Skip to main content

Feeding live yeast (Saccharomyces cerevisiae) improved performance of mid-lactation dairy cows by altering ruminal bacterial communities and functions of serum antioxidation and immune responses

Abstract

Background

The utilization of live yeast (Saccharomyces cerevisiae, YE) in dairy cows is gaining traction in dairy production as a potential strategy to improve feed efficiency and milk yield. However, the effects of YE on dairy cow performance remain inconsistent across studies, leaving the underlying mechanisms unclear. Hence, the primary aim of this study was to investigate the impact of YE supplementation on lactation performance, ruminal microbiota composition and fermentation patterns, as well as serum antioxidant capacity and immune functions in dairy cows.

Results

Supplementation with YE (20 g/d/head) resulted in enhancements in dairy cow’s dry matter intake (DMI) (P = 0.016), as well as increased yields of milk (P = 0.002) and its components, including solids (P = 0.003), fat (P = 0.014), protein (P = 0.002), and lactose (P = 0.001) yields. The addition of YE led to significant increases in the concentrations of ammonia nitrogen (NH3-N) (P = 0.023), acetate (P = 0.005), propionate (P = 0.025), valerate (P = 0.003), and total volatile fatty acids (VFAs) (P < 0.001) in rumen fermentation parameters. The analysis of 16s rRNA gene sequencing data revealed that the administration of YE resulted in a rise in the relative abundances of three primary genera including Ruminococcus_2 (P = 0.010), Rikenellaceae_RC9_gut_group (P = 0.009), and Ruminococcaceae_NK4A214_group (P = 0.054) at the genus level. Furthermore, this increase was accompanied with an enriched pathway related to amino acid metabolism. Additionally, enhanced serum antioxidative (P < 0.05) and immune functionalities (P < 0.05) were also observed in the YE group.

Conclusions

In addition to improving milk performance, YE supplementation also induced changes in ruminal bacterial community composition and fermentation, while enhancing serum antioxidative and immunological responses during the mid-lactation stage. These findings suggest that YE may exert beneficial effects on both rumen and blood metabolism in mid-lactation dairy cows.

Peer Review reports

Background

High-yielding dairy cows during early and mid-lactation encounter challenges in utilizing diets rich in readily fermentable carbohydrates due to physiological limitations [1]. This can lead to a combination of inadequate rumen buffering from fine feed particles and reduced rumination time, ultimately disrupting microbial balance [2]. Consequences include decreased volatile fatty acids (VFAs) production, elevated lactic acid concentrations, and potentially subacute ruminal acidosis (SARA) [2]. Therefore, enhancing dairy cow well-being and productivity remains a key objective in modern dairy production research. The rumen, a complex ecosystem of microbes, plays a crucial role in the digestion and utilization of nutrient, significantly impacting overall cow performance [3]. The composition and function of the rumen bacterial community directly influence rumen fermentation, nutrient availability for the cow, and ultimately, milk yield [4]. Therefore, researchers are increasingly focused on investigating strategies to modulate the rumen microbiota.

Research on live yeast (Saccharomyces cerevisiae, YE) supplementation in dairy cow dates back to the 1950s and continues to be actively investigated [5]. YE is a rich source of nutrients, including vitamins, amino acids, peptides, minerals, organic acids, antioxidants, oligosaccharides, and β-glucans, which have been shown to promote the proliferation of rumen bacteria, protozoa, and fungi [6]. Growing evidence supports the effectiveness of YE supplementation in various aspects, including stimulating cellulolytic bacteria, enhancing the growth of lactate-utilizing bacteria, and mitigating post-feeding rumen pH decline [7]. Multiple studies have reported increased milk yield associated with YE or related products [8,9,10]. However, the observed responses vary significantly due to factors like yeast strain, dose, and mode of action, as well as animal factors such as lactation stage, production level, die composition (energy level), and parity [11]. The presence of these characteristics poses challenges in comparing outcomes and assessing YE efficacy on individual farms. Moreover, prior research has primarily focused on the effects of YE (S. cerevisiae) on rumen fermentation parameters and nutrient metabolism [12,13,14,15]. A few studies have investigated its specific impacts on the composition and functions of the bacterial population within the rumen [16, 17]. However, the findings from these studies regarding the interaction between ruminal microbiota and dairy cows have shown inconsistencies, leading to variations in microbiota composition and functions [18,19,20]. Therefore, we hypothesized that incorporating YE (S. cerevisiae) into diet of mid-lactation high-producing dairy cows would have positive effects on milk performance through its modulation of rumen microbial composition and function. Thus, the object of this study was to investigate the effect of YE supplementation on lactation performance, the rumen bacterial community and its predicted functions, as well as serum antioxidation and immune properties in mid-lactation dairy cows.

Results

Lactation performance

As displayed in Table 1, an evaluation of lactation performance revealed significant treatment effects. Dairy cows supplemented with YE had reduced dry matter intake (DMI) compared to the control (CON) group (P = 0.016), while milk yield significantly increased in the YE group (P = 0.002). Milk protein and lactose percentages were independently affected by treatment (P = 0.019) and time (P < 0.001). Importantly, YE treatment significantly improved the yield of milk fat, protein, lactose, and solids, as well as production efficiency measures such as 3.5% fat corrected milk (FCM) and energy corrected milk (ECM) (P = 0.014, P = 0.002, P = 0.001, P = 0.003, P = 0.001, and P < 0.001). It’s important to note that time did not exert a significant effect on these lactation performance variables, nor were there significant interactions between treatment and time observed for any measured parameters.

Table 1 DMI, milk yield, and composition of dairy cows fed the basal diet (CON) or addition of 20 g/d live yeast (YE).

Ruminal fermentation profiles

Rumen fermentation parameters (Table 2) showed no significant differences between the two groups for pH (P = 0.236), concentrations of butyrate (P = 0.062), isobutyrate (P = 0.069), isovalerate (P = 0.117), or the ratio of acetate to propionate (P = 0.6610). However, the YE group exhibited significantly higher concentrations of ammonia nitrogen (NH3-N), acetate, propionate, valerate, and total VFA (TVFA) compared to the CON group (P = 0.023, P = 0.005, P = 0.025, P = 0.003, and P < 0.001, respectively).

Table 2 Rumen fermentation parameters of dairy cows fed the basal diet (CON) and addition of 20 g/d live yeast (YE).

Serum antioxidation and immune responses

As shown in Table 3, compared to the CON group, the YE group exhibited significantly increased activity of serum antioxidant enzymes, including catalase (CAT), glutathione peroxidase (GSH-Px), and superoxide dismutase (SOD) (P < 0.05). Additionally, the concentration of malondialdehyde (MDA), a marker of oxidative stress, was significantly lower in the YE group (P < 0.05). Regarding serum immune response, YE supplementation significantly increased the concentrations of immunoglobulin A (IgA), G (IgG), and M (IgM) compared to the CON group (P < 0.05). Nevertheless, the contents of soluble CD4 (sCD4) and soluble CD8 (sCD8) were significantly higher in the CON group compared to the YE group (P < 0.05).

Table 3 Serum antioxidation and immune response indices of dairy cows fed the basal diet (CON) and addition of 20 g/d live yeast (YE).

Ruminal bacterial communities

Alpha diversity analyses revealed significant variations in multiple indices between the two groups. The Chao1 and Ace richness indices demonstrated a significant increase in the YE group in comparison to the CON group (Fig. 1a and d). Conversely, there were no significant variations seen in the Shannon and Simpson indices between the two groups (Fig. 1b and c). Beta diversity analysis indicated a significant difference (P = 0.001) in the rumen microbiota compositions between cows fed YE and those in the CON group (Fig. 1e). The taxonomic study provided annotations for a total of 17 bacterial phyla, specifically at the phylum level. The phyla Firmicutes and Bacteroidetes were found to be the most abundant, representing 45.98-46.95% and 44.95-46.12% of the total sequences, respectively (Supplementary file 1). The taxonomic groups Patescibacteria, Actinobacteria, Proteobacteria, Tenericutes, and Spirochaetes accounted for 1.54-2.57%, 1.11-2.27%, 1.12-1.48%, 0.96-1.49%, and 0.50-1.42% of the overall sequences, respectively. The CON group had higher abundance of Tenericutes in this study (P = 0.049). YE addition significantly increased the abundance of Spirochaetes (relative abundance > 0.5%, P = 0.002) (Fig. 2a). The dominating families at the family level were Prevotellaceae (33.76-36.9%), Ruminococcaceae (18.00-22.24%), and Lachnospiraceae (14.51-15.37%). Additional families observed in the study encompassed Acidaminococcaceae (4.17-5.26%), Muribaculaceae (2.65-3.90%), Rikenellaceae (2.01-3.81%), Christensenellaceae (2.13-2.72%), unidentified F082 (2.09-2.49%), Veillonellaceae (1.21-1.62%), and Erysipelotrichaceae (0.47-2.09%) (Supplementary file 1). The CON group had higher abundance of norank_o__Mollicutes_RF39 in this study (P = 0.031). YE addition significantly increased the abundance of Spirochaetaceae (relative abundance > 0.5%, P = 0.002) and Rikenellaceae (P = 0.007) in dairy cows (Fig. 2b). At the taxonomic level of genus (as shown in Supplementary file 1), the dominant genera observed in dairy cows were Prevotella_1, Ruminococcaceae_NK4A214_group, Lachnospiraceae_NK3A20_group, Succiniclasticum, norank_f__Muribaculaceae, Ruminococcaceae_UCG-014, Ruminococcus_2, Rikenellaceae_RC9_gut_group, Christensenellaceae_R-7_group, norank_f__F082, Eubacterium_coprostanoligenes_group, Prevotellaceae_UCG-003, and Acetitomaculum. These genera were considered dominant as their relative abundance exceeded 1%. The CON cows had higher abundances of Prevotella_7 (P < 0.001) and Ruminococcaceae_UCG-014 (P = 0.003). YE addition significantly increased the abundances of Ruminococcus_2 (P = 0.010) and Rikenellaceae_RC9_gut_group (P = 0.009) (Fig. 2c) and trended toward increasing the abundance of Ruminococcaceae_NK4A214_group in dairy cows (P = 0.054). The cladogram presented in Fig. 3 demonstrated the predominant microbiome structure and highlighted the notable variations in taxa between the CON and YE groups, as indicated by liner discriminant analysis effect size (Lefse) analysis. The results acquired from the study revealed that the CON group exhibited a higher abundance of 25 clades, whereas the YE group showed a higher abundance of 61 clades. The CON group had Prevotella_7 as the sole differential biomarker with an LDA score beyond 4. The differential biomarkers in the YE group, as indicated by an LDA score greater than 4, were Ruminococcus_2 and Ruminococcaceae_NK4A214_group (Supplementary file 2).

Fig. 1
figure 1

Effects of live yeast (YE) on ruminal microbiota of dairy cows. (A), (B), (C), and (D) The richness and diversity indices of rumen microbiota in dairy cows fed basal (CON) or live yeast (YE) diet. (E) Principal coordinate analysis (PCoA) of the overall rumen microbiota in dairy cows based on unweighted UniFrac distance

Fig. 2
figure 2

Effects of live yeast (YE) on the rumen bacterial composition at the phylum (A), family (B), and genus (C) levels in dairy cows. “*”, “**”, and “***” indicate the significance level at 0.05, 0.01 and 0.001, respectively

Fig. 3
figure 3

LEFse (Liner discriminant analysis Effect Size) cladogram comparing microbiota communities between the YE and CON groups. Differences are represented by colour, indicating the group where taxa are most abundant: red = taxa abundant in the basal (CON) group, blue = tax abundant in the live yeast (YE) group

The correlation between bacterial populations, lactation performance, and ruminal fermentation profiles

Spearman’s rank correlation analysis revealed associations among lactation performance, ruminal fermentation characteristics, and the predominant ruminal bacterial populations in dairy cows (Fig. 4). As shown in Fig. 4a, there was a negative correlation observed between the relative abundances of Ruminococcaceae_UCG-014 (r = -0.470, P = 0.036), Prevotella_7 (r = -0.464, P = 0.040), and [Ruminococcus]_gauvreauii_group (r = -0.597, P = 0.033) with milk protein percentage. While milk protein (kg/d), milk yield (kg/d), and milk lactose (kg/d) exhibited positive correlations with the relative abundances of Rikenellaceae_RC9_gut_group (r = 0.451, P = 0.046; r = 0.449, P = 0.047; r = 0.476, P = 0.034), Ruminococcaceae_NK4A214_group (r = 0.473, P = 0.035; r = 0.579, P = 0.007; r = 0.501, P = 0.025), and Ruminococcus_2 (r = 0.554, P = 0.011; r = 0.525, P = 0.017; r = 0.519, P = 0.019). Similarly, there was a positive correlation between the yield of milk solids (kg/d) and the relative abundance of Rikenellaceae_RC9_gut_group (r = 0.446, P = 0.049). The study found a favorable correlation between milk fat (kg/d) and the relative abundances of norank_f__Muribaculaceae (r = 0.462, P = 0.040), Ruminococcaceae_NK4A214_group (r = 0.509, P = 0.022), and Ruminococcus_2 (r = 0.537, P = 0.015). The relative abundances of norank_f__Muribaculaceae, Rikenellaceae_RC9_gut_group, Ruminococcaceae_NK4A214_group, and Ruminococcus_2 were found to be positively correlated with the production efficiency of 3.5% FCM (kg/d) and ECM (kg/d) (r = 0.454, P = 0.044; r = 0.463, P = 0.040; r = 0.453, P = 0.045; r = 0.448, P = 0.048; r = 0.583, P = 0.007; r = 0.548, P = 0.012; r = 0.566, P = 0.009; r = 0.573, P = 0.008).

Fig. 4
figure 4

Correlation between the relative abundances of rumen bacteria, lactation performance (A), and fermentation parameters (B). “*”, “**”, and “***” indicate the significance level at 0.05, 0.01 and 0.001, respectively

Analysis correlations between rumen fermentation parameters and bacterial populations revealed distinct associations (Fig. 4b). There was a positive correlation between the concentrations of TVFA and acetate, and the relative abundances of the genera Ruminococcus_2 (r = 0.610, P = 0.004; r = 0.450, P = 0.046), Rikenellaceae_RC9_gut_group (r = 0.474, P = 0.034; r = 0.485, P = 0.030), and Ruminococcaceae_NK4A214_group (r = 0.495, P = 0.026; r = 0.459, P = 0.042). Conversely, there was a negative correlation between the concentrations of TVFA and acetate, and the genera [Ruminococcus]_gauvreauii_group (r = -0.612, P = 0.004; r = -0.614, P = 0.004), Prevotella_7 (r = -0.618, P = 0.004; r = -0.652, P = 0.002), and Olsenella (r = -0.551, P = 0.012; r = -0.656, P = 0.012). Notably, there was a positive correlation observed between the concentration of propionate and the relative abundance of Ruminococcus_2 (r = 0.503, P = 0.024). Interestingly, there was a significant positive correlation between the concentration of valerate and the relative abundance of Prevotellaceae_UCG-003 (r = 0.627, P = 0.003). While there were significant negative correlations between the valerate concentration and the relative abundances of [Ruminococcus]_gauvreauii_group (r = -0.474, P = 0.035), Ruminococcaceae_UCG-014 (r =-0.636, P = 0.003), and norank_o__Mollicutes_RF39 (r = -0.469, P = 0.037).

Function predictions of ruminal bacterial populations

We performed functional predictions using phylogenetic investigation of communities by reconstruction of unobserved states 2 (PICRUSt2) to further understand the ruminal bacteria. All samples shared 46 predicted gene families. As shown in Fig. 5, the YE group exhibited a significantly higher relative abundances of gene families associated with the infectious disease (P = 0.009) and amino acid metabolism (P = 0.045) compared to the CON group. Conversely, the CON group had a significantly greater relative abundance of gene families linked to substance dependence (P < 0.001).

Fig. 5
figure 5

Functional prediction of bacterial populations in ruminal samples of dairy cows fed the basal (CON) or live yeast (YE) diet. Prediction of the differential function of rumen microbes between two groups of dairy cows in Kyoto Encyclopaedia of Genes and Genomes (KEGG) level 2 category based on PICRUSt 2. “*”, “**”, and “***” indicate the significance level at 0.05, 0.01 and 0.001, respectively

Discussion

This study identified significant improvements in lactation performance within the YE group compared to the CON group. These included increased DMI, milk yield, and the yields of milk fat, protein, solids, and lactose. While previous research using YE supplementation in dairy cows reported similar individual variations in performance [21,22,23], the overall impact of YE remains variable across studies [11]. This inconsistency suggests a complex mechanism of action for YE. Multiple factors likely contribute to these differential reactions, including yeast strain, dosage, diet composition, lactation stage, and experimental design employed in various investigations [24]. Further research is crucial to elucidate the specific influence of these factors on YE’s effectiveness. Our findings demonstrate that cows supplemented with YE exhibited superior lactation performance, particularly in milk yield and its components (fat, protein, lactose, and solids). This difference in performance could be largely ascribed to the higher DMI and concentration of ruminal VFAs in the present study. Documented literatures support this connection, as YE supplementation has been shown to elevate DMI and apparent digestibility of nutrients in dairy animals, ultimately increasing digestible energy of the diet available for improved lactation performance [10, 25].

Previous studies highlighted the role of ruminal bacteria in enhancing the feed nutrient utilization efficiency [26], and their potential link to milk composition [27,28,29]. Our study, using 16sRNA gene sequencing, revealed differences in ruminal bacterial richness and diversity (measured by Chao1 and Ace indices) between the YE and CON groups. This aligns with proved evidence suggesting greater bacterial richness and diversity are associated with increased production of ruminal fermentation products, such as VFAs, amino acids, and glucose, all beneficial for milk production [30].

Milk solids components, including fat, protein, lactose, and minerals, is influenced by both feed intake and ruminal fermentation [31, 32]. The current study observed higher concentrations of most individual VFAs and TVFA in the YE group as compared to the CON group. These elevated VFA concentrations, likely due to increased DMI and fermentation rates, may contribute to the higher milk yield and its associated constituents (fat, protein, lactose, and solids) observed in the YE group. VFAs serve as primary energy sources for milk production, and the YE group exhibited a significant increase in acetate and propionate, which aligns with previous findings [30]. The current study also identified correlations between specific bacterial populations and the fluctuations in milk yield and its associated components. The YE group displayed a greater abundance of Rikenellaceae_RC9_gut_group, a prominent bacterial known to generate fermentation end product such as propionate, acetate and/or succinate [33]. This finding indicates a potential role for Rikenellaceae_RC9_gut_group in enhancing milk solids yield by possibly influencing the production of these key fermentation products, thus affecting milk major components. A previous investigation has indicated a positive correlation between the relative abundances of Ruminococcaceae_NK4A214_group and Ruminococcus_2 with milk solids production [30]. In this study, we observed a positive association between the relative abundances of these genera and milk solids yield. However, it is important to note that these correlations did not reach statistically significance (P = 0.058 for Ruminococcaceae_NK4A214_group and P = 0.065 for Ruminococcus_2). Notably, both genera, classified under the Ruminococcaceae family, were identified as distinctive biomarkers in the YE group by LEfse analysis. Furthermore, a positive correlation was observed between these genera and the production of milk fat and protein. The Ruminococcaceae family is recognized for its significant expertise in enzymatic degradation of complex plant substances, particularly cellulose, leading to the production of VFA that serve as energy resources for animals or other rumen-resident bacteria [34]. Acetate serves as a precursor for de novo lipogenesis, contributing to milk fat synthesis in mammary gland epithelial cells of bovine. Additionally, acetate acts as an energy source for microbial protein synthesis [35]. The present investigation revealed significant positive associations between acetate concentration and the relative abundances of Rikenellaceae_RC9_gut_group, Ruminococcaceae_NK4A214_group and Ruminococcus_2. Moreover, KEGG function prediction indicated enhanced amino acid metabolism, supporting by higher NH3-N concentration in the YE group, suggesting sufficient nitrogen resources for protein synthesis. Treponema_2, a prominent genus in the YE group with a relative abundance of 1.36%, is associated with tryptophan metabolism in healthy early lactation Holstein dairy cows [36]. These findings partially elucidated the underlying factors contributing to the higher yields of milk, solids, protein, and fat observed in the YE group compared to the CON group in our study, possibly due to a positive correlation with acetate concentration. Surprisingly, valerate, a key VFA found to be higher in high milk yield cows [37], exhibited a negative correlation with the abundances of [Ruminococcus]_gauvreauii_group and Ruminococcaceae_UCG-014 in the current study. Another investigation also showed that Ruminococcaceae and Ruminococcus had a negative correlation with valerate due to variations in fermentation process (specially protein fermentation) induced by cross feeding mechanisms [38]. Meanwhile, a previous study demonstrated a positive correlation of Ruminococcaceae _UCG-014 with valerate in marine fish [39]. The inconsistency between valerate concentration and these microbiota species or strains indicates the complexity of the interaction between microbiota and VFAs in the host, which requires further research to elucidate the relationships between these microbiotas and valerate. Prevotella is believed to have a significant involvement in the process of starch breakdown [40]. In the current investigation, Prevotella_7 exhibited differential biomarker characteristics exclusively in the CON group, with its relative abundance decreasing following YE supplementation, consistent with previous reports [41, 42]. Prevotella_7 was found to be more abundant in cows with lower efficiency and milk yield, as well as moderate SARA [42,43,44]. Nevertheless, other studies have reported that elevated or unchanged levels of Prevotella_7 following YE supplementation [45, 46]. Further investigation is needed to elucidate the ruminal bacterial community’s responds to YE supplementation.

High concentrate diets in dairy cow are linked to metabolic and systemic dysfunction [47]. Inadequate management practices in commercial dairy farm can further lead to milk production-related diseases such as mastitis and metritis [47, 48]. These factors collectively induce oxidative stress and compromised immune function in dairy cows. Nutritional interventions have been shown to reduce pro-oxidant loads and decrease the incidence of oxidative stress through antioxidant mechanisms. In the current investigation, the serum activities of CAT, GSH-Px, and SOD in the YE group exhibited a significant increase (P < 0.05) compared to the CON group. In contrast, MDA concentrate was significantly reduced, indicating the great potential of YE in mitigating oxidative stress in dairy cows. Numerous studies have reported that both YE and its culture products contribute to increased serum activities of antioxidant enzymes, along with enhanced antioxidant capacity involving radicals scavenging and metal chelating activities [49, 50]. The potential antioxidative effects of YE may be attributed to its antioxidative compounds, including vitamins A, E and C, polysaccharides, and sulfur-containing amino acids [51]. Moreover, YE supplementation appeared to affect the serum levels of immunoglobulins (IgA, IgG, and IgM), as well as sCD4, and sCD8 levels during the study period. Immunoglobulin levels experienced a significant increased, whereas sCD4, and sCD8 levels exhibited a notable decrease. YE and its associated products contain various immunomodulating chemicals that engage in direct and indirect interactions with pathogens and components of the immune system [52, 53]. Notably, polysaccharide β-glucan, a key component of YE, is categorized as a biological response modifier [54]. It stimulates innate immunity by enhancing the function of macrophages and neutrophils [55] and promotes a robust adaptive immune response by increasing antibody production [56]. sCD4 and sCD8 are considered as markers of T lymphocyte activation, which are maintained during the inflammatory process [57]. The induction of SARA can be triggered by a high concentrate diet, which facilitates the pathogenesis of ruminal acidosis by releasing lipopolysaccharides (LPS) into the rumen fluid [58]. Some proportion of LPS can translocate into body’s circulation system through both paracellular and transcellular pathways, leading to an increase in sCD4 and sCD8 levels [59]. The findings of this study indicate that YE supplementation has the potential to enhance cow organism immunity and mitigate inflammation-induced damage to the immune system.

Conclusions

In summary, our results demonstrated that the supplementation of YE to the diet increased DMI and lactation performance, along with enhanced serum antioxidative and immune functionalities in mid-lactation dairy cows. The application of YE also led to modifications in the ruminal microbiota and fermentation processes, resulting in improved amino acid metabolism. These findings significantly contribute to enhancing our overall understanding of the impacts of YE on rumen and blood metabolism in mid-lactation dairy cows.

Methods

Animal, diets, and experimental design

The current investigation was carried out in accordance with the Regulations for the Administration of Affairs Concerning Experimental Animals of the State Council of the People’s Republic of China. The research protocol employed in this study was granted approval by the Committee on Experimental Animal Management of the Chinese Academy of Agricultural Sciences, (Beijing), with the reference No. 39/14.08.2019. All methods were carried out in accordance with relevant guidelines and regulations and were reported in accordance with ARRIVE guidelines for the reporting of animal experiments.

The present study was carried out at Chaoren Dairy Co., in Yuncheng, Shanxi Province. Twenty mid-lactation, multiparous Chinese Holstein cows (mean weight 651 ± 12.47 kg) with similar days in lactation (110 ± 8) and overall health, were randomly assigned to two groups (n = 10) using a computer-generated randomization table. The cows were housed in individual tie-stall barns equipped with overhead fan devices situated above the lying area to ensure free access to water throughout the experiment. The control group (CON) received a total mixed ration (TMR) without yeast as the basal diet, whereas the treatment group (YE) was fed the same TMR supplemented with 20 g/d/cow of YE (strain Y03-0, 2.0 × 1010 CFU/g, ANGEL YEAST Co., Ltd., Yichang, China). The constituents and chemical composition of TMR as basal diet are summarized in Table 4. Feed was provided to the cows twice daily, at 0900 h and 1500 h, and milking procedures were conducted three times a day (0930 h, 1230 h, and 1530 h). The YE supplements was top-dressed on the TMR after initial consumption, followed by the remaining basal diet. The trial lasted for 60 d, divided into three 20-d periods. The experiment began with a 10-d adaption period, followed by a 50-d treatment period with YE. Feed intake measurements were conducted during the final two days of each experimental period. Feed was quantified at each feeding event, and the remaining amount was assessed prior to the morning feeding on the second day. DMI was calculated by subtracting the amount of feed delivered from the amount of residues, both measured on an absolute dried basis. Feed efficiency was determined using 3.5% of FCM and ECM. The nutritional composition of the feed was assessed on a per-period basis using the methods outlined by AOAC (2007). The feed samples obtained from each feeding session within the last 2 d of each experimental period were combined, and a total of 200 g of mixed feed was prepared for subsequent analysis. Detailed descriptions of feed sample analyses are provided in a prior publication [45].

Table 4 Ingredient and chemical composition of the basal diet

Sample collection and analysis

Milk yield and composition were determined on the two consecutive days every period throughout the experimental trial. Milk yield was recorded using the MagStream meter (BouMatic, Madison, WI, USA). The milk composition was measured by a near-infrared absorption analyzer (MilkoScan FT2, Foss Electric, Denmark). A detailed description of milk sample collection and composition analyzes are provided in a previous study [45].

On the final day of the formal experiment, two hours post-feeding in the morning, rumen fluid (100 mL) was collected via oral stomach sampling tube for each cow [49]. Each cow’s rumen fluid sample was snap-frozen in liquid nitrogen, with approximately 5 mL being used. These samples were then stored at -80℃ for the purpose of conducting 16 S rRNA sequencing. The pH of the remaining collected rumen fluid samples was determined using a portable pH meter (Seven2GO S7, Mettler-Toledo, Columbus, OH, USA). Subsequently, the samples were filtered through four layers of cheesecloth. The filtrate was preserved at -20℃ in order to facilitate the examination of rumen fermentation characteristics. These parameters encompassed VFA profiles, which were analyzed using gas chromatography (Agilent 6850, Agilent Technologies Inc., Santa Clara, CA, USA), as well as NH3-N concentrations [7].

Blood samples (n = 10) were collected at 6 h post-feeding (1500 h) via venipuncture of the tail vein and into evacuated tubes. In this study, blood samples of 10 mL each were collected and placed in tubes. These tubes were then allowed to reach room temperature and were held in this condition for a duration of 30 min. Subsequently, the tubes were centrifuged at 3000 g for 10 min at 4℃. The serum samples were carefully stored at 20℃, awaiting further analysis. The immune response and antioxidation parameters of the serum were assessed using a fully automatic biochemistry analyzer (Hitachi 7020, Tokyo, Japan) to determine the levels of IgA, IgM, and IgG. Additionally, the levels of sCD4, sCD8, CAT, GSH-Px, MDA, and SOD were measured using a commercially available diagnostic kit (Nanjing Jiancheng Bioengineering Institute, Nanjing, China) in accordance with the manufacturer’s instructions.

Microbial DNA extraction, sequencing and data analysis

The extraction of microbial DNA from rumen content samples (n = 10) was performed using the HiPure Stool DNA Kits (Angen, Guangzhou, China). The purity of the extracted DNA was assessed using the NanoDrop2000 spectrophotometer (Thermo Scientific, Madison, WI, USA). To generate separate amplicon libraries, the 16S rRNA V3-V4 region was amplified by PCR with the primer pair 341F (5’-CCTACGGGNGGCWGCAG-3’) and 806R (5’-GGACTACHVGGGTATCTAAT-3’). The sequencing process was conducted at a commercial laboratory (Majorbio Biotechnology Co., Ltd., Shanghai, China) using Illumina HiSeq 2500 platform, following established protocols. The raw data underwent processing through the use of quality filters in QIIME. Subsequently, the raw reads underwent a trimming process wherein adapters and low-quality sequences, possessing a quality score above 20, were removed. The study employed UPARSE (version 9.2.64) to conduct operational taxonomic unit (OTU) cluster analysis, using a similarity threshold of 97% as described by Edgar (2013) [60]. The taxonomic categorization of the typical OUT sequences into organisms was conducted using the RDP classifier, employing a confidence threshold value of 0.8. This classification was based on the SILVA database [61]. Bacterial diversity was evaluated using alpha diversity indices (Chao1, ACE, Simpson, and Shannon), and statistical analysis employed the Kruskal-Wallis test with False Discovery Rate (FDR) correction. The analysis of beta diversity was conducted by employing unweighted UniFrac distance metrics and principal coordinates analysis (PCoA) in order to evaluate the differences in bacterial communities across the various samples. The present study utilized LEfSe analysis to identify microbiota biomarker features in each group, utilizing the Kruskal-Wallis test for screening purpose. Additionally, the threshold for the LDA score was set to its default value of 2.0. It was observed that there was a statistically significant difference in the relative abundance between the two groups, as shown by an LDA value greater than 4.

The functional alternations of the microbiota in different samples were predicted through the utilization of PICRUSt2 analysis. This analysis was conducted based on the level 2 pathways of the Kyoto Encyclopedia of Genes and Genomes (KEGG) database. For more information on PICRUSt2, please refer to the following link: https://github.com/picrust/picrust2.

Statistical analysis

The lactation performance data, including DMI, milk yield, and components, were subjected to analysis using the MIXED model of SPSS 20.0 software (SPSS Inc., Chicago, IL, USA). The analysis was conducted for a completely randomized design, with adaption period as a covariate factor. The analysis used fixed effects for treatment, period (time of sample collection), the interaction between treatment and period, and covariate, with the cow serving as the experimental unit. The residual (co)variance matrix was considered to have a compound symmetry structure. The AR1 covariance structure was employed in accordance with the methodology outlined in a prior research study [7]. Before analysis, normality of variance in the data was checked, and it was determined that no data changes were necessary. The data pertaining to rumen fermentation parameters and serum chemistry were subjected to analysis using the Student’s t test. The nonparametric test (Kruskal-Wallis) was employed to assess the variations in the relative abundance of bacterial communities. The significance level was determined to be P < 0.05. The statistical analysis to determine the relationship between lactation performance, rumen VFA concentrations and bacteria abundance was conducted using Spearman’s correlation test in SPSS 20.0. A significant level of P < 0.05 and a correlation coefficient (r) with an absolute value greater than 0.8 were considered indicators of significant correlations.

Data availability

The datasets generated and/or analyzed during the current study are available in the Sequence Read Archive of the National Center for Biotechnology database repository with accession project number PRJNA1034486.

References

  1. McDonald P, Edwards RA, Greenhalgh JFD, Morgan CA. Animal nutrition. 6th ed. Harlow, UK: Pearson Education Limited; 2002.

    Google Scholar 

  2. Ricci S, Rivera-Chacon R, Petri RM, et al. Supplementation with phytogenic compounds modulates salivation and salivary physico-chemical composition in cattle fed a high-concentrate diet. Front Physiol. 2021;12:645529.

    Article  PubMed  PubMed Central  Google Scholar 

  3. Jize Z, Zhuoga D, Xiaoqing Z, et al. Different feeding strategies can affect growth performance and rumen functions in Gangba sheep as revealed by integrated transcriptome and microbiome analyses. Front Microbiol. 2022;13:908326.

    Article  PubMed  PubMed Central  Google Scholar 

  4. Golder HM, Thomson J, Rehberger J, Smith AH, Block E, Lean IJ. Associations among the genome, rumen metabolome, ruminal bacteria, and milk production in early-lactation Holsteins. J Dairy Sci. 2023;106(5):3176–91.

    Article  CAS  PubMed  Google Scholar 

  5. Hansen HH, El-Bordeny NE, Ebeid HM. Response of primiparous and multiparous buffaloes to yeast culture supplementation during early and mid-lactation. Anim Nutr. 2017;3(4):411–8.

    Article  PubMed  PubMed Central  Google Scholar 

  6. Sousa D, Oliveira CA, Velásquez AV, et al. Live yeast supplementation improves rumen fibre degradation in cattle grazing tropical pastures throughout the year. Anim Feed Sci Technol. 2018;236:149–58.

    Article  CAS  Google Scholar 

  7. Kumprechtová D, Illek J, Julien C, Homolka P, Jančík F, Auclair E. Effect of live yeast (Saccharomyces cerevisiae) supplementation on rumen fermentation and metabolic profile of dairy cows in early lactation. J Anim Physiol Anim Nutr (Berl). 2019;103(2):447–55.

    Article  PubMed  Google Scholar 

  8. Nocek JE, Holt MG, Oppy J. Effects of supplementation with yeast culture and enzymatically hydrolyzed yeast on performance of early lactation dairy cattle. J Dairy Sci. 2011;94(8):4046–56.

    Article  CAS  PubMed  Google Scholar 

  9. Moallem U, Lehrer H, Livshitz L, Zachut M, Yakoby S. The effects of live yeast supplementation to dairy cows during the hot season on production, feed efficiency, and digestibility. J Dairy Sci. 2009;92(1):343–51.

    Article  CAS  PubMed  Google Scholar 

  10. Desnoyers M, Giger-Reverdin S, Bertin G, Duvaux-Ponter C, Sauvant D. Meta-analysis of the influence of Saccharomyces cerevisiae supplementation on ruminal parameters and milk production of ruminants. J Dairy Sci. 2009;92(4):1620–32.

    Article  CAS  PubMed  Google Scholar 

  11. Baker LM, Kraft J, Karnezos TP, Greenwood SL. Review: the effects of dietary yeast and yeast-derived extracts on rumen microbiota and their function. Anim Feed Sci Technol. 2022;294:115476.

    Article  CAS  Google Scholar 

  12. Ambriz-Vilchis V, Jessop NS, Fawcett RH, et al. Effect of yeast supplementation on performance, rumination time, and rumen pH of dairy cows in commercial farm environments. J Dairy Sci. 2017;100(7):5449–61.

    Article  CAS  PubMed  Google Scholar 

  13. Thrune M, Bach A, Ruiz-Moreno M, Stern MD, Linn JG. Effects of Saccharomyces cerevisiae on ruminal pH and microbial fermentation in dairy cows: yeast supplementation on rumen fermentation. Livest Sci. 2009;124:261–5.

    Article  Google Scholar 

  14. de Poppi AC, Lazzari G, Gomes ALM, et al. Effects of feeding a live yeast on rumen fermentation and fiber degradability of tropical and subtropical forages. J Sci Food Agric. 2021;101(15):6220–27.

    Article  PubMed  Google Scholar 

  15. Cattaneo L, Lopreiato V, Piccioli-Cappelli F, Trevisi E, Minuti A. Effect of supplementing live Saccharomyces cerevisiae yeast on performance, rumen function, and metabolism during the transition period in Holstein dairy cows. J Dairy Sci. 2023;106(6):4353–65.

    Article  CAS  PubMed  Google Scholar 

  16. Jiang Y, Ogunade IM, Qi S, Hackmann TJ, Staples CR, Adesogan AT. Effects of the dose and viability of Saccharomyces cerevisiae. 1. Diversity of ruminal microbes as analyzed by Illumina MiSeq sequencing and quantitative PCR. J Dairy Sci. 2017;100(1):325–42.

    Article  CAS  PubMed  Google Scholar 

  17. AlZahal O, Dionissopoulos L, Laarman AH, Walker N, McBride BW. Active dry Saccharomyces cerevisiae can alleviate the effect of subacute ruminal acidosis in lactating dairy cows. J Dairy Sci. 2014;97(12):7751–63.

    Article  CAS  PubMed  Google Scholar 

  18. Castillo GA, Burrola BM, Domínguez VJ, Chávez MA. Rumen microorganisms and fermentation. Arch Med Vet. 2014;46:349–61.

    Article  Google Scholar 

  19. Bach A, López-García A, González-Recio O, et al. Changes in the rumen and colon microbiota and effects of live yeast dietary supplementation during the transition from the dry period to lactation of dairy cows. J Dairy Sci. 2019;102(7):6180–98.

    Article  CAS  PubMed  Google Scholar 

  20. Fonty G, Chaucheyras-Durand F. Effects and modes of action of live yeasts in the rumen. Biologia. 2006;61:741–50.

    Article  CAS  Google Scholar 

  21. Alshaikh MA, Alsiadi MY, Zahran SM, Mogawer HH, Aalshowime TA. Effect of feeding yeast culture from different sources on the performance of lactating Holstein cows in Saudi Arabia. Asian Austral J Anim. 2022;15:352–56.

    Article  Google Scholar 

  22. Dehghan-banadaky M, Ebrahimi M, Motameny R, Heidari S. Effects of live yeast supplementation on mid-lactation dairy cows performances, milk composition, rumen digestion and plasma metabolites during hot season. J Appl Anim Res. 2013;41:137–42.

    Article  CAS  Google Scholar 

  23. Kumprechtová D, Illek J, Julien C, Homolka P, Jančík F, Auclair E. Effect of live yeast (Saccharomyces cerevisiae) supplementation on rumen fermentation and metabolic profile of dairy cows in early lactation. J Anim Physiol Anim Nutr (Berl). 2019;103(2):447–55.

    Article  PubMed  Google Scholar 

  24. Amin AB, Mao S. Influence of yeast on rumen fermentation, growth performance and quality of products in ruminants: a review. Anim Nutr. 2021;7(1):31–41.

    Article  CAS  PubMed  Google Scholar 

  25. Perdomo MC, Marsola RS, Favoreto MG, Adesogan A, Staples CR, Santos JEP. Effects of feeding live yeast at 2 dosages on performance and feeding behavior of dairy cows under heat stress. J Dairy Sci. 2020;103(1):325–39.

    Article  CAS  PubMed  Google Scholar 

  26. Clemmons BA, Martino C, Powers JB, et al. Rumen bacteria and serum metabolites predictive of feed efficiency phenotypes in beef cattle. Sci Rep. 2019;9(1):19265.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Sun P, Wang JQ, Deng LF. Effects of Bacillus subtilis natto on milk production, rumen fermentation and ruminal microbiome of dairy cows. Animal. 2013;7(2):216–22.

    Article  CAS  PubMed  Google Scholar 

  28. Golder HM, Denman SE, McSweeney C, et al. Effects of partial mixed rations and supplement amounts on milk production and composition, ruminal fermentation, bacterial communities, and ruminal acidosis. J Dairy Sci. 2014;97(9):5763–85.

    Article  CAS  PubMed  Google Scholar 

  29. Wang X, Zeng H, Xu J, et al. Characteristics of ruminal microbiota and metabolome in Holstein cows differing in milk protein concentrations. J Anim Sci. 2022;100(11):skac253.

    Article  PubMed  PubMed Central  Google Scholar 

  30. Liu K, Zhang Y, Huang G, Zheng N, Zhao S, Wang J. Ruminal bacterial community is associated with the variations of total milk solid content in Holstein lactating cows. Anim Nutr. 2022;9:175–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Lucey JA, Otter D, Horne DS. A 100-Year review: progress on the chemistry of milk and its components. J Dairy Sci. 2017;100(12):9916–32.

    Article  CAS  PubMed  Google Scholar 

  32. Côrtes C, da Silva-Kazama DC, Kazama R, et al. Milk composition, milk fatty acid profile, digestion, and ruminal fermentation in dairy cows fed whole flaxseed and calcium salts of flaxseed oil. J Dairy Sci. 2010;93(7):3146–57.

    Article  PubMed  Google Scholar 

  33. Graf J. The family rikenellaceae. Berlin: Springer; 2014. pp. 857–9.

    Google Scholar 

  34. Ren H, Su X, Bai H, et al. Specific enrichment of microbes and increased ruminal propionate production: the potential mechanism underlying the high energy efficiency of Holstein heifers fed steam-flaked corn. AMB Express. 2019;9(1):209.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Xue MY, Sun HZ, Wu XH, Liu JX, Guan LL. Multi-omics reveals that the rumen microbiome and its metabolome together with the host metabolome contribute to individualized dairy cow performance. Microbiome. 2020;8(1):64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Chuang ST, Li KY, Tu PW, et al. Investigating the reciprocal interrelationships among the ruminal microbiota, metabolome, and mastitis in early lactating holstein dairy cows. Anim (Basel). 2021;11(11):3108.

    Google Scholar 

  37. Xue MY, Sun HZ, Wu XH, Guan LL, Liu JX. Assessment of rumen bacteria in dairy cows with varied milk protein yield. J Dairy Sci. 2019;102(6):5031–41.

    Article  CAS  PubMed  Google Scholar 

  38. Mônica MS, Ahles S, Verhoeven J, Verbruggen S, Stevens Y, Venema K. A citrus fruit extract high in polyphenols beneficially modulates the gut microbiota of healthy human volunteers in a validated in vitro model of the colon. Nutrients. 2021;13(11):3915.

    Article  Google Scholar 

  39. Escalas A, Auguet JC, Avouac A, et al. Shift and homogenization of gut microbiome during invasion in marine fishes. Anim Microbiome. 2022;4(1):37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Bui AT, Williams BA, Hoedt EC, Morrison M, Mikkelsen D, Gidley MJ. High amylose wheat starch structures display unique fermentability characteristics, microbial community shifts and enzyme degradation profiles. Food Funct. 2020;11(6):5635–46.

    Article  CAS  PubMed  Google Scholar 

  41. AlZahal O, Li F, Guan LL, Walker ND, McBride BW. Factors influencing ruminal bacterial community diversity and composition and microbial fibrolytic enzyme abundance in lactating dairy cows with a focus on the role of active dry yeast. J Dairy Sci. 2017;100(6):4377–93.

    Article  CAS  PubMed  Google Scholar 

  42. Pinloche E, McEwan N, Marden JP, Bayourthe C, Auclair E, Newbold CJ. The effects of a probiotic yeast on the bacterial diversity and population structure in the rumen of cattle. PLoS ONE. 2013;8(7):e67824.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Cox MS, Deblois CL, Suen G. Assessing the response of ruminal bacterial and fungal microbiota to whole-rumen contents exchange in dairy cows. Front Microbiol. 2021;12:665776.

    Article  PubMed  PubMed Central  Google Scholar 

  44. Mu Y, Lin X, Wang Z, Hou Q, Wang Y, Hu Z. High-production dairy cattle exhibit different rumen and fecal bacterial community and rumen metabolite profile than low-production cattle. Microbiologyopen. 2019;8(4):e00673.

    Article  PubMed  Google Scholar 

  45. Li Z, Fan Y, Bai H, Zhang J, Mao S, Jin W. Live yeast supplementation altered the bacterial community’s composition and function in rumen and hindgut and alleviated the detrimental effects of heat stress on dairy cows. J Anim Sci. 2023;101:skac410.

    Article  PubMed  Google Scholar 

  46. Sandri M, Manfrin C, Pallavicini A, Stefanon B. Microbial biodiversity of the liquid fraction of rumen content from lactating cows. Animal. 2014;8(4):572–9.

    Article  CAS  PubMed  Google Scholar 

  47. Seegers H, Fourichon C, Beaudeau F. Production effects related to mastitis and mastitis economics in dairy cattle herds. Vet Res. 2003;34(5):475–91.

    Article  PubMed  Google Scholar 

  48. Halasa T, Huijps K, Østerås O, Hogeveen H. Economic effects of bovine mastitis and mastitis management: a review. Vet Q. 2007;29(1):18–31.

    Article  CAS  PubMed  Google Scholar 

  49. Du D, Feng L, Chen P, Jiang W, Zhang Y, Liu W, Zhai R, Hu Z. Effects of Saccharomyces Cerevisiae cultures on performance and immune performance of dairy cows during heat stress. Front Vet Sci. 2022;9:851184.

    Article  PubMed  PubMed Central  Google Scholar 

  50. Asil AK, Mohammadabadi T, Chaji M, Direkvandi E. Evaluating the effects of direct-fed microbial supplementation on the performance, milk quality and fatty acid of mid-lactating dairy cows. Vet Med Sci. 2023;00:1–9.

    Google Scholar 

  51. Tun HM, Li S, Yoon I, et al. Saccharomyces cerevisiae fermentation products (SCFP) stabilize the ruminal microbiota of lactating dairy cows during periods of a depressed rumen pH. BMC Vet Res. 2020;16(1):237.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Kogan G, Kocher A. Role of yeast cell wall polysaccharides in pig nutrition and health protection. Livest Sci. 2007;109:161–5.

    Article  Google Scholar 

  53. Beutler B. Innate immunity: an overview. Mol Immunol. 2004;40(12):845–59.

    Article  CAS  PubMed  Google Scholar 

  54. Bohn JA. James N. BeMiller. (1→3)-β-d-Glucans as biological response modifiers: a review of structure-functional activity relationships. Carbohyd Polym. 1995;28:3–14.

    Article  CAS  Google Scholar 

  55. Williams DL, Mueller A, Browder W. Glucan-based macrophage stimulators. Clin Immunother. 1996;5:392–9.

    Article  Google Scholar 

  56. Burdick Sanchez NC, Broadway PR, Carroll JA. Influence of yeast products on modulating metabolism and immunity in cattle and swine. Anim (Basel). 2021;11(2):371.

    Google Scholar 

  57. Wang X, Li L, Shang H, et al. Effects of duck circovirus on immune function and secondary infection of Avian Pathogenic Escherichia coli. Poult Sci. 2022;101(5):101799.

    Article  PubMed  PubMed Central  Google Scholar 

  58. Mulakala BK, Smith KM, Snider MA, Ayers A, Honan MC, Greenwood SL. Use of milk proteins as biomarkers of changes in the rumen metaproteome of Holstein cows fed low fiber, high starch diets. J Dairy Sci. 2023:S0022-0302(23)00237-0.

  59. Monteiro HF, Faciola AP. Ruminal acidosis, bacterial changes, and lipopolysaccharides. J Anim Sci. 2020;98(8):skaa248.

    Article  PubMed  PubMed Central  Google Scholar 

  60. Edgar RC. UPARSE: highly accurate OTU sequences from microbial amplicon reads. Nat Methods. 2013;10:996–8.

    Article  CAS  PubMed  Google Scholar 

  61. Caporaso JG, Bittinger K, Bushman FD, DeSantis TZ, Andersen GL, Knight R. PyNAST: a flexible tool for aligning sequences to a template alignment. Bioinformatics. 2010;26(2):266–7.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank Majorbio Biotechnology (Shanghai, China) provided the technical support in 16s rRNA sequencing analysis.

Funding

This work was supported by grants from the Central Public-interest Scientific Institution Basal Research Fund (1610332022002), the Natural Science Foundation of Inner Mongolia (2021LHBS03008 and 2022LHMS03005), and the National Center of Technology Innovation for Dairy (2022 Open Project-2). The funders had no role in study design, data collection and analysis, decision to publish, or the manuscript preparation.

Author information

Authors and Affiliations

Authors

Contributions

QZ, JZ, and JW developed and framed the research questions. QZ, JZ, and JW analyzed RNA sequencing data. LM, HJ, and YG contributed to serum biochemical analysis and subsequent data analysis. XZ and TN contributed to revise the entire manuscript. QZ, JZ, and JW drafted the manuscript. All authors have read and approved the manuscript.

Corresponding authors

Correspondence to Jize Zhang or Jianlong Wang.

Ethics declarations

Ethics approval and consent to participate

This study was conducted in strict accordance with the Regulations for the Administration of Affairs Concerning Experimental Animals of the State Council of the People’s Republic of China. The protocol was approved by the Committee on Experimental Animal Management of the Chinese Academy of Agricultural Sciences as the official ethical committee of the affiliation. Participating farm was informed of the present study and consented to participate the entire animal experiment.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary Material 1

Supplementary Material 2

Additional file 1, Figure S1 Comparison of ruminal bacteria in dairy cows fed basal (CON) or live yeast (YE) diet. (A) Relative abundances of bacterial communities at the phylum level. (B) Relative abundances of bacterial communities at the family level. (C) Relative abundances of bacterial communities at the genus level.

Supplementary Material 3

Additional file 2, Figure S2 LDA value distribution histogram. LDA value > 4, and the length of the bar chart represents the influence of different species.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhang, Q., Ma, L., Zhang, X. et al. Feeding live yeast (Saccharomyces cerevisiae) improved performance of mid-lactation dairy cows by altering ruminal bacterial communities and functions of serum antioxidation and immune responses. BMC Vet Res 20, 245 (2024). https://doi.org/10.1186/s12917-024-04073-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12917-024-04073-0

Keywords