837
Views
1
CrossRef citations to date
0
Altmetric
Research Article

Initial effects of supplemental forages and feedstuffs on bovine rumen ecology in vitro as determined by DNA-based molecular procedures

ORCID Icon, , &
Pages 268-280 | Received 25 Sep 2019, Accepted 04 Jun 2020, Published online: 22 Jun 2020

ABSTRACT

This research aided in determining the impacts on rumen microbial ecology when supplemental forages and feedstuffs were added for 48 h after an initial 24-h in vitro rumen fermentation of orchardgrass (Dactylis glomerata L.) hay. Short-term shifts in bovine rumen community structure (bacteria, archea, protozoa, and fungi) resulting from each separate forage and feedstuff addition were measured using unique operational taxonomic units (OTU)s contained in terminal-restriction fragment length polymorphism (T-RFLP) profiles. The Tukey Vacuum Cleaner Analysis (TVCA) model of bacteria accounted for 66.3% of the treatment variance and OTUs clustered into 3 distinct groupings. The model developed for TVCA analysis of archaea accounted for 76.6% of the treatment variance and OTUs clustered into 2 distinct groupings. The TVCA treatment variance of protozoa accounted for 75.8% and OTUs clustered into 3 distinct groupings. Additions of various forages and feedstuffs did not result in shifts in fungal community structure in the short-term experiment. Our results demonstrate that the rumen ecology using initial T-RFLP profiles in vitro of bacteria, archaea, and protozoa and a TVCA analysis can identify pattern groupings of forages and feedstuffs. Groupings of forages and feedstuffs can help refine supplementation for improved nutritional management of ruminants.

Introduction

A diverse array of forages including herbages (grasses), forbs (broadleaf legumes, weeds, and herbs), and browse is used in ruminant livestock production systems (Turner and Belesky Citation2010). Supplements including energy and protein-rich forages and feedstuffs are used to meet dietary requirements and improve nutrient-use efficiency in ruminants such as cattle (Bos sp), sheep (Ovis aries), and goats (Capra hircus). Energy supplements include corn (Zea mays L.) and various cereal grains. Protein supplements include high crude protein (CP) legumes, such as alfalfa (Medicago sativa L.), and high protein feedstuffs and by-product feeds, such as whole cottonseed (Gossypium hirsutum L.). Some forages contain secondary plant metabolites that can influence grazing behavior, selectivity, feed intake (Acamovic and Brooker Citation2005), and rumen microbial populations (Wallace et al. Citation2002). Many of these metabolites are being re-evaluated to identify opportunities to influence ruminant performance through improved feed intake, digestion, nutrient-use efficiency, and health (Rochfort et al. Citation2008). Improved nutrient-use efficiency in grazing livestock can help to reduce methane emissions, enhance carbon sequestration, and reduce fertilizer inputs in grassland agriculture (Abberton et al. Citation2007).

The rumen of cattle contains a complex mixture of anaerobic bacteria, archea, protozoa, and fungi that responds to short- and long-term changes in diet. High starch feedstuffs can shift the ruminal microbial population from fiber-degrading to a starch-degrading microbial population. Excessive amounts or sudden diet changes with limited rumen microbial population adaptation to high starch diets is detrimental to maintaining ruminant health (i.e. rumen acidosis). Usually when supplementing feedstuffs high in starch, the supplement is generally limited to 0.25 - 0.5% of body weight (BW) to maintain a diverse microbial ecosystem for sustained fiber utilization by the animal and improved forage intake and overall performance (Horn and McCollum Citation1987; Bowman and Sanson Citation1996). Protein supplementation is usually required when forage chemical composition contains < 8% CP. Both supplementation strategies are aimed at maintaining a desirable energy (as total digestible nutrients; TDN) to protein ratio in the diet (TDN:CP ratio ∼7.0; Moore et al. Citation1999) to maximize microbial protein synthesis, especially from bacteria, in the rumen for improved nutrient-use efficiency and ultimately animal performance.

Metagenomics involves the analysis of natural assemblies of organisms based on the DNA signature (Handelsman Citation2004). These techniques can include high throughput sequencing platforms (next generation sequencing; NGS), but there is evidence that terminal-restriction fragment length polymorphism (T-RFLP) data is comparable (de la Fuente et al. Citation2014). Specifically, the T-RFLP fingerprint technique provides a snapshot of rumen microbiota and can be used to compare relative differences between or among treatments (de la Fuente et al. Citation2014; Fang et al. Citation2018). This information can be used to better understand how rumen communities adapt to diet and dietary supplements (Firkins et al. Citation2008).

The rumen fluid from two different beef steers was mixed together and used in this experiment. Our goal was not to evaluate or relate the rumen microbiome of each animal, but to use an initial rumen fluid inoculum in vitro that was representative of an orchardgrass forage-based diet. Our aim was to profile the modifying effects from different supplemental forages or feedstuffs on the resultant rumen community pattern in vitro to help group feedstuffs with similar fermentation patterns. Our specific objective was to use T-RFLP fingerprinting to profile bovine rumen community structure 48 h after addition of a supplemental forage or feedstuff to a 24-h basal cool-season grass hay fermentation in vitro.

Materials and methods

Rumen fluid and sample feedstuffs

All applicable international, national, and/or institutional guidelines for the care and use of animals were followed. All animal care and use procedures were reviewed and approved by the Institutional Animal Care and Use Committee, Appalachian Farming Systems Research Center, Beaver, WV, USA. Rumen fluid donor steers were cared for in accordance with the standards of the Guide for the Care and Use of Agricultural Animals in Agricultural Research and Teaching (FASS Citation2010).

Rumen fluid was collected and combined from two ruminally cannulated beef steers and transported to the lab in closed containers. A grab sample of particulate material was also included with the raw rumen fluid (initial pH∼6.8) to account for the particulate-adhering microbes. Rumen fluid, McDougall’s (Citation1948) buffer solution and in vitro rumen fermentation incubations were maintained under anaerobic conditions at 39°C throughout the experiment. Carbon dioxide (CO2) was bubbled through the rumen fluid to minimize oxygen and to agitate the rumen fluid containing particulate material as to help release microbes adhering to fiber in order to create a more heterogenous rumen fluid inoculum. The rumen in vitro dry matter disappearance (IVDMD) of forage and feedstuffs was determined by using the Tilley and Terry (Citation1963) procedure as modified by Moore (Citation1970). All tubes contained 0.6 g (dry weight basis) orchardgrass (Dactylis glomerata L.; OGR) hay and were inoculated with 50 ml of 2:3 mixture of rumen fluid to McDougall’s buffer on d 1, fitted with a stopper containing a one-way valve to vent excess gas, and incubated in a water bath maintained at 39°C. All tubes were incubated for 24 h to establish the initial baseline fermentation. This helped to create an OGR baseline fermentation with an established short-term rumen ecological community. A BLANK was included which contained rumen fluid plus buffer solution (no substrate).

In order to model the offering of supplemental feedstuffs to ruminants already being fed OGR hay, supplemental feedstuffs (dried plant materials and energy feeds; 0.40 g on a dry weight basis) listed in were added to tubes on h 24 to maintain a 1g-total amount (dry weight basis) of substrate in each tube at a ratio of 60:40 (OGR: supplemental feedstuff ratio). Tubes were incubated for an additional 48 h and were swirled while maintaining anaerobicity 3 times per d. Each treatment (supplemental forages, feedstuffs, and blank) was represented by four parallel replicate tubes (each replicated set contained all treatments plus a blank).

Table 1. Dried plant samples, hays, and feedstuffs used in experiment.

After the supplemental forage or feedstuff was added and the additional 48 h of incubation completed, each 100-ml tube was removed from the water bath, thoroughly mixed via vortex for 20 sec, contents filtered through 2-layers of cheesecloth in a funnel, and the liquid captured in a 50-ml screw-top centrifuge tube. Tubes were centrifuged using a Jouan KR22i centrifuge with AK500-11 fixed angle rotor with inserts for 50-ml conical, screw-cap centrifuge tubes spun at 9400 × g for 10 min at 5°C to isolate the microbial mass. The vortexing of tubes prior to filtering helped to dislodge any fiber adhering microbes for a more homogenous sample.

Following centrifugation, supernatant was decanted into a 20-ml vial for volatile fatty acid (VFA) analysis; the remainder of the supernatant was decanted into a 60-ml Nalgene bottle maintained on ice for determination of pH using a hydrogen ion electrode (Fisher Accumet Basic-AB15; Thermo Fisher Scientific, Hampton, NH, USA) with calibrated probe and recorded within 1 h after collection.

All vials and bottles were placed on ice until moved and stored in the cooler (3o C) or freezer (−20°C). Centrifuge tubes containing microbial pellets were frozen at −20°C until genomic DNA (gDNA) was isolated.

Composition of forages and feedstuffs

Forages and feedstuffs used in the experiment were ground to pass a 1-mm screen in a Wiley mill or a 0.5-mm screen in a cyclone mill. The 1-mm ground subsamples of feedstuffs were analyzed for dry matter (DM) and ash (AOAC Citation1990); neutral detergent fiber (NDF; procedure included amylase; final values contained ash residues) and acid detergent fiber (ADF; final values contained ash residues) (Goering and Van Soest Citation1970; Van Soest et al. Citation1991) using a Fiber Analyzer (Ankom Technology Corp., Fairport, NY, USA); and IVOMD (Tilley and Terry Citation1963; Moore Citation1970). A set of 0.5-mm ground subsamples were analyzed for total N using an elemental analyzer (Flash EA, 1112 Series, Thermo Electron Corp., Rodano, MI, Italy) and CP calculated as total N% × 6.25. Total non-structural carbohydrates (TNC) were determined by an automated hydrolysis method (Denison et al. Citation1990).

The supernatant samples were thawed, a one ml-subsample was centrifuged at 10,000 × g for 10 min, then the supernatant was removed and mixed with metaphosphoric acid, cooled on ice, and centrifuged at 10,000 × g for 10 min (Goetsch and Galyean Citation1993). The VFA concentrations were then quantified via gas liquid chromatography (Hewlett-Packard GC 6890, Hewlett-Packard Co., Wilmington, DE, USA) using a Supelco glass column (Supelco/Sigma-Aldrich Inc., Bellefonte, PA, USA) with packing material as outlined in the 1988 Supelco Bulletin 856B for rumen fluid VFA.

gDNA extraction

Samples were thawed and mixed thoroughly after which gDNA was extracted from a 250-mg subsample using a PowerSoil DNA isolation kit (MoBio Laboratories Inc., Carlsbad, CA, USA) (Johnson et al. Citation2009) following the manufacturer’s instructions with the exception that the final volume eluted was 60 μL. The gDNA concentration in all samples was determined using a CARY 100 UV/VIS spectrophotometer (Mulgrave, Australia) after which samples were stored at −20°C until amplification via polymerase chain reaction (PCR).

PCR amplification for T-RFLP analysis

Genomic DNA from bacteria, archaea, protozoa, and fungi were each amplified separately using nested primer sets. A nested PCR approach is necessary when copy numbers of the gDNA of interest are low or in regions difficult to amplify due to non-specific binding of primers to amplicons requiring secondary PCR amplification to obtain sufficient amplicons for analysis. The secondary primer set is designed to target a smaller region within the amplicon obtained from the initial PCR reaction. All primer pairs were previously determined to provide coverage of organisms present in rumen fluid (see individual primer references for details). In this experiment, the second amplification was performed using 1 μL of the first PCR product as template. The forward primer for the second PCR reaction was labeled on the 5′ end with HEX (hexachloro-fluorescein) and FAM (5-carboxy-fluorescein) fluorescent markers [1.5 μL (10 μM)] for terminal fragment tracking. The PCR reactions containing gDNA and primary PCR amplicons were amplified using a GeneAmp9600 Thermocycler (Perkin Elmer, Norwalk, CT, USA). The thermocycler conditions used for amplification were the same irrespective of the template organism or PCR master mix. Conditions included an initial 2-min denaturing step at 94° C, followed by 40 cycles of a 92° C-denaturing step for 45 sec, a 55° C-annealing step for 1 min, a 72° C-extension step for 2 min, and a final 10-min extension step at 72° C.

Master mix for bacteria

The 16S ribosomal DNA (rDNA) was amplified from the isolated gDNA samples using a set of universal bacterial-primers that target 16S rDNA primers, 8F (5′-AGA GTT TGA TCM TGG CTC AG-3′) and 1492R (5′-GGT TAC CTT GTT ACG ACT T-3′) (Johnson et al. Citation2009). The gDNA sample was amplified in a 50-μL PCR reaction that combined 50–200 ng of gDNA, 25 μL of AccuPrime SuperMixII (Invitrogen, Carlsbad, CA, USA), 1.5 μL (10 μM) of reverse primer, and 1.5 μL (10 μM) of forward primer, and PCR grade water sufficient to bring the final volume to 50 μL. Bacterial re-amplification employed nested primer sets 530F (HEX 5′-GTC CCA GCM GCC GCG G-3′) (Winker and Woese C Citation1991) and 1392R (FAM 5′-ACG GGC GGT GTG TRC-3′) (Scully et al. Citation2005).

Master mix for archea

Initial 16S rDNA amplicons were obtained from gDNA samples using the archaeal-specific primers Ar109F (5′-ACK GCT CAG TAA CAC GT-3′) (Grosskopf et al. Citation1998) and AR1383R (5′-CGG TGT GTG CAA GGA GCA-3′) (Shlimon et al. Citation2004). The second reaction used to amplify the amplicon from the above reaction contained primers, Ar349F (HEX 5′-GYG CAS CAG KCG MGA AW-3′) (Hristov et al. Citation2012) and Ar912F (5′-CTC CCC CGC CAA TTC CTT TA-3′) (Lueders and Friedrich Citation2000). The forward primer was labeled on the 5′ end with a HEX fluorescent marker for terminal fragment tracking.

Master mix for protozoa

Procedures as outlined above for bacterial amplification were used. The initial 16S rDNA amplicon was amplified from the isolated gDNA samples using the protozoal-specific primer set Euk82F (5′-AAA CTG CGA ATG GCT C-3′) and MedlinB (5′-TGA TCC TTC TGC AGC AGG TTC ACC TAC-3′) (Skillman et al. Citation2006). The second amplicon was obtained using the forward primer Euk342F (HEX 5′-CTT TCG ATG GTA GTG TAT TGG ACT AC-3′) (Egan Citation2005) for terminal fragment tracking and reverse primer MedlinB (5′-TGA TCC TTC TGC AGC AGG TTC ACC TAC-3′).

Master mix for fungi

Initial 18S rDNA from fungi were amplified from isolated gDNA samples using primers NS1F (5′-GTA GTC ATA TGC TTG TCT C-3′) (Fliegerova et al. Citation2006) and SR6R (5′-TGT TAC GAC TTT TAC TT-3′) (Vilgalys and Hester Citation1990). The second amplification used a nested set of primers NS19F (HEX 5′-CCG GAG AAG GAG CCT GAG AAA C-3′) and NS6R (5′-GCA TCA CAG ACC TGT TAT TGC CTC-3′) (White et al. Citation1990; James et al. Citation2000) the forward primer was labeled on the 5′ end with a HEX fluorescent marker for terminal fragment tracking and used.

Enzymatic digestion

All PCR products were cleaned using a Qiagen PCR Purification Kit (Qiagen, Valencia, CA, USA). The concentrations and purity of each individual PCR product were verified using A260/A280 ratios measured on a spectrophotometer. Each tube contained 3 μg of re-amplified PCR reaction, 1.5 μL of MspI restriction enzyme, cuts at recognition site C^CGG (Invitrogen, Carlsbad, CA), 5 μL of buffer, 1 μL BSA, and enough PCR-grade water to bring final volume to 50 μL. The restriction enzyme was selected to optimize genus-specific terminal restriction fragments (T-RF) associated with phylogenetically informative species. Digestion reactions were incubated overnight in a 37° C water bath and heat inactivated by immersing in a 60° C water bath for 10 min. Digested amplicons were desalted using AmiconUltra 30 kDA SpinFilters (Millipore, Billerica, MA, USA) and stored at −80° C until gene analysis.

Gene analyses and indices

Digested and desalted fungal, archaeal, bacterial, and protozoal samples were shipped to the Genomic Technology Support Facility at Michigan State University (East Lansing, MI, USA) for analysis on an Applied Biosystems Prism 3130 Genetic Analyzer (Applied Biosystems, Foster City, CA, USA). The fragment data files were analyzed and filtered using GeneMarker V1.6 (SoftGeneTics LLC, State College, PA, USA) so only information associated with the tagged PCR products remained. This technique permits an automated quantification of the fluorescence signal intensities of the individual terminal restriction fragments in a given T-RFLP profile (i.e. community fingerprint). The resulting fragment patterns were exported to a MicroSoft Excel-based software package T-Align (Smith et al. Citation2005) for further determination of gene indices.

Each T-RFLP profile community was compared using T-Align. A unique terminal restriction fragments (T-RF) in a given profile that varied by 0.5 bp or less among all the T-RFLP community profiles was considered the same fragment. A file was generated that showed the presence or absence of a T-RF, its average size, and the percentage fluorescence intensity of total fluorescence in each profile. Estimates S (Colwell Citation2019), a MicroSoft Excel software add-in program, was employed to calculate diversity indices for the T-RFLP profiles using output files imported from T-Align. These indices include the Shannon-Weaver index (H′), Simpson diversity (reported as 1-D), the reciprocal Simpson (1/D), and evenness (E). The diversity equations used have been defined and described previously (Krause et al. Citation1995) and represent diversity measures based on fragment numbers in each treatment.

Statistical analyses

The VFA and diversity indices (H′, 1-D, 1/D, and E) data were analyzed according to a one-way analysis of variance using ANOVA procedure (SAS Institute, Cary, NC, USA). Sources of variations in the model included supplemental feedstuff and replicate (n = 4). Significance was reported at P < 0.05, unless otherwise stated.

Statistical analyses of T-RFLP profiles

Statistical analyses were performed with the GLM model of the SAS Statistical Package Version 9.1 (SAS Institute, Cary, NC, USA). Supplemental forages and feedstuffs were fixed effects and T-RFLP profiles random. Data from T-RFLP were transformed ((X + 0.5)0.05) before analysis to satisfy the assumption of normality and equality of variances. Analysis of variance tests were performed on univariate data of T-RFLP profiles for each individual community (bacterial, archaeal, protozoal, and fungal), and mean comparisons were performed using Fisher’s protected least significant difference (LSD;α = 0.10). Interaction between supplemental feedstuffs and microbial populations were carried out on the DNA similarity search peak height data matrices using a modified Additive Main Effects and Multiplicative Interaction (AMMI) using a model described by Zobel and Wallace (Citation1995). The model was modified to use the two-factor analysis of variance model proposed by Wright (Citation1971) supplemented by Singular Value Decomposition of the residual (modified Principal Component Analysis). Non-significant terms were eliminated from the model and the analysis was re-run to produce the final tables (variance components were ordered by size) of variance; largest first; Tukey’s vacuum cleaner analysis (TVCA; Tukey Citation1962a, Citation1962b; Lee and Johnson Citation2006).

Results

Feedstuff chemical composition

Composition of supplemental forages and feedstuffs used in this experiment is shown in . Generalized trends in the chemical composition of these forages and feedstuffs were: (1) the total N (CP) was greatest for ALF, ART, and CCL relative to the other forages and feedstuffs; (2) the NDF was generally highest for OGR, RZP, and SLZ; (3) the ADF was generally least for ART and CCL when compared to the other feedstuffs; (4) the soluble carbohydrates were greatest in CORN, BFT, and PPN relative to other feedstuffs; and (5) the overall in vitro organic matter disappearance (IVOMD) of pure feedstuffs was generally greatest for CORN, CCL, BFT, and CHIC in comparison to other feedstuffs.

Table 2. Chemical composition (dry matter basis) of amendmentsa that include dried plant samples, hays, and feedstuffs used in the experiment.

Diversity indices

There were no differences (P > 0.01) in the gene indices for bacteria, archaea, or fungal (data not shown) in the short-term rumen ecological community established by the OGR baseline. Gene index differences (P < 0.01) for protozoa are presented in . Mean richness and percent richness indices of CORN, ORG, and BLANK were similar; all were higher (P < 0.001) when compared to the other treatments. The BLANK represented fresh rumen fluid with no additions of OGR or other supplements used in the experiment. Shannon’s index mean for ART, BFT, CCL, CFEAST, CHIC, CPUNA, PNT, RZP, SLZ, and WCS were similar; all were higher (P < 0.001) when compared to ALF, BLANK, CORN, ORG, and RCL. Simpson’s index means for ART, BFT, CCL, CFEAST, CHIC, CPUNA, and RCL were similar; all were higher (P < 0.001) compared to the remaining feedstuffs.

Table 3. gDNA indicies for Protozoa. Means ± 1 SEMg.

Interaction analysis

The data analysis is based on the variation of T-RF peak heights in each forage and feedstuff treatment relative to the 1-gm ORG baseline with the established short-term rumen community. Two assumptions were made in reporting the results: (1) the supplemental feedstuffs are independent and the T-RF are dependent; (2) all results are relative to the OGR baseline. Specifically, data are modeled as the variance around the T-RF values obtained from the ORG treatments.

Tukey’s vacuum cleaner analysis

Peak heights of each set of community profiles (bacteria, archaea, protozoa, and fungi) were analyzed via TVCA separately and are presented in the order of the AMMI analyses (, , and ) followed by tables of mean interactions for forage and feedstuff treatments (, , and Supplemental Table S3) and mean interactions for individual T-RF (Supplemental Tables S1, S2, and S4). Culman et al. (Citation2008) have successfully applied AMMI analysis to T-RFLP data. The clustering of the T-RFLP profiles resulting from forage and feedstuff applications within each community (bacteria, archaea, protozoa, and fungi) is presented in .

Table 4. TVCAa for Bacteria. Row (effects) are in Sums of Squares order (except the error term). The percentage of Tmtb variance due to model effect is shown in the last column. Grand Mean = 0.81579.

Table 5. Feedstuff mean and interaction for bacteria.

Table 6. TVCAa for Archea. This is a typical table for AMMIb analysis. Row (effects) are in Sums of Squares order (except the error term). The percentage of treatment variance due to model effect is shown in the last column. Treatment Grand Mean = 0.87596.

Table 7. Feedstuff mean and interaction for archaea.

Table 8. TVCAa for Protozoa. Row (effects) are in Sums of Squares order (except the error term). The percentage of treatment variance due to model effect is shown in the last column. Treatment Grand Mean = 0.80530.

Table 9. TVCAa for Fungi. Results obtained by removing non-significant factors from the analysis model and rerunning the TVCA analysis. Grand Mean = 0.83083.

Table 10. AMMIa model analysis of feedstuffb grouping patterns (action with T-RFc) for an individual rumen ecological group (bacteria, archaea, protozoa, and fungi).

Bacteria

The TVCA () includes a row of feedstuff regressions that are associated with T-RF means. Removal of PNT treatment from the data set eliminates the regression significance, demonstrating that the positive slope of the PNT regression is responsible for this significant interaction. Specifically, the DNA similarity search peak value of all T-RF is increased with PNT feedstuff, and T-RF with higher means are increased to a greater degree by addition of PNT. When CPUNA and ORG are removed from the data set, the regression becomes non-significant most likely due to the negative response slopes associated with these treatments (). The interaction of these two feedstuffs was opposite to that of the PNT treatment. Specifically, a decrease in peak height with increasing T-RF mean. The ART, BFT, SLZ and WCS treatments had the highest peak heights of the forage and feedstuffs (). In the bacteria data, T-RF 2, 4, 5, and 24 have the highest average peak heights, while T-RF 29 has the lowest (Table S1).

Interaction axis significance is predicated by PNT feedstuff addition, and T-RF 13. The removal of either of these from the dataset removes the significance of the interaction axis, thus indicating that they have a broad interaction with the opposing factor elements (positive with some, negative with others, and slight with the rest). If CHIC and T-RF14 or BFT and T-RF 21 are removed, the interaction axis also becomes non-significant, indicating specific interactions (i.e. CHIC specifically suppresses T-RF 14 and BFT specifically suppresses T-RF 21).

In the full model, there were three distinct pattern groupings (1) BFT, PNT, RCL, CFEAST; (2) CHIC, RZP, WCS, ART; and (3) CPUNA, ORG. The remaining feedstuffs slightly mimic one of these three patterns, but at very low variance (Weak; ). The developed TVCA model accounted for 66.3% () of the treatment variance and demonstrated that BFT and PNT dramatically increase the peak height of T-RF 2, 5, 12, 14, and 24, while PNT suppressed the peak heights of T-RF 6-11, 16-18, 23, 26, 28, and 29. Feedstuffs CHIC and RZP suppressed peak heights of T-RF 1, 3, 12, 14-15, 22, 27, and 30–32 whereas CHIC increased the peak heights of T-RF 13 and 21. In contrast, addition of CHIC and RZP reduced the peak heights of T-RF 12, 13, and 14 and additions of BFT and PNT has the opposite effect.

Archaea

The TVCA demonstrated a typical AMMI analysis pattern (i.e. significant additive variation within the two sets of factors and one significant interaction axis). The feedstuff ORG decreased T-RF peak height of Archaea (lowest average additive variance T-RF peak height by 2 LSDs; ). The dominant factor feature in the additive feedstuff variance was the low value of T-RF peak heights in the ORG treatment (), and higher values for T-RF 37, 38, and 67 (Table S2). The T-RF peak heights 37, 38 and 67 are high in additive variance (Table S2). The interaction component was characterized by strong opposite interactions of the feedstuffs CORN and ART; and T-RF 4, any of which when eliminated from the data set, increased the probability above 0.05. This indicates that these factor elements have a broad interaction with the elements of the opposing factor (positive with some, negative with others, and slight with the rest).

The signs of the Interaction Scores (IAS) were used to determine the nature of the interaction [e.g. feedstuff CORN interacted negatively with the Archaea represented by T-RF 4 (−0.45812 * 0.38769 = −0.17761)]. Thus, the average DNA similarity search peak for T-RF 4 when CORN was used [0.87618 + (−0.00618) + 0.10832 = 0.97832] decreased by 0.17761 (0.97832–0.17761 = 0.80071) to a value that is less than the grand mean of 0.87618. In contrast, the average DNA similarity search peak height for T-RF 4 increased after addition of ART. The interaction of feedstuff CORN with the average DNA similarity search of T-RF peak heights revealed an inhibitory effect when the IAS was positive and a stimulatory effect when the IAS was negative, while feedstuff ART addition resulted in the opposite response.

By using paired elimination of characters from the data set, it was determined that the removal of RCL and T-RF 1; RZP and T-RF 1; and CHIC and T-RF 1, each eliminated the significant interaction axis. These pairs represented specific interactions (in this case negative interactions), where ART, CORN and T-RF 4 generated broad interactions.

The TVCA procedure was designed to remove noise from data sets. The simplified reduced full model accounted for 76.6% of the treatment variance in archaea T-RFLP profiles, and it was found that there were distinct patterns in the model (). Pattern 1 represented the response of T-RF to treatments ART, CCL, CFEAST, and to a lesser extent, to BFT and PNT. Pattern 2 represented the response to T-RF to treatments CORN, RZP, and to a lesser extent RCL and CHIC additions. The third pattern represented the response of T-RF to treatments ORG, CPUNA, WCS, SLZ, and ALF, all of which affected the T-RF equally (Weak; ). Because the last four treatment additions of the third pattern gave similar peak height values (), it was assumed that ORG generally suppressed archaeal development. The strong interaction variance of ART addition raised the model variances of T-RF 1, 4, 49, 50, 53, and 69 well above those of the other feedstuffs, while strongly suppressing T-RF 12, 20, and 22. The CORN addition had exactly the opposite response pattern. Variance in T-RF patterns after addition of BFT, RCL, and ALF mimicked to a lesser effect that of the ART treatment.

Protozoa

The model derived from TVCA analysis accounted for 75.8% of the treatment variance in T-RF peak heights () and described 3 distinct patterns: ORG, CORN, and CCL (pattern 1); RCL, ALF, WCS, CHIC (pattern 2); PNT, ART, and CPUN (pattern 3); while pattern 4 consisted of BFT, RZP, CFEA, and SLZ, but represented low or intermediate T-RF values with no distinctive peaks (). The CORN and ORG feedstuffs had the greatest effect on increased T-RF peak heights (Table S3) and the mean peak heights of T-RF 18 was greater than that of any other T-RF (Table S4). In the protozoa data set, there was a significant T-RF regression associated with the feedstuff means. The average peak height values of T-RF 7 and 8 increased with feedstuff additions relative to the remaining T-RF while peak height values of T-RF 14, 15, 16, and 17 decreased with feedstuff additions (Table S4). If WCS, RCL, or T-RF 15 were eliminated from the data set, then the significant interaction axis was removed. Therefore, these factor elements had a broad interaction with the elements of the opposing factor (positive with some, negative with others, and slight with the rest). Conversely, if ALF and T-RF 14 were removed, then significant interaction axis was also removed; this effect represented a specific interaction where ALF selectively suppressed the peak height of T-RF 14.

Patterns are presented in . In pattern 1, both ORG and CORN additions increased T-RF peak heights with the exception of T-RF 14, 15, and 17 (Table S4). Both feedstuffs suppressed T-RF 17, but only ORG suppressed T-RF 15. In the non-distinct group, BFT suppressed the peak height of T-RF 15 while the RCL group (pattern 2) stimulated it. The RCL group (pattern 2) suppressed peak heights of T-RF 2, 3, 10 and 14, but stimulated T-RF 5 and 11 to nearly the same degree as ORG and CORN additions (pattern 1). The BFT addition (pattern 4; Weak) generally suppressed T-RF peak heights except for T-RF 1, 14, 16, and 18. The PNT group (pattern 3) followed a similar less distinct pattern but did not suppress T-RF 15 peak height. The CHIC (pattern 1) and CCL groups (pattern 2) were intermediate and were distinct from one another. The CCL (pattern 2) mirrored ORG pattern 1 while the CHIC treatment mirrored the RCL group ().

Fungi

There were no significant mean interactions with forage and feedstuff treatments or variation among T-RF within the fungal community. Therefore, only the AMMI analysis () is presented. The DSS revealed no interactions. All pattern groupings for the forages and feedstuffs were Weak ().

pH and VFA profiles

The pH and VFA concentrations are presented in . The pH at the end of the 48-h incubation ranged from 6.4–7. The pH was lowest for CORN compared to the other feedstuffs. Total VFA concentrations were lowest for BLANK compared to all feedstuffs. The OGR, WCS, ALF, and SLZ tended to have lower total VFA in comparison to the majority of the feedstuffs.

Table 11. Volatile fatty acid concentrations from in vitro rumen fermentation of feedstuffs and forages used in the experiment. Data are Mean ± 1 SEMh.

Acetate concentration in the fresh rumen fluid BLANK was higher (P < 0.001) than ALF, RZP, BFT, CFEAST, PNT, CPUNA, CCL, WCS, OGR, and CORN; CORN, OGR and WCS were similar. The acetate concentration in BLANK was similar to ART, CHIC, SLZ, and RCL. Acetate was similar for ART, CHIC, SLZ, RCL, ALF, RZP, BFT, CFEAST, PNT, CPUNA and CCL.

Propionate concentrations were higher (P < 0.001) for WCS and OGR when compared to all treatments except CORN, and CCL. The propionate produced from supplemental CORN, CCL, CFEAST, CPUNA, PNT, RZP, RCL and BFT were similar, but higher (P < 0.001) than ART and BLANK. Propionate produced from added BFT, ALF, SLZ, and CHIC were similar, but higher (P < 0.001) than ART and BLANK.

Butyrate concentration was highest (P < 0.001) for CORN compared to the other treatments. The OGR was higher (P < 0.001) than WCS, RZP, SLZ, CCL, CFEAST, ALF, CHIC and RCL. The BLANK was similar to OGR, BFT, PNT, ART, CPUNA, WCS, RZP, SLZ, CCL, CFEAST, ALF, CHIC, and RCL.

Iso-butyrate was highest (P < 0.001) for BLANK compared to all other treatments. The WCS was similar to ALF, OGR, SLZ, and RZP, but higher (P < 0.001) than the remaining treatments while ALF was similar to OGR, SLZ, RZP and CHIC, but higher (P < 0.001) than the remaining treatments. The OGR, was similar to SLZ, RZP, CHIC, ART, RCL, CCL, CPUNA, CORN, CFEAST, BFT and PNT.

Valerate in ART was similar to ALF and WCS, but higher (P < 0.001) than the other treatments. ALF was similar to WCS and CCL and higher (P < 0.001) than the remaining treatments. The CCL, CORN, and BFT were similar.

Iso-valerate followed the trend of BLANK > ALF, WCS > remaining treatments.

The acetate:propionate (A/P) ratio of the BLANK was greater (P < 0.001) compared to all the other treatments. Feedstuff CHIC and ART had similar ratios. The A/P ratio in CHIC, SLZ, ALF, RCL, BFT, RZP, PNT were similar, but greater (P < 0.001) than WCS, CORN, and OGR which were similar.

The acetate:butyrate (A/B) ratio for RCL was similar to CHIC, SLZ, ALF, CFEAST, CCL, RZP, and ART, but greater (P < 0.001) than the remaining treatments. The CORN had the lowest (P < 0.001) A/B ratio compared to the other treatments while BFT, OGR, and BLANK were similar to PNT, but lower than all treatments except CORN.

Discussion

Potential effects of forages and feedstuffs on rumen diversity

Forages and supplemental feedstuffs used in the experiment represent typical conserved forages and feedstuff supplements to cattle diets or forages grazed by cattle, except for Artemisia annua, and represent variations in protein, soluble carbohydrates, fiber, and secondary plant metabolites.

Orchardgrass (OGR) is a cool-season perennial grass (Christie and McElroy Citation1995) while corn grain (CORN) is a starchy feed grain (Theurer Citation1986). Whole cottonseed (WCS) is a by-product feedstuff of the cotton ginning industry often used as a high protein feed supplement for ruminants and contains condensed tannins (Terrill et al. Citation1992; Feng et al. Citation1993). Alfalfa (ALF), red clover (RCL), crimson clover (CCL), peanut hay (PPN), rhizome peanut hay (RZP), birdsfoot trefoil (BFT) and sericea lespedeza hay (SLZ) are all classified as high protein legume forages (McGraw and Nelson Citation2003; Sollenberger and Collins Citation2003). Chicory (CHIC), chicory cv. Puna (CPUNA), and chicory cv. Forage Feast (CFEAST) are classified as forbs and are generally used for grazing; each cultivar differs in the amount of secondary plant compounds (i.e. sesquiterpene lactones, Foster et al. Citation2011; and condensed tannins, Terrill et al., 1992). Artemesia (ART) is a forb containing secondary plant compounds (i.e. sesquiterpene lactones; Jessing et al. Citation2014).

Stimulation or suppression of T-RF or peak heights in our study is most likely a result of specific nutrients or secondary metabolites contained in the feedstuffs affecting the rumen ecology. The secondary metabolites were not quantified in the specific feedstuffs or hays used in the present study, but the discussion included here is relevant, published research to support overall trends of these metabolites.

In our study, treatments receiving addition of CORN and OGR had greater diversity of protozoa compared to the other treatments. Addition of rapidly fermentable carbohydrates (i.e. starch in CORN) likely stimulated protozoa; the OGR treatment represented the fresh microbial and protozoan populations in the rumen fluid used in the current experiment. Kocherginskaya et al. (Citation2001) reported greater microbial diversity and richness in grain-based diets compared to hay diets. In addition, results of the present study represent rapid (48 h) utilization of added nutrients by microorganisms and protozoa, but do not represent long-term adaptation of overall rumen populations in vivo. Thus, only initial response of the rumen to changes are represented in vitro. Because of the synergism and antagonism among the rumen microbes and protozoa, the ability to quantify and identify the specific relationships/interactions among microbial groups is difficult (Kamra Citation2005). Short-term in vitro of grass-legume mixture have been used to evaluate rumen fermentation parameters (Tilley and Terry Citation1963; Niderkorn et al. Citation2011) Overall livestock fed high-grain (starch) diets have lower bacterial diversity, especially fibrolytic microbes (Belanche et al. Citation2012). Mainly amylolytic bacteria (and protozoa and fungi to a lesser degree) degrade starch in the rumen (Huntington Citation1997). The community species richness (bacteria) in the rumen was higher when alfalfa hay compared to triticale (Triticum aestivum) straw was fed to cows (Kong et al. Citation2010). These authors suggest that the greater density and complexity of overall nutrients available in alfalfa vs. triticale was responsible for the higher rumen bacterial species richness. Thus, overall chemical composition differences among feedstuffs is an important factor impacting rumen microbial ecology profiles. Zhu et al. (Citation2018) reported that the relative bacterial abundance of Ruminococcus, Butyrivibrio, Clostridium, Coprococcus, and Pseudobutyrivibrio were not impacted when dairy cattle were transitioned from low-grain (pre-partum) to high-grain (post-partum) diets. These researchers also reported that there were distinct changes in the relative rumen abundance of Prevotella, Ruminococcaceae, and Succinibibrionaceae as well as in profiles of VFA.

Bacteria

Primary fibrolytic bacteria in the rumen include Fibrobacter succinogenes, Ruminococcus albus and R. flavefaciens while F. succinogenes and R. flavefaciens are associated with the particulate fraction in the rumen (Firkins and Yu Citation2006). Fang et al. (Citation2018) reported that F. succinogenes concentration and overall NDF and ADF concentrations were higher in perennial ryegrass-white clover fermentation cultures compared to herbal plintain and fodder beet forage rumen fermentation cultures using mixed rumen fluids from different cattle and PCR techniques. In a hay feeding study with cattle, total counts of ruminal bacteria were greatest when fed alfalfa, intermediate for switchgrass (Panicum virgatum L; warm-season grass), and least for smooth bromegrass (Bromus inermis Leyss.; cool-season grass) (Jung and Varel Citation1988). Bowman and Firkins (Citation1993) reported that red clover had more cellulolytic bacteria associated with forage particles in early rumen incubation in situ compared with a cool-season grass (orchardgrass) and a warm-season grass (gamagrass; Tripsacum dactyloides [L.] L). In our study, the initial rumen fluid along with a sample of the particulate fraction were collected from ruminally cannulated steers. The collected rumen fluid was agitated to weaken attachment of bacteria to fiber with the intent of obtaining a more homogenous inoculum for tubes.

Secondary plant metabolites can suppress or stimulate rumen microbial populations. Growth of the rumen hyper-ammonia producing bacteria Clostridium sticklandii has been shown to be inhibited by red clover (Flythe and Kagan Citation2010). Red clover contains the enzyme polyphenol oxidase which protects plant protein from degradation in the rumen (Lee Citation2014). Min et al. (Citation2002) reported that condensed tannin in birdsfoot trefoil offered to sheep decreased populations of the proteolytic rumen bacteria Clostridium proteoclasticum B316, Eubacterium sp. C12b, Streptococcus bovis B315, and Butyrivibrio fibrisolvens C211a. Sun et al. (Citation2008) reported that Lachnospira multiparus and Fibrobactor succinogenes but not Ruminococcus flavefaciens or Butyrivibrio hungatei degraded fresh forage chicory in vitro; L. multiparus degraded high pectin fiber in chicory faster than F. succinogenes. Chicory (CHIC, CPUNA and CFEAST in the present study) contains sesquiterpene lactones which are reported to be antifungal, antibacterial, and antiprotozoal (Picman Citation1986). Plants containing sesquiterpene lactones have been reported to have anthelmintic (deworming) activity against Haemonchus contortus (a gastrointestinal nematode of sheep and goats) eggs in vitro (Foster et al. Citation2011).

Protozoa

Rumen protozoa are generally more sensitive to dietary changes than bacteria. Protozoa are difficult to isolate from the rumen environment, and are generally found in higher numbers when diets contain more readily fermentable carbohydrates (Yokoyama and Johnson Citation1988). Fang et al. (Citation2018) reported the protozoa abundance in in vitro fermentation cultures using mixed rumen fluids from different cattle and PCR techniques were greater for fodder beet (Beta vulgaris L.) compared to herbal plantain (Plantago lanceolata L.) and perennial ryegrass (Lolium perenne L.)-white clover forages. Fodder beet in that experiment had the highest water soluble cabohydrates compared to the other two forages. In our study, CORN supplementation (starch and a more readily fermentable carbohydrate) resulted in a greater protozoan Richness index. However, protozoa alone are not responsible for degrading the cell wall fraction (Lee et al. Citation2000). Cellulolysis of orchardgrass cell walls by bacteria was reported to be highest during early stages (Lee et al. Citation2000). Using rumen fluid collected from animals offered different forages (cool- vs warm-season grasses) and supplemental grain would likely have influenced our results. Additionally, dietary fat and dietary oil and essential oils have been shown to suppress protozoal numbers (Loor et al. Citation2004; Hristov and Jouany Citation2005) which may be the case with oil-containing. WCS supplementation in our study. Saponins in plants (high in ALF) would not have eliminated all protozoa in the rumen, but probably reduced total protozoal numbers (Lu and Jorgensen Citation1987) and specific protozoal species (Ivan et al. Citation2004). Saponins in alfalfa (Sen et al. Citation1998) is dependent upon cultivar and growth stage (including number of leaves in early growth stages) (Hanson et al. Citation1963).

Archaea

Bacteria in the genus Methanobrevibacter are primarily responsible for methane production in the rumen (Miller Citation1995). Ouwerkerk et al. (Citation2008) suggested that grain feeding reduced rumen methanogen species diversity in Bos indicus cattle. Sheep offered alfalfa pellets had a more diverse Archaea community than Bos indicus cattle and may have been the result of differences in host-animal eating behavior or greater pectin (saponins) in alfalfa that supports Archaeal growth (Ouwerkerk et al. Citation2008). Rumen methanogens were directly inhibited by condensed tannins in Lotus pedunculatus (Tavendale et al. Citation2005). Condensed tannins were shown to indirectly reduce hydrogen production via decreased forage digestibility in the rumen (Tavendale et al. Citation2005). Orchardgrass in the present study significantly reduced peak heights of T-RF of archaea. These results may be due to increases in propionate and butyrate production resulting from competition for hydrogen in the rumen (Abberton et al. Citation2007). Propionate and butyrate concentrations in OGR in the present study were higher in comparison to many of the other feedstuffs used in the study. The aspect of using orchardgrass pasture and hays in livestock production to help reduce methane emissions needs further investigation.

Fungi

Rumen fungi have cellulolytic and hemicellulolytic activity (Bauchop Citation1981). Although the anaerobic fungi make up a very small percentage of the total microbial population they are thought to act as the initial colonizers of plant material and play an important role in plant cell wall weakening due to their high hemicellulase and cellulase activity and tend to be reduced on high concentrate diets. Cellulolysis of orchardgrass cell wall by the bacterial fraction was highest during early incubation stages, but cellulolysis by the fugal fraction was highest during the later stages of incubation (Lee et al. Citation2000). In the current study there was no variation in the fungal levels among the feedstuffs, whereas in contrast, Fang et al. (Citation2018) reported that anaerobic fungi levels were higher in perennial ryegrass-white clover forage samples following in vitro rumen fermentation compared to herbal plantain or fodder beet forages. These authors further related the higher fungi levels to higher fiber concentration in the ryegrass-white clover forages compared to plantain or fodder beet forage samples. Specific fungi effects in the current study were not evident.

pH and volatile fatty acids

The range in pH (6.4–7) was considered conducive to rumen microbial ecology activity (Roger et al. Citation1990). In an in-vitro study, Mouriño et al. (Citation2001) reported that rumen pH below 6.0-6.3 (considered a critical value) impaired fiber breakdown. A low rumen pH reduced primary fibrolytic bacteria (Erdman Citation1988). As VFA are produced by the rumen ecosystem in vivo or in vitro can reduce the overall pH if the VFA are not buffered or removed from the system. In the present study the pH was buffered with McDougall’s buffer which should have been effective for approx. 4 d. Volatile fatty acid concentrations and ratios are the result of a mixed community in the rumen that represents interactions among bacteria, archaea, protozoa, and fungi which creates a unique VFA profile dependent on the overall diet or feedstuff being fermented. The main fermentation products of bacterial spp. include acetate, butyrate, propionate, valerate, and iso-butyrate, and iso-valerate while the main fermentation products of protozoal spp. are acetate and butyrate (Van Soest Citation1983). Cheng et al. (Citation2009) reported that acetate with minimum quantities of lactate and formate were the main end products when rumen-sourced anaerobic fungi and methanogens digested lignocellulose in batch culture.

Forage-based diets typically result in the production of acetate as the predominant VFA while grain-based (CORN or high carbohydrate) diets result in the production of propionate as the predominant VFA in the rumen. High concentrate (starch as in corn) in ruminant diets typically have the lowest A:P ratio (Bannink et al. Citation2008; Dijkstra et al. Citation2012). Zhu et al. (Citation2018) reported that when dairy cows were transitioned from low-grain to high-grain diets, the change in specific bacterial abundance resulted in changes in VFA profiles. Forage-based diets with supplemental fats generally inhibit protozoa and increase rumen propionate concentrations in the rumen (Firkins Citation1996). In the present study, addition of WCS (typically contains high oil; Moore et al. Citation1986) did not increase propionate concentrations over those of ORG or CORN, but all three were higher in propionate relative to the other supplemental feedstuffs. Saponins in plants such as alfalfa have been reported to reduce total short-chain VFA but increase the molar proportions of propionate to decrease the acetate to propionate ratio (Lu and Jorgensen Citation1987). Saponins can target specific rumen bacteria, archea, protozoa, and fungi to impact H ion flow thus impacting VFA and methane production (Patra and Saxena Citation2009). In the present study total VFA and propionate concentrations in ALF were not different from other protein legume feedstuffs suggesting that saponins did not have an impact on VFA. Waghorn and Shelton (Citation1995, Citation1997) reported 1.8% condensed tannin in the dietary DM had minimal effect on rumen fermentation, mainly on minor VFA (isobutyrate, valerate, and isovalerate). In the present study, SLZ (condensed tannin containing; Muir et al. Citation2018), and WCS (oil; Moore et al., Citation1986, and condensed tannin containing; Yu et al., Citation1993) resulted in similar VFA patterns while the BFT (contains condensed tannin; Min and Hart Citation2003) treatment exhibited higher overall VFA concentrations. Condensed tannins in SLZ forage have been reported to have a higher affinity for protein than tannins in BFT forages (Aerts et al. Citation1999). Thus, in our study, condensed tannins in SLZ could bind protein more tightly than condensed tannins in BFT resulting in the protein being protected from rumen fermentation and may have resulted in lower VFA concentrations overall.

Conclusions

The information presented here is an application of the T-RFLP method to a rumen in vitro digestibility technique commonly used to evaluate forage and feedstuff chemical composition for ruminants. Our results demonstrate that the ecological profile resulting from T-RFLP methods and then refined using a simplified TVCA analysis can be used to determine effects of feedstuffs by identifying groupings of T-RF of rumen bacteria, archaea, and protozoa from rumen fermentation in vitro. Nutritionists and producers typically provide supplemental feedstuffs and mixed rations to correct nutrient deficiencies for improved overall feed- and forage-use efficiency and improved animal growth and performance, all aimed at improving economic return. Information in the present study could be used to allot forages and feedstuffs into ecological groups with similar taxonomic patterns. Ecological profiling could potentially be used to create forage mixtures and feedstuff supplements (energy, protein, or a combination) for improved nutritional management and nutrient use in livestock, However, more research is needed on integration of ecological groupings based on a single feedstuff having commonality with two or more of the bacterial, archeal, protozoal, and fungal groups.

Supplemental material

TAAR_1781648_Supplementary_Tables

Download MS Word (20.3 KB)

Acknowledgments

The authors thank Jerry M. Carter, Jeffrey B. Ellison, Kenneth N. Harless, J. Mark Peele (posthumously), Joyce M. Ruckle, and John P. Snuffer for their invaluable laboratory support and efforts. The authors are grateful to Drs. David Burner, Kim Cassida, Jorge Ferreira, Joyce Foster, Will Getz (posthumously), and Tom Terrill for supplying some of the plant samples used in this study. Mention of trade names or commercial products in this publication is solely for the purpose of providing specific information and does not imply recommendation or endorsement by the US Department of Agriculture. The US Department of Agriculture (USDA) prohibits discrimination in all its programs and activities on the basis of race, color, national origin, age, disability, and where applicable, sex, marital status, familial status, parental status, religion, sexual orientation, genetic information, political beliefs, reprisal, or because all or part of an individual’s income is derived from any public assistance program. USDA is an equal opportunity provider and employer.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

Funding for this project was provided from Congressional appropriations to the United States Department of Agriculture, Agricultural Research Service.

References

  • Abberton MT, MacDuff JH, Marshall AH, Humphreys MW. 2007. The genetic improvement of forage grasses and legumes to reduce greenhouse gas emissions. Rome, Italy: Food and Agricultural Organization of the United Nations; pp. 1–15. Available at: https://www.uncclearn.org/sites/default/files/inventory/fao143.pdf.
  • Acamovic T, Brooker JD. 2005. Biochemistry of plant secondary metabolites and their effects on animals. Proc Nutr Soc. 64:403–412.
  • Aerts RJ, McNabb WC, Molan A, Brand A, Peters JS, Barry TN. 1999. Condensed tannins from Lotus corniculatus and Lotus pedunculatus affect the degradation of ribulose 1,5-bisphosphate carboxylase (Rubisco) protein in the rumen differently. J Sci Food Agric. 79:79–85.
  • AOAC. 1990. Association of Official Analytical Chemists. Official methods of analysis, 15th ed. Washington, DC, USA: AOAC. 1298 p.
  • Bannink A, France J, Lopez S, Gerrits WJJ, Kebreab E, Tamminga S, Dijkstra J. 2008. Modelling the implications of feeding strategy on rumen fermentation and functioning of the rumen wall. Anim Feed Sci Technol. 143:3–26.
  • Bauchop T. 1981. The anaerobic fungi in rumen fibre digestion. Agric Environ. 6:339–348.
  • Belanche A, Doreau M, Edwards JE, Moorby JM, Pinloche E, Newbold CJ. 2012. Shifts in the rumen microbiota due to the type of carbohydrate and level of protein ingested by dairy cattle are associated with changes in rumen fermentation. J Nutr. 142:1684–1692.
  • Bowman JGP, Firkins JL. 1993. Effects of forage species and particle size on bacterial cellulolytic activity and colonization in situ. J Anim Sci. 71:1623–1633.
  • Bowman JGP, Sanson DW. 1996. Starch- or fiber-based energy supplements for grazing ruminants. In: Judkins MB and McCollum III FT (Ed.) Proc 3rd Grazing Livestock Nutrition Conf, West Sect Am Soc Anim Sci (suppl 1):118-135 (abstr).
  • Cheng YF, Edwards JE, Allison GG, Zhu W-Y, Theodorou MK. 2009. Diversity and activity of enriched ruminal cultures of anaerobic fungi and methanogens grown together on lignocellulose in consecutive batch culture. Bioresour Technol. 100:4821–4828.
  • Christie BR, McElroy AR. 1995. Orchardgrass. In: Barnes RF, Miller DA, Nelson CJ, editor. Forages, Vol I. An introduction to grassland agriculture. 5th ed. Ames, IA, USA: Iowa State University Press; p. 325–344.
  • Colwell RK. 2019. User’s guide to EstimateS 9.1.0. Statistical estmation of species richness and shared species from samples. [accessed 2020 Apr 9]. http://viceroy.colorado.edu/estimates.
  • Culman SW, Gauch HG, Blackwood CB, Thies JE. 2008. Analysis of T-RFLP data using analysis of variance and ordination method: A comparative study. J Microbiol Methods. 75:55–63.
  • de la Fuente G, Belanche A, Girwood SE, Pinloche E, Wilkinson T, Newbold CJ. 2014. Pros and cons of ion-torrent next generation sequencing versus terminal restriction fragment length polymorphism T-RFLP for studying the rumen bacterial community. PLoS ONE. 9(7):e101435. doi:10.1371/journal.pone.0101435.
  • Denison RF, Fedders JM, Tong CBS. 1990. Amyloglucosidase hydrolysis can over estimate starch concentration of plants. Agron J. 82:869–873.
  • Dijkstra J, Ellis JL, Kebreab E, Strathe AB, Lopez S, France J, Bannink A. 2012. Rumial pH regulation and nutritional consequences of low pH. Anim Feed Sci Technol. 172:22–33.
  • Egan AR. 2005. Experimental designs for rumen microbiology. In: Makkar HP, McSweeney CS, editor. Methods in gut microbial ecology for ruminants. Dordrecht, Netherlands: Springer; p. 3–19.
  • Erdman RA. 1988. Dietary buffering requirements of the lactating dairy cow: A review. J Dairy Sci. 71:3246–3266.
  • Fang H, Al-Marashdeh O, Zhou H, Podolyan A, Hickford JG, Edwards GR, Cameron KC, Cheng L. 2018. Ex-vivo rumen fluid fermentation: changes in microbial populations and fermentation products with different forages. J Appl Anim Res. 46:1272–1279. doi:10.1080/09712119.2018.1495082.
  • FASS. 2010. Guide for care and Use of Agricultural animals in Research and Teaching, 3rd ed. Champaign, IL, USA: Federation of Animal Science Societies.
  • Feng Y, Barry TN, Moughan PJ, Wilson GF. 1993. Condensed tannin and gossypol concentrations in cottonseed and in processes cottonseed meal. J Sci Food Agric. 63:7–15.
  • Fliegerova K, Mrazek J, Voigt K. 2006. Differentiation of anaerobic polycentric fungi by rDNA PCR-RFLP. Folia Microbiol. 51(4):273–277.
  • Firkins JL. 1996. Maximizing microbial protein synthesis in the rumen. J Nutr. 126:1347S–1354S.
  • Firkins JL, Karnati SKR, Yu Z. 2008. Linking rumen function to animal response by application of metagenomics techniques. Austral J Exp Agric. 48:711–721.
  • Firkins JL, Yu Z. 2006. Characterisation and quantification of the microbial populations in the rumen. In: Sejrsen K, Hvelplund T, Nielsen MO, editor. Ruminant physiology, digestion, metabolism and impact of nutrition on gene expression, immunology and stress. Wageningen, The Netherlands: Academic Publishers; p. 19–54.
  • Flythe M, Kagan I. 2010. Antimicrobial effect of red clover (Trifolium pratense) phenolic extract on the rumen hyper ammonia-producing bacterium, Clostridium sticklandii. Curr Microbiol. 61:125–131.
  • Foster JG, Cassida KA, Turner KE. 2011. In vitro analysis of anthelmintic activity of forage chicory (Cichorium intybus L.) sesquiterpene lactones against a predominately Haemonchus contortus egg population. Vet Parasitol. 180:298–306.
  • Goering H, Van Soest PJ. 1970. Forage fiber analysis (apparatus, reagents, procedures, and some applications). Agric. Handbook 379. Washington, DC, USA: USDA-ARS; pp. 1–20.
  • Goetsch AL, Galyean ML. 1993. Influence of feeding frequency on passage of fluid and particle markers in steers fed a concentrate diet. Can J Anim Sci. 63:727–730.
  • Grosskopf R, Janssen PH, Liesack W. 1998. Diversity and structure of the methanogenic community in anoxic rice paddy soil microcosms as examined by cultivation and direct 16S rRNA gene sequence analysis. Appl Environ Microbiol. 64:668–675.
  • Handelsman J. 2004. Metagenomics: application of genomics to uncultured microorganisms. Microbiol Mol Biol Rev. 68:669–685.
  • Hanson CH, Kohler GO, Dudley JW, Sorensen EL, Van Atta GR, Taylor KW, Pedersen MW, Carnahan HL, Wilsie CP, Kehr WR, et al. 1963. Saponin content of alfalfa as related to location, cutting, variety, and other variables. USDA Res Rep ARS. 1963:34–44.
  • Horn GW, McCollum FT. 1987. Energy supplementation of grazing ruminants. In: Judkins M, editor. Proc grazing livestock Nutrition Conf. Jackson, WY: Univ. of Wyoming, Laramie; p. 125–136.
  • Hristov AN, Callaway TR, Lee C, Dowd SE. 2012. Rumen bacterial, archaeal, and fungal diversity of dairy cows in response to ingestion of lauric or myristic acid. J Anim Sci. 90:4449–4457.
  • Hristov AN, Jouany J-P. 2005. Factors affecting the efficiency of nitrogen utilization in the rumen. In: Hristov AN, Pfeffer E, editor. Nitrogen and phosphorus nutrition of cattle and environment. Wallingford, UK: CAB International; p. 117–166.
  • Huntington GB. 1997. Starch utilization by ruminants: from basics to the bunk. J Anim Sci. 75:852–867.
  • Ivan M, Mir PS, Mir Z, Entz T, He ML, McAllister TA. 2004. Effects of dietary sunflower seeds on rumen protozoa and growth of lambs. Br J Nutr. 9:303–310.
  • James TY, Porter D, Leander CA, Vilgalys R, Longcore JE. 2000. Molecular phylogenetics of the Chytridiomycota supports the utility of ultrastructural data in chytrid systematics. Can J Bot. 78:336–350.
  • Jessing KK, Duke SO, Cedergreeen N. 2014. Potential ecological roles of artemisinin produced by artemisia annua L. J Chem Ecol. 40:100–117.
  • Johnson MC, Devine AA, Ellis JC, Grunden AM, Fellner V. 2009. Effects of antibiotics and oil on microbial profiles and fermentation in mixed cultures of ruminal microorganisms. J Dairy Sci. 92:4467–4480.
  • Jung HG, Varel VH. 1988. Influence of forage type on ruminal bacterial populations and subsequent in vitro fiber digestion. J Dairy Sci. 71:1526–1535.
  • Kamra DN. 2005. Rumen microbial ecosystem. Microbiology Section, Centre of Advanced Studies in animal Nutrition, India Veterinary Research Institute, Izatnagar 243 122, India. Curr Sci. 89(1):124–135.
  • Kocherginskaya SA, Aminov RI, White BA. 2001. Analysis of the rumen bacterial diversity under two different diet conditions using denaturing gel electrophoresis, random sequencing, and statistical ecology approaches. Anaerobe. 7:119–134.
  • Kong Y, Teather R, Forster R. 2010. Composition, spatial distribution, and diversity of the bacterial communities in the rumen of cows fed different forages. FEMS Microbiol Ecol. 74(3):612–622.
  • Krause DO, Easte RA, White BA, Mackie RI. 1995. Effect of weaning diet on the ecology of adherent lactobacilli in the gastrointestinal tract of the pig. J Anim Sci. 73:2347–2354.
  • Lee E-J, Johnson DE. 2006. AMMI macros for multiplicative interaction models. Proceedings, SAS Users Group International Conference; March 26-29; San Francisco, CA, USA. Paper 049-31. [Accessed 2018 Aug 21]. http://www2.sas.com/proceedings/sugi31/toc.html.
  • Lee MRF. 2014. Forage polyphenol oxidase and ruminant livestock nutrition. Front Plant Sci. 5:694. doi:10.3389/fpls.2014.00694.
  • Lee S, Ha JK, Cheng K-J. 2000. Relative contributions of bacteria, protozoa, and fungi to in vitro degradation of orchardgrass cell walls and their interaction. Appl Environ Microbiol. 66(9):3807–3813.
  • Lueders T, Friedrich M. 2000. Archaeal population dynamics during sequential reduction processes in rice field soil. Appl Environ Microbiol. 66:2732–2742.
  • Loor JJ, Ueda K, Ferlay A, Chilliard Y, Doreau M. 2004. Biohydrogenation, duodenal flow, and intestinal digestibility of trans fatty acids and conjugated linoleic acids in response to dietary forage:concentrate ratio and linseed oil in dairy cows. J Dairy Sci. 87:2472–2485.
  • Lu CD, Jorgensen NA. 1987. Alfalfa saponins affect site and extent of nutrient digestion in ruminants. J Nutr. 117:919–927.
  • McDougall EI. 1948. Studies on ruminant saliva. 1. The composition and output of sheep’s saliva. Biochem J. 43:99–109.
  • McGraw RL, Nelson CJ. 2003. Legumes for northern areas. In: Barnes RF, Nelson CJ, Collins M, Moore KJ, editor. Forages: An Introduction to grassland Agriculture. Vol 1. 6th Ed. Ames, IA: Iowa State Univ Press; p. 171–190.
  • Miller TL. 1995. Ecology of methane production and hydrogen sinks in the rumen. In: Engelhardt WV, Leonhardt-Mark S, Breves G, Gieseke D, editor. Ruminant physiology: digestion, metabolism, growth and reproduction. Stuttgart, Germany: Ferdinand Enke Verlag; p. 317–331.
  • Min BR, Attwood GT, Reilly K, Sun W, Peters JS, Barry TN, McNabb WC. 2002. Condensed tannins decrease in vivo populations of proteolytic bacteria and affect nitrogen metabolism in the rumen of sheep. Can J Microbiol. 48:911–921.
  • Min BR, Hart SP. 2003. Tannins for suppression of internal parasites. J Anim Sci. 81(E. Suppl):E102–E109.
  • Moore JA, Swingle RS, Hale WH. 1986. Effects of whole cottonseed, cottonseed oil, or animal fat on digestibility of wheat straw diets by steers. J Anim Sci. 63:1267–1273.
  • Moore JE. 1970. Procedures for the two-stage in vitro digestion of forages. In: Harris LE, editor. Nutrition research techniques for domestic and wild animals. Vol. I. Logan, UT, USA: Utah State University; p. 5001.1–5001.3.
  • Moore JE, Brant MH, Kunkle WE, Hopkins DI. 1999. Effects of supplementation on voluntary intake, diet digestibility, and animal performance. J Anim Sci. 77(Suppl 2):122–135.
  • Mouriño F, Akkarawongsa RA, Weimer PJ. 2001. Initial pH as a determinant of cellulose digestion rate by mixed ruminal microorganisms in vitro. J Dairy Sci. 84:848–859.
  • Muir JP, Terrill TH, Mosjidis JA, Luginbuhl J-M, Miller JE, Burke JM, Coleman SW. 2018. Harvest regimen changes sericea lespedeza condensed tannin, fiber, and protein concentrations. Grassl Sci. 64:137–144.
  • Niderkorn V, Baumont R, LeMorvan A, Macheboeuf D. 2011. Occurrence of associative effects between grasses and legumes in binary mixtures on in vitro rumen fermentation characteristics. J Anim Sci. 89:1138–1145.
  • Ouwerkerk D, Turner AF, Klieve AV. 2008. Diversity of methanogens in ruminants in Queensland. Aust J Exp Agric. 48:722–725.
  • Patra AK, Saxena J. 2009. The effect and mode of action of saponins on the microbial populations and fermentation in the rumen and ruminant production. Nutr Rep Rev. 22:204–219.
  • Picman AK. 1986. Biological activities of sesquiterpene lactones. Biochem Syst Ecol. 14:255–281.
  • Rochfort S, Parker AJ, Duhshea FR. 2008. Plant bioactives for ruminant health and productivity. Phytochem. 69:299–322.
  • Roger V, Fonty G, Komisarczuk-Bony S, Gouet P. 1990. Effects of physiochemical factors on the adhesion to cellulose avicel of the rumen bacteria Ruminococcus flavefaciens and Fibrobacter succinogenes subsp. succinogenes. Appl Environ Microbiol. 56:3081–3087.
  • Scully C, Collins G, O'Flaherty V. 2005. Assessment of anaerobic wastewater treatment failure using terminal restriction fragment length polymorphism analysis. J Appl Microbiol. 99(6):1463–1471.
  • Sen S, Makkar HPS, Becker K. 1998. Alfalfa saponins and their implication in animal nutrition. J Agric Food Chem. 46:131–140.
  • Shlimon A, Friedric MW, Niemann H, Ramsing NB, Finster K. 2004. Methanobacterium aarhusense sp. nov., a novel methanogen isolated from marine sediment (Aarhus Bay, Denmark). Int J Syst Evol Microbiol. 54:759–763.
  • Skillman LC, Toovey AF, Williams AJ, Wright A-DG. 2006. Development and validation of a real-time PCR method to quantify rumen protozoa and examination of variability between Entodinium populations in sheep offered a hay-based diet. Appl Environ Microbiol. 72(1):200–206.
  • Smith CJ, Danoilowicz BS, Clear AK, Costello FJ, Wilson B, Meijer WG. 2005. T-Align, a web-based tool for comparison of multiple terminal restriction fragment length polymorphism profiles. FEMS Microbiol Ecol. 54:375–380.
  • Sun XZ, Joblin KN, Andrew IG, Hoskin SO, Harris PJ. 2008. Degradation of forage chicory by ruminal fibrolytic bacteria. J Appl Microbiol. 105:1286–1297.
  • Sollenberger LE, Collins M. 2003. Legumes for northern areas. In: Barnes RF, Nelson CJ, Collins M, Moore KJ, editor. Forages: An Introduction to grassland Agriculture. Vol 1. 6th Ed. Ames, IA: Iowa State Univ Press; p. 191–213.
  • Tavendale MH, Meagher LP, Pacheco D, Walker N, Attwood GT, Sivakumaran S. 2005. Methane production from in vitro rumen incubations with Lotus pedunculatus and Medicago sativa, and effects of extractable condensed tannin fractions on methanogenesis. Anim Feed Sci Technol. 123-124:403–419.
  • Terrill TH, Rowan AM, Douglas GB, Barry TN. 1992. Determination of extractable and bound condensed tannin concentrations in forage plants, protein concentrate meals, and cereal grain. J Sci Food Agric. 58:321–329.
  • Theurer CB. 1986. Grain processing effects on starch utilization by ruminants. J Anim Sci. 63:1649–1662.
  • Tilley JMA, Terry RA. 1963. A two-stage technique for the in vitro digestion of forage crops. J Br Grassl Soc. 18:104–111.
  • Tukey JW. 1962a. The future of data analysis. Ann of Math Stat. 33(1):1–67.
  • Tukey JW. 1962b. Correction notes: Correction to “The future of data analysis”. Ann Math Stat. 33(2):812.
  • Turner KE, Belesky DP. 2010. Terminology revisited: effective communications for the agricultural community. J Extension [On-line]. 48(2): Article No. 2FEA5. http://www.joe.org/joe/2010april/a5.php.
  • Van Soest PJ. 1983. Rumen microbes. In: Van Soest PJ, editor. Nutritional ecology of the ruminant. Corvallis, OR: O&B Books, Inc.; p. 152–177.
  • Van Soest PJ, Robertson JB, Lewis BA. 1991. Methods for dietary fiber, neutral detergent fiber, and nonstarch polysaccharides in relation to animal nutrition. J Dairy Sci. 74:3583–3597.
  • Vilgalys R, Hester M. 1990. Rapid genetic identification and mapping of enzymatically amplified ribosomal DNA from several Cryptococcus species. J Bacteriol. 172:4238–4246.
  • Waghorn GC, Shelton ID. 1995. Effect of condensed tannins in Lotus pedunculatus on the nutritive value of ryegrass (Lolium perenne) fed sheep. J Agric Sci (Camb.). 125:291–297.
  • Waghorn GC, Shelton ID. 1997. Effect of condensed tannins in Lotus corniculatus on the nutritive value of pasture for sheep. J Agric Sci (Camb.). 128:365–372.
  • Wallace RJ, McEwan NR, McIntosh FM, Teferedegne B, Newbold CJ. 2002. Natural products as manipulators of rumen fermentation. Asian-Aust Anim Sci. 15:1458–1468.
  • White TJ, Bruns TD, Lee SB, Taylor JW. 1990. Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In: Innis MA, Gelfand H, Sninsky JS, White T, editor. PCR-Protocols and applications: A laboratory manual. San Diego, CA, USA: Academic Press; p. 315–322.
  • Winker S, Woese C R. 1991. A definition of the domains archaea, cacteria and eucarya in terms of small subunit ribosomal-RNA characteristics. Syst Appl Microbiol. 14(4):305–310.
  • Wright AJ. 1971. The analysis and prediction of some two factor interactions in grass breeding. J Agric Sci Camb. 76:301–306.
  • Yokoyama MT, Johnson KA. 1988. Microbiology of the rumen and intestine. In: Church DC, editor. The ruminant animal: Digestive physiology and nutrition. Englewood Cliffs, NJ, USA: Prentice Hall; p. 125–144.
  • Yu F, Barry TN, Moughan PJ, Wilson GF. 1993. Condensed tannin and gossypol concentrations in cottonseed and in processed cottonseed meal. J Sci Food. 63:7–15.
  • Zhu Z, Kristensen L, Difford GF, Poulsen M, Noel SJ, Al-Soud WA, Sørensen SJ, Lassen J, Løvendahl P, Højberg O. 2018. Changes in rumen bacterial and archaeal communities over the transition period in primiparous Holstein dairy cows. J Dairy Sci. 101:9847–9862.
  • Zobel RW, Wallace DH. 1995. AMMI statistical model and interaction analysis. In: Pessarakli M, editor. Handbook of plant and crop physiology. New York: Marcel Dekker, Inc; p. 849–862.