You are on page 1of 8

Systematic and Applied Microbiology 40 (2017) 492–499

Contents lists available at ScienceDirect

Systematic and Applied Microbiology


journal homepage: www.elsevier.de/syapm

Assessing the impact of rumen microbial communities on methane


emissions and production traits in Holstein cows in a tropical climate
Camila S. Cunha a , Cristina M. Veloso a , Marcos I. Marcondes a , Hilario C. Mantovani b ,
Thierry R. Tomich c , Luiz Gustavo R. Pereira c , Matheus F.L. Ferreira a ,
Kimberly A. Dill-McFarland d,e,∗ , Garret Suen d,∗
a
Department of Animal Science, Universidade Federal de Viçosa, Peter Henry Rolfs Avenue, University Campus, Viçosa, Minas Gerais 36570-900,Brazil
b
Department of Microbiology, Universidade Federal de Viçosa, Peter Henry Rolfs Avenue, University Campus, Viçosa, Minas Gerais 36570-900, Brazil
c
Brazilian Agricultural Research Corporation, Embrapa Dairy Cattle, Eugênio do Nascimento Avenue, 610, Cascatinha, Juiz de Fora, Minas Gerais
36038-330, Brazil
d
Department of Bacteriology, University of Wisconsin-Madison, 1550 Linden Dr, Madison, Wisconsin 53706, USA
e
Department of Microbiology & Immunology, Life Sciences Institute, University of British Columbia, 2350 Health Sciences Mall, Vancouver, British
Columbia V6T 1Z3, Canada

a r t i c l e i n f o a b s t r a c t

Article history: The evaluation of how the gut microbiota affects both methane emissions and animal production is nec-
Received 28 May 2017 essary in order to achieve methane mitigation without production losses. Toward this goal, the aim of
Received in revised form 27 July 2017 this study was to correlate the rumen microbial communities (bacteria, archaea, and fungi) of high (HP),
Accepted 31 July 2017
medium (MP), and low milk producing (LP), as well as dry (DC), Holstein dairy cows in an actual tropical
production system with methane emissions and animal production traits. Overall, DC cows emitted more
Keywords:
methane, followed by MP, HP and LP cows, although HP and LP cow emissions were similar. Using next-
Methane emissions
generation sequencing, it was found that bacteria affiliated with Christensenellaceae, Mogibacteriaceae,
Milk production
Rumen microbiology
S24-7, Butyrivibrio, Schwartzia, and Treponema were negatively correlated with methane emissions and
Bacteria showed positive correlations with digestible dry matter intake (dDMI) and digestible organic matter
Archaea intake (dOMI). Similar findings were observed for archaea in the genus Methanosphaera. The bacte-
Fungi rial groups Coriobacteriaceae, RFP12, and Clostridium were negatively correlated with methane, but did
not correlate with dDMI and dOMI. For anaerobic fungal communities, no significant correlations with
methane or animal production traits were found. Based on these findings, it is suggested that manipula-
tion of the abundances of these microbial taxa may be useful for modulating methane emissions without
negatively affecting animal production.
© 2017 Elsevier GmbH. All rights reserved.

Introduction fungi and bacteria degrade host-indigestible fiber, and the result-
ing breakdown products are fermented mainly by bacteria into
The rumen is inhabited by a diversity of microorganisms, includ- important nutrients, such as volatile fatty acids (VFAs) [4,19]. Thus,
ing bacteria, archaea, and fungi. These microbes interact with each the rumen microbiota is responsible for producing the nutrients
other to coordinate the breakdown and fermentation of differ- used by the host for growth, reproduction, and ultimately, the gen-
ent feed components ingested by the ruminant host. In particular, eration of agriculturally-relevant products like milk and meat. In
addition to the production of nutrients such as VFAs, byproducts of
fiber degradation include hydrogen (H2 ) and carbon dioxide (CO2 ),
which are utilized by methanogenic archaea to produce methane
∗ Corresponding authors at: Department of Bacteriology, University of Wisconsin- (CH4 ) [43]. The release of methane not only deprives the host of
Madison 5159 MSB, 1550 Linden Drive, Madison, WI, 53706, USA. important carbon resources [17] but also significantly contributes
E-mail addresses: camilasoares4@yahoo.com.br (C.S. Cunha),
to greenhouse gas emissions and global warming [15].
cristina.veloso@ufv.br (C.M. Veloso), marcos.marcondes@ufv.br (M.I. Marcondes),
hcm6@ufv.br (H.C. Mantovani), thierry.tomich@embrapa.br (T.R. Tomich), The diet is influenced by the type of feed [31] and quality of
luiz.gustavo@embrapa.br (L.G.R. Pereira), matheus.fellipe234@gmail.com forage [29], which are primary factors in modulating the avail-
(M.F.L. Ferreira), dillmcfarlan@wisc.edu (K.A. Dill-McFarland), gsuen@wisc.edu ability of hydrogen used by methanogens in the rumen. However,
(G. Suen).

http://dx.doi.org/10.1016/j.syapm.2017.07.008
0723-2020/© 2017 Elsevier GmbH. All rights reserved.
C.S. Cunha et al. / Systematic and Applied Microbiology 40 (2017) 492–499 493

dietary strategies for mitigating methane emissions are costly, and sequentially over a unique 10-day period, due to limited methane
different feeds and forages are often used according to milk pro- emission equipment.
duction levels (high, low, dry) to meet the dairy cow’s nutritional All cows were housed in individual tie stalls and kept under
requirements at lower costs [44]. Considering this, there is a need the feeding management protocol they were already adapted to in
for studies that mirror what occurs on commercial dairy farms in the production system. These animals were not maintained on the
tropical conditions, since feeding animals at different diets accord- same diet because we purposefully wanted to mirror what occurred
ing to the lactation stage and milk production levels is a common on an actual dairy farm in tropical conditions. All diets consisted
practice. For example, cheaper, more fibrous diets fed to dry and of corn silage and concentrate comprising corn, soybean meal and
low-milk producing cows lead to more residual hydrogen because minerals (Table S1). The forage:concentrate ratios were 50:50 for
of the greater amounts of acetate produced, while more expensive, HP cows, 70:30 for the MP cows, 80:20 for LP cows, and 90:10 for
high-milk producing diets based on high starch content induce a DC cows. High-producing cows also received Coastcross grass (Cyn-
greater production of propionate, thereby lowering the produc- odon sp.) hay in the proportion of 8.5% of the total dry matter (DM)
tion of hydrogen and methane in the rumen [4]. Differences in (41.5% of corn silage). Animals were fed twice a day, at 7:00 am and
methane production due to diet are further impacted by climate 4:00 pm, and water was available ad libitum to all animals. Cows
and geography, as forages in tropical conditions often present lower in lactation were mechanically milked twice a day, at 6:00 am and
digestibility and their degradation in the rumen results in a higher 3:00 pm.
proportion of acetate and methane [10,27]. Thus, adjusted methane
mitigation practices may be necessary for animals with different
production levels fed different diets in different environments.
The composition of the rumen microbial community is also Methane emission measurements
thought to contribute significantly to methane production. There
are a number of studies that have focused on the relationship The trial for each production group was performed separately
between methane emissions and the rumen bacterial or archaeal for a 10-day period. Animals were adapted to tie stall housing and
microbiota, since these communities contribute directly to hydro- yokes from days 1 to 5. The adaptation period was appropriate
gen or methane production in the rumen [8,38,48]. However, there since slight interferences of feed intake were observed only in the
is little information regarding how the overall community of bacte- first two days of this period. Gas collections from animal eructa-
ria, archaea, and fungi in the same animal contributes to methane tion for evaluation of methane (CH4 ) emissions were performed
emissions, and this knowledge could be used to modulate methane on days 6–10 in 24 h periods. In situations where problems were
emissions through manipulation of the ruminal microbiota. This is encountered in the sampling process, the collection for that day
of particular importance as direct modulation of archaeal popula- was repeated after day 10.
tions to reduce CH4 emissions is difficult and can result in decreased Animal methane emission measurements were performed by
feed digestion due to ruminal hydrogen accumulation [28]. the tracer gas sulfur hexafluoride (SF6 ) technique [18] with the
Thus, feed degradation and methanogenesis in the rumen are adaptations described in Oss et al. [34]. Briefly, the flow of CH4
dependent on a combination of diet, the composition of the rumen released by the animal was calculated in relation to the flow of
microbiota [7], and the host [37]. However, it remains unclear SF6 , measured from the SF6 release rate from a permeation capsule
which attributes (diet, microbiota, genetics, etc.) result in the com- lodged in the rumen and from the concentrations of CH4 and SF6 in
bination of high production with low methane emissions in the gas samples [16]. Analysis of CH4 and SF6 levels were performed at
same animal. Therefore, it was hypothesized for this study that the Laboratory of Gas Chromatography at the Empresa Brasileira de
differences in bacterial, archaeal and/or fungal community com- Pesquisa Agropecuária (EMBRAPA) Dairy Cattle, Juiz de Fora, Minas
position would contribute to differences in the methane emissions Gerais, Brazil. The concentrations of CH4 and SF6 in the samples
of Holstein cows under a feeding management strategy based on were evaluated by gas chromatograph, as described in Schloss et al.
milk production level. To address this, the relationship between [40]. Raw CH4 emissions were corrected for digestible fractions
the overall ruminal microbial community, enteric methane pro- of dry matter intake (dDMI) and digestible organic matter intake
duction, and milk production traits in dairy cows was investigated (dOMI) by dividing raw values by dDMI and dOMI (Table 1 and
in a tropical production system. Table S3). As these characteristics were highly variable between
animals, the corrections were required to make the comparison
between animals more reliable, considering that microorganisms
Materials and methods would degrade only a fraction of ingested food.

Animals and management

All experiments were conducted at the Unit for Teaching, Food, refusals and feces sampling
Research, and Extension on Dairy Cattle (UEPE-GL), Universidade
Federal de Viçosa (UFV), Viçosa, Minas Gerais, Brazil. Microbiolog- From days 6 to 10, forage and refusals from each animal were
ical analyses were performed at the Department of Bacteriology, sampled. The quantities of food provided and the weight of the
University of Wisconsin-Madison, Madison, WI, USA. The project refusals were recorded. Concentrate ingredients were sampled dur-
was approved by the Ethics Committee in the Use of Production ing mixing. At day 6, feces were collected during a 24-h period,
Animals from the UFV (protocol number 99/2014). weighed, and a subsample was then taken from this collection.
In a functioning dairy system at UEPE-GL, five dry cows (DC, The food, refusal and feces samples of each animal were grouped
746 ± 38.61 kg; 5.24 ± 1.84 years) and 14 lactating cows were per trial and analyzed for dry matter (DM), according to Detmann
selected according to their milk production. Lactating cows were et al. [11] method INCT-CA n◦ G-003/1, as well as the crude protein
split into high-producing (HP, 25 ± 2.74 L milk d−1 ; 540 ± 70.26 kg; (CP), ether extract (EE), neutral detergent fiber (NDF), non-fiber car-
2.67 ± 0.67 years; n = 5), medium-producing (MP, 15 ± 0.85 L milk bohydrates (NFC) and ash (MM) contents [2]. These results were
d−1 ; 611 ± 57.29 kg; 4.41 ± 1.20 years; n = 5) or low-producing used for the calculation of the dry matter intake (DMI), dry mat-
(LP, 10 ± 0.76 L milk d−1 ; 624 ± 122.85 kg; 3.91 ± 1.46 years; n = 4) ter digestibility (DMD), organic matter intake (OMI) and organic
groups. Each of the four cohorts was measured and sampled matter digestibility (OMD) (Table S3) [3,24].
494 C.S. Cunha et al. / Systematic and Applied Microbiology 40 (2017) 492–499

Table 1
Means and standard deviations for methane emissions and production traits of Holstein cows.

Group CH4 g day−1 CH4 g kg−1 DMI CH4 g kg−1 dDMI CH4 g kg−1 dOMI DMI (kg) DMD dDMI (kg) dOMI (kg) MY (L day−1 )

HP 286.22 ± 67.43B 19.13 ± 4.16B 21.70 ± 4.68C 23.60 ± 5.04C 15.15 ± 2.79B 0.76 ± 0.02A 13.39 ± 2.66A 12.31 ± 2.40A 25.86 ± 6.73A
MP 369.92 ± 41.24A 23.08 ± 2.21AB 29.25 ± 3.07B 31.44 ± 3.43B 16.09 ± 1.99A 0.71 ± 0.02B 12.72 ± 1.70A 11.84 ± 1.63A 15.54 ± 1.91B
LP 174.68 ± 31.80C 13.29 ± 2.11C 19.52 ± 3.47C 20.64 ± 3.38C 13.22 ± 2.05BC 0.68 ± 0.05C 9.07 ± 1.88B 8.56 ± 1.76B 10.55 ± 1.51C
DC 277.35 ± 22.22B 25.09 ± 3.81A 38.10 ± 5.79A 40.09 ± 6.26A 11.16 ± 0.03C 0.66 ± 0.03C 7.35 ± 0.62B 6.99 ± 0.62B NA

CH4 = methane; DMI = dry matter intake; DMD = dry matter digestibility; dDMI = digestible dry matter intake; dOMI = digestible organic matter intake; MY = milk yield;
HP = high-producing (n = 5); MP = medium-producing (n = 5); LP = low-producing (n = 4); DC = dry cows (n = 5). Means followed by different letters in a column are different
by Tukey’s test (P < 0.10).

Ruminal sampling and fermentation pattern paired-end sequencing kit (v3) according to the manufacturer’s
guidelines.
Ruminal contents were sampled 4 h after morning feeding using
the stomach tube technique [13,23,26]. The initial 200 mL of col- DNA sequence processing analysis
lected fluid were discarded to avoid saliva contamination, and
rumen samples therefore represented only the liquid phase of The program mothur, version 1.36.1 [40], was used to analyze
ruminal content. A sample of approximately 70 mL was filtered the sequenced libraries following a protocol adapted from [22].
through four layers of cheesecloth and split into two aliquots. The Sequences that could not be aligned were chimeric, and all con-
first one was used for DNA extraction (50 mL), and the other (20 mL) taminants were removed. The sequences from bacteria and archaea
for volatile fatty acid (VFA) concentration analysis. Both samples were aligned against the SILVA 16S rRNA gene database [35] and
were frozen at −80 ◦ C until they were analyzed. classified using the Greengenes database [9]. For anaerobic fungal
VFA concentrations were analyzed using high-performance liq- analysis, sequences were classified using the UNITE database [21]
uid chromatography (HPLC). Briefly, after a centrifugation step, and contaminants were also removed. A sequence similarity of 97%
1.5 mL of a cell-free supernatant was treated as described by was used as a cut-off to classify operational taxonomic units (OTUs)
Siegfried et al. [41] and analyzed in a Dionex Ultimate 3000 Dual at the species level [39]. Taxonomic assignment of OTUs was also
chromatograph (Dionex Corporation, Sunnyvale, CA, USA) system performed with mothur. Microbial communities were normalized
with a Shodex RI-101 refractive index detector at 45 ◦ C, using a to equal sequence counts before further analyses. All sequences
Phenomenex Rezex ROA, 300 × 7.8 mm ion-exclusion column, also used for this study were publicly deposited in the NCBI Short Read
at 45 ◦ C. The mobile phase was H2 SO4 (5 mM) at a flow rate of Archive under BioProject PRJNA 380769.
0.7 mL min−1 .
Statistical analysis

Relative abundances were calculated as the number of


DNA extraction, amplicon library preparation and sequencing sequences of the taxa divided by the number of total sequences
present in the sample multiplied by 100. Alpha diversity indices
DNA extractions were performed on 25 mL rumen samples, as for richness (Chao1) and diversity (Shannon), as well as Good’s
described by Stevenson and Weimer [42], and then treated with coverage, were calculated in mothur.
RNAse Ribonuclease A (10 mg mL−1 ). Total DNA was quantified The program R [36] was used to test differences in CH4 emis-
®
(Table S2) using a Qubit 2.0 Fluoremeter (Invitrogen, San Diego, sions from each group using analysis of variance (ANOVA), followed
CA, USA). A two-step PCR was employed to amplify the V3 and V4 by Tukey’s honest significance test (TukeyHSD) for multiple com-
regions of the 16S ribosomal RNA gene (16S rRNA) for bacteria, the parisons, when necessary. Differences in richness and diversity of
V6 to V8 regions of 16S rRNA for archaea, and the ITS1 gene for fungi bacterial, archaeal and anaerobic fungal communities were also
[20], which corresponded to amplicon lengths of 571 bp, 608 bp and tested with ANOVA and TukeyHSD. For these tests, P < 0.10 was
366 bp, respectively. accepted as significant due to the high variability of the methane
The first PCR aimed to amplify the region of interest using spe- measurement technique [46].
cific primers attached to Illumina sequencing primers, and the Bacterial, archaeal and anaerobic fungal communities were
second one to attach dual indices and Illumina sequencing adapters. visualized using non-metric multidimensional scaling (nMDS)
The reactions were amplified under the following conditions: 95 ◦ C plots of the Bray-Curtis dissimilarity index in the vegan package
for 3 min, 95 ◦ C for 30 s, 55 ◦ C for 30 s, 72 ◦ C for 30 s and 72 ◦ C for [33]. Analysis of similarity (ANOSIM) [5] was performed to deter-
5 min. A total of 25 cycles were performed for bacterial and archaeal mine differences in the overall microbial community structure
amplicons, 35 cycles for fungal, and eight cycles for the second (Bray–Curtis) and composition (Jaccard) between groups (HP, MP,
PCR for all amplicons. The first PCR was performed using 5–200 ng LP and DC). Similarity percentage analysis (SIMPER) [6] was used
DNA, 0.4 ␮M of each primer, 1X KAPA HiFi HotStart ReadyMix (Kapa to evaluate the contribution of each OTU to the differences seen in
®
Biosystems , Wilmington, MA, United States) and ultrapure water the ANOSIM. The relative abundances of OTUs identified by SIMPER
to a total of 25 ␮L. This PCR product was purified using an Invitro- (at least 3% for bacteria and 8% for fungi) were compared using the
gen PureLink Pro96 PCR Purification Kit and 5 ␮L of product was Kruskal–Wallis non-parametric test (Agricolae package) [30]. Other
used in the second PCR. The products of the second PCR were differences in relative abundance of interest were also tested using
run on a 1% AquaPor LM low-melt agarose gel and the gel extrac- Kruskal–Wallis. For these tests, P < 0.05 was considered significant
tion was performed using the ZR-96 Zymoclean Gel DNA Recovery and P < 0.10 was a trend.
Kit (Zymo Research, Irvine, CA). DNA was then quantified using a To investigate if microbial community data and VFA concen-
®
Qubit 2.0 Fluoremeter, diluted, and equimolar pooled to 2 nM. The trations co-varied, a Mantel test [27] was performed. Mantel was
library was denatured at 96 ◦ C using 0.2 N NaOH, then diluted with also used to evaluate if bacterial, archaeal and anaerobic fungal
hybridization buffer to 20 pM, and combined with the PhiX con- communities co-varied. All these analyses were performed with
trol library (20 pM). The library was loaded into an Illumina MiSeq the aid of the vegan package. For these analyses, P < 0.05 was con-
cartridge and sequenced on an Illumina MiSeq using a 2 × 300 bp sidered significant. Spearman’s correlation was used to correlate
C.S. Cunha et al. / Systematic and Applied Microbiology 40 (2017) 492–499 495

Table 2
Means and standard deviations for volatile fatty acids in the rumen fluid of Holstein cows.

Group Total VFA (mmol L−1 ) Acetate (mmol L−1 ) Propionate (mmol L−1 ) Butyrate (mmol L−1 ) A:P

HP 124.47 ± 11.20A 72.47 ± 4.61A 31.90 ± 5.38A 12.15 ± 1.15 2.31 ± 0.26B
MP 108.27 ± 11.80AB 68.39 ± 7.45A 22.22 ± 1.67B 10.76 ± 1.86 3.07 ± 0.13A
LP 75.51 ± 21.84C 57.77 ± 5.89B 19.04 ± 2.52B 12.29 ± 4.96 3.10 ± 0.70A
DC 101.33 ± 3.60B 65.87 ± 1.90AB 20.98 ± 1.42B 8.98 ± 0.29 3.15 ± 0.20A

VFA = volatile fatty acid (acetate, propionate, butyrate, lactate, isobutyrate, isovalerate, valerate); A:P = acetate:propionate ratio. HP = high-producing; MP = medium-
producing; LP = low-producing; DC = dry cows. Values are mean ± standard deviation. Means followed by different letters in a column are different by Tukey’s test (P < 0.10).

continuous variables (CH4 emissions, dDMI, dOMI) with bacterial phylum Firmicutes, eight to Bacteroidetes, two to TM7 and one to
families and genera, as well as archaeal and fungal genera rela- Chloroflexi. At the genus level, the most abundant representatives
tive abundances, using the corrplot package [50]. Only correlations of this core bacterial microbiota included Prevotella, Ruminococcus,
significant at P < 0.05 were added to the resultant correlogram. Butyrivibrio, Clostridium, Coprococcus, Shuttleworthia, CF231 and
SHD-231. Within the archaeal community, two OTUs belonging to
the genera Methanobrevibacter and Methanosphaera were present
Results
in all samples. None of the anaerobic fungi OTUs were present in
all samples.
Methane emissions and physiological traits related to milk
production
Differences of rumen bacterial, archaeal and fungal diversity and
Raw CH4 emissions (Table 1) differed by production group composition related to milk production groups
(ANOVA, P < 0.01) and were highest in medium-producing cows
(MP, TukeyHSD P < 0.05) (Table S4). High-producing cows (HP) and The number of species and the distribution of taxa across all
dry cows (DC) had similar emission levels, but they were less than samples were evaluated using the Chao1 index for richness and
MP cows and low-producing cows (LP) (TukeyHSD P < 0.05). When Shannon’s index for diversity. Bacterial richness (ANOVA P = 0.04)
CH4 emissions were corrected for digestible dry matter intake but not diversity (ANOVA P = 0.26) varied by production group, with
(dDMI) and digestible organic matter intake (dOMI) (Table S4), DC LP cows showing higher richness (501.85 ± 58.44) than DC cows
had the highest CH4 emissions (TukeyHSD P < 0.05), followed by (366.22 ± 79.86, TukeyHSD P = 0.02) (Table S5). The richness and
MP, and HP/LP cows. MP was greater than LP (TukeyHSD P < 0.05) diversity of the archaeal and anaerobic fungal communities were
but a trend was only observed when compared to HP (TukeyHSD similar between groups (ANOVA P > 0.05).
P < 0.10). When comparing the dDMI and dOMI, HP and MP cows Physiological and production traits were significant in cluster-
were similar (TukeyHSD P > 0.10) but higher than LP and DC cows ing animals only in the bacterial and anaerobic fungal communities
(TukeyHSD P < 0.05, Table 1 and Table S4). (Fig. 2). Cows were observed to have similar bacterial community
VFA concentrations in rumen fluid (Table 2 and Table S4) were composition (Jaccard, ANOSIM P = 0.58) but different bacterial com-
lower for LP (TukeyHSD P < 0.05) cows, and HP cows had the highest munity structure (Bray–Curtis, ANOSIM P = 0.02) (Table S6). Several
propionate concentration (TukeyHSD P < 0.05), which led to a lower OTUs were found to influence beta-diversity (Tables S7 and S8), and
acetate:propionate ratio (TukeyHSD P < 0.05). OTUs within the order Clostridiales were mainly responsible for the
observed differences (SIMPER, Table S7).
Archaeal community composition (Jaccard, ANOSIM P = 0.40)
Rumen liquid fraction microbiota characterization
and structure (Bray-Curtis, ANOSIM P = 0.23) did not differ between
groups (Table S6). Similarly, no differences were observed between
After all filtering analyses, the number of sequences obtained
groups for the anaerobic fungal community composition (Jaccard,
was 214,676 for bacteria (mean = 11,298 ± 5109), 64,774 for
ANOSIM P = 0.40), but the community structure was influenced by
archaea (mean = 3409 ± 1555) and 12,544 for anaerobic fungi
the production group (Bray–Curtis, ANOSIM P = 0.01) (Table S6).
(mean = 660 ± 505). For all amplicons, Good’s coverage of all sam-
Differences in the relative abundances for specific taxa were
ples was >0.98.
also determined, as shown in Table 3 and Table S9. The rela-
The most abundant bacterial phyla within all samples included
tive abundance of Firmicutes varied largely among the production
the Firmicutes (49.18 ± 6.47%), Bacteroidetes (31.55 ± 9.07%), and
groups but the relative abundances of Bacteroidetes did not dif-
TM7 (8.36 ± 8.21%) (Fig. 1A). Unidentified sequences accounted
fer between groups (Kruskal–Wallis P > 0.05). DC cows had a
for 0.37 ± 0.32% of the total sequence counts. Among the Firmi-
greater relative abundance of Methanobrevibacter (95.75 ± 1.89%)
cutes, the genera Ruminococcus, Butyrivibrio and Coprococcus were
and lower abundance of Methanosphaera (3.01 ± 1.47%) than HP
the most abundant in all animals (8.25 ± 5.34%, 6.97 ± 3.75% and
cows (90.04 ± 3.84%, 7.79 ± 3.81%, respectively, Kruskal–Wallis
3.46 ± 2.28%). Within the Bacteroidetes, the genus Prevotella had
P < 0.05). The relative abundances of the main fungal genera were
the highest relative abundance at 62.48 ± 19.88%, followed by
similar between production groups (Kruskal–Wallis P > 0.05).
CF231 in the family Paraprevotellaceae with a relative abundance
of 2.72 ± 1.02%.
Methanobrevibacter was the most abundant genus in the Correlation between bacterial, archaeal and anaerobic fungal taxa
archaeal community (93.43 ± 3.18%), followed by Methanosphaera and physiological traits
(5.32 ± 2.84%) and Vadin CA11 (1.23 ± 1.56%) (Fig. 1B). For the
anaerobic fungal community, 73.19 ± 19.99% of the sequences were The strongest negative correlations found between CH4 emis-
unclassified at the phylum level, 21.95 ± 17.40% were assigned sions and bacteria were for sequences belonging to the family
to the genus Caecomyces, 4.16 ± 7.46% to genus Orpinomyces and Coriobacteriaceae (r = −0.86, Spearman’s test P = 0.01), as well as
0.70 ± 1.30% to genus Neocallimastix (Fig. 1C). the genera Butyrivibrio (r = −0.65, Spearman’s test P = 0.01) and
A total of 32 bacterial OTUs were found to be ubiquitous across Schwartzia (r = −0.62, Spearman’s test P = 0.01) (Table S10). Fam-
all samples in this study and altogether represented 3.14% of the ilies RF16 and Succinivibrionaceae showed positive correlations of
relative abundance of the entire dataset. Of these, 21 belonged to 0.49 to 0.60 (Spearman’s test P < 0.05) with CH4 emissions, while
496 C.S. Cunha et al. / Systematic and Applied Microbiology 40 (2017) 492–499

Fig. 1. Percentage relative abundance of bacterial phyla (A), archaeal genera (B) and anaerobic fungal genera (C) in the rumen of Holstein cows at different production levels.
HP = high-producing; MP = medium-producing; LP = low-producing; DC = dry cows.

Table 3
Means and standard deviations for relative abundances of ruminal microbial groups related to production groups.

Microbial taxa HP MP LP DC

Firmicutes 47.87 ± 4.21B 50.62 ± 3.97A 56.38 ± 5.80A 42.15 ± 5.61B


Bacteroidetes 33.19 ± 10.45 34.10 ± 11.79 26.81 ± 7.53 31.15 ± 6.70
Coriobacteriaceae 1.10 ± 1.53A 0.32 ± 0.21B 0.87 ± 0.41B 0.12 ± 0.23C
Succinivibrionaceae 0.52 ± 0.51B 1.68 ± 2.46AB 0.25 ± 0.23B 3.63 ± 2.09A
OTU08 Ruminococcaceae 0.55 ± 0.18C 4.89 ± 2.38A 4.46 ± 2.45A 1.79 ± 1.73B
Methanobrevibacter 90.04 ± 3.84A 94.54 ± 0.74AB 94.07 ± 2.52AB 95.75 ± 1.89B
Methanosphaera 7.79 ± 3.81A 4.85 ± 0.75B 5.75 ± 2.39B 3.01 ± 1.47B
Vadin CA11 2.16 ± 2.60 1.13 ± 0.58 0.18 ± 0.17 1.24 ± 1.26
Caecomyces 23.85 ± 22.66 32.12 ± 14.76 16.17 ± 16.69 14.51 ± 13.59
Orpinomyces 0.38 ± 0.84 1.5 ± 1.55 7.08 ± 6.71 8.25 ± 12.53
Neocallimastix 1.54 ± 1.63 0.74 ± 1.65 0.00 ± 0.00 0.37 ± 0.82
OTU0003 Neocallimastigaceae 28.3 ± 11.57A 34 ± 18.47A 6.67 ± 8.68B 6.06 ± 6.77B

HP = high-producing; MP = medium-producing; LP = low-producing; DC = dry cows. Means followed by different letters in a row are different by the Kruskal-Wallis test
(P < 0.10).

Methanobrevibacter showed weak correlations with this variable The families Christensenellaceae, Mogibacteriaceae and S24-7, as
(r = 0.35, Spearman’s test P > 0.05) (Fig. 3; Table S10). well as the genera Butyrivibrio, Schwartzia and Treponema, were
negatively correlated with CH4 emissions, with correlations rang-
C.S. Cunha et al. / Systematic and Applied Microbiology 40 (2017) 492–499 497

Fig. 2. Non-metric multidimensional scaling plot of Bray-Curtis dissimilarity indices of bacterial (A) and anaerobic fungal (B) community structure. Variables that had a
significant fit in envfit were fitted to the x-y plane as arrows. HP = high-producing; MP = medium-producing; LP = low-producing; DC = dry cows. CH4 = methane emissions;
dDMI = digestible dry matter intake; dOMI = digestible organic matter intake; A:P = acetate to propionate ratio.

Fig. 3. Correlation analysis between CH4 emissions, production parameters, and relative abundances of ruminal microbial taxa at the family and genus taxonomic levels.
Spearman’s linear correlation values for families/genera across all samples. Only significant correlations (P < 0.05) for at least one of the analyzed variables are shown. Names
in bold indicate families, whereas non-bold names indicate genera. Names in black and green indicate bacterial and archaeal taxa, respectively. CH4 = methane emissions;
dDMI = digestible dry matter intake; dOMI = digestible organic matter intake.

ing from −0.58 to −0.73 (Spearman’s test P < 0.01), but positively production precursors used by archaea. The study sought to deter-
correlated with the digestible fractions of DMI and OMI (r = 0.52 mine the interplay between the microbiota, methane and animal
to 0.73, Spearman’s test P < 0.05) (Table S10). OTUs affiliated to production, since efforts to decrease methane have often resulted
Coriobacteriaceae, RFP12, Clostridium, and Methanosphaera were in decreases in animal production. Understanding this complex
negatively correlated with CH4 emissions, with correlations rang- microbiota could lead to novel ways for mitigating methane pro-
ing from −0.86 to −0.55 (Spearman’s test P < 0.05) but did not duction without the concomitant losses in animal production.
show significant correlations with dDMI and dOMI (Spearman’s test Within the archaeal community, no significant correlations
P > 0.05). Correlations between CH4 emissions corrected for dDMI were found between Methanobrevibacter and CH4 emissions. This
or dOMI and anaerobic fungi were not found to be significant. was unexpected since Methanobrevibacter has previously been
Finally, none of the bacteria, archaea or anaerobic fungi that cor- associated with higher CH4 emission in cattle [48]. The lack of
related with CH4 emissions showed significant correlations (Mantel association found here was likely due to differences in abso-
test, P > 0.01) with total VFAs or the individual concentrations of lute versus relative Methanobrevibacter abundance, since it was
acetate, propionate or butyrate. previously determined that the absolute abundance of particular
methanogens determines CH4 production in the rumen [8,51]. In
addition, there may have been species level differences within the
Discussion Methanobrevibacter that contributed to methane emissions [7]. The
findings also revealed that archaea in the genus Methanosphaera
In this study, the liquid fraction of the rumen microbiota of had negative correlations with CH4 emissions, which corroborated
Holstein cows and its relationship to both CH4 emissions and the findings of Kittelmann et al. [19]. Comparisons of the pathways
production traits in a functioning, tropical dairy system were inves- for CH4 production by Methanobrevibacter and Methanosphaera
tigated. Since the aim was to model commercial dairy farms, the have indicated that Methanobrevibacter produces one mole of CH4
feed strategy used reproduced common practices on tropical farms. for each mole of CO2 [14], while Methanosphaera requires four
A recent study [7] showed that differences in methane emissions moles of methanol to produce three moles of CH4 [12]. This shows
from cows at similar physiological stages could be linked to dif- that, for any given carbon input, Methanobrevibacter produces more
ferences in the ruminal archaeal and bacterial communities. In the CH4 than Methanosphaera. Thus, the findings from the current
present study, animals were considered at different physiological study that Methanosphaera negatively correlated with CH4 may
states, including dry cows, and the focus was on archaea, in addi- reflect its lower CH4 production per unit of carbon, relative to
tion to bacteria and fungi, given their role in mediating the methane
498 C.S. Cunha et al. / Systematic and Applied Microbiology 40 (2017) 492–499

Methanobrevibacter. An alternative explanation may be the avail- increase production, their association with decreased CH4 emis-
ability of methylotrophic substrates (e.g. methanol), which would sions is desirable, and future work pinpointing their role in CH4
strongly influence CH4 production by both Methanobrevibacter and emissions should be considered. Interestingly, none of the bacterial
Methanosphaera. taxa previously found to correlate with lower methane emissions,
In addition, archaea from the genus Methanobrevibacter were namely those in the family Succinivibrionaceae or the genus Pre-
found to be negatively correlated with dDMI and dOMI, indicat- votella [7], were found to be associated with lower methane and
ing this genus’s association with lower production efficiency. This increased or unchanged production in this study. This could be due
was confirmed by the higher relative abundance of this genus in to numerous factors, including diet and the impact of a tropical
high forage fed animals, such as DC cows. In contrast, archaea in management system on these animals.
the genus Methanosphaera showed a moderate positive correlation Finally, the study also documented the ruminal anaerobic fungi
with dDMI and dOMI. Previous work found no association between and their relationship to both production and CH4 emissions. Given
production traits and specific ruminal archaea [7]; thus, the corre- their role in degrading fibrous feed consumed by the host, fungi may
lations identifying specific genera of archaea in this study provided indirectly affect CH4 emissions by modulating fermentation pat-
new information in this research area. terns. In line with other studies [23,25], the majority of the rumen
As part of CH4 production, archaea are responsible for removing liquid-associated fungi were identified as unclassified (73.19%).
H2 produced during the degradation of feed by other microorgan- Due to their primary function, anaerobic fungi are predominantly
isms. This also facilitates the reduction of available CO2 [43]. This found attached to the particles of the solid phase of ruminal content
is of particular importance, as H2 accumulation is harmful to the [1] and the findings of the present study may be due to the sam-
correct functioning of the rumen, since it inhibits upstream fermen- pling approach used, which only considered the liquid portion of
tation and reduces overall production efficiency [28]. As a result, the rumen. This may explain why no significant associations were
depleting archaea from the rumen without an associated strategy found between the anaerobic fungal community and CH4 emis-
for reducing H2 production is not a desirable CH4 mitigation option sions. Due to the non-significant correlations with fungi found in
[28,49]. Given that the bacterial and fungal communities provide the study, it is suggested that future work on the ruminal anaerobic
the substrates for CH4 production by ruminal archaea, understand- fungi should include those populations found in the solid portion
ing the interplay between all three ruminal kingdoms is necessary of the rumen in order to understand their role better in mediating
for gaining insights into overall CH4 production. CH4 emissions.
The analysis of the ruminal bacterial microbiota indicated
that bacteria in the families Christensenellaceae, Mogibacteriaceae
Conclusions
and S24-7, as well as in the genera Butyrivibrio, Schwartzia, and
Treponema, were negatively correlated with CH4 emissions and
By considering the ruminal bacteria, fungi, and archaea within
positively correlated with dDMI and dOMI. These attributes are
the context of CH4 emissions and animal production, the findings
highly desirable, as they demonstrate the feasibility of high milk
identified a number of bacterial families and genera associated
production while lowering CH4 emissions. Therefore, it is specu-
with decreased CH4 emissions without reducing animal produc-
lated that bacteria within these families and genera are likely to
tion. The study specifically mirrored tropical commercial dairy
be able to increase DMI because their members include known
farms in an effort to understand more fully how such manage-
degraders of fibrous feed. Given that fiber is primarily responsi-
ment systems, which work to maximize production, impact CH4
ble for ruminal distention and filling, and that these attributes are
emissions. Based on the results, it is suggested that attempts to
related to the satiety response in cows [45], one hypothesis is that
increase the abundances of microbial groups, such as Christensenel-
increasing the rate of fiber degradation could lead to more DM and
laceae, Mogibacteriaceae, S24-7, Butyrivibrio, Shwartzia, Treponema
OM intake, which in turn could stimulate the flow of digesta and
and Methanosphaera, may result in a decrease in CH4 emissions
the stimulus to eat.
and an improvement in animal production traits, while decreasing
In addition, genera such as Butyrivibrio are known for produc-
taxa affiliated to Coriobacteriaceae, RFP12 and Clostridium is likely
ing butyrate [32], and members of this genus can direct ruminal
to reduce CH4 emissions without any impact on animal production.
fermentation towards more butyrate and less acetate, which leads
Future work considering these taxa should be undertaken to assess
to lower levels of H2 and CH4 . Indeed, it was observed that among
their impact more fully on increasing production, such as meat and
cows that diverged the most in relation to CH4 emission (HP versus
milk, while limiting CH4 emissions.
DC), butyrate concentrations were greater for HP cows (Table 2),
which also emitted less CH4 than DC cows. Butyrate production
in the rumen consumes hydrogen, so these bacteria may serve Acknowledgements
as competitors for hydrogen with methanogenic archaea. This is
in contrast to previous reports where low methane emitters had This work was supported by the Brazilian governmental
lower butyrate proportions than high methane emitters on the agencies CNPq, CAPES, FAPEMIG, INCT-CA, Embrapa, and PECUS-
same diet [7]. However, this difference is unsurprising when con- RumenGases. The work was also supported, in part, by United
sidering that dietary differences existed in this study and the lowest States Department of Agriculture, National Institute of Food and
emitting group (LC, 175 g d−1 ) was much lower than that reported Agriculture grant #WIS01729 to G.S. We thank all members of the
by Danielsson et al. [7] (301 g d−1 ). Veloso, Marcondes, Mantovani and Suen laboratories for their sup-
Bacterial taxa were also found that negatively correlated with port and helpful discussions. We would also like to acknowledge
CH4 emission and had no significant associations to production the CAPES Doctorate Sandwich Program that facilitated the collab-
traits. These included the bacterial OTUs within the Coriobacteri- oration enabling this study.
aceae, RFP12, and Clostridium. Previous work confirms the findings
of the current study that Clostridium was negatively correlated with
CH4 emissions [47], although its actual role in CH4 is likely to be Appendix A. Supplementary data
variable given that this genus encompasses a wide variety of species
with numerous metabolic abilities. However, the other associated Supplementary data associated with this article can be found,
taxa have been less well studied and their role in methane pro- in the online version, at http://dx.doi.org/10.1016/j.syapm.2017.07.
duction in the rumen is unknown. While these bacteria may not 008.
C.S. Cunha et al. / Systematic and Applied Microbiology 40 (2017) 492–499 499

References Neocallimastigomycota) in ruminant and non-ruminant herbivores. ISME J. 4,


1225–1235.
[1] Akin, D.E., Borneman, W.S. (1990) Role of rumen fungi in fiber degradation. J. [26] Lodge-Ivey, S.L., Browne-Silva, J., Horvath, M.B. (2009) Technical note: Bacterial
Dairy. Sci. 73, 3023–3032. diversity and fermentation end products in rumen fluid samples collected via
[2] Association of Official Analytical Chemists (AOAC), 1990 Official methods of oral lavage or rumen cannula. J. Anim. Sci. 87, 2333–2337.
analyses of the AOAC, Fifteenth edition, AOAC International, Washington, pp. , [27] Mantel, N. (1967) The detection of disease clustering and a generalized regres-
1–1230. sion approach. Cancer Res. 27, 209–220.
[3] (2006) Nutrição de Ruminantes. In: Berchiellli, T.T., Pires, A.V., Oliveira, S.G. [28] Martin, C., Morgavi, D.P., Doreau, M. (2010) Methane mitigation in ruminants:
(Eds.), FUNEP, Jaboticabal, 583p. from microbe to the farm scale. Animal 4, 351–365.
[4] Boadi, D., Benchaar, C., Chiquette, J., Massé, D. (2004) Mitigation strategies [29] McCaughey, W.P., Wittenberg, K., Corrigan, D. (1999) Impact of pasture type on
to reduce enteric methane emissions from dairy cows: update review. Can. methane production by lactating beef cows. Can. J. Anim. Sci. 79, 221–226.
J. Anim. Sci. 84, 319–335. [30] Mendiburu, F., (2016) Agricolae: Statistical Procedures for Agricul-
[5] Clarke, K.R. (1993) Non-parametric multivariate analyses of changes in com- tural Research. R package version 1. 2–4. http://CRAN.R-project.org/
munity structure. Aust. J. Ecol. 18, 117–143. package=agricolae.
[6] Clarke, K.R., Warwick, R.M. 1994 Change in marine communities: an approach [31] Menezes, A.B., Lewis, E., O’Donovan, M., O’Neill, B.F., Clipson, N., Doyle, E.M.
to statistical analysis and interpretation, Natural Environment Research Coun- (2011) Microbiome analysis of dairy cows fed pasture or total mixed ration
cil, UK. diets. FEMS Microbiol. Ecol. 78, 256–265.
[7] Danielsson, R., Dicksved, J., Sun, L., Gonda, H., Müller, B., Schnürer, A., Bertils- [32] Mohammed, R., Brink, G.E., Stevenson, D.M., Neumann, A.P., Beauchemin, K.A.,
son, J. (2017) Methane production in dairy cows correlates with rumen Suen, G., Weimer, P.J. (2014) Bacterial communities in the rumen of Holstein
methanogenic and bacterial community structure. Front. Microbiol. 8, 226. heifers differ when fed orchardgrass as pasture vs hay. Front. Microbiol. 5, 689.
[8] Danielson, R., Schnürer, A., Arthurson, V., Bertilsson, J. (2012) Methanogenic [33] Oksanen, J., Blanchet, F.G., Kindt, R., Legendre, P., Minchin, P.R., O’Hara, R.B.,
population and methane production in Swedish dairy cows fed different levels Simpson, G.L., Solymos, P., Stevens, M.H.H., Wagner, H. 2015 Vegan: community
of forage. Appl. Environ. Microbiol. 78, 6172–6179. ecology package http://vegan.r-forge.r-project.org/.
[9] DeSantis, T.Z., Hugenholtz, P., Larsen, N., Rojas, M., Brodie, E.L., Keller, K., Huber, [34] Oss, D.B., Marcondes, M.I., Machado, F.S., Pereira, L.G.R., Tomich, T.R., Ribeiro,
T., Dalevi, D., Hu, P., Andersen, G.L. (2006) Greengenes, a chimera-checked 16S G.O., Jr, Chizzotti, M.L., Ferreira, A.L., Campos, M.M., Maurício, R.M., Chaves, A.V.,
rRNA gene database and 666 workbench compatible with ARB. Appl. Environ. McAllister, T.A. (2016) An evaluation of the face mask system based on short-
Microbiol. 72, 5069–5072. term measurements compared with the sulfur hexafluoride (SF6 ) tracer, and
[10] Detmann, E., Paulino, M.F., Valadares Filho, S.C., Huhtanen, P. (2014) Nutritional respiration chamber techniques for measuring CH4 emissions. Anim. Feed Sci.
aspects applied to grazing cattle in the tropics: a review based on Brazilian Technol. 216, 49–57.
results. Semina: Ciências Agrárias, Londrina, v.35, n.4, Suppl., pp. 2829–2854. [35] Pruesse, E., Quast, C., Knittel, K., Fuchs, B.M., Ludwig, W., Peplies, J., Glockner,
[11] Detmann, E., Souza, M.A., Valadares Filho, S.C., Queiroz, A.C., Berchielli, T.T., F.O. (2007) SILVA: a comprehensive online resource for quality checked and
Saliba, E.O.S., Cabral, L.S., Pina, D.S., Ladeira, M.M., Azevêdo, J.A.G. 2012 Méto- 660 aligned ribosomal RNA sequence data compatible with ARB. Nucl. Acids
dos para análise de alimentos, first ed., Suprema, Visconde do Rio Branco, MG, Res. 35, 7188–7196.
Brazil. [36] R Core Team, 2015 R: a language and environment for statistical computing, R
[12] Fricke, W.F., Seedorf, H., Henne, A., Kruer, M., Liesegang, H., Hedderich, R., Foundation for Statistical Computing, Vienna, Austria.
Gottschalk, G., Thauer, R.K. (2006) The genome sequence of Methanosphaera [37] Roehe, R., Dewhurst, R.J., Duthie, C.A., Rooke, J.A., McKain, N., Ross, D.W., Hylop,
stadtmanae reveals why this human intestinal archaeon is restricted to J.J., Waterhouse, A., Freeman, T.C., Watson, M., Wallace, R.J. (2016) Bovine host
methanol and H2 for methane formation and ATP synthesis. J. Bacteriol. 188, genetic variation influences rumen microbial methane production with best
624–658. selection criterion for low methane emitting and efficiently feed converting
[13] Henderson, G., Cox, F., Ganesh, S., Jonker, A., Young, W., Global Rumen Census hosts based on metagenomic gene abundance. PLoS Genet. 12, e1005846.
Collaborators, Hansen, P.H. (2015) Rumen microbial community composition [38] Ross, E.M., Moate, P.J., Marett, L., Cocks, B.G., Hayes, B.J. (2013) Investigating
varies with diet and host, but a core microbiome is found across a wide geo- the effect of two methane-mitigating diets on the rumen microbiome using
graphical range. Sci. Rep. 5, 14567. massively parallel sequencing. J. Dairy Sci. 96, 6030–6046.
[14] Hook, S.E., Wright, A.-D.G., McBride, B.W. (2010) Methanogens: methane pro- [39] Schloss, P.D., Handelsman, J. (2005) Introducing DOTUR, a computer program
ducers of the rumen and mitigation strategies. Archaea, 945785. for defining operational taxonomic units and estimating species richness. Appl.
[15] Intergovernmental Panel on Climate Change. Climatechange 2014 Synthesis Environ. Microbiol. 71, 1501–1506.
report 2014. [40] Schloss, P.D., Westcott, S.L., Ryabin, T., Hall, J.R., Hartmann, M., Hollister,
[16] Johnson, K.A., Huyler, M., Westberg, H.H., Lamb, B., Zimmerman, P. (1994) E.B., Lesniewski, R.A., Oakley, B.B., Parks, D.H., Robinson, C.J., Sahl, J.W.,
Measurement of methane emissions from ruminant livestock using a sulphur Stres, B., Thallinger, G.G., Van Horn, D.J., Weber, C.F. (2009) Introducing
hexafluoride tracer technique. Environ. Sci. Technol. 28, 59–362. Mothur: open-source, platform-independent, community supported software
[17] Johnson, K.A., Johnson, D.E. (1995) Methane emissions from cattle. J. Anim. Sci. for describing and comparing microbial communities. Appl. Environ. Microbiol.
73, 2483–2492. 75, 7537–7541.
[18] Johnson, K.A., Westberg, H.H., Michal, J.J., Cossalman, M.W. (2007) The SF6 [41] Siegfried, B.R., Ruckemann, H., Stumpf, G. 1984 Method for the determination
tracer technique: methane measurement from ruminants. In: Makkar, H.S., of organic acids in silage by high performance liquid chromatography, Vol. 37,
Vercoe, P. (Eds.), Measuring methane production from ruminants, Springer, Landwirtsch. Forsch, p. 298.
Netherlands, pp. 33–67. [42] Stevenson, D.A., Weimer, P.J. (2007) Dominance of Prevotella and low abun-
[19] Kittelmann, S., Pinares-Patiño, C.S., Seedorf, H., Kirk, M.R., Ganesh, S., McEwan, dance of classical ruminal bacterial species in the bovine rumen revealed by
J.C., Janssen, P. (2014) Two different bacterial community types are linked with relative quantification real-time PCR. Appl. Microbiol. Biotechnol. 75, 165–174.
the low-methane emission trait in sheep. PLoS One 9, e103171. [43] Stewart, C.S., Flynt, H.J., Bryant, M.P. (1997) The rumen bacteria. In: Hobson,
[20] Kittelmann, S., Seedorf, H., Walters, W.A., Clemente, J.C., Knight, R., Gordon, J.I., P.N., Stewart, C.S. (Eds.), The rumen microbial ecosystem, Blackie Academic
Janssen, P. (2013) Simultaneous amplicon sequencing to explore co-occurrence and Professional, New York, NY, pp. 10–72.
patterns of bacterial, archaeal and eukaryotic microorganisms in rumen micro- [44] St-Pierre, N.R., Thraen, C.S. (1999) Animal grouping strategies, sources of vari-
bial communities. PLoS One 8, e47879. ation, and economic factors affecting nutrient balance on dairy farms. J. Anim.
[21] Kõljalg, U., Nilsson, R.H., Abarenkov, K., Tedersoo, L., Taylor, A.F.S., Bahram, Sci. 77, 72–83.
M., Bates, S.T., Bruns, T.D., Bengtsson-Palme, J., Callaghan, T.M., Douglas, B., [45] Van Soest, P.J. 1982 Gastrointestinal fermentations. In: Nutritional ecology of
Drenkhan, T., Eberhardt, U., Dueñas, M., Grebenc, T., Griffith, G.W., Hartmann, the ruminant. Ruminant metabolism, nutritional strategies, the cellulolytic fer-
M., Kirk, P.M., Kohout, P., Larsson, E., Lindahl, B.D., Lücking, R., Martín, M.P., mentation and the chemistry of forages and plant fibers, Cornell University
Matheny, P.B., Nguyen, N.H., Niskanen, T., Oja, J., Peay, K.G., Peintner, U., Peter- Press, Ithaca, NY, pp. , 152–177.
son, M., Põldmaa, K., Saag, L., Saar, I., Schüßler, A., Scott, J.A., Senés, C., Smith, [46] Vlaming, J.B., Lopez-Villalobos, N., Brookes, I.M., Hoskin, S.O., Clark, H. (2008)
M.E., Suija, A., Taylor, D.L., Telleria, M.T., Weiss, M., Larsson, K.-H. (2013) Within- and between-animal variance in methane emissions in non-lactating
Towards a unified paradigm for sequence-based identification of fungi. Mol. dairy cows. Aust. J. Exp. Agric. 48, 124–127.
Ecol. 22, 5271–5277. [47] Wallace, R.J., Rooke, J.A., Duthie, C., Hyslop, J.J., Ross, D.W., McKain, N., Souza,
[22] Kozich, J.J., Westcott, S.L., Baxter, N.T., Highlander, S.K., Schloss, P.D. (2013) S.M., Snelling, T.J., Waterhouse, A., Roehe, R.1 (2014) Archaeal abundance in
Development of a dual-index sequencing strategy and curation pipeline for post-mortem ruminal digesta may help predict methane emissions from beef
analyzing amplicon sequence data on the MiSeq Illumina sequencing platform. cattle. Sci. Rep. 4, 5892.
Appl. Environ. Microbiol. 79, 5112–5120. [48] Wallace, R.J., Rooke, J.A., McKain, N., Duthie, C.A., Hyslop, J.J., Ross, D.W., Water-
[23] Kumar, S., Indugu, N., Vecchiarelli, B., Pitta, D.W. (2015) Associative patterns house, A., Watson, M., Roehe, R. (2015) The rumen microbial metagenome
among anaerobic fungi, methanogenic archaea, and bacterial communities in associated with high methane production in cattle. BMC Genomic 16, 839.
response to changes in diet and age in the rumen of dairy cows. Front. Microbiol. [49] Wedlock, D.N., Janssen, P.H., Leahy, S., Shu, D., Buddle, B.M. (2013) Progress in
6, 781. the development of vaccines against rumen methanogens. Animal 7, 244–252.
[24] Lana, R.P. 2005 Nutrição e Alimentação animal: mitos e realidades, UFV, Viçosa, [50] Wei, T. (2012) Package ‘Corrplot’.
pp. 344. [51] Zhou, M., Chung, Y.-H., Beauchemin, K.A., Holtshausen, L., Oba, M., McAllister,
[25] Liggenstoffer, A.S., Youssef, N.H., Couger, M., Elshahed, M.S. (2010) Phylo- T.A., Guan, L.L. (2011) Relationship between rumen methanogens and methane
genetic diversity and community structure of anaerobic gut fungi (phylum production in dairy cows fed diets supplemented with a feed enzyme additive.
J. Appl. Microbiol. 111, 1148–1158.

You might also like