen.

1.2.2

Anaerobic Fungi

The rumen is a storehouse of anaerobic fungi that


are not found elsewhere and play a very dominant
role in the degradation of lignocellulosic components of the feed particles. These constitute the
smallest in numbers and represent only about
20 % (Rezaeian et al. 2004) of the rumen microbial biomass but are considered to be the most
efficient fiber degrader (Akin et al. 1988).
According to the recent classification system
(Hibbett et al. 2007), anaerobic fungi have been
described under the order Neocallimastigales,
class Neocallimastigomycetes in the phylum
Neocallimastigomycota (Dagar et al. 2011; Sirohi
et al. 2013a; Gruninger et al. 2014). The thallus
may be monocentric (single reproductive body,
i.e., sporangium from single zoospore) or polycentric (numerous sporangia from single zoospore) (Ho 1995). These forms are determined in
the earliest stages of growth, soon after zoosporogenesis. The zoospores may be posteriorly
uniflagellated or polyflagellated in both forms.
The life cycle of anaerobic fungi is asexual, and
no sexual stage has been described, as only

mitotic nuclear divisions have been observed


(Heath et al. 1986). It alters between a motile
zoospore, encysted zoospore, and vegetative
zoosporangial stage. Once released from zoosporangium, the motile zoospores move by chemotaxis to colonize the plant fibrous material and
shed its flagella to get transformed into cyst
(encystment). The cyst germinates by producing
a germ tube, which ultimately gives rise to a
rhizoidal system. In monocentric forms, the
development may further be classified as endogenous or exogenous (Sirohi et al. 2012b).
These fungi differ from their aerobic counterparts in having hydrogenosomes as a site for
energy production, instead of mitochondria
(Yarlett et al. 1986). These fungi are not exclusively found in the rumen of the animals, but
throughout the entire digestive tract (Davies et al.
1993). Moreover, these have also been isolated
from the saliva and feces (Lowe et al. 1987), suggesting the presence of a resistant structure tolerant to the oxygen and desiccation. Brookman
et al. (2000) identified the formation of spores in
the polycentric anaerobic gut fungi, which
enhance their ability to survive and extend a possibility toward their use as direct-fed microbial
(Tripathi et al. 2007a; Saxena et al. 2010; Puniya
et al. 2014). Anaerobic fungi have been isolated
from a number of ruminant and nonruminant herbivorous mammals (Ljungdahl 2008; Paul et al.
2010). These contribute significantly to overall
metabolism of host with their high cellulolytic
activity and play a greater role in degradation of
lignified plant tissues (Akin and Borneman 1990;
Dey et al. 2004; Thareja et al. 2006; Sirohi et al.
2012b) with the help of a wide range of hydrolytic enzymes (Tripathi et al. 2007b; Shelke et al.
2009; Nagpal et al. 2009, 2010, 2011) including
cellulases (Barichievich and Calza 1990), hemicellulases (Novotna et al. 2010; Mountfort and
Asher 1989), proteases (Michel et al. 1993),
amylases, amyloglycosidases (Paul et al. 2004),
feruloyl and p-coumaroyl esterases (Borneman
et al. 1990), various disaccharidases (Chen et al.
1995), and pectinases (Kopecny and Hodrova
1995). To date, only six genera (Table 1.2) have
been described (Griffith et al. 2009; Sirohi et al.
2012b). However, certain uncharacterized iso-

P.K. Choudhury et al.

lates have also been reported (Phillips and


Gordon 1988; Ho 1995).
Since these fungi are obligatory anaerobic,
these obtain energy by fermentation of carbohydrates (Gordon and Phillips 1998; Trinci et al.
1994), a process in which the energy source acts
as both the electron acceptor and the electron
donor. These lack mitochondria (Yarlett et al.
1986), cytochromes, and other biochemical
features of the oxidative phosphorylation pathway. Instead of mitochondria, hydrogenosomes
are present, which couple the metabolism of
glucose to cellular energy production. In the
beginning, hydrogenosomes were thought to be
novel inventions to cope with anaerobic conditions (Cavalier-Smith 1987), but now it has
become quite clear that hydrogenosomes are
either specialized or derived mitochondria
(Embley et al. 1997). These organelles are not
unique to anaerobic fungi and were discovered in
parasitic trichomonads (Lindmark and Mller
1973). Since then, these have been found in a
wide variety of anaerobic eukaryotes. Typical
mitochondrial features like genome, electron
transport chain, and cristae are missing from fungal hydrogenosomes and contain unusual
enzymes like hydrogenase and pyruvate/ferredoxin oxidoreductase instead of usual mitochondrial pyruvate dehydrogenase (van der Giezen
2002). The common properties of hydrogenosomes and mitochondria include the existence of
a transmembrane pH gradient and an alkaline
lumen. Just like mitochondria, hydrogenosomes
also serve as the site for conversion of pyruvate to
acetyl CoA and ATP production, whereas the
presence of other metabolic processes like amino
acid metabolism and -oxidation is presently
unknown. Hydrogenosomes contain hydrogenase
and produce H2, CO2, and acetate as metabolic
waste products (Brul and Stumm 1994;
Theodorou et al. 1996).

1.2.3

Protozoa

Rumen protozoa constitute about 50 % of the


viable biomass in the rumen and are dependent.

Majority are ciliates and few flagellates and are


very motile. Unicellular organisms with 20200
m size are not attached to feed particles. They
engulf bacteria and feed particles and digest carbohydrates, proteins, and fats (Williams and
Coleman 1992). Two groups, i.e., holotrich and
entodiniomorphid protozoa, have been studied
inside the rumen (Table 1.2). Morphological
studies have identified more than 250 species of
ciliates living in the various ruminants (Williams
and Coleman 1992; Imai 1998). These ciliates
play an important role in fiber digestion and the
modulation of the fermentation profiles. The
rumen protozoa produce fermentation end
products similar to those made by the bacteria,
particularly acetate, butyrate, and H2. Rumen
methane bacteria actually attach and live on the
surface of rumen protozoa for immediate access
to H2. They utilize large amounts of starch at one
time and can store it in their bodies. This may
help to slow down the production of acids that
lower rumen pH, benefiting the rumen. Rumen
protozoa multiply very slowly in the rumen over
1524 h as opposed to the bacteria. For this reason, the rumen protozoa hide out in the slowermoving fiber mat of the rumen, so that these are
not washed out before these have a chance to
multiply. Low-roughage diets reduce the retention of fiber in the rumen and may decrease the
number of protozoa in a cows rumen.
The enzymatic profile of holotrich protozoa
indicates that these have amylase, invertase, pectin esterase, and polygalacturonase in sufficiently
large quantities for using starch, pectin, and soluble sugars as energy source (Bailey and Howard
1963; Williams 1979). The enzymes responsible
for cellulose and hemicellulose degradation have
also been reported in the holotrich protozoa, but
the levels are very low compared to those present
in the entodiniomorphid protozoa (Williams and
Coleman 1985). The rumen ciliates are proteolytic producing ammonia and amino acids as end
products (Warner 1956). Their N metabolism is
based largely on the digestion of engulfed bacteria (Coleman 1975), although all rumen ciliates
contain enzymes capable of digesting plant proteins (Coleman 1983). The utilization of these

Rumen Microbiology: An Overview

protein substrates is inefficient, as a significant


proportion of the total amount engulfed by the
protozoa is excreted as amino acids (Coleman
1975). Degradation of protein or amino acids of
microbial origin reduces the net yield of microbial amino acids available for intestinal digestion
and hence, the net efficiency of microbial growth.
Further, a large proportion of the protein incorporated in the biomass of protozoa in the rumen
may not be available for digestion in the intestines. Direct enumeration of protozoa in the
rumen fluid entering the omasum (Weller and
Pilgrim 1974), concentration of a protozoa
marker (i.e., phosphatidylcholine) in abomasal
digesta, and measurement of the turnover rates of
the protozoa (Leng 1982) all indicate a low outflow rate of ciliates from the rumen in relation to
their concentration in the rumen. The synthesis of
protozoa nitrogenous compounds and their
breakdown in the rumen, therefore, contribute
considerably to the N recycling within the rumen,
as the apparent turnover of protozoa protein was
calculated to be 0.67 of the total protein synthesized by protozoa (Leng 1982).

1.2.4

Methanogens

Methanogenic archaea, considered as the earliest


organisms originated in our planet, are phyllogenetically widespread and represent three of the
seven classes in the phylum Euryarchaeota
(Sirohi et al. 2010; Singh and Sirohi 2012). These
rumen microorganisms utilize the H2 and CO2
produced by the protozoa, fungi, and bacteria
from the catabolism of hexoses to produce CH4
and generate ATP (Albers et al. 2007; Ferry and
Kastead 2007), which benefits the donors by providing an electron sink for reducing equivalents
to minimize the partial pressure of H2 (Wolin and
Miller 1988; Lange et al. 2005) inside the rumen.
These are comprised of about 70 species among
31 genera as reviewed in Bergeys Manual of
Systematic Bacteriology (Whitman et al. 2001)
and The Prokaryotes. These microorganisms
specifically possess three coenzymes, i.e., coenzyme F-420, involved in electron transfer; coenzyme M, involved in transfer of methyl groups;

and a low molecular weight, oxygen-labile, heatstable coenzyme B, involved in the final reaction
of CH4 production. Methanogens are found in a
symbiotic association with rumen bacteria (Wolin
and Miller 1988) and protozoa (Lange et al.
2005). The establishment and maintenance of the
stable population of methanogens is affected by
the type of diet and level and frequency of feeding (Kumar et al. 2012, 2013a, b; Sirohi et al.
2013b). The symbiotic association of hydrophobic methanogens with H2 producers is usually
realized by the attachment or by floc formation
(Thiele et al. 1988; Lange et al. 2005). Among
these ciliates, protozoa are the only ones for
which such interaction can be microscopically
demonstrated (Vogels et al. 1980). The symbiotic
relation between methanogens and ciliates may
generate up to 37 % of rumen CH4 emission
(Finlay et al 1994). In the rumen, H2 is produced
during plant cell wall degradation as an intermediate compound by cellulolytic bacteria (i.e.,
Ruminococcus albus, Ruminococcus flavifaciens)
and anaerobic fungi (Stewart and Bryant 1988). It
never gets accumulated, as it is rapidly utilized
by methanogens in ruminants (Wolin and Miller
1988) to produce CH4 and generate ATP (Albers
et al. 2007; Ferry and Kastead 2007). Details of
different genera are given in Table 1.2.

1.2.5

Bacteriophages

Bacteriophages are one of the most important


component of the rumen microbial community
and are present typically at >109 particles per
mL. These are specific for different bacteria present in the rumen. These are also considered to be
obligate pathogens for the bacteria, as bacteriophages are capable of lysing bacteria. These
phages help in bacterial mass turnover in the
rumen, which may be considered not so useful
for the animals on different feeding schedules
(Klieve and Swain 1993), but by lysing the bacterial cells, the bacterial protein is easily made
available to the animals as a source of amino
acids. The specificity of the bacteriophages for a
particular rumen bacterium may be exploited for
removal or killing by lysis of unwanted rumen

P.K. Choudhury et al.

10

bacteria from the ecosystem like Streptococcus


bovis and methanogens (Klieve et al. 1999; Bach
et al. 2002). A little information is available on
the genetic blueprint and gene functionality of
archaeal, particularly methanogenic, phages, but
more are being discovered using electron microscopy (Ackermann 2007) and in vitro techniques
(Stanton 2007). McAllister and Newbold (2008)
reported siphophages that can infect methanogens (Methanobacterium, Methanobrevibacter,
and Methanococcus spp.) (Table 1.2), although
these phages have not been isolated from the
rumen. Despite the lack of knowledge, ruminant
phage and their enzymes that are involved in
lysing host cell represent significant opportunities for controlling both methanogen population
and other community members in the rumen. In
addition, it is possible to identify key host enzyme
targets that are susceptible to inhibition by phage
proteins understanding the life cycles of phage.

1.3

Microbial Interactions

The bionetwork inside the rumen is responsible


for the complete degradation of organic matter
ingested by the animals. There is no single organism that has the capacity for a complete degradation of complex substrates in the rumen (Bladen
et al. 1961). A complex succession of microorganisms takes part in the cooperative catabolism
of substrates in the rumen and production of fermentative end products. The complex feedstuffs
are broken down into simple sugars, transported,
and fermented by other members of the microbial
population that in turn produce branched-chain
fatty acids, vitamins, or other cofactors for other
bacteria responsible for degrading feedstuffs
(Allison et al. 1961; Bryant 1973). The community analysis approach to the breakdown of nutrients by a microbial consortium led to the
discovery of the concept of cross feeding of
nutrients across and within environmental niches
and can support the overall health of the host animal. The steady supply of food and constant
removal of digested feed material and end products along with the abovementioned rumen conditions allow a dense population of

microorganisms to grow inside the rumen


(Hungate 1966). Some bacteria and protozoa
(i.e., holotrichs) are closely associated with the
rumen wall (McAllister et al. 1994; Dehority
2003). The rumen wall-associated bacteria scavenge oxygen and break down urea into CO2 and
NH3 that enter into the rumen via feed, saliva, and
blood (McAllister et al. 1994). Further, some of
those degrade the sloughing off of epithelial
cells. The NH3 is a source of nitrogen to the
rumen bacteria, while CO2 is an essential growth
factor for the fiber-digesting bacteria. The extramural bacteria (99 % of the rumen bacteria) can
be either free floating or attached to protozoa
(i.e., symbiotic methanogens) or feed particles
(Koike and Kobayashi 2009). About 75 % of the
bacteria are associated with feed particles (Craig
et al. 1987). Out of the feed particle-associated
bacteria, approximately 8090 % is responsible
for fiber digestion, 75 % for protein digestion,
and 70 % for starch digestion (Brock et al. 1982;
Minato et al. 1993). The free-floating bacteria or
bacteria associated with liquid material inside the
rumen are about 30 % of the total bacteria inside
the rumen (Legay-Carmier and Bauchart 1989).
These can either become attached to the feed particles or simply washed away into the lower gastrointestinal tract and then become available as a
good source of protein for the host animal.
Production of B vitamins by rumen bacteria is
important to ensure bacterial growth and, particularly, fiber degradation and is also responsible for
ensuring animal health and well-being (Scott and
Dehority 1965). For example, a strain of the
prominent
rumen
cellulolytic
bacterium
(Ruminococcus albus) was found to require phenylpropanoic acid (PPA) to digest cellulose (Stack
and Hungate 1984). Rumen microbes produce
PPA from the fermentation of phenolic compounds, and the lack of PPA in cultures meant that
R. albus could not adhere to cellulose. The PPA
was thought to act as a glue that holds the cellulolytic enzyme complex in a stable conformation, allowing for cellulose degradation (Stack
and Hungate 1984). As researchers delineated
bacterial functions and nutritional requirements, it
became clear that there are significant interactions
between microbial populations, both within an

Rumen Microbiology: An Overview

environmental niche and between niches.


Application of this concept spreads into analysis
of other microbial habitats (Schultz and Breznak
1979) and has shaped the view of how microbes
interact with the world around those. From the
time of these, it has been discovered that quorum
signals play a role in other natural microbial environments (Khan et al. 1995; Miller and Bassler
2001; Sperandio et al. 2003) and that these biochemical messenger molecules can cause changes
in microbial growth and fermentation.
Furthermore, it now appears that a 2-way
communication between the host and microbial
population exists, meaning that the endocrine status of the host can have a direct impact on the gut
microbial population (Sperandio et al. 2003; Lyte
and Freestone 2009; Freestone and Lyte 2010;
Bailey et al. 2011). This has, in turn, led to the
developing concept of the microbial organ playing a critical role in animal production and health
(Lyte 2010) and the theory that probiotics or the
rumen microbial population could be used as a
drug delivery vehicle to improve animal performance for higher productivity.
The increase in internal H2 concentration has
several harmful effects in lowering the rumen
pH and overall inhibits the biodegradation by
creating an unfavorable equilibrium (Ragsdale
and Pierce 2008). But, some of the groups
involved in the transfer of produced H2 as reducing power. The process of syntrophic H2 transfer
was first hypothesized by Hungate (1966).
Methanogens, a group of archaeal microbes,
combine this molecular H2 with CO2 to produce
CH4. Both from environment and nutrient utilization point of view, methanogenesis is not a
suitable sink as the end product is a greenhouse
gas with 23 times global warming (Eggleston
2006) potential than CO2 with a loss of 813 %
digestible energy ingested (Moss et al. 2000).
Ruminococcus albus, a major cellulose-digesting species in the rumen, produces ethanol, acetate, H2, and CO2 in the absence of methanogen
but produces more acetate and less or no ethanol
in the presence of methanogens (Iannotti et al.
1973). Because acetate production results in
ATP synthesis, the interaction results in
increased growth and cellulose digestion. This

11

startling concept demonstrated that although


excessive rumen CH4 production was negative
from the perspective of the host animal and the
environment, a certain level of CH4 production
is necessary to remove reducing the equivalents,
thereby, allowing the overall rumen fermentation to proceed. As a result, rumen fermentations will likely always continue to produce at
least some CH4. Major precursors of methanogens are acetate, H2 and CO2, the question of
how propionate and butyrate would become
CH4 in anaerobic ecosystems other than the
rumen arose. Several alternate approaches from
diet to animal management have been employed
since long, but most of them failed as being
costly and noxious to host. Accordingly, microbial interventions that employ biological routes
are ideal approaches to sequester these gaseous
end products from the rumen.

1.4

Future of Rumen
Microbiology

The rumen microbial systems have now been


examined in detail for some decades by microbiologists, nutritionist, and biotechnologists.
Always, the focus is to enhance the animal productivity by manipulating the dietary composition or by altering the microbial ecosystems. The
upcoming strategies in discovery of novel gene/
microbes with genomic approaches will pave a
manipulating way toward the developments of indepth knowledge on microbial sequence signature. Application of highly active enzymes (i.e.,
endoglucanase, xylanase, esterases, etc.) from
the rumen source for commercial applications
will provide a new dimension in agro-industries.
Genetic engineering, by adding a desired activity
to an established rumen microorganism, may
overcome some of the problems. But there still
remain the few problems of how long a new
microorganism, once established, will remain
active. The control of rumen methanogenesis by
inhibiting the active sites of different enzymes by
the in silico prediction tools from the genome
sequences will limit the research inputs in terms
of manpower, energy, and time. Bioinformatics

P.K. Choudhury et al.

12

approaches used for the inhibitor prediction


against the F420-dependent NADP oxidoreductase
enzyme that catalyzes an important electron
transfer step in the methanogenesis from
Methanobrevibacter smithii reveal that lovastatin
and compactin had high affinity to the enzyme
and can act as the potential inhibitors (Sharma
et al. 2011). Both in silico approaches and in vitro
enzyme assays may be useful for screening
chemical inhibitors of methanogenesis. Most
recent report of Lee et al. (2013) that only very
few rumen methanogens are cultured as pure
isolates and 13 genome projects are completed as
yet (Morgavi et al. 2013). Most of these genome
sequences are from the genus Methanobrevibacter,
which is considered to be dominating rumen
methanogen, as per the global data set of rumen
microbes (Jeyanathan et al. 2011; Kumar et al.
2012). This will provide knowledge on how
microbes and their metabolic products affect the
gastrointestinal physiology and immune
responses in the host and also will provide genetics of syntrophic interactions, e.g., interspecies
hydrogen transfer, microbe-microbe interactions
relevant to host nutrition and health, optimization
of fiber utilization, and probiotic and prebiotic
(synbiotic) developments.

1.5

Conclusions

The rumen microbial ecosystem has a significant impact on understanding the digestive
physiology and also provides a test bed for other
microbial environments in technology development and validation. This will also have a potential application in industry for a cost-effective
ruminant feed production and management of
enteric fermentation. Recent developments in
microbial ecology have been adapted again and
validated in the rumen, providing the ability to
understand the gastrointestinal consortium at a
level never dreamed of in the times of Hungate.
Application of rumen microbial ecosystems will
contribute much to the understanding of the
ecology of rumen microorganism including the
potential to elucidate and overcome the biochemical, ecological, or physiological processes

that are rate limiting for rumen fiber degradation. By linking these data with animal diets and
in vitro analysis, it may be possible to define
key microbes and their role in the fiber degradation. The learning novel pathways will also lead
to the knowledge about the currently non-culturable bacteria to culture. The modern genomics-based strategies to elucidate the microbial
community structure, novel gene discoveries,
and recombinant DNA technology will provide
information for eco-friendly animal production
and management.

References
Ackermann HW (2007) 5500 phages examined in the
electron microscope. Arch Virol 152:227243
Akin DE, Borneman WS (1990) Role of rumen fungi in
fiber degradation. J Dairy Sci 73:30233032
Akin DE, Borneman WS, Windham WR (1988) Rumen
fungi: morphological types from Georgia cattle and
the attack on forage cell walls. Biosystems 2:385391
Albers SV, Konings WN, Driessen JM (2007) Solute
transport. In: Cavicchioli R (ed) Archaea: molecular
and cellular biology. ASM Press, Washington, DC,
pp 354368
Allison MJ, Bryant MP, Doetsch RN (1961) Studies on
the metabolic function of branched-chain volatile fatty
acids, growth factors for Ruminocci. J Gen Microbiol
5:869879
Asanuma N, Hino T (2000) Activity and properties of
fumarate reductase in ruminal bacteria. J Gen Appl
Microbiol 49:119125
Bach SJ, Mcallister TA, Veira DM et al (2002)
Transmission and control of Escherichia coli O157:
H7 a review. Can J Anim Sci 82:475490
Bailey RW, Howard BH (1963) The biochemistry of
rumen protozoa: the maltases of Dasytricha ruminantium, Epidinium ecaudatum (Crawley) and Entodinium
caudatum. Biochem J 86:446
Bailey MT, Dowd SE, Galley JD et al (2011) Exposure to
a social stressor alters the structure of the intestinal
microbiota: implications for stressor-induced immunomodulation. Brain Behav Immun 25:397407
Barichievich EM, Calza RE (1990) Supernatant protein
and cellulase activities of the anaerobic ruminal fungus Neocallimastix frontalis EB188. Appl Environ
Microbiol 56:4348
Bladen HA, Bryant MP, Doetsch RN (1961) A study of
bacterial species from the rumen which produce
ammonia from protein hydrolyzate. Appl Microbiol
9:175180
Borneman WS, Hartley RD, Himmelsbach DS et al (1990)
Assay for trans-p-coumaroyl esterase using a specific

Rumen Microbiology: An Overview

substrate from plant cell walls. Anal Biochem


190:129133
Brock FM, Forsberg CW, Buchanan-Smith JG (1982)
Proteolytic activity of rumen microorganisms and
effects of proteinase inhibitors. Appl Environ
Microbiol 44:561569
Brookman JL, Ozkose E, Rogers S et al (2000)
Identification of spores in the polycentric anaerobic
gut fungi which enhance their ability to survive. FEMS
Microbiol Ecol 31:261267
Brul S, Stumm CK (1994) Symbionts and organelles in
anaerobic protozoa and fungi. Trends Ecol Evol
9:319324
Bryant MP (1973) Nutritional requirements of the predominant rumen cellulolytic bacteria. Fed Proc
32:18091813
Castro-Montoya JM, Makkar HPS, Becker K (2011)
Chemical composition of rumen microbial fraction
and fermentation parameters as affected by tannins
and saponins using an in vitro rumen fermentation system rumen. Can J Anim Sci 91:433448
Cavalier-Smith T (1987) Eukaryotes with no mitochondria. Nature 326:332333
Chen H, Li XL, Ljungdahl LG (1995) Biomass degrading
enzymes from anaerobic rumen fungi. SAAS Bull
Biochem Biotechnol 8:16
Choudhury PK, Sirohi SK, Puniya AK et al (2012)
Harnessing the diversity of rumen microbes using
molecular approaches. In: Sirohi SK, Walli TK, Singh
B, Singh N (eds) Livestock Green House Gases: emission and options for mitigation. Satish Serial
Publishing House, Delhi, Section-B, Chapter-6,
pp 6582. ISBN 978-93-81226-52-065
Church DC (1969) Digestive physiology and nutrition of
ruminants. OSU. Book Stores, Corvallis
Clarke RTJ (1977) Protozoa in the rumen ecosystem. In:
Clarke RTJ, Bauchop T (eds) Microbial ecology of the
gut. Academic, London, pp 251275
Coleman GS (1975) The interrelationship between rumen
ciliate protozoa and bacteria. In: McDonald IW,
Warner ACI (eds) Digestion and metabolism in the
ruminant. University of New England Publishing Unit,
Armidale, pp 149164
Coleman GS (1983) The cellulolytic activity of thirteen
species of rumen entodiniomorphid protozoa. J
Protozool 30:36A
Craig WM, Broderick GA, Ricker DB (1987) Quantitation
of microorganisms associated with the particulate
phase of ruminal ingesta. J Nutr 117:5662
Dagar SS, Kumar S, Mudgil P, Singh R, Puniya AK
(2011) D1/D2 domain of large subunit rDNA for differentiation of Orpinomyces spp. Appl Environ
Microbiol 77(18):67226725
Davies DR, Theodorou MK, Lawrence MI (1993)
Distribution of anaerobic fungi in the digestive tract of
cattle and their survival in faeces. J Gen Microbiol
139:13951400
Dehority BA (2003) Rumen microbiology. Nottingham
University Press, Nottingham
Dey A, Sehgal JP, Puniya AK, Singh K (2004) Influence
of anaerobic fungal culture (Orpinomyces sp) admin-

13
istration on growth rate, ruminal fermentation and
nutrient digestion in calves. Asian Aust J Anim Sci
17:820824
Eggleston H (2006) 2006 IPCC guidelines for national
greenhouse gas inventories. Forestry 5:112
Embley TM, Horner DA, Hirt RP (1997) Anaerobic eukaryote evolution: hydrogenosomes as biochemically modified mitochondria? Trends Ecol Evol 12:437441
Ferry JG, Kastead KA (2007) Methanogenesis. In:
Cavicchioli R (ed) Archaea: molecular and cellular
biology. ASM Press, Washington, DC, pp 288314
Finlay BJ, Esteban G, Clarke KJ et al (1994) Some rumen
ciliates have endosymbiotic methanogens. FEMS
Microbiol Lett 117:157162
Freestone P, Lyte M (2010) Stress and microbial endocrinology: prospects for ruminant nutrition. Animal
4:12481257
Goel G, Puniya AK, Singh K (2005a) A study on tannic
acid resistance in ruminal streptococcal isolates. J
Basic Microbiol 45:243245
Goel G, Puniya AK, Singh K (2005b) Xylanolytic activity
of ruminal Streptococcus bovis in presence of tannic
acid. Ann Microbiol 4:295298
Goel G, Kumar A, Beniwal V, Raghav M, Puniya AK,
Singh K (2011) Degradation of tannic acid and
purification and characterization of tannase from
Enterococcus faecalis. Int Biodeterior Biodegrad
65:10611065
Gordon GLR, Phillips MW (1998) The role of anaerobic
gut fungi in ruminants. Nutr Res Rev 11:136
Griffith GW, Ozkose E, Theodorou MK et al (2009)
Diversity of anaerobic fungal populations in cattle
revealed by selective enrichment culture using different carbon sources. Fungal Ecol 2:8797
Gruninger RJ, Puniya AK, Callaghanc TM, Edwardsc JE,
Youssef N, Dagar SS, Fliegerova K, Griffith GW,
Forster R, Tsang A, McAllister T, Elshahed MS (2014)
Anaerobic fungi (Phylum Neocallimastigomycota):
advances in understanding of their taxonomy, life
cycle, ecology, role, and biotechnological potential.
FEMS Microbiol Ecol. doi:10.1111/1574-6941.12383
Heath IB, Kaminskyj SG, Bauchop T (1986) Basal body
loss during fungal zoospore encystment: evidence
against centriole autonomy. J Cell Sci 83:135140
Hibbett DS, Binder M, Bischoff JF et al (2007) A higherlevel phylogenetic classification of the fungi. Mycol
Res 111:509547
Ho YWBD (1995) Classification of anaerobic gut fungi
from herbivores with emphasis on rumen fungi from
Malaysia. Mycologia 87:655677
Hungate RE (1966) The rumen and its microbes.
Academic, New York, p 41
Iannotti EL, Kafkewitz D, Wolin MJ et al (1973) Glucose
fermentation products of Ruminococcus albus grown
in continuous culture with Vibrio succinogenes:
changes caused by interspecies transfer of H2. J
Bacteriol 114:12311240
Imai S (1998) Phylogenetic taxonomy of rumen ciliate
protozoa based on their morphology and distribution.
J Appl Anim Res 13:1736

14
Janssen PH, Kirs M (2008) Structure of the archaeal community of the rumen. Appl Environ Microbiol
74:36193625
Jeyanathan J, Kirs M, Rominus RS et al (2011)
Methanogen community structure in the rumens of
farmed sheep, cattle and red deer fed different diets.
FEMS Microbiol Ecol 74:311326
Joblin KN (1996) Options for reducing methane emissions from ruminants in New Zealand and Australia.
In: Bouma WJ, Pearman GI, Manning MR (eds)
Greenhouse: coping with climate change. CSIRO
Publishing, Collingwood, pp 437449
Joblin KN (1999) Ruminal acetogens and their potential
to lower ruminant methane emissions. Aust J Agr Res
50:13071313
Kamra DN (2005) Rumen microbial ecosystem. Curr Sci
89:124135
Khan S, Spudich JL, McCray JA et al (1995) Chemotactic
signal integration in bacteria. Proc Natl Acad Sci U S
A 92:97579976
Klieve AV, Swain RA (1993) Estimation of ruminal bacteriophage numbers by pulsed field gel electrophoresis
and laser densitometry. Appl Environ Microbiol
59:22992303
Klieve AV, Heck GL, Prance MA (1999) Genetic homogeneity and phage susceptibility of ruminal strains of
Streptococcus bovis isolated in Australia. Lett Appl
Microbiol 29:108112
Koike S, Kobayashi Y (2009) Fibrolytic rumen bacteria:
their ecology and function. Asian Aust J Anim Sci
22:131138
Koike S, Yoshitani S, Kobayashi Y (2003) Phylogenetic
analysis of fiber-associated rumen bacterial community and PCR detection of uncultured bacteria. FEMS
Microbiol Lett 229:2330
Kopecny J, Hodrova B (1995) Pectinolytic enzymes of
anaerobic fungi. Lett Appl Microbiol 20:312316
Krause D, Denman AE, Mackie RI (2003) Opportunities
to improve fiber degradation in the rumen: microbiology, ecology, and genomics. FEMS Microbiol Rev
27:663693
Kumar S, Dagar SS, Puniya AK (2012) Isolation and characterization of methanogens from rumen of Murrah
buffalo. Ann Microbiol 62(1):345350
Kumar S, Dagar SS, Puniya AK, Upadhyay RC (2013a)
Changes in methane emission, rumen fermentation in
response to diet and microbial interactions. Res Vet
Sci 94(2):263268
Kumar S, Dagar SS, Sirohi SK, Upadhyay RC, Puniya AK
(2013b) Microbial profiles, in vitro gas production and
dry matter digestibility based on various ratios of roughage to concentrate. Ann Microbiol 63(2):541545
Kumar S, Choudhury PK, Carro MD et al (2014) New
aspects and strategies for methane mitigation from
ruminants. Appl Microbiol Biotechnol 98:3144
Lange M, Westermann P, Ahring BK (2005) Archaea in
protozoa and metazoan. Appl Microbiol Biotechnol
66:465474
Larue R, Yu Z, Parisi VA (2005) Novel microbial diversity
adherent to plant biomass in the herbivore gastrointestinal tract, as revealed by ribosomal intergenic spacer

P.K. Choudhury et al.


analysis and rrs gene sequencing. Environ Microbiol
7:530543
Lee JH, Kumar S, Lee GH et al (2013) Methanobrevibacter
boviskoreani sp. nov., isolated from the rumen of
Korean native cattle. Int J Syst Evol Microbiol
63:4196-4201
Legay-Carmier F, Bauchart D (1989) Distribution of bacteria in the rumen contents of dairy cows given a diet
supplemented with soya-bean oil. Br J Nutr
61:725740
Leng RA (1982) Modification of rumen fermentation. In:
Hacker JB (ed) Nutritional limits to animal production
from pastures. CAB, Farnham Royal, pp 427453
Lindmark DG, Mller M (1973) Hydrogenosome, a cytoplasmic organelle of the anaerobic flagellate
Tritrichomonas foetus, and its role in pyruvate metabolism. J Biol Chem 248:77247728
Ljungdahl LG (2008) The cellulase/hemicellulase system
of the anaerobic fungus Orpinomyces PC-2 and aspects
of its applied use. Ann N Y Acad Sci 1125:308321
Lowe SE, Theodorou MK, Trinci APJ (1987) Isolation of
anaerobic fungi from saliva and faeces of sheep. J Gen
Microbiol 133:18291834
Luo Y, Pfister P, Leisinger T (2001) The genome of
archaeal prophage m100 encodes the lytic enzyme
responsible for autolysis of Methanothermobacter
wolfeii. J Bacteriol 183:57885792
Lyte M (2010) The microbial organ in the gut as a driver of
homeostasis and disease. Med Hypotheses 74:634638
Lyte M, Freestone P (2009) Microbial endocrinology
comes of age. Microbe 4:169175
Mackie RI, McSweeney CS, Aminov RI (1999) Rumen.
In: Encyclopedia of life sciences. Nature Publishing
Company, London
McAllister TA, Newbold CJ (2008) Redirecting rumen
fermentation to reduce methanogenesis. Aust J Exp
Agric 48:713
McAllister TA, Bae HD, Jones GA et al (1994) Microbial
attachment and feed digestion in the rumen. J Anim
Sci 72:30043018
Michel V, Fonty G, Millet L et al (1993) In vitro study of
the proteolytic activity of rumen anaerobic fungi.
FEMS Microbiol Lett 110:59
Miller MB, Bassler BL (2001) Quorum sensing in bacteria. Annu Rev Microbiol 55:165199
Minato H, Endo A, Higuchi M et al (1966) Ecological
treatise on the rumen fermentation. I. The fractionation of bacteria attached to the rumen digesta solids.
J Gen Appl Microbiol 12:3952
Minato H, Ishizaki S, Adachi Y et al (1989) Effect on
rumen microbial populations of ammonia treatment of
rice straw forage for steers. J Gen Appl Microbiol
35:113124
Minato H, Mitsumori M, Cheng KJ (1993) Attachment of
microorganisms to solid substrates in the rumen. In:
Proceedings of the MIE bioforum on genetics, biochemistry and ecology of lignocellulose degradation.
Institut Pasteur, Paris, pp 139145
Mitsumori M, Minato H (1997) Cellulose-binding proteins from rumen microorganisms. In: Onodera R,
Itabashi H, Ushida K, Yano H, Sasaki Y (eds) Rumen

Rumen Microbiology: An Overview

microbes and digestive physiology in ruminants.


Japan Scientific Societies Press/S. Karger, Tokyo/
Basel, pp 3545
Morgavi DP, Kelly WJ, Janssen PH et al (2013) Rumen
microbial (meta)genomics and its application to ruminant production. Animal 7:184201
Morvan B, Bonnemoy F, Fonty G et al (1996) Quantitative
determination of H2-utilizing acetogenic and sulfatereducing bacteria and methanogenic archaea from
digestive tract of different mammals. Curr Microbiol
32:129133
Moss AR, Jouany JP, Newbold J (2000) Methane production by ruminants: its contribution to global warming.
Ann Zootech 49:231253
Mountfort DO, Asher RA (1989) Production of xylanase
by the ruminal anaerobic fungus Neocallimastix frontalis. Appl Environ Microbiol 55:10161022
Nagpal R, Puniya AK, Singh K (2009) In-vitro activities
of immobilized anaerobic rumen fungus Caecomyces
sp for its use as direct-fed microbials. J Anim Feed Sci
18:758768
Nagpal R, Puniya AK, Singh K, Sehgal JP (2010)
Influence of bacteria and protozoa from the rumen of
buffalo on in-vitro activities of anaerobic fungus
Caecomyces sp isolated from the feces of elephant. J
Yeast Fungal Res 1(8):152156
Nagpal R, Puniya A, Sehgal J et al (2011) In vitro fibrolytic
potential of anaerobic rumen fungi from ruminants and
non-ruminant herbivores. Mycoscience 52:3138
Novotna Z, Prochazka J, Simunek J et al (2010) Xylanases
of anaerobic fungus Anaeromyces mucronatus. Folia
Microbiol (Praha) 55:363367
Odenyo AA, Osuji PO (1998) Tannin-tolerant ruminal
bacteria from East African ruminants. Can J Microbiol
44:905909
Orpin CG, Mathiesen SD, Greenwood Y et al (1985)
Seasonal changes in the ruminal microflora of the
high-arctic Svalbard reindeer (Rangifer tarandus platyrhynchus). Appl Environ Microbiol 50:144151
Paul SS, Kamra DN, Sastry VR et al (2004) Effect of
anaerobic fungi on in vitro feed digestion by mixed
rumen microflora of buffalo. Reprod Nutr Dev
44:313319
Paul SS, Deb SM, Punia BS et al (2010) Fibrolytic potential of anaerobic fungi (Piromyces sp.) isolated from
wild cattle and blue bulls in pure culture and effect of
their addition on in vitro fermentation of wheat straw
and methane emission by rumen fluid of buffaloes. J
Sci Food Agric 9:12181226
Pfister P, Wasserfallen A, Stettler R et al (1998) Molecular
analysis of Methanobacterium phage M2. Mol
Microbiol 30:233244
Phillips MW, Gordon GL (1988) Sugar and polysaccharide fermentation by rumen anaerobic fungi from
Australia, Britain and New Zealand. Biosystems
21:377383
Puniya AK, Salem AZM, Kumar S, Dagar SS, Griffith GW,
Puniya M, Ravella SR, Kumar N, Dhewa T, Kumar R
(2014) Role of live microbial feed supplements with reference to anaerobic fungi in ruminant productivity. J
Integr Agric. doi:10.1016/S2095-3119(14)60837-6

15
Ragsdale SW, Pierce E (2008) Acetogenesis and the
Wood-Ljungdahl pathway of CO (2) fixation. Biochim
Biophys Acta 1784:18731898
Rezaeian M, Beakes GW, Parker DS (2004) Distribution
and estimation of anaerobic zoosporic fungi along
the digestive tracts of sheep. Mycol Res
108:12271233
Saxena S, Sehgal JP, Puniya AK, Singh K (2010) Effect of
administration of rumen fungi on production performance of lactating buffaloes. Benefic Microbes
1(2):183188
Schultz JE, Breznak JA (1979) Cross-feeding of lactate
between Streptococcus lactis and Bacteroides sp. isolated from termite hindguts. Appl Environ Microbiol
37:12061210
Scott HW, Dehority BA (1965) Vitamin requirements of
several cellulolytic rumen bacteria. J Bacteriol
89:11691175
Sharma A, Chaudhary PP, Sirohi SK et al (2011) Structure
modeling and prediction of NADP oxidoreductase
enzyme
from
Methanobrevibacter
smithii.
Bioinformation 6:1519
Shelke S, Chhabra A, Puniya A et al (2009) In vitro degradation of sugarcane bagasse based ruminant rations
using anaerobic fungi. Ann Microbiol 59:415418
Singh N, Sirohi SK (2012) Biochemistry, bioenergetics
and genetics of methanogenesis in rumen methanogenic archaea. In: Sirohi SK, Walli TK, Singh B,
Singh N Livestock Green House Gases: Emission
and options for mitigation. Satish Serial Publishing
House, Delhi, Section-B, Chapter-8, pp 103120.
ISBN 978-93-81226-52-065
Sirohi SK, Pandey N, Singh B, Puniya AK (2010)
Rumen methanogens: a review. Indian J Microbiol
50:253262
Sirohi SK, Choudhury PK, Dagar SS et al (2012a)
Isolation, characterization and fibre degradation
potential of anaerobic rumen fungi from cattle. Ann
Microbiol 63:11871194
Sirohi SK, Singh N, Dagar SS, Puniya AK (2012b)
Molecular tools for deciphering the microbial community structure and diversity in rumen ecosystem. Appl
Microbiol Biotechnol 95(5):11351154
Sirohi SK, Chaudhary PP, Singh N, Singh D, Puniya AK
(2013a) The 16S rRNA and mcrA gene based comparative diversity of methanogens in cattle fed on high
fibre based diet. Gene 523(2):161166
Sirohi SK, Choudhury PK, Puniya AK, Singh D, Dagar
SS, Singh N (2013b) Ribosomal ITS1 sequencebased diversity analysis of anaerobic rumen fungi in
cattle fed on high fiber diet. Ann Microbiol
63(4):15711577
Sperandio V, Torres AG, Jarvis B et al (2003) Bacteriahost communication: the language of hormones. Proc
Natl Acad Sci U S A 100:89518956
Stack RJ, Hungate RE (1984) Effect of 3 phenylpropanoic
acid on capsule and cellulases of Ruminococcus albus
8. Appl Environ Microbiol 48:218223
Stanton TB (2007) Prophage-like gene transfer agents:
novel mechanisms of gene exchange for

16
Methanococcus, Desulfovibrio, Brachyspira, and
Rhodobacter species. Anaerobe 13:4349
Stewart CS, Bryant MP (1988) The rumen bacteria. In:
Hobson PN (ed) The rumen microbial ecosystem.
Elsevier, New York, pp 2175
Thareja A, Puniya AK, Goel G, Nagpal R, Sehgal JP,
Singh PK, Singh K (2006) In-vitro degradation of
wheat straw by anaerobic fungi from small ruminants.
Arch Anim Nutr 60:412417
Theodorou MK, Mennim G, Davies DR et al (1996)
Anaerobic fungi in the digestive tract of mammalian
herbivores and their potential for exploitation. Proc
Nutr Soc 55:913926
Thiele JH, Chartrain M, Zeikus JG (1988) Control of
interspecies electron flow during anaerobic digestion:
role of floc formation in syntrophic methanogenesis.
Appl Environ Microbiol 54:1019
Trinci APJ, Davies DR, Gull K et al (1994) Anaerobic
fungi in herbivorous animals. Mycol Res 98:129152
Tripathi VK, Sehgal JP, Puniya AK, Singh K (2007a)
Effect of administration of anaerobic fungi isolated
from cattle and wild blue bull (Boselaphus tragocamelus) on growth rate and fibre utilization in buffalo
calves. Arch Anim Nutr 61:416423
Tripathi VK, Sehgal JP, Puniya AK, Singh K (2007b)
Hydrolytic activities of anaerobic fungi from wild blue
bull (Boselaphus tragocamelus). Anaerobe 13:3639
Van der Giezen M (2002) Strange fungi with even stranger
insides. Mycologist 16:129131
Vogels GD, Hoppe WF, Stumm CK (1980) Association of
methanogenic bacteria with rumen ciliates. Appl
Environ Microbiol 40:608612

P.K. Choudhury et al.


Warner ACI (1956) Proteolysis by rumen microorganisms. J Gen Microbiol 14:749
Weller RA, Pilgrim AF (1974) Passage of protozoa and volatile
fatty acids from the rumen of the sheep and from a continuous in vitro fermentation system. Br J Nutr 32:341351
Whitman WB, Boone DR, Koga Y et al (2001) Taxonomy
of methanogenic archaea. In: Garrity GM (ed)
Bergeys manual of systematic bacteriology. Springer
Verlag, New York, pp 211213
Williams AG (1979) The selectivity of carbohydrate
assimilation in the anaerobic rumen ciliate Dasytricha
ruminantium. J Appl Bacteriol 47:511520
Williams AG, Coleman GS (1985) Hemicellulose degrading enzymes in rumen ciliate protozoa. Curr Microbiol
12:8590
Williams AG, Coleman GS (1992) The rumen protozoa.
Springer, New York, pp 483
Wolin MJ, Miller TL (1988) Microbe-microbe interactions. In: Hobson PN (ed) The rumen microbial ecosystems. Elsevier, London, pp 343459
Wollin MJ, Miller TL, Stewart CS (1997) Microbemicrobe interactions. In: Hobson PN, Stewart CS (eds)
The rumen microbial ecosystem. Chapman and Hall,
London, pp 467488
Wright ADG, Klieve AV (2011) Does the complexity
of the rumen microbial ecology preclude methane
mitigation.
Anim
Feed
Sci
Technol
166167:248253
Yarlett N, Orpin CG, Munn EA et al (1986)
Hydrogenosomes in the rumen fungus Neocallimastix
patriciarum. Biochem J 236:729739

Вам также может понравиться