首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Information available on the role of Mg for growth and cellulose degradation by rumen bacteria is both limited and inconsistent. In this study, the Mg requirements for two strains each of the cellulolytic rumen species Fibrobacter succinogenes (A3c and S85), Ruminococcus albus (7 and 8) and Ruminococcus flavefaciens (B34b and C94) were investigated. Maximum growth, rate of growth and lag time were all measured using a complete factorial design, 2(3)×6; factors were: strains (2), within species (3) and Mg concentrations (6). R. flavefaciens was the only species that did not grow when Mg was singly deleted from the media, and both strains exhibited a linear growth response to increasing Mg concentrations (P<0.001). The requirement for R. flavefaciens B34b was estimated as 0.54 mM; whereas the requirement for R. flavefaciens C94 was >0.82 as there was no plateau in growth. Although not an absolute requirement for growth, strains of the two other species of cellulolytic bacteria all responded to increasing Mg concentrations. For F. succinogenes S85, R. albus 7 and R. albus 8, their requirement estimated from maximum growth was 0.56, 0.52 and 0.51, respectively. A requirement for F. succinogenes A3c could not be calculated because there was no solution for contrasts. Whether R. flavefaciens had a Mg requirement for cellulose degradation was determined in NH3-free cellulose media, using a 2×4 factorial design, 2 strains and 4 treatments. Both strains of R. flavefaciens were found to have an absolute Mg requirement for cellulose degradation. Based on reported concentrations of Mg in the rumen, 1.0 to 10.1 mM, it seems unlikely that an in vivo deficiency of this element would occur.  相似文献   

2.

Aims

The objective was to determine the effect of the isoflavone biochanin A (BCA) on rumen cellulolytic bacteria and consequent fermentative activity.

Methods and Results

When bovine microbial rumen cell suspensions (n = 3) were incubated (24 h, 39°C) with ground hay, cellulolytic bacteria proliferated, short‐chain fatty acids were produced and pH declined. BCA (30 μg ml?1) had no effect on the number of cellulolytic bacteria or pH, but increased acetate, propionate and total SCFA production. Addition of BCA improved total digestibility when cell suspensions (n = 3) were incubated (48 h, 39°C) with ground hay, Avicel, or filter paper. Fibrobacter succinogenes S85, Ruminococcus flavefaciens 8 and Ruminococcus albus 8 were directly inhibited by BCA. Synergistic antimicrobial activity was observed with BCA and heat killed cultures of cellulolytic bacteria, but the effects were species dependent.

Conclusions

These results indicate that BCA improves fibre degradation by influencing cellulolytic bacteria competition and guild composition.

Significance and Impact of the Study

BCA could serve as a feed additive to improve cellulosis when cattle are consuming high‐fibre diets. Future research is needed to evaluate the effect of BCA on fibre degradation and utilization in vivo.  相似文献   

3.
Competition for adhesion to cellulose among the three main ruminal cellulolytic bacterial species was studied using differential radiolabeling (14C/3H) of cells. When added simultaneously to cellulose, Ruminococcus flavefaciens FD1 and Fibrobacter succinogenes S85 showed some competition; however, both species were surpassed competitively by Ruminococcus albus 20. When R. flavefaciens FD1 and F. succinogenes S85 were already adherent, R. albus 20 adhesion occurred without inhibition but involved R. flavefaciens FD1 detachment. Received: 28 October 1996 / Accepted: 28 January 1997  相似文献   

4.
Detection and quantification of cellulolytic bacteria with oligonucleotide probes showed that Ruminococcus flavefaciens was the predominant species in the pony and donkey cecum. Fibrobacter succinogenes and Ruminococcus albus were present at low levels. Four isolates, morphologically resembling R. flavefaciens, differed from ruminal strains by their carbohydrate utilization and their end products of cellobiose fermentation.  相似文献   

5.
Three rumen anaerobic fungi—Neocallinastix frontalis MCH3,Piromyces (Piromonas) communis FL, andCaecomyces (Sphaeromonas) communis FG10—were cultured on cellulose filter paper alone or in association with one of two rumen cellulolytic bacteria,Ruminococcus flavefaciens 007 andFibrobacter succinogenes S85. Cocultures ofN. frontalis orP. communis andR. flavefaciens were markedly less effective than the fungal monocultures in degrading cellulose but more effective than the bacterial monocultures.R. flavefaciens had an antagonistic effect against both of the fungal species. In contrast, no interaction was observed between the two fungal species andF. succinogenes. Cellulose was more effectively degraded by the cocultureC. communis-R. flavefaciens than by the corresponding fungal and bacterial monocultures. The effectiveness of degradation of the cocultureC. communis-F. succinogenes was comparable to that of the bacterial strains but greater than that of the fungi; no interaction was observed between these two microorganisms.  相似文献   

6.
Three predominant ruminal cellulolytic organisms, Fibrobacter succinogenes S85, Ruminococcus albus 8, and R. flavefaciens FD-1, were cultured with a methanogen, Methanobrevibacter smithii. Growth rates, bacterial protein, organic acids, and methane production were measured. When grown in diculture with the methanogen, a fermentative advantage was observed with F. succinogenes S85 as seen by an increase in specific rate of ATP production and organic acid concentration. The introduction of the methanogen did not improve the growth rate, organic acid yield, or specific rate of ATP production for R. albus 8. The growth rate and amount of organic acid end products increased when R. flavefaciens FD-1 was cultured with the methanogen; however, the specific activity of ATP production did not increase. Received: 3 August 1999 / Accepted: 4 September 1999  相似文献   

7.
Isolated rumen bacteria were examined for growth and, where appropriate, for their ability to degrade cellulose in the presence of the hydroxycinnamic acids trans-p-coumaric acid and trans-ferulic acid and the hydroxybenzoic acids vanillic acid and 4-hydroxybenzoic acid. Ferulic and p-coumaric acids proved to be the most toxic of the acids examined and suppressed the growth of the cellulolytic strains Ruminococcus albus, Ruminococcus flavefaciens, and Bacteroides succinogenes when included in a simple sugars medium at concentrations of >5 mM. The extent of cellulose digestion by R. flavefaciens and B. succinogenes but not R. albus was also substantially reduced. Examination of rumen fluid from sheep maintained on dried grass containing 0.51% phenolic acids showed the presence of phloretic acid (0.1 mM) and 3-methoxyphloretic acid (trace) produced by hydrogenation of the 2-propenoic side chain of p-coumaric and ferulic acids, respectively. The parent acids were found in trace amounts only, although they represented the major phenolic acids ingested. Phloretic and 3-methoxyphloretic acids proved to be considerably less toxic than their parent acids. All of the cellulolytic strains (and Streptococcus bovis) showed at least a limited ability to hydrogenate hydroxycinnamic acids, with Ruminococcus spp. proving the most effective. No further modification of hydroxycinnamic acids was produced by the single strains of bacteria examined. However, a considerable shortfall in the recovery of added phenolic acids was noted in media inoculated with rumen fluid. It is suggested that hydrogenation may serve to protect cellulolytic strains from hydroxycinnamic acids.  相似文献   

8.
Zhou Z  Yu Z  Meng Q 《Bioresource technology》2012,103(1):173-179
The objective of this study was to examine the effects of nitrate on methane production, important fermentation characteristics, Fibrobacter succinogenes, Ruminococcus albus, Ruminococcus flavefaciens, total bacteria, and methanogens using in vitro ruminal cultures. Potential adaptation of the above microbes and persistency of nitrate to mitigate CH4 production were also evaluated. Methane production was reduced by 70% at 12 μmol ml−1 and nearly completely at ?24 μmol ml−1 nitrate. Production of volatile fatty acids (VFAs) was affected to different extents at different nitrate concentrations. Over a series of six consecutive cultures receiving 12 μmol ml−1nitrate, production of CH4 and VFA did not change significantly. R. albus and R. flavefaciens seemed to adapt to nitrate, while F. succinogenes and methanogens did not. Nitrate may be used in achieving persistent mitigation of CH4 production by ruminants.  相似文献   

9.
The purpose of this study was to detect three fibrolytic bacteria, Fibrobacter succinogenes, Ruminococcus flavefaciens, and Ruminococcus albus, in the cecal digesta of the ostrich (Struthio camelus) by PCR using a species-specific primer set for each 16S ribosomal RNA gene (16S rDNA). Although amplified DNA fragments obtained from each primer set had the expected size, the clone library derived from the amplimer contained non-specific sequences. The F. succinogenes-specific primer set recovered a partial 16S rDNA sequence of an uncultivated Fibrobacter with low similarity (<95%) and distantly related phylogenetic positioning to Fibrobacter sequences deposited in the databases, indicating a novel species of Fibrobacter. The sequence was considered to be identical to a clone detected in our previous experiment. Thus, we confirm that the gastrointestinal tract of the ostrich is one of the habitats of Fibrobacter species. The clone library derived from the R. flavefaciens-specific primer set contained a 16S rDNA sequence with 97% similarity to R. flavefaciens, indicating it could be one of a major fibrolytic bacterium in the ostrich ceca. No R. albus 16S rDNA sequence was found in the clone library of the R. albus-specific primer set.  相似文献   

10.
The effect of glycerol on the growth, adhesion, and cellulolytic activity of two rumen cellulolytic bacterial species,Ruminococcus flavefaciens andFibrobacter succinogenes subsp.succinogenes, and of an anaerobic fungal species,Neocallimastix frontalis, was studied. At low concentrations (0.1–1%), glycerol had no effect on the growth, adhesion, and cellulolytic activity of the two bacterial species. However, at a concentration of 5%, it greatly inhibited their growth and cellulolytic activity. Glycerol did not affect the adhesion of bacteria to cellulose. The growth and cellulolytic activity ofN. frontalis were inhibited by glycerol, increasingly so at higher concentrations. At a concentration of 5%, glycerol totally inhibited the cellulolytic activity of the fungus. Thus, glycerol can be added to animal feed at low concentrations.  相似文献   

11.
Abstract The establishment of microorganisms in the rumen is a critical step if rumen manipulation is to be accomplished by use of microbial inoculants. Microbial populations in the maturing rumen undergo successional changes and, while in a state of flux, provide a possible opportunity for the introduction of specific strains of bacteria. While the rumen of the young lamb was maturing, we measured changes in several microbial populations with 16S-rRNA specific oligonucleotides: Rumincoccus, Fibrobacter, eukaryotes, Gram-positive bacteria, the Bacteroides–Porphromonas–Prevotella group, and anaerobic rumen fungi. In this study we repeatedly dosed 15 lambs with approximately 3.4 × 108 to 0.8 × 109 Ruminococcus cells dose-1, twice a week, for 7 wk from 23 d to 63 d of age. Of the five Ruminococcus strains dosed (R. albus SY3 and AR67, and R. flavefaciens Y1, LP9155, and AR72) the most specific primers (based on 16S rDNA) were obtained for strain SY3. There was an increase in the eukaryotic population during dosing, and it was hypothesized that protozoal predation contributed to the disappearance of strain SY3. At the end of dosing PCR amplification showed that SY3 were approximately 109 cells ml-1, but decreased to below the detection limit of the PCR system (8.6 × 104 ml-1) within 28 d postdosing. These experiments showed that fibrolytic populations increased significantly (P < 0.1) above the controls during the dosing period and were elevated for several days postdosing. This suggests that dosing of highly fibrolytic bacteria makes more of the fiber available to other organisms able to degrade fiber, and in so doing increases the overall fibrolytic activity of the rumen. Examination of the succession of gram-positive bacteria and the Bacteroides–Porphromonas–Prevotella group showed a decline in relative abundance as the lambs matured. Received: 13 April 1999; Accepted: 14 July 1999; Online Publication: 15 February 2000  相似文献   

12.
Volatile Fatty Acid Requirements of Cellulolytic Rumen Bacteria   总被引:12,自引:1,他引:11  
A gas chromatographic method was developed which could separate the isomers isovaleric and 2-methylbutyric acid. Subsequent analyses revealed that most commercially available samples of these acids were cross-contaminated; however, one sample of each acid was found to be pure by this criterion. The growth response of seven strains of cellulolytic rumen bacteria (three strains of Bacteroides succinogenes, three strains of Ruminococcus flavefaciens, and one strain of R. albus) to additions of isobutyric, isovaleric, 2-methylbutyric, valeric, and combinations of valeric and a branched-chain acid was determined. Strains of B. succinogenes required a combination of valeric plus either isobutyric or 2-methylbutyric acid. Isovaleric acid was completely inactive. Either isobutyric or 2-methylbutyric acid was required for the growth of R. albus 7. Strain C-94 of R. flavefaciens grew slowly in the presence of any one of the three branched-chain acids, but a combination of isobutyric and 2-methylbutyric acids appeared to satisfy this organism's growth requirements. None of the individual acids or mixtures of straight- and branched-chain acids allowed growth of R. flavefaciens strain C1a which would approach the response obtained from the total mixture of acids. Further work indicated that all three branched-chain acids were required for optimal growth by this strain, although isovaleric acid only influenced the rate of maximal growth. Either 2-methylbutyric or isovaleric acid allowed growth of nearly the same magnitude as that of the positive control for R. flavefaciens B34b. The presence of acetic acid had little influence on the rate or extent of growth of any of the strains except R. albus 7, for which the extent of growth was markedly increased. Determination of the quantitative fatty acid requirements for the three B. succinogenes strains indicated that 0.1 μmole of valeric per ml and 0.05 μmole of 2-methylbutyric per ml permitted maximal growth. However, with isobutyric acid as the branched-chain component, strains A3c and B21a required 0.1 μmole/ml in contrast to S-85 which exhibited optimal growth at the 0.05 μmole/ml level. By use of mixtures of isobutyric and 2-methylbutyric acids, good growth of C-94 was obtained at concentrations of 0.1 and 0.01 μmole/ml, respectively. About 0.3 μmole/ml of each acid was required for satisfactory growth of C1a.  相似文献   

13.
Twenty-one ruminal bacteria species were tested for their ability to degrade 2,4,6-trinitrotoluene (TNT) within 24 h. Butyrivibrio fibrisolvens, Fibrobacter succinogenes, Lactobacillus vitulinus, Selenomonas ruminantium, Streptococcus caprinus, and Succinivibrio dextrinosolvens were able to completely degrade 100 mg/L TNT, with <5% of the original TNT recovered as diaminonitrotoluene metabolites. Eubacterium ruminantium, Lactobacillus ruminis, Ruminobacter amylophilus, Streptococcus bovis, and Wolinella succinogenes were able to completely degrade 100 mg/L TNT, with 23–60% of the TNT recovered as aminodinitrotoluene and/or diaminonitrotoluene metabolites. Clostridium polysaccharolyticum, Megasphaera elsdenii, Prevotella bryantii, Prevotella ruminicola, Ruminococcus albus, and Ruminococcus flavefaciens were able to degrade 80–90% of 100 mg/L TNT. Desulfovibrio desulfuricans subsp. desulfuricans, Prevotella albensis, and Treponema bryantii degraded 50–80% of the TNT. Anaerovibrio lipolytica was completely inhibited by 100 mg/L TNT. These results indicate that a variety of rumen bacteria is capable of transforming TNT.  相似文献   

14.
Summary Pure cultures of the cellulolytic rumen bacterial strains Bacteroides succinogenes S85, Ruminococcus flavefaciens FD1 and Ruminococcus albus 7 were grown on lucerne cell walls (CW) or on cellobiose as the sole added carbohydrate substrate. Scanning electron microscopy visualization using cationized-feritin pretreatment have shown that cell surface topology of these strains grown on and attached to CW particles was specified by a dense coat of characteristic protuberant structures. In contrast, when grown on cellobiose, the surface topology of these bacterial strains was smoother, and contained fewer protuberant structures. The ability of these bacterial strains to attach to cellulose was higher for bacteria previously adapted to lucerne CW compared to cellobiose adaptation. Bacteroides succinogenes S85 was the best digester of lucerne CW (46.5%) and also had the best adhesion capability (65.6%) after adaption to grow on CW.Contribution from the Agricultural Research Organization, The Volcani Center, Bet Dagan, Israel. No. 2599-E, 1989 seriesOffprint requests to: J. Miron  相似文献   

15.
The influence of slow-release urea ( urea–calcium sulphate mixture; U–CaS) in feed blocks on rumen micro-organisms, predominant cellulolytic bacteria, microbial protein synthesis and ecology was studied in Thai native beef cattle. Four animals with an initial body weight of 100 ± 3.0 kg were randomly assigned to a 4 × 4 Latin square design with four dietary treatments (U–CaS in iso-nitrogen feed blocks at 0, 120, 150 and 180 g/kg dry matter (DM), respectively; U–CaS replaced urea). After 21 days of experimental feeding, rumen fluid was collected at 0 and 4 h after feeding. The mean intake of feed blocks and other feedstuffs offered (rice straw and concentrates) amounted to 0.3, 2.3 and 0.6 kg DM/day, respectively. Inclusion of U–CaS did not altered pH and temperature in the rumen. However, ruminal NH3–N concentration decreased quadratically (p < 0.05) in response to U–CaS inclusion, with the lowest value at 180 g U–CaS per kg feed block. With inclusion of U–CaS, the populations of rumen bacteria increased quadratically (p < 0.05) and counts of fungal zoospores were linearly enhanced (p < 0.05), being highest at 180 g U–CaS per kg feed block. Supplementation of U–CaS increased the concentration of total bacteria linearly (p < 0.05) and of Fibrobacter succinogenes quadratically (p < 0.05), whereas Ruminococcus flavefaciens and Ruminococcus albus were not affected by dietary treatments. Microbial crude protein yield and efficiency of microbial nitrogen (N) synthesis were linearly increased with different levels of U–CaS addition. Furthermore, current data clearly indicate that inclusion of U–CaS in feed blocks can affect micro-organism diversity and major cellulolytic bacteria.  相似文献   

16.
The origin of cell nitrogen and amino acid nitrogen during growth of ruminal cellulolytic bacteria in different growth media was investigated by using 15NH3. At high concentrations of peptides (Trypticase, 10 g/liter) and amino acids (15.5 g/liter), significant amounts of cell nitrogen of Fibrobacter succinogenes BL2 (51%), Ruminococcus flavefaciens 17 (43%), and Ruminococcus albus SY3 (46%) were derived from non-NH3-N. With peptides at 1 g/liter, a mean of 80% of cell nitrogen was from NH3. More cell nitrogen was formed from NH3 during growth on cellobiose compared with growth on cellulose in all media. Phenylalanine was essential for F. succinogenes, and its 15N enrichment declined more than that of other amino acids in all species when amino acids were added to the medium.  相似文献   

17.
In vitro cultures of ruminal microorganisms were used to determine the effect of cinnamic acid and vanillin on the digestibility of cellulose and xylan. Cinnamic acid and vanillin depressed in vitro dry matter disappearance of cellulose 14 and 49%, respectively, when rumen fluid was the inoculum. The number of viable Bacteroides succinogenes cells, the predominant cellulolytic organism, was threefold higher for fermentations which contained vanillin than for control fermentations. When xylan replaced cellulose as the substrate, a 14% decrease in the digestibility of xylan was observed with vanillin added; however, the number of viable xylanolytic bacteria cultured from the batch fermentation was 10-fold greater than that of control fermentations. The doubling time of B. succinogenes was increased from 2.32 to 2.58 h when vanillin was added to cellobiose medium, and absorbance was one-half that of controls after 18 h. The growth rate of Ruminococcus albus and Ruminococcus flavefaciens was inhibited more by p-coumaric acid than by vanillin, although no reduction of final absorbance was observed in their growth cycles. Vanillin, and to a lesser extent cinnamic acid, appeared to prevent the attachment of B. succinogenes cells to cellulose particles, but did not affect dissociation of cells from the particles. B. succinogenes, R. albus, R. flavefaciens, and Butyrivibrio fibrisolvens all modified the parent monomers cinnamic acid, p-coumaric acid, ferulic acid, and vanillin, with B. fibrisolvens causing the most extensive modification. These results suggest that phenolic monomers can inhibit digestibility of cellulose and xylan, possibly by influencing attachment of the fibrolytic microorganisms to fiber particles. The reduced bacterial attachment to structural carbohydrates in the presence of vanillin may generate more free-floating fibrolytic organisms, thus giving a deceptively higher viable count.  相似文献   

18.
Six matured male Yaks (Bos grunniens) with a mean live weight of 450 ± 23 kg (mean ± SD), were housed indoors in metabolism cages and fed pelleted lucerne (Medicago sativum). After an adjustment period of 24 days of feeding the diet, samples of rumen content were obtained for analysis of the bacteria in the liquor. The diversity of rumen bacteria was investigated by constructing a 16S rRNA gene clone library using the general bacterial primers F27 and R1492. A total of 130 clones, comprising nearly full length sequences (approx. 1.5 kb) were sequenced and submitted to BLAST and phylogenetic analysis. Using the criterion that similarity of 97% or greater with the sequences of cultivated bacteria, 16 clones were identified as Butyrivibrio fibrisolvens, Pseudobutyrivibrio ruminis, Ruminococcus flavefaciens, Succiniclasticum ruminis, Selenomonas ruminantium and Prevotella ruminicola, respectively. A further 10 clones shared similarity ranging from 90 to 97% with cultivated bacteria but the similarity in sequences for the remaining 104 clones were less than 90% of those of cultivated bacteria. Using a phylogenetic analysis it was found that the majority of the clones identified (63.8%) were located in the Low G + C Subdivision, with most of the remainder (35.4% of clones) located in the Cytophaga-Flexibacter-Bacteroides phylum and one clone (0.8%) was identified as a Proteobacteria. It was apparent that Yaks have a large and diverse range of bacteria in the rumen content which differ from those of cattle and other ruminants.  相似文献   

19.

Background

A complex community of microorganisms is responsible for efficient plant cell wall digestion by many herbivores, notably the ruminants. Understanding the different fibrolytic mechanisms utilized by these bacteria has been of great interest in agricultural and technological fields, reinforced more recently by current efforts to convert cellulosic biomass to biofuels.

Methodology/Principal Findings

Here, we have used a bioinformatics-based approach to explore the cellulosome-related components of six genomes from two of the primary fiber-degrading bacteria in the rumen: Ruminococcus flavefaciens (strains FD-1, 007c and 17) and Ruminococcus albus (strains 7, 8 and SY3). The genomes of two of these strains are reported for the first time herein. The data reveal that the three R. flavefaciens strains encode for an elaborate reservoir of cohesin- and dockerin-containing proteins, whereas the three R. albus strains are cohesin-deficient and encode mainly dockerins and a unique family of cell-anchoring carbohydrate-binding modules (family 37).

Conclusions/Significance

Our comparative genome-wide analysis pinpoints rare and novel strain-specific protein architectures and provides an exhaustive profile of their numerous lignocellulose-degrading enzymes. This work provides blueprints of the divergent cellulolytic systems in these two prominent fibrolytic rumen bacterial species, each of which reflects a distinct mechanistic model for efficient degradation of cellulosic biomass.  相似文献   

20.
The genetic diversity of protozoa in Surti buffalo rumen was studied by amplified ribosomal DNA restriction analysis, 18S rDNA sequence homology and phylogenetic and Real-time PCR analysis methods. Three animals were fed diet comprised green fodder Napier bajra 21 (Pennisetum purpureum), mature pasture grass (Dicanthium annulatum) and concentrate mixture (20% crude protein, 65% total digestible nutrients). A protozoa-specific primer (P-SSU-342f) and a eukarya-specific primer (Medlin B) were used to amplify a 1,360 bp fragment of DNA encoding protozoal small subunit (SSU) ribosomal RNA from rumen fluid. A total of 91 clones were examined and identified 14 different 18S RNA sequences based on PCR–RFLP pattern. These 14 phylotypes were distributed into four genera-based 18S rDNA database sequences and identified as Dasytricha (57 clones), Isotricha (14 clones), Ostracodinium (11 clones) and Polyplastron (9 clones). Phylogenetic analyses were also used to infer the makeup of protozoa communities in the rumen of Surti buffalo. Out of 14 sequences, 8 sequences (69 clones) clustered with the Dasytricha ruminantium-like clone and 4 sequences (13 clones) were also phylogenetically placed with the Isotricha prostoma-like clone. Moreover, 2 phylotypes (9 clones) were related to Polyplastron multivesiculatum-like clone. In addition, the number of 18S rDNA gene copies of Dasytricha ruminantium (0.05% to ciliate protozoa) was higher than Entodinium sp. (2.0 × 105 vs. 1.3 × 104) in per ml ruminal fluid.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号