ArticlePDF Available

Identification of the core bacteria in rectums of diarrheic and non-diarrheic piglets

Authors:

Abstract and Figures

Porcine diarrhea is a global problem that leads to large economic losses of the porcine industry. There are numerous factors related to piglet diarrhea, and compelling evidence suggests that gut microbiota is vital to host health. However, the key bacterial differences between non-diarrheic and diarrheic piglets are not well understood. In the present study, a total of 85 commercial piglets at three pig farms in Sichuan Province and Chongqing Municipality, China were investigated. To accomplish this, anal swab samples were collected from piglets during the lactation (0–19 days old in this study), weaning (20–21 days old), and post-weaning periods (22–40 days), and fecal microbiota were assessed by 16S rRNA gene V4 region sequencing using the Illumina Miseq platform. We found age-related biomarker microbes in the fecal microbiota of diarrheic piglets. Specifically, the family Enterobacteriaceae was a biomarker of diarrheic piglets during lactation (cluster A, 7–12 days old), whereas the Bacteroidales family S24–7 group was found to be a biomarker of diarrheic pigs during weaning (cluster B, 20–21 days old). Co-correlation network analysis revealed that the genus Escherichia-Shigella was the core component of diarrheic microbiota, while the genus Prevotellacea UCG-003 was the key bacterium in non-diarrheic microbiota of piglets in Southwest China. Furthermore, changes in bacterial metabolic function between diarrheic piglets and non-diarrheic piglets were estimated by PICRUSt analysis, which revealed that the dominant functions of fecal microbes were membrane transport, carbohydrate metabolism, amino acid metabolism, and energy metabolism. Remarkably, genes related to transporters, DNA repair and recombination proteins, purine metabolism, ribosome, secretion systems, transcription factors, and pyrimidine metabolism were decreased in diarrheic piglets, but no significant biomarkers were found between groups using LEfSe analysis.
This content is subject to copyright. Terms and conditions apply.
1
SCIENTIFIC REPORTS | (2019) 9:18675 | https://doi.org/10.1038/s41598-019-55328-y
www.nature.com/scientificreports
Identication of the core bacteria
in rectums of diarrheic and non-
diarrheic piglets
Jing Sun1,3,4,5*, Lei Du1,5, XiaoLei Li2,5, Hang Zhong1, Yuchun Ding1,3,4, Zuohua Liu1,3,4 &
Liangpeng Ge1,3,4*
Porcine diarrhea is a global problem that leads to large economic losses of the porcine industry. There
are numerous factors related to piglet diarrhea, and compelling evidence suggests that gut microbiota
is vital to host health. However, the key bacterial dierences between non-diarrheic and diarrheic
piglets are not well understood. In the present study, a total of 85 commercial piglets at three pig farms
in Sichuan Province and Chongqing Municipality, China were investigated. To accomplish this, anal
swab samples were collected from piglets during the lactation (0–19 days old in this study), weaning
(20–21 days old), and post-weaning periods (22–40 days), and fecal microbiota were assessed by 16S
rRNA gene V4 region sequencing using the Illumina Miseq platform. We found age-related biomarker
microbes in the fecal microbiota of diarrheic piglets. Specically, the family Enterobacteriaceae was
a biomarker of diarrheic piglets during lactation (cluster A, 7–12 days old), whereas the Bacteroidales
family S24–7 group was found to be a biomarker of diarrheic pigs during weaning (cluster B, 20–21
days old). Co-correlation network analysis revealed that the genus Escherichia-Shigella was the core
component of diarrheic microbiota, while the genus Prevotellacea UCG-003 was the key bacterium in
non-diarrheic microbiota of piglets in Southwest China. Furthermore, changes in bacterial metabolic
function between diarrheic piglets and non-diarrheic piglets were estimated by PICRUSt analysis,
which revealed that the dominant functions of fecal microbes were membrane transport, carbohydrate
metabolism, amino acid metabolism, and energy metabolism. Remarkably, genes related to
transporters, DNA repair and recombination proteins, purine metabolism, ribosome, secretion systems,
transcription factors, and pyrimidine metabolism were decreased in diarrheic piglets, but no signicant
biomarkers were found between groups using LEfSe analysis.
Diarrhea of neonatal piglets has long been a problem aicting global piglets production. During the last few dec-
ades, reports have described diarrhea in neonatal pigs belonging to various age groups13. Porcine diarrhea leads
directly to economic losses because of increased morbidity and mortality, reduced average daily gain (ADG),
and the consumption of extra medication4,5. Intestinal microbes have a profound impact on health and disease
through programming of immune and metabolic pathways6. Diarrhea has various causes, including porcine
parvovirus, porcine kobuvirus, and enterotoxigenic Escherichia coli (ETEC)710, all of which have been linked
to imbalances of normal intestinal ora as well as extra-intestinal microecological imbalance1113. A number of
recent studies have utilized high-throughput sequencing of the 16S rRNA gene to characterize gut microbiota of
diarrheic piglets. Neonatal pigletdiarrheawas associated with increases in the relative abundance of Prevotella
(Bacteroidetes), Sutterella and Campylobacter (Proteobacteria)14. e percentage of Enterococcus (Firmicutes) was
also more abundant in new neonatal porcine diarrhea (NNPD) aectedpiglets13. Genus Veillonella (Firmicutes)
was the dominant bacteria in fecal microbiota in porcine epidemic diarrhea virus (PEDV)-infected piglets
during the suckling transition stage15, while higher Escherichia-Shigella (Proteobacteria) in the feces was in
Enterotoxigenic Escherichia coli-induced diarrhea in piglets16. Although Holman and the colleagues used a
meta-analysis to dene a “core” microbiota in the swine gut17, the key microbial populations related to diarrhea
1Chongqing Academy of Animal Sciences, Chongqing, 402460, China. 2Institute of Animal Genetics and Breeding,
Sichuan Agricultural University, Chengdu, 611130, China. 3Key Laboratory of Pig Industry Sciences, Ministry of
Agriculture, Chongqing, 402460, China. 4Chongqing Key Laboratory of Pig Industry Sciences, Chongqing, 402460,
China. 5These authors contributed equally: Jing Sun, Lei Du and XiaoLei Li. *email: sunjing85026@163.com;
geliangpeng1982@163.com
OPEN
Content courtesy of Springer Nature, terms of use apply. Rights reserved
2
SCIENTIFIC REPORTS | (2019) 9:18675 | https://doi.org/10.1038/s41598-019-55328-y
www.nature.com/scientificreports
www.nature.com/scientificreports/
in piglets being poorly understood. us, we conducted a survey of porcine diarrhea in three medium-scale pig
farms in Southwest China to investigate the eects of diarrhea on fecal microbiota. e cause of diarrhea was
not considered when sampling, and a total of 52 and 33 swab samples were collected from diarrheic piglets and
non-diarrheic piglets, respectively, of the same or similar age in the same hog house for 16S ribosomal RNA gene
V4 region sequencing using the Illumina Miseq platform. We then compared and analyzed bacterial changes
in the composition and function of the feces of piglets that were suering from diarrhea and those that did not
develop diarrhea to identify key dierences in the fecal microbiota of piglets to reveal diarrhea-related bacteria.
Results
Overall information regarding the fecal microbiota of piglets. No signicant dierences in gen-
der or sample location were discerned between diarrheic and non-diarrheic groups (P > 0.05; Table1). Illumina
Miseq sequencing of the V4 region of the bacterial 16S rRNA gene generated 6,868,150 high-quality sequences.
Aer removal of chimeras, ltered high-quality sequences were grouped into 75,943 OTUs based on 97% species
similarity (detail information of OTUs was shown in Supplementary Table1).
Pairwise comparisons between groups were detected and values at P = 0.001, representing the grouping
(D group and ND group), were valid. e four most abundant phyla in the fecal microbiota of diarrheic and
non-diarrheic piglets were Firmicutes, Proteobacteria, Bacteroidetes, and Fusobacteria (Table2). Firmicutes and
Bacteroidetes constituted the top two phyla in the piglet gut microbiota in the ND group, whereas Firmicutes and
Proteobacteria constituted the two predominant phyla in the gut microbiota of diarrheic piglets (D group). A sim-
ilar abundance of Firmicutes was shown in the gut microbiota of piglets in groups D and ND (42.06% vs. 43.09%,
P > 0.05). Diarrheic piglets showed a signicantly lower percentage of Bacteroidetes and a higher percentage of
Proteobacteria than non-diarrheic individuals (P < 0.05). Moreover, the Proteobacteria-Bacteroidetes ratio in the
diarrheic group was 1.96, whereas the ratio in the non-diarrheic group was 0.36 (on average, Table2).
e OTUs were also used to compare the dierences in abundance between D and ND piglets (Table3). e
total abundance of 2 families, 11 genera, and 8 species diered signicantly in the gut microbiota of D and ND
piglets. For example, levels of the genera Bacteroides, Ruminococcaceae, and Prevotella in the fecal microbiota of
diarrheic piglets were signicantly lower than those in non-diarrheic piglets (P < 0.05). Diarrheic piglets also
contained a signicantly higher percentage of several species in the phylum Proteobacteria, including Pasteurella
aerogenes, Enterococcus cecorum, Enterococcus durans, and Escherichia coli (P < 0.05).
Major microbial dierences in dierent stages of piglet diarrhea. e experimental piglets used in
the present study were early-weaned at 21 days of age. To evaluate overall dierences in beta-diversity, we used
principal coordinate analysis (PCoA) to identify discrepancies between groups. As shown in Fig.1A, four distinct
clusters were evident (Clusters A–D). e fecal microbiota of the ND group was distinct from that of group D, and
the fecal microbiota of diarrheic piglets was distinct from the feeding phases. Specically, the gut microbiota of 14
piglets (ranging in age from 7–12 days old) was gathered in cluster A, and these piglets were still in their lactation
Information
Group
P valueDiarrheic (D) Non-diarrheic (ND)
Gender 29 (Female), 23 (Male) 24 (Female), 9 (Male) 0.12
Sampling location 16 (Sichuan), 36 (Chongqing) 15 (Sichuan), 18 (Chongqing) 0.22
Number of samples 52 33
Average age 15 days-old 23 days-old
Clean reads 76,206.58 ± 14,461.35 79,481.82 ± 12,609.14
OTU 879.88 ± 343.35 914.82 ± 368.90
Table 1. Overall microbiological and gene sequencing information regarding stool samples in this study.
Group Firmicutes Proteobacteria Bacteroidetes Fusobacteria
D group 42.06 ± 18.37 32.78 ± 28.21 16.75 ± 17.75 6.31 ± 8.24
ND group 43.09 ± 10.42 11.20 ± 9.69 31.53 ± 8.39 6.64 ± 5.28
P value 0.74 0.00 0.00 0.82
Cluster1
Cluster A 40.24 ± 20.03 56.68 ± 19.54 1.24 ± 0.81 1.38 ± 3.46
Cluster B 52.37 ± 12.93 5.49 ± 2.21 32.44 ± 11.72 7.74 ± 4.85
Cluster C 43.08 ± 11.19 12.27 ± 10.91 30.92 ± 8.50 5.84 ± 5.10
Cluster D 43.12 ± 8.19 7.86 ± 2.08 33.44 ± 8.30 9.12 ± 5.43
Table 2. Percentage of the top four phyla in the gut microbiota of piglets in the diarrheic group (D group) and
the non-diarrheic group (ND group). 1e experimental piglets used in the present study were early-weaned at
21 days of age. To evaluate overall dierences in beta-diversity, we used principal coordinate analysis (PCoA) to
identify discrepancies between groups. As shown in Fig.1A, four distinct clusters were evident (Clusters A–D),
and the relative abundance of the top four phyla in piglet microbiota were calculated.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
3
SCIENTIFIC REPORTS | (2019) 9:18675 | https://doi.org/10.1038/s41598-019-55328-y
www.nature.com/scientificreports
www.nature.com/scientificreports/
period. Cluster B contained the gut microbiota of 23 piglets (ranging in age from 20–21 days old) that were in the
early weaning period. Cluster A was clearly dierentiated from cluster B (Fig.1A). Moreover, 52.17% of samples
in cluster C were from piglets in the post-weaning period (average age = 33 days), and the gut microbiota of the
D and ND piglets were indistinguishable, suggesting that the beta-diversity of their gut microbiota tended to be
more similar across groups with age.
We used LEfSe analysis to identify biomarkers of fecal microbiota of diarrheic piglets (Fig.1B) and found
that the family Enterobacteriaceae was a biomarker of diarrheic piglets in cluster A (7–12 days old), whereas
the Bacteroidales family S24–7 was found to be a biomarker of diarrheic pigs in cluster B (20–21 days old). e
Wilcoxon-rank-sum test was used to identify bacterial genera with signicant dierences in relative abun-
dance in the fecal microbiota diarrheic piglets between clusters A and B. As shown in Fig.1C,D, the genus
Escherichia-Shigella in the family Enterobacteriacae was most abundant in cluster A, whereas the uncultured
genus in the Bacteroidales family S24–7 was the biomarker for cluster B.
Core bacterial genera by co-occurrence network analysis. To identify the potential interactions that
occur in response to diarrhea, co-correlative network analysis of the top 20 taxa was conducted for diarrheic and
non-diarrheic piglets based on Spearmans correlation coecient (Fig.2). Interestingly, we found that the genus
Escherichia-Shigella was the core node in diarrheic samples, and that it tended to be positively correlated with
aerobes and facultative anaerobes, such as the genera Actinobacillus, Pasteurella, Enterococcus, and Lactobacillus;
however, it was negatively correlated with anaerobes, including the genera Fusobacterium, Eubacterium copros-
tanoligenes group, Prevotella 2, Prevotella 9, Lachnospira, Rumniococcaceae NK4A214 group, Rikenellaceae RC9
gut group, and Alloprevotella (Fig.2A). e genus Prevotellaceae UCG-003 was the core node in non-diarrheic
piglets, and only positive correlations were found between Prevotellaceae UCG-003 and anaerobes and facultative
anaerobes, including the genera Pasteurella, Prevotella, Phascolarctobacterium, Ruminococcaceae UCG-002, and
Rikenellaceae RC9 gut group (Fig.2C). Among these marker genera, diarrheic samples comprised a signicantly
higher percentage of Escherichia-Shigella (22.92% vs.5.73%, P < 0.05), whereas non-diarrheic piglets contained a
higher percentage of Prevotella (4.50% vs. 1.44%, P < 0.05) (Fig.2B,D). e dierent core genera and the transi-
tion from negative correlations in diarrheic samples to positive correlations in non-diarrheic samples appeared to
indicate that there was a correlation between bacterial competition for oxygen and the intestinal health of piglets.
We also found that members of the phylum Proteobacteria were reduced from four genera (Escherichia-Shigella,
Actinobacillus, Pasteurella, and Sutterella) in the diarrheic group to only one genus (Pasteurella) in the
non-diarrheic group, suggesting that an increase in the abundance and diversity of the phylum Proteobacteria
played a pivotal role in piglet diarrhea.
Figure 1. Comparison of fecal microbiota between diarrheic and non-diarrheic piglets. (A) Principal
coordinate analysis (PCoA) shows the fecal microbiota of diarrheic (D) and non-diarrheic (ND) piglets. Red
triangles, ND; green dots, D. (B) Identication of bacterial biomarkers in the fecal microbiota of diarrheic
piglets in cluster A and cluster B using LEfSe analysis, and LDA scores >4.0. Comparison of the top four
bacterial phyla (C) and the top een bacteria genera (D) in the fecal microbiota of diarrheic piglets indierent
stages of development based on the Wilcoxon-rank-sum test are shown in the box plot (Cluster A: 7–12 day-
old piglets; Cluster B: 20–21-day-old piglets). Samples in cluster A are in red, samples in cluster B are in green;
*P < 0.05, **P < 0.01, ***P < 0.001.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
4
SCIENTIFIC REPORTS | (2019) 9:18675 | https://doi.org/10.1038/s41598-019-55328-y
www.nature.com/scientificreports
www.nature.com/scientificreports/
KEGG pathway analysis. To determine if enrichment of gut microbiota was associated with enrichment
of specic metabolic activity associated with piglet diarrhea, the functional contributions of the gut microbiota
were assessed using the PICRUSt tool. We found that KEGG pathways involved in membrane transport, carbo-
hydrate metabolism, amino acid metabolism, and DNA replication and repair were predominant in both groups
(Fig.3A). Overall, 38 pathways related to membrane transport at level 2 were obtained, and membrane transport,
carbohydrate metabolism, amino acid metabolism, and energy metabolism were major KEGG pathways in the
fecal microbiota of piglets in this study (Fig.3B). Interestingly, we also found that membrane transport was the
most abundant pathway in the fecal microbiota of diarrheic piglets during lactation (cluster A) and weaning
(cluster B) based on analysis of the functional contributions of the gut microbiota in cluster A and cluster B. We
used LEfSe analysis to identify biomarkers of the KEGG pathways that diered signicantly between diarrheic
and non-diarrheic microbiota, as well as the number of signicantly discriminative features with an LDA score
>4.0. Similarly, no dierentially abundant features of the KEGG pathways were found in the fecal microbiota of
diarrheic piglets between cluster A and B (LDA score > 4.0). ese ndings clearly indicated that the occurrence
of diarrhea in this study did not aect ecosystem processes of the fecal microbiota.
Moreover, we found that multiple KEGG (level 3) categories were disturbed when piglets had diarrhea. e
gut microbiota of diarrheal piglets were characterized by a reduced representation of proteins involved in metab-
olism of pyrimidine and purine, transporters of the ATP-binding cassette, secretion systems as well as DNA repair
and recombination (Table4).
Discussion
Our study investigated variations in the composition and function of fecal microbiota between diarrheic piglets
and non-diarrheic piglets. Consistent with the results of previous studies, Firmicutes was the dominant phylum
in the piglet gut microbiota1820, and there were no signicant dierences in relative abundance between groups
(P > 0.05). Proteobacteria constituted the second most common phylum in the gut microbiota of diarrheic piglets,
whereas Bacteroidetes was the second most abundant phylum in the fecal microbiota of non-diarrheic piglets
(Fig.1A and Table2). When compared with non-diarrheic piglets, the abundance of the phylum Proteobacteria
was signicantly higher in samples from diarrheic piglets, while that of the phylum Bacteroidetes decreased sig-
nicantly. Analysis of variations in bacterial genera between groups indicated that the genera Prevotella and
Ruminococcus, which are known to be ubiquitous in the fecal microbiota of piglets17, were signicantly lower
in diarrheic samples (Fig.1B and Table3). Moreover, opportunistic bacteria in the phylum Proteobacteria21,
including Escherichia coli22, Pasteurella aerogenes23, Enterococcus cecorum24,25, and Enterococcus durans2426, were
signicantly higher in fecal samples from diarrheic piglets.
Taxonomic name1
Average%
P value
Tendency in diarrheic piglets
compared with non-diarrheic
samplesD piglets ND piglets
Family
Clostridiales vadinBB60 group 0.514% 2.220% 0.003
Erysipelotrichaceae 0.780% 1.802% 0.018
Genus
Allisonella 0.994% 1.465% 0.033
Lactobacillus 1.674% 0.393% 0.013
Bacteroides 0.841% 1.705% 0.000
Ruminococcaceae NK4A214 group 0.776% 1.807% 0.009
Ruminococcaceae UCG-002 0.532% 2.193% 0.000
Ruminiclostridium 9 0.823% 1.726% 0.000
Anaerotruncus 0.637% 2.026% 0.000
Eubacterium coprostanoligenes group 0.659% 1.993% 0.007
Family XIII AD3011 group 0.750% 1.849% 0.000
Prevotella2 0.789% 1.787% 0.005
Prevotella9 0.849% 1.692% 0.015
Species
Lactobacillus salivarius 1.478% 0.708% 0.002
Lactobacillus vaginalis 1.702% 0.349% 0.001
Lactobacillus gasseri 0.445% 2.330% 0.020
Lactobacillus amylovorus 1.588% 0.529% 0.003
Pasteurella aerogenes 1.667% 0.404% 0.013
Enterococcus cecorum 1.685% 0.381% 0.010
Enterococcus durans 1.699% 0.353% 0.019
Escherichia coli 1.670% 0.399% 0.000
Table 3. Comparison of the relative abundance of OTUs in the gut microbiota of D and ND piglets. 1OTUs
for which the overall number in each sample was greater than 1000 and the number in half of the samples was
greater than 100 were used to compare dierences in abundances between D and ND piglets.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
5
SCIENTIFIC REPORTS | (2019) 9:18675 | https://doi.org/10.1038/s41598-019-55328-y
www.nature.com/scientificreports
www.nature.com/scientificreports/
In the present study, we ignored the cause of piglet diarrhea, and instead focused on dierences in the compo-
sition of fecal microbiota between groups. Surprisingly, beta-diversity analysis revealed that the fecal microbiota
of diarrheic piglets was also dierentiated by growth phases. Since piglets used in this study were early-weaned at
21 days of age, those aged less than 2 weeks were still in lactation. When combined with LEfSe analysis, the family
Enterobacteriaceae was identied as a biomarker in diarrhetic piglets during lactation (from 7–12 days old in this
study). An increase in Proteobacteria was previously reported as a marker for intestinal microbial community
dysbiosis and a potential diagnostic criterion for disease21. A wide variety of opportunistic pathogens that belong
to Proteobacteria are facultative anaerobes, and changes in the abundance of Proteobacteria might inuence oxy-
gen homeostasis or concentration in the gut27. Enrichment of Proteobacteria, such as Enterobacteriaceae, has also
been observed in response to imbalances in the intestinal community and changes in animal health28,29.
e abundance of Escherichia-Shigella has been reported to decrease sharply as piglets mature from the suck-
ling period to the weaning period19. Several species of Escherichia have been reported to be important to piglet
diarrhea and to have a severe impact on animal intestinal barrier function30,31. Interestingly, in this study, a signi-
cant increase in Escherichia-Shigella that belong to the family Enterobacteriaceae was shown in microbial commu-
nity of diarrheic piglets (Fig.1D), which was assigned as the core node in diarrheic piglets (Fig.2). Prevotellacecea
UCG-003 was identied as a key node in non-diarrheic piglets upon co-correlation network analysis, and dier-
ences in the core genus and the transition from negative correlations in diarrheic samples to positive correlations
in non-diarrheic samples indicate that there is a correlation between bacterial competition for oxygen and the
intestinal health of piglets.
In this study, the average abundance of the Bacteroidales family S24–7 and Escherichia-Shigella in diarrheic
piglets (D group) was 4.94% and 24.50%, whereas their average abundance in non-diarrheic piglets (ND group)
was 7.41% and 5.99%, respectively. is change in fecal microbiota reected the dierent causes of swine diarrhea
Figure 2. Co-correlation network analysis of bacterial genera constructed in diarrheic and non-diarrheic
piglets. Co-correlation networks were deduced from the top 20 genera identied upon16S rRNA sequencing.
Each node represents a genus, the size of each node is proportional to the relative abundance and the color of
the nodes indicates their taxonomic assignment. e width of the lines indicates the correlation magnitude,
while red represents a positive correlation and green a negative correlation. Only lines corresponding to
correlations with a magnitude greater than 0.5 are shown. Co-correlation network of (A) Escherichia-Shigella
genus in the diarrheic group (clustering = 0.82, closeness centrality = 0.86) and (C) Prevotellaceae UCG-003
genus in the non-diarrheic group (clustering = 0.30, closeness centrality = 0.34). Comparison of the relative
abundance of (B) the genus Escherichia-Shigella and (D) the genus Prevotellaceae UCG-003 between diarrheic
and non-diarrheic samples, which were visualized based on the means ± SEM. An independent t test was used
to identify dierences between groups. *P < 0.05; **P < 0.01; ***P < 0.001.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
6
SCIENTIFIC REPORTS | (2019) 9:18675 | https://doi.org/10.1038/s41598-019-55328-y
www.nature.com/scientificreports
www.nature.com/scientificreports/
in dierent stages aer birth. One important reason for piglet diarrhea in lactation in this study was the expan-
sion of swine enteric pathogens (e.g., Escherichia-Shigella). However, when grown, the average abundance of
Escherichia-Shigella in the gut microbiota of diarrheic piglets during weaning was only 1.80% (cluster B, shown
in Fig.1D), suggesting that these enteric pathogens were weakly correlated with diarrhea in weaning pigs in this
study. Abrupt changes in the diet and environment of piglets have been reported as the leading causes of weaning
diarrhea32,33. Interestingly, there was an enormous increase in members of the ber-degrader Bacteroidales family
S24–734,35 when piglets grew up (less than 1.00% in cluster A versus 20.04% in cluster B). However, very little work
regarding Bacteroidales family S24–7 has been conducted to date. In short, it is necessary to conduct ongoing
Figure 3. Comparison of variations in abundance of known KEGG pathways. e functional contributions
of the gut microbiota were assessed using the PICRUSt tool. (A,C) Pathways at level 1 were obtained; (B,D)
Pathways at level 2 were obtained. Diarrheic group (D group); non-diarrheic group (ND group); Cluster A: 7–12
day-old piglets; Cluster B: 20–21-day-old piglets.
Category Level 3 Diarrheic piglets
(D group) Non-diarrheic
piglets (ND group)
Transporters 736755 1199406
General function prediction only 391357 688770
ABC transporters 377209 595634
DNA repair and recombination
proteins 309864 562249
Purine metabolism 255530 443130
Ribosome 236991 481353
Two-component system 202776 279138
Secretion system 201378 242878
Transcription factors 199787 293331
Pyrimidine metabolism 194493 373833
Table 4. e 10 most abundant KEGG pathways (at level 3) in the fecal microbiota of non-diarrheic and
diarrheic piglets based on PICRUSt prediction.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
7
SCIENTIFIC REPORTS | (2019) 9:18675 | https://doi.org/10.1038/s41598-019-55328-y
www.nature.com/scientificreports
www.nature.com/scientificreports/
research regarding its biological function and usage. Nevertheless, the above results suggest that the focus of early
weaning syndrome in piglets should be shied from intestinal pathogens to moderate changes in diet and better
feeding and management.
In the present study, we also found a dysbiosis of intestinal microbiota in diarrheic samples, especially the
higher percentage of several Lactobacillus strains, which was consistent with the results of a previous study36.
e increased abundance of the GABA-producing Lactococcus lactis led to increased expression of IL-17 dur-
ing piglet ETEC infection37. In the present study, several Lactobacillus strains, including Lactobacillus salivar-
ius, Lactobacillus vaginalis, and Lactobacillus amylovorus, were higher in the diarrheic microbiota (Table3).
Lactobacillus salivarius is known for its ability to produce lactic acid. In addition to lactic acids, Lactobacillus sal-
ivarius also produced γ-aminobutyric acid38. Similar to Lactococcus lactis, we believe that this GABA-producing
strain may have increased GABA signaling to actively aect host health and disease states. Future studies should
be conducted to investigate this and explore the mechanisms responsible for the increased abundance of specic
Lactobacillus strain(s) during piglet diarrhea.
e ABC transporters are primary transporters that couple the energy stored in adenosine triphosphate (ATP)
to the movement of molecules across the membrane, which link with multi-drug resistance in both bacteria
and eukaryotes39. A general overview of the DNA damage response pathway in humans indicated that decient
DNA repair could aect genome stability, which could induce tumorigenesis40. In this study, PICRUSt prediction
revealed that the relative abundance of ABC transporters, DNA repair, and recombination proteins were down-
regulated in the fecal microbiota of diarrheic piglets, implying multi-drug resistance and DNA in swine cells
was damaged when diarrhea occurred. However, no dierentially abundant KEGG pathways were found in the
fecal microbiota of diarrheic and non-diarrheic piglets with a LDA score >4.0 (Fig.3). A reliable reason for why
changes in microbial composition did not aect their functional contributions is that the taxa in the microbial
community of diarrheic piglets were functionally redundant41 with the taxa in the community of non-diarrheic
piglets.
Conclusion
We revealed the main variations in the composition of fecal microbiota of diarrheic piglets and non-diarrheic
piglets. Proteobacteria was the second most abundant phylum in intestinal microbiota of diarrheic piglets. We
found that the fecal microbiota of diarrheic piglets was dierentiated by animal growth phases, and the family
Enterobacteriaceae was a biomarker in piglets during lactation, but the Bacteroidales family S24–7 group was a
biomarker in later stages of growth. In addition, Escherichia-Shigella was the core in diarrheic gut microbiota,
whereas Provteollaceace UCG-003 was the core in the fecal microbiota of non-diarrheic piglets.
Materials and Methods
Ethics statement. All animal experiments were conducted pursuant to the Regulations for the
Administration of Aairs Concerning Experimental Animals (Ministry of Science and Technology, Beijing,
China, revised June 2014). All guidelines related to the care of laboratory animals were followed. e institutional
ethics committee of the Chongqing Academy of Animal Sciences (Chongqing, China) reviewed the relevant eth-
ical issues and approved this study (permit number xky-20150113). Only fresh stool samples collected by rectal
swabs were analyzed, and no animals were killed or injured in this study. e preparation of total genomic DNA
was conducted at the Experimental Swine Engineering Center of the Chongqing Academy of Animal Sciences
(CMA No. 162221340234; Rongchang, Chongqing, China).
Sample collection. In the present study, piglets were early-weaned at 21 days of age. We collected a total
of 85 piglet fecal samples during January of 2016. Specically, 31 samples were collected from Shuangjia Farm
(Longchang County, Sichuan Province, China), 41 were obtained from Taoranju Farm (Rongchang District,
Chongqing, China), and 13 were obtained from Pengkang Farm (Yongchuan District, Chongqing, China).
Overall, 52 piglets had diarrhea (diarrhea group or D group), which was characterized by liquid, yellow-green
or taupe feces with a foul smell or stench that stuck around the anus. Thirty-three piglets had no diarrhea
(non-diarrhea group or ND group), as indicated by solid feces with no blood or mucus and no waste attached
around the anal area (non-diarrhea group or ND group).
About 0.5 g of freshly passed stool from the swab samples was transferred into a sterile Eppendorf tube
(Axygen Inc., Union City, CA, USA), aer which 10% glycerol (vol/vol) in sterile pre-reduced saline was added
to each tube. e samples were then homogenized and then immediately frozen at80 °C until needed for 16S
ribosomal RNA gene sequencing.
Sequencing and Analysis. 16S rRNA gene sequencing. Total genomic DNA was extracted from samples
using the CTAB/SDS method, aer which the 16S rRNA gene of the distinct 16S V4 region was amplied using
specic primers (515F–806 R) with a barcode. e microbial diversity and composition were then determined by
16S rRNA gene sequencing and analysis as previously described6.
LDA eect size (LEfSe). To identify the genomic features of taxa diering in abundance between two or more
biological conditions or classes, the LEfSe (Linear Discriminant Analysis Eect Size) algorithm was used with
the online interface Galaxy (http://huttenhower.sph.harvard.edu/lefse/)42. A size-eect threshold of 4.0 on the
logarithmic LDA score was used for discriminative functional biomarkers.
Co-correlation statistics. According to the calculation method developed by Hartmann et al.43, co-correlation
networks were generated using the python package NetworkX (https://github.com/networkx/networkx) and the
Content courtesy of Springer Nature, terms of use apply. Rights reserved
8
SCIENTIFIC REPORTS | (2019) 9:18675 | https://doi.org/10.1038/s41598-019-55328-y
www.nature.com/scientificreports
www.nature.com/scientificreports/
OTUs as target nodes, with edges (e.g., connecting nodes) representing signicant negative (green) or positive
(red) Spearman’s correlations. We retained OTUs when they had a Spearmans correlation coecient >0.5.
Predicted functionality of the differently grouped samples. Phylogenetic Investigation of Communities by
Reconstruction of Unobserved States (PICRUSt) (http://galaxy.morganlangille.com/)44 was used to predict the
functional gene content in the fecal microbiota based on taxonomy from the Greengenes reference database
(http://greengenes.lbl.gov/cgi-bin/nph-index.cgi). First, a collection of closed-reference OTUs was obtained from
the ltered reads using QIIME (v 1.7.0, http://qiime.org/scripts/split_libraries_fastq.html)45, and by querying the
data against a reference collection (Greengenes), aer which OTUs were assigned at 97% identity. e resulting
OTUs were then employed for microbial community metagenome prediction with PICRUSt using the online
Galaxy interface (http://huttenhower.sph.harvard.edu/galaxy/). Supervised analysis was conducted using LEfSe
to elicit the microbial functional pathways that were dierentially expressed among samples. PICRUSt was used
to derive relative Kyoto Encyclopedia of Genes and Genomes (KEGG) Pathway abundance.
Statistical analysis. Data proportions of sites and gender were regarded as categorical variables and com-
pared by the Chi-square test. Pairwise comparisons between groups were assessed by analysis of similarity
(ANOSIM). Values represent the pairwise test statistic (R) for ANOSIM. e permutation-based level of sig-
nicance was adjusted for multiple comparisons using the Benjamini-Hochberg false discovery rate (FDR) pro-
cedure. A P < 0.05 indicates the dierence between groups is greater than the dierence within the group. e
Wilcoxon-rank-sum test was used to detect the dierent populated bacterial genera between groups. e relative
abundances of bacterial taxa are presented as the means ± SD, and dierences between groups were identied by
the independent-sample t test (for normally distributed data) or the Mann-Whitney U-test (for non-normally
distributed data). A p-value <0.05 was considered statistically signicant, and a p-value <0.01 indicated extreme
signicance. e raw sequences obtained in the present study have been submitted to the NCBI Sequence Read
Archive (accession number SRP134239).
Received: 19 June 2019; Accepted: 26 November 2019;
Published: xx xx xxxx
References
1. Jung, ., Annamalai, T., Lu, Z. & Saif, L. J. Comparative pathogenesis of US porcine epidemic diarrhea virus (PEDV) strain PC21A
in conventional 9-day-old nursing piglets vs. 26-day-old weaned pigs. Veterinary microbiology 178, 31–40, https://doi.org/10.1016/j.
vetmic.2015.04.022 (2015).
2. omas, J. T. et al. Eect of Porcine Epidemic Diarrhea Virus Infectious Doses on Infection Outcomes in Naive Conventional
Neonatal and Weaned Pigs. PloS one 10, e0139266, https://doi.org/10.1371/journal.pone.0139266 (2015).
3. Dean, E. A., Whipp, S. C. & Moon, H. W. Age-specic colonization of porcine intestinal epithelium by 987P-piliated enterotoxigenic
Escherichia coli. Infection and immunit y 57, 82–87 (1989).
4. Panel, E. A. Scientic Opinion on porcine epidemic diarrhoea and emerging pig deltacoronavirus. EFSA J 12, 3877 (2014).
5. ongsted, H., Stege, H., To, N. & Nielsen, J. P. e eect of New Neonatal Porcine Diarrhoea Syndrome (NNPDS) on average daily
gain and mortality in 4 Danish pig herds. BMC veterinary research 10, 90, https://doi.org/10.1186/1746-6148-10-90 (2014).
6. Collado, M. C., autava, S., Aao, J., Isolauri, E. & Salminen, S. Human gut colonisation may be initiated in utero by distinct
microbial communities in the placenta and amniotic uid. Scientic reports 6, 23129, https://doi.org/10.1038/srep23129 (2016).
7. Jacova, A. et al. Porcine obuvirus 1 in healthy and diarrheic pigs: Genetic detection and characterization of virus and co-infection
with rotavirus A. Infection, genetics and evolution: journal of molecular epidemiology and evolutionary genetics in infectious diseases
49, 73–77, https://doi.org/10.1016/j.meegid.2017.01.011 (2017).
8. Opriessnig, T., Gerber, P. F., Matzinger, S. ., Meng, X. J. & Halbur, P. G. Maredly dierent immune responses and virus inetics in
littermates infected with porcine circovirus type 2 or porcine parvovirus type 1. Veterinary immunology and immunopathology 191,
51–59, https://doi.org/10.1016/j.vetimm.2017.08.003 (2017).
9. Huang, G. et al. Lysozyme improves gut performance and protects against enterotoxigenic Escherichia coli infection in neonatal
piglets. Vet. es. 49, 20 (2018).
10. De Filippo, C. et al. Impact of diet in shaping gut microbiota revealed by a comparative study in children from Europe and rural
Africa. Proceedings of the National Academy of Sciences of the United States of America 107, 14691–14696, https://doi.org/10.1073/
pnas.1005963107 (2010).
11. oh, H.-W., im, M. S., Lee, J.-S., im, H. & Par, S.-J. Changes in the Swine Gut Microbiota in esponse to Porcine Epidemic
Diarrhea Infection. Microbes and Environments 30, 284 (2015).
12. Gao, Y. et al. Changes in gut microbial populations, intestinal morphology, expression of tight junction proteins, and cytoine
production between two pig breeds aer challenge with 88: A comparative study. Journal of animal science 91, 5614–5625 (2013).
13. Hermann-Ban, M. L. et al. Characterization of the bacterial gut microbiota of piglets suffering from new neonatal porcine
diarrhoea. BMC veterinary research 11, 1 (2015).
14. Yang, Q. et al. Structure and Function of the Fecal Microbiota in Diarrheic Neonatal Piglets. Frontiers in microbiology 8, 502, https://
doi.org/10.3389/fmicb.2017.00502 (2017).
15. Huang, A. et al. Dynamic Change of Gut Microbiota During Porcine Epidemic Diarrhea Virus Infection in Sucling Piglets.
Frontiers in microbiology 10, 322, https://doi.org/10.3389/fmicb.2019.00322 (2019).
16. Bin, P. et al. Intestinal microbiota mediates Enterotoxigenic Escherichia coli-induced diarrhea in piglets. BMC veterinary research
14, 385, https://doi.org/10.1186/s12917-018-1704-9 (2018).
17. Holman, D. B., Brunelle, B. W., Trachsel, J. & Allen, H. . Meta-analysis To Dene a Core Microbiota in the Swine Gut. mSystems 2,
https://doi.org/10.1128/mSystems.00004-17 (2017).
18. Hu, J. et al. Gradual Changes of Gut Microbiota in Weaned Miniature Piglets. Frontiers in microbiology 7, 1727, https://doi.
org/10.3389/fmicb.2016.01727 (2016).
19. Chen, L. et al. e Maturing Development of Gut Microbiota in Commercial Piglets during the Weaning Transition. Frontiers in
microbiology 8, 1688, https://doi.org/10.3389/fmicb.2017.01688 (2017).
20. Loo, T. et al. In-feed antibiotic eects on the swine intestinal microbiome. Proceedings of the National Academy of Sciences of the
United States of America 109, 1691–1696, https://doi.org/10.1073/pnas.1120238109 (2012).
21. Shin, N. ., Whon, T. W. & Bae, J. W. Proteobacteria: microbial signature of dysbiosis in gut microbiota. Trends in biotechnology 33,
496–503, https://doi.org/10.1016/j.tibtech.2015.06.011 (2015).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
9
SCIENTIFIC REPORTS | (2019) 9:18675 | https://doi.org/10.1038/s41598-019-55328-y
www.nature.com/scientificreports
www.nature.com/scientificreports/
22. Curcio, L. et al. Detection of the colistin resistance gene mcr-1 in pathogenic Escherichia coli from pigs aected by post-weaning
diarrhoea in Italy. Journal of global antimicrobial resistance 10, 80–83, https://doi.org/10.1016/j.jgar.2017.03.014 (2017).
23. Confer, A. W. Immunogens of Pasteurella. Veterinary microbiology 37, 353–368 (1993).
24. Delaunay, E., Abat, C. & olain, J. M. Enterococcus cecorum human infection, France. New microbes and new infections 7, 50–51,
https://doi.org/10.1016/j.nmni.2015.06.004 (2015).
25. Jung, A., Metzner, M. & yll, M. Comparison of pathogenic and non-pathogenic Enterococcus cecorum strains from dierent
animal species. BMC microbiology 17, 33, https://doi.org/10.1186/s12866-017-0949-y (2017).
26. Agudelo Higuita, N. I. & Huyce, M. M. In Enterococci: From Commensals to Leading Causes of Drug esistant Infection (eds M. S.,
Gilmore, D. B., Clewell, Y. Ie, & N. Shanar) (2014).
27. Guaraldi, F. & Salvatori, G. Eect of breast and formula feeding on gut microbiota shaping in newborns. Frontiers in cellular and
infection microbiology 2, 94, https://doi.org/10.3389/fcimb.2012.00094 (2012).
28. Fei, N. & Zhao, L. An opportunistic pathogen isolated from the gut of an obese human causes obesity in germfree mice. e ISME
journal 7, 880–884, https://doi.org/10.1038/ismej.2012.153 (2013).
29. Morgan, X. C. et al. Dysfunction of the intestinal microbiome in inammatory bowel disease and treatment. Genome biology 13,
79, https://doi.org/10.1186/gb-2012-13-9-r79 (2012).
30. Xu, C., Li, C. Y. & ong, A. N. Induction of phase I, II and III drug metabolism/transport by xenobiotics. Archives of pharmacal
research 28, 249–268 (2005).
31. Yang, . M., Jiang, Z. Y., Zheng, C. T., Wang, L. & Yang, X. F. Eect of Lactobacillus plantarum on diarrhea and intestinal barrier
function of young piglets challenged with enterotoxigenic Escherichia coli 88. J Anim Sci 92, 1496–1503, https://doi.org/10.2527/
jas.2013-6619 (2014).
32. Gilbert, H. et al. esponses to weaning in two pig lines divergently selected for residual feed intae depending on diet. J Anim Sci 97,
43–54, https://doi.org/10.1093/jas/sy416 (2019).
33. Gresse, . et al. Gut Microbiota Dysbiosis in Postweaning Piglets: Understanding the eys to Health. Trends in microbiology 25,
851–873, https://doi.org/10.1016/j.tim.2017.05.004 (2017).
34. Ormerod, . L. et al. Genomic characterization of the uncultured Bacteroidales family S24-7 inhabiting the guts of homeothermic
animals. Microbiome 4, 36, https://doi.org/10.1186/s40168-016-0181-2 (2016).
35. Garcia-Mazcorro, J. F., Mills, D. A., Murphy, . & Noratto, G. Eect of barley supplementation on the fecal microbiota, caecal
biochemistry, and ey biomarers of obesity and inammation in obese db/db mice. European journal of nutrition 57, 2513–2528,
https://doi.org/10.1007/s00394-017-1523-y (2018).
36. Mach, N. et al. Early-life establishment of the swine gut microbiome and impact on host phenotypes. Environmental microbiology
reports 7, 554–569, https://doi.org/10.1111/1758-2229.12285 (2015).
37. en, W. et al. Intestinal Microbiota-Derived GABA Mediates Interleuin-17 Expression during Enterotoxigenic Escherichia coli
Infection. Frontiers in immunology 7, 685, https://doi.org/10.3389/mmu.2016.00685 (2016).
38. Chiu, T. H., Tsai, S. J., Wu, T. Y., Fu, S. C. & Hwang, Y. T. Improvement in antioxidant activity, angiotensin-converting enzyme
inhibitory activity and in vitro cellular properties of fermented pepino mil by Lactobacillus strains containing the glutamate
decarboxylase gene. Journal of the science of food and agriculture 93, 859–866, https://doi.org/10.1002/jsfa.5809 (2013).
39. Beis, . Structural basis for the mechanism of ABC transporters. Biochemical Society transactions 43, 889–893, https://doi.
org/10.1042/BST20150047 (2015).
40. Mateo, J. et al. DNA epair in Prostate Cancer: Biology and Clinical Implications. European urology 71, 417–425, https://doi.
org/10.1016/j.eururo.2016.08.037 (2017).
41. Allison, S. D. & Martiny, J. B. Colloquium paper: resistance, resilience, and redundancy in microbial communities. Proceedings of the
National Academy of Sciences of the United States of America 105(Suppl 1), 11512–11519, https://doi.org/10.1073/pnas.0801925105
(2008).
42. Segata, N. et al. Metagenomic biomarer discovery and explanation. Genome biology 12, 60, https://doi.org/10.1186/gb-2011-12-
6-r60 (2011).
43. Hartmann, M., Frey, B., Mayer, J., Mader, P. & Widmer, F. Distinct soil microbial diversity under long-term organic and conventional
farming. e ISME journal 9, 1177–1194, https://doi.org/10.1038/ismej.2014.210 (2015).
44. Langille, M. G. et al. Predictive functional proling of microbial communities using 16S rNA marer gene sequences. Nature
biotechnology 31, 814–821, https://doi.org/10.1038/nbt.2676 (2013).
45. Caporaso, J. G. et al. QIIME allows analysis of high-throughput community sequencing data. Nature methods 7, 335–336, https://
doi.org/10.1038/nmeth.f.303 (2010).
Acknowledgements
is work was funded by the National Key R&D Program of China (2017yfd0500501), Performance Incentive Guidance
for Scientic Research Institution of Chongqing Science & Technology Commission (cstc2019jxjl0035),Chongqing
Postdoctoral Research Special Funding Project (Xm2016031), and Chongqing Basic Scientic Research (grant no.
17408).
Author contributions
J.S., L.P.G. and Z.H.L. designed the experiments. J.S., L.D. and H.Z. analyzed the data and draed the manuscript.
X.L.L. and Y.C.D. collected the samples.
Competing interests
e authors declare no competing interests.
Additional information
Supplementary information is available for this paper at https://doi.org/10.1038/s41598-019-55328-y.
Correspondence and requests for materials should be addressed to J.S. or L.G.
Reprints and permissions information is available at www.nature.com/reprints.
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional aliations.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
10
SCIENTIFIC REPORTS | (2019) 9:18675 | https://doi.org/10.1038/s41598-019-55328-y
www.nature.com/scientificreports
www.nature.com/scientificreports/
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 Cre-
ative Commons license, and indicate if changes were made. e images or other third party material in this
article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons license and your intended use is not per-
mitted 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 license, visit http://creativecommons.org/licenses/by/4.0/.
© e Author(s) 2019
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... Moreover, from an analytical point of view, the amount of feces can be limiting in very young piglets and may not enable an additional dry matter analysis. Only few studies have been conducted that investigated early microbial colonization patterns in piglets in relation to fecal consistency, often making the main difference between diarrheic and non-diarrheic piglets [18]. Diarrhea has various causes, all of which are reflected in imbalances of the commensal gut microbiome [18]. ...
... Only few studies have been conducted that investigated early microbial colonization patterns in piglets in relation to fecal consistency, often making the main difference between diarrheic and non-diarrheic piglets [18]. Diarrhea has various causes, all of which are reflected in imbalances of the commensal gut microbiome [18]. A similar description is missing for feces of different consistency in non-diarrheic piglets. ...
... Compositional differences in the bacterial community of feces that vary in color and consistency were mainly reported for the "extreme conditions" in piglets with and without diarrhea [18,39]. Our results provide evidence that feces differing in color and consistency diverge in their total microbial load and taxonomic bacterial composition. ...
Article
Full-text available
Feces enable frequent samplings for the same animal, which is valuable in studies investigating the development of the gut microbiome in piglets. Creep feed should prepare the piglet’s gut for the postweaning period and shape the microbiome accordingly. Little is known about the variation that is caused by differences in fecal color and consistency and different sample types (feces versus swab samples). Therefore, this study evaluated the age-related alterations in the microbiome composition (16S rRNA gene) in feces of suckling and newly weaned piglets in the context of nutrition and fecal consistency, color and sample type from day 2 to 34 of life. Feces from 40 healthy piglets (2 each from 20 litters) were collected on days 2, 6, 13, 20, 27, 30 and 34. Weaning occurred on day 28. Half of the litters only drank sow milk during the suckling phase, whereas the other half had access to creep feed from day 10. Creep feeding during the suckling phase influenced the age-related total bacterial and archaeal abundances but had less of an influence on the relative bacterial composition. Results further showed different taxonomic compositions in feces of different consistency, color and sample type, emphasizing the need to consider these characteristics in comprehensive microbiome studies.
... The elevated abundance of Fusobacterota in the E. coli-infected group confirmed its association with pathogenic conditions and inflammation in the intestines. In contrast, recovered pigs once again demonstrated the dominance of Firmicutes and Bacteroidetes at the phylum level, indicating their active involvement in carbohydrate metabolism [36,37]. ...
Article
Full-text available
This study aimed to evaluate the disruption of the swine gut microbiota and histopathological changes caused by infection with enterotoxigenic E. coli. Fecal samples were collected from piglets suffering from diarrhea post-recovery and healthy animals. Intestinal tissues were collected for histopathological changes. The results revealed histopathological changes mainly in the ileum of the infected animals compared to those in the ileum of the control and recovered animals. The operational taxonomic units (OTUs) revealed that the E. coli diarrheal group exhibited the highest bacterial richness. Principal coordinate analysis (PCoA) corroborated the presence of dysbiosis in the gut microbiota following E. coli-induced diarrhea. While the normal control and infected groups displayed slight clustering, the recovery group formed a distinct cluster with a distinct flora. Bacteroidetes, Firmicutes, and Fusobacteria were the dominant phyla in both the healthy and recovered piglets and in the diarrheal group. LEfSe and the associated LDA score analysis revealed that the recovered group exhibited dominance of the phyla Euryarchaeota and Bacteroidota, while groups N and I showed dominance of the phyla Firmicutes and Fusobacteriota, respectively. The LDA scores highlighted a significant expression of the Muribaculacea family in group R. The obtained findings will help in understanding the microbiome during swine colibacillosis, which will support control of the outbreaks.
... An increase in Proteobacteria has also been observed in the ileum of NE infected birds (Xu et al., 2018). An increase in Proteobacteria has also been noted in other species with gastrointestinal disturbances (Sun et al., 2019), and is therefore, not specific to NE or coccidiosis, but could be considered a hallmark of dysbiosis. The phylum Firmicutes contains over 250 genera, including Clostridia and Bacilli. ...
Article
Full-text available
Necrotic enteritis (NE) is a disease of the gastrointestinal tract that is common in broiler chickens and causes substantial economic losses to the poultry industry worldwide. The removal of many antimicrobials in poultry diets has driven the search for alternatives. The purpose of this study was to determine the microbiota changes in the cecal luminal (CE-L) and mucosal (CE-M) populations of broiler chickens undergoing clinical NE (co-infected with Eimeria maxima and Clostridium perfringens) while fed a diet containing a flavonoid rich corn (PennHFD1) or control diet using commercial corns. It was previously shown that chickens fed a diet high in flavonoids had improved performance parameters, lower mortality rate, and lower incidence of intestinal lesions. Flavonoids have been shown to have anti-bacterial, immuno-modulatory, and anti-inflammatory activity. The current study included four experimental groups: infected chickens fed commercial corn diet (CTRL-A) or PennHFD1 (CTRL-B) and infected chickens fed commercial corn diet (IF-A) or PennHFD1 (IF-B). We found that most of the microbiota changes were due to infection rather than diet. The alpha diversity in the IF chickens was lower in both CE-L and CE-M. The beta diversity of microbial communities was different between IF and CTRL chickens, as well as between CTRL-A and CTRL-B. The beta diversity of CTRL birds was more homogenous compared to IF samples. Taxonomic analysis showed a decrease in short chain fatty acid producing bacteria in IF birds. An increase in lactic acid producing bacteria, Escherichia coli, and Enterococcus cecorum was also observed in IF birds. It is possible that the effect of the high flavonoid corn on the microbiota was overcome by the effect of NE, or that the positive effects of increased flavonoids in NE-challenged birds are a result of mechanisms which do not involve the microbiota. The effects of high flavonoid corn on NE infections may be further investigated as a possible alternative to antimicrobials.
... To more specifically assess whether ROM supplementation modulated gut microbiota development at genus level over time, the GAMLSS-BEZI model identified that ROM treated piglets had a significantly higher relative abundance of [Eubacterium]_coprostanoligenes_group, whereas Peptostreptococcus was significantly more abundant in Ctrl piglets during pre-weaning. The abundance of [Eubacterium]_coprostanoligenes_group has previously been reported to positively respond to dietary Bacillus subtilis supplementation for piglets around weaning [52], and be negatively associated with diarrhoea incidence of piglets [53]. In contrast, the genus Peptostreptococcus may include potential porcine intestinal pathogens and was associated with the level of Clostridioides difficile in suckling piglets [54]. ...
Article
Full-text available
Background Agaricus subrufescens is considered as one of the most important culinary-medicinal mushrooms around the world. It has been widely suggested to be used for the development of functional food ingredients to promote human health ascribed to the various properties (e.g., anti-inflammatory, antioxidant, and immunomodulatory activities). In this context, the interest in A. subrufescens based feed ingredients as alternatives for antibiotics has also been fuelled during an era of reduced/banned antibiotics use. This study aimed to investigate the effects of a fermented feed additive -rye overgrown with mycelium (ROM) of A. subrufescens—on pig intestinal microbiota, mucosal gene expression and local and systemic immunity during early life. Piglets received ROM or a tap water placebo (Ctrl) perorally every other day from day 2 after birth until 2 weeks post-weaning. Eight animals per treatment were euthanized and dissected on days 27, 44 and 70. Results The results showed ROM piglets had a lower inter-individual variation of faecal microbiota composition before weaning and a lower relative abundance of proteobacterial genera in jejunum (Undibacterium and Solobacterium) and caecum (Intestinibacter and Succinivibrionaceae_UCG_001) on day 70, as compared to Ctrl piglets. ROM supplementation also influenced gut mucosal gene expression in both ileum and caecum on day 44. In ileum, ROM pigs showed increased expression of TJP1/ZO1 but decreased expression of CLDN3, CLDN5 and MUC2 than Ctrl pigs. Genes involved in TLR signalling (e.g., TICAM2, IRAK4 and LY96) were more expressed but MYD88 and TOLLIP were less expressed in ROM pigs than Ctrl animals. NOS2 and HIF1A involved in redox signalling were either decreased or increased in ROM pigs, respectively. In caecum, differentially expressed genes between two groups were mainly shown as increased expression (e.g., MUC2, PDGFRB, TOLLIP, TNFAIP3 and MYD88) in ROM pigs. Moreover, ROM animals showed higher NK cell activation in blood and enhanced IL-10 production in ex vivo stimulated MLN cells before weaning. Conclusions Collectively, these results suggest that ROM supplementation in early life modulates gut microbiota and (local) immune system development. Consequently, ROM supplementation may contribute to improving health of pigs during the weaning transition period and reducing antibiotics use.
Chapter
The microbiome undoubtedly plays a major role in stabilizing health and ensuring the highest possible related resilience against environmental stressors and infections. Pigs have distinctly developed microbiomes depending on age, husbandry, and diet. Furthermore, the microbial composition in the digestive tract differs depending on the localization. While the stomach and small intestine are largely dominated by lactic acid bacteria and enterobacteria, the spectrum in the large intestine changes significantly. In addition, there are also important differences between the mucosa-associated and the luminal microbiomes in different intestinal sections. Animal health is supported by a balanced colonization of microorganisms in the sense of eubiosis. Disturbances of the microbial ecological system are referred to as dysbiosis and can lead to an increased incidence of diarrheal diseases in particular as a result of infection with pathogens. An important factor in the control of the microbiome is the diet, in addition to the husbandry and genetics of the animals. A balanced diet formulation is an important basis; additionally, stabilization can be achieved by using feed additives. In this context, it is important to ensure that macronutrients are within nutritional recommendations, because protein, carbohydrates, fiber, and fats can influence the microbiome. Supplementation of minerals and feed additives also plays a major role. The latter have been studied very intensively in recent years; stabilizing effects are described for probiotics, prebiotics, synbiotics, organic acids, enzymes, and plant additives. High pharmacological concentrations of zinc oxide that were commonly used in the EU have been banned since 2022, so that the importance of alternative approaches became much greater. Improved knowledge of mechanisms, signaling pathways, and molecular interactions with the digestive tract and host organism is necessary to exploit the potential of the microbiome to stabilize animal health.
Article
High crude protein (CP; 21% to 26%) diets fed during the first 21 to 28 d postweaning are viewed negatively because of a perceived increase in the incidence rates of diarrhea due to increased intestinal protein fermentation and/or augmented enteric pathogen burden. This is thought to antagonize nursery pig health and growth performance. Therefore, our objective was to evaluate the impact of low vs. high dietary CP on 21-day postweaned pig intestinal function. Analyzed parameters included ex vivo intestinal barrier integrity (ileum and colon), ileal nutrient transport, tissue inflammation, and fecal DM. One hundred and twenty gilts and barrows (average body weight) were randomly assigned to one of two diets postweaning. Diets were fed for 21 d, in two phases. Phase 1 diets: low CP (17%) with a 1.4% standardized ileal digestible (SID) Lys (LCP), or high CP (24%) with a 1.4% SID Lysine (HCP). Phase 2: LCP (17%) and a 1.35% SID lysine, or HCP (24%) formulated to a 1.35% SID lysine. Pig growth rates, feed intakes, and fecal consistency did not differ (P > 0.05) due to dietary treatment. Six animals per treatment were euthanized for additional analyses. There were no differences in colonic epithelial barrier function as measured by transepithelial electrical resistance (TER) and fluorescein isothiocyanate (FITC)-dextran transport between treatments (P > 0.05). Interleukins (IL)-1α, IL-1β, IL-1ra, IL-2 IL-4, IL-6, and IL-12 were not different between treatments (P > 0.05). However, IL-8 and IL-18 were higher in HCP- vs. LCP-fed pigs (P < 0.05). There were no differences in fecal dry matter (DM; P > 0.05) between treatments. In the ileum, there was a tendency (P = 0.06) for TER to be higher in HCP-fed pigs, suggesting a more robust barrier. Interestingly, glucose and glutamine transport were decreased in HCP- vs. LCP-fed pigs (P < 0.05). FITC-dextran transport was not different between treatments (P > 0.05). There were also no differences in ileal cytokine concentrations between diets (P > 0.05). Taken together, the data show that low CP does not negatively impact colonic barrier function, fecal DM, or inflammation. In contrast, ileal barrier function and nutrient transport were altered, suggesting a regional effect of diet on overall intestinal function.
Article
Full-text available
Weaning is a critical transitional point in the life cycle of piglets. Early weaning can lead to post-weaning syndrome, destroy the intestinal barrier function and microbiota homeostasis, cause diarrhea and threaten the health of piglets. The nutritional components of milk and solid foods consumed by newborn animals can affect the diversity and structure of their intestinal microbiota, and regulate post-weaning diarrhea in piglets. Therefore, this paper reviews the effects and mechanisms of different nutrients, including protein, dietary fiber, dietary fatty acids and dietary electrolyte balance, on diarrhea and health of piglets by regulating intestinal function. Protein is an essential nutrient for the growth of piglets; however, excessive intake will cause many harmful effects, such as allergic reactions, intestinal barrier dysfunction and pathogenic growth, eventually aggravating piglet diarrhea. Dietary fiber is a nutrient that alleviates post-weaning diarrhea in piglets, which is related to its promotion of intestinal epithelial integrity, microbial homeostasis and the production of short-chain fatty acids. In addition, dietary fatty acids and dietary electrolyte balance can also facilitate the growth, function and health of piglets by regulating intestinal epithelial function, immune system and microbiota. Thus, a targeted control of dietary components to promote the establishment of a healthy bacterial community is a significant method for preventing nutritional diarrhea in weaned piglets.
Article
Full-text available
This study examined the effects of maternal and/or post-weaning Bacillus altitudinis supplementation on the microbiota in sow colostrum and faeces, and offspring digesta and faeces. Sows (n = 12/group) were assigned to: (1) standard diet (CON), or (2) CON supplemented with probiotic B. altitudinis spores (PRO) from day (d)100 of gestation to weaning (d26 of lactation). At weaning, offspring were assigned to CON or PRO for 28d, resulting in: (1) CON/CON, (2) CON/PRO, (3) PRO/CON, and (4) PRO/PRO, after which all received CON. Samples were collected from sows and selected offspring (n = 10/group) for 16S rRNA gene sequencing. Rothia was more abundant in PRO sow colostrum. Sow faeces were not impacted but differences were identified in offspring faeces and digesta. Most were in the ileal digesta between PRO/CON and CON/CON on d8 post-weaning; i.e. Bacteroidota, Alloprevotella, Prevotella, Prevotellaceae, Turicibacter, Catenibacterium and Blautia were more abundant in PRO/CON, with Firmicutes and Blautia more abundant in PRO/PRO compared with CON/CON. Lactobacillus was more abundant in PRO/CON faeces on d118 post-weaning. This increased abundance of polysaccharide-fermenters (Prevotella, Alloprevotella, Prevotellaceae), butyrate-producers (Blautia) and Lactobacillus likely contributed to previously reported improvements in growth performance. Overall, maternal, rather than post-weaning, probiotic supplementation had the greatest impact on intestinal microbiota.
Article
Full-text available
Diarrhea is a global problem that causes economic losses in the pig industry. There is a growing attention on finding new alternatives to antibiotics to solve this problem. Hence, this study aimed to compare the prebiotic activity of low-molecular-weight hydrolyzed guar gum (GMPS) with commercial manno-oligosaccharide (MOS) and galacto-oligosaccharide (GOS). We further identified their combined effects along with probiotic Clostridium butyricum on regulating the intestinal microbiota of diarrheal piglet by in vitro fermentation. All the tested non-digestible carbohydrates (NDCs) showed favorable short-chain fatty acid-producing activity, and GOS and GMPS showed the highest production of lactate and butyrate, respectively. After 48 h of fermentation, the greatest enhancement in the abundance of Clostridium sensu stricto 1 was observed with the combination of GMPS and C. butyricum. Notably, all the selected NDCs significantly decreased the abundances of pathogenic bacteria genera Escherichia-Shigella and Fusobacterium and reduced the production of potentially toxic metabolites, including ammonia nitrogen, indole, and skatole. These findings demonstrated that by associating with the chemical structure, GMPS exhibited butyrogenic effects in stimulating the proliferation of C. butyricum. Thus, our results provided a theoretical foundation for further application of galactosyl and mannosyl NDCs in the livestock industry. Key points • Galactosyl and mannosyl NDCs showed selective prebiotic effects. • GMPS, GOS, and MOS reduced pathogenic bacteria and toxic metabolites production. • GMPS specifically enhanced the Clostridium sensu stricto 1 and butyrate production. Graphical Abstract
Article
Full-text available
Porcine epidemic diarrhea (PED) is a disease that has a devastating effect on livestock. Currently, most studies are focused on comparing gut microbiota of healthy piglets and piglets with PED, resulting in gut microbial populations related to dynamic change in diarrheal piglets being poorly understood. The current study analyzed the characteristics of gut microbiota in porcine epidemic diarrhea virus (PEDV)-infected piglets during the suckling transition stage. Fresh fecal samples were collected from 1 to 3-week-old healthy piglets (n = 20) and PEDV infected piglets (n = 18) from the same swine farm. Total DNA was extracted from each sample and the V3–V4 hypervariable region of the 16S rRNA gene was amplified and sequenced using the Illumina MiSeq platform. Statistically significant differences were observed in bacterial diversity and richness between the healthy and diarrheal piglets. Principal coordinates analysis (PCoA) showed structural segregation between diseased and healthy groups, as well as among 3 different age groups. The abundance of Escherichia-Shigella, Enterococcus, Fusobacterium, and Veillonella increased due to dysbiosis induced by PEDV infection. Notably, there was a remarkable age-related increase in Fusobacterium and Veillonella in diarrheal piglets. Certain SCFA-producing bacteria, such as Ruminococcaceae_UCG-002, Butyricimonas, and Alistipes, were shared by all healthy piglets, but were not identified in various age groups of diarrheal piglets. In addition, significant differences were observed between clusters of orthologous groups (COG) functional categories of healthy and PEDV-infected piglets. Our findings demonstrated that PEDV infection caused severe perturbations in porcine gut microbiota. Therefore, regulating gut microbiota in an age-related manner may be a promising method for the prevention or treatment of PEDV.
Article
Full-text available
Background Enterotoxigenic Escherichia coli (ETEC) causes diarrhea in humans, cows, and pigs. The gut microbiota underlies pathology of several infectious diseases yet the role of the gut microbiota in the pathogenesis of ETEC-induced diarrhea is unknown. Results By using an ETEC induced diarrheal model in piglet, we profiled the jejunal and fecal microbiota using metagenomics and 16S rRNA sequencing. A jejunal microbiota transplantation experiment was conducted to determine the role of the gut microbiota in ETEC-induced diarrhea. ETEC-induced diarrhea influenced the structure and function of gut microbiota. Diarrheal piglets had lower Bacteroidetes: Firmicutes ratio and microbiota diversity in the jejunum and feces, and lower percentage of Prevotella in the feces, but higher Lactococcus in the jejunum and higher Escherichia-Shigella in the feces. The transplantation of the jejunal microbiota from diarrheal piglets to uninfected piglets leaded to diarrhea after transplantation. Microbiota transplantation experiments also supported the notion that dysbiosis of gut microbiota is involved in the immune responses in ETEC-induced diarrhea. Conclusion We conclude that ETEC infection influences the gut microbiota and the dysbiosis of gut microbiota after ETEC infection mediates the immune responses in ETEC infection. Electronic supplementary material The online version of this article (10.1186/s12917-018-1704-9) contains supplementary material, which is available to authorized users.
Article
Full-text available
Diarrhea remains one of the leading causes of morbidity and mortality globally, with enterotoxigenic Escherichia coli (ETEC) constituting a major causative pathogen. The development of alternative treatments for diarrhea that do not involve chemotherapeutic drugs or result in antibiotic resistance is critical. Considering that lysozyme is a naturally occurring antimicrobial peptide, in a previous study we developed a transgenic pig line that expresses recombinant human lysozyme (hLZ) in its milk. In the present study, we examined the protective effects of the consumption of this milk against ETEC infection in neonatal piglets. We found that consuming hLZ milk facilitated faster recovery from infection and decreased mortality and morbidity following an ETEC oral inoculation or infection acquired by contact-exposure. The protective effect of hLZ was associated with the enrichment of intestinal bacteria that improve gut health, such as Lactobacillus, and the enhancement of the mucosal IgA response to the ETEC-induced diarrhea. Our study revealed potential protective mechanisms underlying the antimicrobial activity of human lysozyme, validating the use of lysozyme as an effective preventive measure for diarrhea.
Article
Full-text available
Early weaned piglets are vulnerable to diarrhea because of weaning stress and immaturity of intestinal tract. Compelling evidence suggests that gut microbiota is vital to host health. However, it is not well understood on the composition and succession of piglet gut microbiota during the weaning transition. In our two trials, total 17 commercial piglets were studied in a pig farm in Jiangxi Province, China. Fresh feces were collected for four times (10 days before weaned, weaned day, 10 days after weaned, 21 days after weaned) by rectal massage. Fecal bacterial composition was assessed by 16S rRNA gene V3–V4 regions sequencing by Illumina Miseq platform. The results showed that the gut microbiota of piglets shifted quickly after weaned and reached relatively stable level in 10 days after weaned. The alpha diversity increased significantly with the age of piglets. The microbiota of suckling piglets was mainly represented by Fusobacterium, Lactobacillus, Bacteroides, Escherichia/Shigella, and Megasphaera. This pattern contrasted with that of Clostridium sensu stricto, Roseburia, Paraprevotella, Clostridium XIVa, and Blautia, which were major representative genera after weaned. In summary, we delineated the development of piglet gut microbiota during the weaning transition. This study helps us understand the maturing development of gut microbiota in commercial piglets.
Article
Full-text available
PurposeBarley is a low-glycemic index grain that can help diabetic and obese patients. The effect of barley intake depends on the host and the associated gut microbiota. This study investigated the effect of barley intake on the fecal microbiota, caecal biochemistry, and key biomarkers of obesity and inflammation. Methods Obese db/db mice were fed diets with and without barley during 8 weeks; lean mice were used as lean controls. Fecal microbiota was evaluated using 16S marker gene sequencing in a MiSeq instrument; several markers of caecal biochemistry, obesity, and inflammation were also evaluated using standard techniques. ResultsBacterial richness (i.e., Operational Taxonomic Units) and Shannon diversity indexes were similar in all obese mice (with and without barley) and higher compared to lean controls. Barley intake was associated with increased abundances of Prevotella, Lactobacillus, and the fiber-degraders S24-7 (Candidatus Homeothermaceae) compared to both lean and obese controls. The analysis of unweighted UniFrac distances showed a separate clustering of samples for each experimental group, suggesting that consumption of barley contributed to a phylogenetically unique microbiota distinct from both obese and lean controls. Caecal butyrate concentrations were similar in all obese mice, while succinic acid was lower in the barley group compared to obese controls. Barley intake was also associated with lower plasma insulin and resistin levels compared to obese controls. Conclusions This study shows that barley intake is associated with a different fecal microbiota, caecal biochemistry, and obesity biomarkers in db/db mice that tend to be more similar to lean controls.
Article
Full-text available
Objectives: The aim of this study was to investigate the presence of plasmid-mediated colistin resistance genes in Escherichia coli from pigs affected by post-weaning diarrhoea (PWD). Methods: DNA samples collected from 51 E. coli isolates from Italian pigs affected by PWD in 2015-2016 were studied. Isolates were classified as presumptively resistant to colistin by routine susceptibility testing and were investigated for the presence of the mcr-1 gene of plasmid origin by PCR. Escherichia coli isolates testing negative for mcr-1 were analysed for the presence of a novel plasmid-mediated gene, mcr-2. Isolates were characterised for fimbrial [F4 (k88), F5 (k99), F6 (987P), F18 and F41] and toxin (LT, STa, STb and Stx2e) determinants by PCR as well as for the occurrence of haemolysis by phenotypic observation. Susceptibility to apramycin, cefquinome, enrofloxacin, florfenicol, gentamicin, tetracycline and trimethoprim/sulfamethoxazole (SXT) was also determined by disk diffusion. Results: Most of the isolates showed the presence of at least one virulence factor, confirming their pathogenic potential. The presence of mcr-1 was shown in 37 (72.5%) of the 51 isolates. All of the mcr-1-negative isolates tested negative for the mcr-2 gene. Moreover, 80.4% of the isolates were resistant to apramycin, 9.8% to cefquinome, 54.9% to enrofloxacin, 52.9% to florfenicol, 76.5% to gentamicin, 96.1% to tetracycline and 80.4% to SXT. Conclusions: This is the first report documenting the presence of the mcr-1 gene in pathogenic E. coli isolated from pigs affected by PWD in Italy.
Article
Full-text available
The swine gut microbiota encompasses a large and diverse population of bacteria that play a significant role in pig health. As such, a number of recent studies have utilized high-throughput sequencing of the 16S rRNA gene to characterize the composition and structure of the swine gut microbiota, often in response to dietary feed additives. It is important to determine which factors shape the composition of the gut microbiota among multiple studies and if certain bacteria are always present in the gut microbiota of swine, independently of study variables such as country of origin and experimental design. Therefore, we performed a metaanalysis using 20 publically available data sets from high-throughput 16S rRNA gene sequence studies of the swine gut microbiota. Next to the "study" itself, the gastrointestinal (GI) tract section that was sampled had the greatest effect on the composition and structure of the swine gut microbiota (P = 0.0001). Technical variation among studies, particularly the 16S rRNA gene hypervariable region sequenced, also significantly affected the composition of the swine gut microbiota (P = 0.0001). Despite this, numerous commonalities were discovered. Among fecal samples, the genera Prevotella, Clostridium, Alloprevotella, and Ruminococcus and the RC9 gut group were found in 99% of all fecal samples. Additionally, Clostridium, Blautia, Lactobacillus, Prevotella, Ruminococcus, Roseburia, the RC9 gut group, and Subdoligranulum were shared by >90% of all GI samples, suggesting a so-called "core" microbiota for commercial swine worldwide.
Article
Weaning is a stress every piglet has to face. It is a main cause of antibiotic uses due to digestive disorders. In this study, response to weaning was analyzed in pigs from 2 lines divergently selected for residual feed intake (RFI) during growth. A total of 132 pigs from each line, housed per line and diet in conventional post-weaning units of 12 castrated males and 12 females, were fed either a conventional control (2 successive diets) or a complex (3 successive diets) dietary sequence during the post-weaning period (4 to 10 wk of age). Body weights (BW) were recorded at weaning (d 0, 28 d of age), d 1, d 2, d 6, d 12, d 19, d 26, d 42 (10 wk of age), and at 23 wk of age. Feces texture was examined before weaning (d -1), at d 1, d 2, d 6, d 12, and d 19. Feed intake was recorded at pen level from d 0 to d 42 after weaning, and individually thereafter. Plasma was collected after blood samplings at d -1, d 6, d 19, and d 42 on half of the piglets: all piglets of a given sex in each pen were sampled, to achieve a balanced number across factors. Pigs of the low RFI line (LRFI) were heavier at weaning, had greater glucose concentration and lower levels of diarrhea at d 1 and d 2 than pigs from the high RFI line (HRFI) (P < 0.01). At d 42, there was no BW difference between lines, and gain-to-feed ratio did not differ between lines (P = 0.40). The LRFI pigs had lower feed intake and growth rate from d 0 to d 19 (P < 0.005), and greater plasma concentration of non-esterified fatty acid (P < 0.001), indicating an increased mobilization of body lipids and proteins immediately after weaning compared with HRFI pigs. They also had greater levels of diarrhea at d 6 (22% for LRFI vs. 14% for HRFI, P = 0.002), but the concentration of plasma haptoglobin did not indicate acute inflammation. The complex diet sequence improved feed intake and growth, and reduced diarrhea, mainly in the LRFI line (P < 0.001). To conclude, pigs from the LRFI line were more negatively affected by weaning stress, but managed to recover afterwards. The complex diet sequence ameliorated some of the negative effects that weaning had on the LRFI pigs, but limited effects of nursery period feeding sequence on growth performance during the growing-finishing period were observed.
Article
Porcine parvovirus type 1 (PPV1) and porcine circovirus type 2 (PCV2) are small single-stranded DNA viruses with high prevalence in the global pig population. The aim of this study was to compare and contrast PCV2 and PPV1 infections in high-health status pigs and to describe PCV2 long-term infection dynamics. Six caesarian-derived colostrum-deprived pigs were randomly divided into two groups and were experimentally infected with PCV2 or PPV1 at 5 weeks of age. All pigs had detectable viremia by day (D) 3 post-infection. Pigs infected with PPV1 had a detectable INF-α response by D3 followed by a high IFN-γ response by D6. The PPV1 pigs developed antibodies against PPV1 by D6 resulting in decreasing virus titers until PPV1 DNA became undetectable from D28 until D42. In contrast, PCV2-infected pigs had no detectable INF-α or IFN-γ response after PCV2 infection. PCV2-infected pigs had no detectable anti-PCV2 humoral response until D49 and had a sustained high level of PCV2 DNA for the duration of the study. While PPV1-infected pigs were clinically normal, PCV2-infected pigs developed severe clinical illness including fatal systemic porcine circovirus associated disease (PCVAD) by D28, fatal enteric PCVAD by D56 and chronic PCVAD manifested as decreased weight gain and periods of diarrhea. Microscopically, all three PCV2-infected pigs had lymphoid lesions consistent with PCVAD and associated with low (chronic disease) to high (acute disease) levels of PCV2 antigen. Under the study conditions, there was a lack of early IFN-γ and INF-α activation followed by a delayed and low humoral immune response and persisting viremia with PCV2 infection. In contrast, PPV1-infected pigs had IFN-γ and INF-α activation and an effective immune response to the PPV1 infection.
Article
Weaning is a critical event in the pig’s life cycle, frequently associated with severe enteric infections and overuse of antibiotics; this raises serious economic and public health concerns. In this review, we explain why gut microbiota dysbiosis, induced by abrupt changes in the diet and environment of piglets, emerges as a leading cause of post-weaning diarrhea, even if the exact underlying mechanisms remain unclear. Then, we focus on nonantimicrobial alternatives, such as zinc oxide, essential oils, and prebiotics or probiotics, which are currently evaluated to restore intestinal balance and allow a better management of the crucial weaning transition. Finally, we discuss how in vitro models of the piglet gut could be advantageously used as a complement to ex vivo and in vivo studies for the development and testing of new feed additives.