ArticlePDF Available

Abstract and Figures

Background Bovine TB (bTB), caused by infection with Mycobacterium bovis , is a major endemic disease affecting global cattle production. The key innate immune cell that first encounters the pathogen is the alveolar macrophage, previously shown to be substantially reprogrammed during intracellular infection by the pathogen. Here we use differential expression, and correlation- and interaction-based network approaches to analyse the host response to infection with M. bovis at the transcriptome level to identify core infection response pathways and gene modules. These outputs were then integrated with genome-wide association study (GWAS) data sets to enhance detection of genomic variants for susceptibility/resistance to M. bovis infection. Results The host gene expression data consisted of RNA-seq data from bovine alveolar macrophages (bAM) infected with M. bovis at 24 and 48 h post-infection (hpi) compared to non-infected control bAM. These RNA-seq data were analysed using three distinct computational pipelines to produce six separate gene sets: 1) DE genes filtered using stringent fold-change and P -value thresholds (DEG-24: 378 genes, DEG-48: 390 genes); 2) genes obtained from expression correlation networks (CON-24: 460 genes, CON-48: 416 genes); and 3) genes obtained from differential expression networks (DEN-24: 339 genes, DEN-48: 495 genes). These six gene sets were integrated with three bTB breed GWAS data sets by employing a new genomics data integration tool— gwinteR . Using GWAS summary statistics, this methodology enabled detection of 36, 102 and 921 prioritised SNPs for Charolais, Limousin and Holstein-Friesian, respectively. Conclusions The results from the three parallel analyses showed that the three computational approaches could identify genes significantly enriched for SNPs associated with susceptibility/resistance to M. bovis infection. Results indicate distinct and significant overlap in SNP discovery, demonstrating that network-based integration of biologically relevant transcriptomics data can leverage substantial additional information from GWAS data sets. These analyses also demonstrated significant differences among breeds, with the Holstein-Friesian breed GWAS proving most useful for prioritising SNPS through data integration. Because the functional genomics data were generated using bAM from this population, this suggests that the genomic architecture of bTB resilience traits may be more breed-specific than previously assumed.
This content is subject to copyright. Terms and conditions apply.
R E S E A R C H A R T I C L E Open Access
Integrative genomics of the mammalian
alveolar macrophage response to
intracellular mycobacteria
Thomas J. Hall
1
, Michael P. Mullen
2
, Gillian P. McHugo
1
, Kate E. Killick
1,3
, Siobhán C. Ring
4
, Donagh P. Berry
5
,
Carolina N. Correia
1
, John A. Browne
1
, Stephen V. Gordon
6,7
and David E. MacHugh
1,7*
Abstract
Background: Bovine TB (bTB), caused by infection with Mycobacterium bovis, is a major endemic disease affecting
global cattle production. The key innate immune cell that first encounters the pathogen is the alveolar
macrophage, previously shown to be substantially reprogrammed during intracellular infection by the pathogen.
Here we use differential expression, and correlation- and interaction-based network approaches to analyse the host
response to infection with M. bovis at the transcriptome level to identify core infection response pathways and
gene modules. These outputs were then integrated with genome-wide association study (GWAS) data sets to
enhance detection of genomic variants for susceptibility/resistance to M. bovis infection.
Results: The host gene expression data consisted of RNA-seq data from bovine alveolar macrophages (bAM)
infected with M. bovis at 24 and 48 h post-infection (hpi) compared to non-infected control bAM. These RNA-seq
data were analysed using three distinct computational pipelines to produce six separate gene sets: 1) DE genes
filtered using stringent fold-change and P-value thresholds (DEG-24: 378 genes, DEG-48: 390 genes); 2) genes
obtained from expression correlation networks (CON-24: 460 genes, CON-48: 416 genes); and 3) genes obtained
from differential expression networks (DEN-24: 339 genes, DEN-48: 495 genes). These six gene sets were integrated
with three bTB breed GWAS data sets by employing a new genomics data integration toolgwinteR. Using GWAS
summary statistics, this methodology enabled detection of 36, 102 and 921 prioritised SNPs for Charolais, Limousin
and Holstein-Friesian, respectively.
(Continued on next page)
© The Author(s). 2021 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.
* Correspondence: david.machugh@ucd.ie
1
Animal Genomics Laboratory, UCD School of Agriculture and Food Science,
UCD College of Health and Agricultural Sciences, University College Dublin,
Belfield, Dublin D04 V1W8, Ireland
7
UCD Conway Institute of Biomolecular and Biomedical Research, University
College Dublin, Belfield, Dublin D04 V1W8, Ireland
Full list of author information is available at the end of the article
Hall et al. BMC Genomics (2021) 22:343
https://doi.org/10.1186/s12864-021-07643-w
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
(Continued from previous page)
Conclusions: The results from the three parallel analyses showed that the three computational approaches could
identify genes significantly enriched for SNPs associated with susceptibility/resistance to M. bovis infection. Results
indicate distinct and significant overlap in SNP discovery, demonstrating that network-based integration of
biologically relevant transcriptomics data can leverage substantial additional information from GWAS data sets.
These analyses also demonstrated significant differences among breeds, with the Holstein-Friesian breed GWAS
proving most useful for prioritising SNPS through data integration. Because the functional genomics data were
generated using bAM from this population, this suggests that the genomic architecture of bTB resilience traits may
be more breed-specific than previously assumed.
Keywords: Alveolar macrophage, GWAS, Integrative genomics, Mycobacterium bovis, Network, RNA-seq,
Tuberculosis
Background
Bovine tuberculosis (bTB) is a chronic disease of live-
stock, particularly among domestic dairy and beef cattle,
which has been conservatively estimated to cause more
than $3 billion annual losses to global agriculture [1,2].
The disease can also establish across a large variety of
wildlife species including, for example, American bison
(Bison bison), African buffalo (Syncerus caffer), the
brushtail possum (Trichosurus vulpecula), red deer (Cer-
vus elaphus), wild boar (Sus scrofa), and the European
badger (Meles meles)[36]. The aetiological agent of
bTB, Mycobacterium bovis, is a member of the Mycobac-
terium tuberculosis complex (MTBC) and has a genome
sequence 99.95% identical to M. tuberculosis, the pri-
mary cause of human tuberculosis (TB) [7] and the lead-
ing cause of human deaths from a single infectious
agentapproximately 1.25 million in 2018 [8]. In
addition, for several low- and middle-income countries,
the human TB disease burden is increased by zoonotic
TB (zTB) caused by infection with M. bovis [912].
Scientific understanding of bTB and human TB has
been synergistically intertwined since the nineteenth
century and the foundational research work of Theobald
Smith and others [5,13,14]. The pathogenesis of bTB
disease in cattle is comparable with human TB disease
and many aspects of M. bovis infection are also charac-
teristic of M. tuberculosis infection [1519]. Conse-
quently, it is now widely recognised that M. bovis
infection of cattle and bTB disease represent an import-
ant comparative system for understanding human TB
caused by M. tuberculosis [2024].
Inhalation of aerosolized bacteria is the main route of
transmission for M. bovis in cattle and the primary site of
infection is normally the lungs [17,25,26]. Here the bacilli
are phagocytosed by alveolar macrophages (AM)key
effector cells of the innate immune system, which provide
surveillance of pulmonary surfaces and can normally des-
troy or restrict inhaled intracellular bacilli [27,28]. M. bovis
and other facultative intracellular MTBC pathogens have
evolved a complex range of mechanisms to evade, subvert,
and exploit innate immune responses, thereby facilitating
colonisation, persistence and replication within host macro-
phages [2932]. These mechanisms include: recruitment of
cell surface receptors on the host macrophage through mo-
lecular mimicry; restricting phagosome maturation and au-
tophagy; detoxification of reactive oxygen species and
reactive nitrogen intermediates (ROSs and RNIs); modula-
tion of type I interferon (IFN) signalling; suppression of
antigen presentation; rewiring and short-circuiting of
macrophage signal transduction pathways; manipulation of
host macrophage metabolism; egress of bacilli into the
macrophage cytosol; and inhibition of apoptosis with con-
comitant induction of necrosis leading to immunopathol-
ogyandsheddingbythehosttocompletethepathogenic
life cycle [3339]. Hence after infection, a two-way re-
sponse is triggered between the pathogen and macrophage,
the outcome of which ultimately leads to establishment of
infection or clearance of the pathogen. The latter outcome
of clearance may, or may not, require engagement of the
adaptive immune system. As the detection of M. bovis in-
fection in cattle generally relies on detecting an adaptive
immune response to the pathogen, the outcome of which is
slaughter of positive animals (reactors), identifying genes
that underpin efficacious innate responses promises to
reveal favourable genomic variants for incorporation into
breeding programmes.
Since 2005, substantial efforts have been made to
better understand host-pathogen interaction for bTB
using transcriptomics technologies such as gene expres-
sion microarrays and RNA sequencing (RNA-seq) at the
host cellular levelspecifically the bovine alveolar
macrophage (bAM) and initial innate immune responses
to infection by M. bovis [4046]. These studies have
helped to define a pathogenic signature[31,47]ofM.
bovis infection in bAM, which reflects the tension be-
tween macrophage responses to contain and kill intra-
cellular pathogens and evasion and avoidance
mechanisms evolved by these mycobacteria. Using func-
tional genomics data mining of transcriptomics data, it
has also been shown that bAM responses to M. bovis
Hall et al. BMC Genomics (2021) 22:343 Page 2 of 20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
infection can be clearly differentiated from infection with
M. tuberculosis, the primary cause of human TB [45]. In
addition, these studies have been expanded to encom-
pass surveys of the bAM epigenome using methylome
sequencing and chromatin immunoprecipitation sequen-
cing (ChIP-seq). This work has demonstrated that the
transcriptional reprogramming of bAM caused by M.
bovis infection is profoundly shaped by chromatin re-
modelling at gene loci associated with critical compo-
nents of host-pathogen interaction [46,48].
In parallel to functional genomics studies of bTB,
genome-wide association studies (GWAS) have been
performed in Irish and UK cattle populations using
estimated breeding values (EBVs; estimate of genetic
merit of an animal derived from a statistical model)
for several M. bovis infection resistance traits with
heritabilities ranging from 0.04 to 0.37, depending on
the phenotype used [4955]. These GWAS have used
medium- and high-density single-nucleotide poly-
morphism (SNP) arrays and, more recently, imputed
whole-genome sequence (WGS) data sets for a large
multi-breed GWAS on 7346 bulls, which identified 64
quantitative trait loci (QTLs) associated with resist-
ance to M. bovis infection [55]; the association study
was based on phenotypic data from 781,270 individual
animals.
We have recently shown that integration of bAM
functional genomics data setsRNA-seq, microRNA-seq
and ChIP-seqwith a GWAS data set for resistance to
M. bovis infection can be used to enhance detection of
genomic regions associated with reduced incidence of
bTB disease [46]. For the present study, we substantially
expand this work by leveraging gene-focused network-
and pathway-based methods under a statistical frame-
work based on a new software tool, gwinteR, to integrate
transcriptomics data from M. bovis-challenged bAM
with WGS-based GWAS results for resistance to M.
bovis infection [55]. Functional genomics data and
downstream data mining (e.g., to generate lists of differ-
entially expressed genes and outputs from network and
pathway analyses) can be used to obtain prioritised sub-
sets of genes that are likely to be important for a specific
biological process or phenomenon [56,57]. The gwinteR
tool can leverage these prioritised gene subsets and com-
bine them with summary statistics from biologically rele-
vant GWAS data sets. Biologically meaningful SNP-
phenotype associations can therefore be identified and
enriched that would otherwise be filtered out because of
stringent multiple test correction for the very large num-
bers of observations in a typical GWAS. The primary
aim of this work was to evaluate whether this approach
can systematically enhance detection of genomic se-
quence variants and genes underpinning bTB disease re-
sistance in cattle populations.
Results
Differential gene expression and pathway analyses of M.
bovis-infected bovine AM
Quality filtering of RNA-seq read pairs yielded a mean
of 22,681,828 ± 3,508,710 reads per individual library
(n= 78 libraries). A mean of 19,582,959 ± 3,021,333 read
pairs (86.17%) were uniquely mapped to locations in the
ARS-UCD1.2 bovine genome assembly. Detailed filtering
and mapping statistics are provided in a data file avail-
able from the Dryad Digital Repository (https://doi.org/
10.5061/dryad.83bk3j9q6) and multivariate PCA analysis
of the individual animal sample expression data using
DESeq2 revealed separation of the control and M. bovis-
infected bAM groups at the 24 and 48 h post infection
(hpi) time points, but not at the 2 and 6 hpi time points
(S1Fig).
Using default criteria for differential expression (FDR
P
adj.
< 0.05; |log
2
FC| > 1), and considering the M. bovis-
infected bAM relative to the control non-infected mac-
rophages, three DE genes were detected at 2 hpi (all
three exhibited increased expression in the M. bovis-in-
fected group); 97 DE genes were detected at 6 hpi (40
increased and 57 decreased); 1345 were detected at 24
hpi (764 increased and 581 decreased); and 2915 were
detected at 48 hpi (1528 increased and 1387 decreased)
(Fig. 2a; Additional file 1). Figure 2b shows that 2982
genes were differentially expressed across the 24 and 48
hpi time points. Table 1shows a breakdown of DE genes
across the infection time course for a range of statistical
thresholds and fold-change cut-offs, including the de-
fault criteria (FDR P
adj.
< 0.05; |log
2
FC| > 1). To ensure
manageable computational loads, the DE gene sets that
were used for GWAS integration with gwinteR were fil-
tered with |log
2
FC| > 2, and P
adj.
< 0.01 and P
adj.
<
0.000001 for 24 and 48 hpi, respectively. With these cri-
teria, there were 378 input genes for GWAS integration
identified at 24 hpi and 390 input genes at 48 hpi. (24
hpi and 48 hpi DEG gene sets). In addition, 210 genes
overlapped between the two time points. The two DEG
gwinteR input gene sets (DEG-24 and DEG-48 see
Fig. 1) are also detailed in Additional file 1.
To produce gene sets for the IPA Core Analysis within
the recommended range for the number of input entities
[58,59] and to include DE genes with small fold-change
values, gene sets were filtered using only P
adj.
thresholds
of 0.05 and 0.01 at 24 hpi and 48 hpi, respectively. This
resulted in 1957 input genes (1071 upregulated and 886
downregulated) from a background detectable set of 16,
084 at 24 hpi and 2492 input genes (1401 upregulated
and 1091 downregulated) from a background detectable
set of 17,492 genes at 48 hpi. The IPA analysis was fo-
cused on the 24 and 48 hpi time points because a rela-
tively small numbers of DE genes were detected at 2 and
6 hpi (Fig. 2and Table 1).
Hall et al. BMC Genomics (2021) 22:343 Page 3 of 20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Using the B-H method for multiple test correction in
IPA (P
adj.
< 0.05), there were 68 and 48 statistically sig-
nificant enriched IPA canonical pathways at 24 hpi and
48 hpi, respectively (Additional file 2). Enriched path-
ways at 24 hpi included Role of Pattern Recognition Re-
ceptors in Recognition of Bacteria and Viruses,IL-6
Signalling,TNFR2 Signalling,Role of RIG1-like Receptors
in Antiviral Innate Immunity,Role of Cytokines in Medi-
ating Communication between Immune Cells,Communi-
cation between Innate and Adaptive Immune Cells,IL-
12 Signalling and Production in Macrophages,IL-10 Sig-
nalling,Protein Ubiquitination Pathway,Toll-like Recep-
tor Signalling,NF-κB Signalling,PI3K/AKT Signalling,
and TNFR1 Signalling. The most highly activated path-
way at 24 hpi was PI3K/AKT Signalling. Enriched path-
ways at 48 hpi included Protein Ubiquitination Pathway,
Role of Cytokines in Mediating Communication between
Immune Cells,IL-12 Signalling and Production in Mac-
rophages,Role of RIG1-like Receptors in Antiviral Innate
Immunity,Role of Pattern Recognition Receptors in
Table 1 Differentially expressed genes detected in M. bovis-infected bovine AM relevant to control non-infected bAM
Post-infection time point P
adj.
< 0.05; |log
2
FC| > 0
(increased/decreased)
P
adj.
< 0.05; |log
2
FC| > 1
(increased/decreased)
P
adj.
< 0.01; |log
2
FC| > 0
(increased/decreased)
P
adj.
< 0.01; |log
2
FC| > 1
(increased/decreased)
2 hpi 14 (7/7) 3 (3/0) 8 (4/4) 2 (2/0)
6 hpi 410 (203/207) 97 (40/57) 119 (63/56) 32 (14/18)
24 hpi 3620 (1898/1722) 1345 (764/581) 2059 (1168/891) 933 (577/356)
48 hpi 6442 (3295/3147) 2915 (1528/1387) 4737 (2516/2221) 2386 (1294/1092)
Fig. 1 Schematic showing the experimental and computational workflow use to integrate bAM transcriptomics outputs and M. bovis infection
resistance trait GWAS data (some figure components created with a BioRender.com licence)
Hall et al. BMC Genomics (2021) 22:343 Page 4 of 20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Fig. 2 Differentially expressed genes in M. bovis-infected bAM at 2, 6, 24, and 48 hpi. aVolcano plots of differentially expressed genes with FDR
P
adj.
value thresholds of 0.05 and absolute log
2
fold-change > 1. bUpSet plot showing the intersection of shared differentially expressed genes
across the four post-infection time points
Hall et al. BMC Genomics (2021) 22:343 Page 5 of 20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Recognition of Bacteria and Viruses,Communication be-
tween Innate and Adaptive Immune Cells,TNFR2 Sig-
nalling,Role of PI3K/AKT Signalling in the Pathogenesis
of Influenza,IL-10 Signalling, and Toll-like Receptor Sig-
nalling. The SIGORA software tool [60] has been previ-
ously used to identify biological pathways associated
with a robust corebAM response to infection with both
M. bovis and M. tuberculosis [45]. It is therefore reassur-
ing that many of these pathwaysincluding PI3K-Akt
Signalling Pathway,RIG-I-like Receptor Signalling Path-
way,Toll-like Receptor Signalling and Protein Ubiquiti-
nation Pathwaywere also enriched using the IPA
methodology at 24 and 48 hpi.
Differential co-expression correlation networks and
identification of functional gene modules
For the generation of bAM differential co-expression
correlation networks, filtering of genes with low
measure of central tendency, which reduces the
number of potential spurious correlations [61], re-
sulted in 11,354 and 11,170 genes at 24 and 48 hpi,
respectively. Following this step, differential correl-
ation analysis using DGCA with an empirical P
adj.
value threshold of 0.10 resulted in 3507 differentially
correlated gene pairs out of 128,913,316 total pair-
wise correlations at 24 hpi; and 1135 from a total of
124,768,900 at 48 hpi (Additional file 3). The correl-
ation networks generated at 24 hpi and 48 hpi
(Fig. 3a) yielded a total of 22 and 14 functional gene
modules, respectively (Fig. 3b and c, and Additional
file 3). After removal of duplicates, consolidated to-
tals of 460 genes and 416 genes were contained in
the functional modules at 24 hpi and 48 hpi, re-
spectively. There were also 26 genes that overlapped
between the functional modules for the two time
points. The two correlation network (CON) gwinteR
input gene sets (CON-24 and CON-48 see Fig. 1)
are also detailed in Additional file 3.
GO term enrichment was also performed for each
functional module at 24 hpi and 48 hpi, with the top
three GO terms retained for each functional module
(S2Fig and S3Fig). The top five GO terms at 24 hpi
(ranked by P
adj.
)weretranslation (GO:0006412), pep-
tide biosynthetic process (GO:0043043), amide biosyn-
thetic process (GO:0043604), structural constituent of
ribosome (GO:0003735), and cellular amide metabolic
process (GO:0043603) (S2Fig). The top five enriched
GO terms at 48 hpi (ranked by P
adj.
)weresignalling
receptor activity (GO:0038023) and molecular trans-
ducer activity (GO:0060089), transforming growth fac-
tor beta activation (GO:0036363), chemokine activity
(GO:0008009), and signalling receptor binding (GO:
0005102) (S3Fig).
Differential expression network analysis and identification
of activated modular subnetworks
To provide a computationally manageable number of genes
for an InnateDB input data set [62], a GeneCards Relevance
Score (GCRS) threshold was used (GCRS > 2.5). This GCRS
cut-off produced an input list of 258 functionally prioritised
genes for generation of an InnateDB gene interaction net-
work (GIN) and the top ten genes from this list ranked by
GCRS were: interferon gamma receptor 1 (IFNGR1), inter-
leukin 12 receptor subunit beta 1 (IL12RB1), toll like recep-
tor 2 (TLR2), solute carrier family 11 member 1 (SLC11A1),
signal transducer and activator of transcription 1 (STAT1),
interleukin 12B (IL12B), cytochrome b-245 beta chain
(CYBB), tumour necrosis factor (TNF), interferon gamma re-
ceptor 2 (IFNGR2), and interferon gamma (IFNG).
The large GIN produced by InnateDB starting with
the input list of 258 functionally prioritised genes was
visualised using Cytoscape and consisted of 7001 nodes
(individual genes) and 19,713 edges (gene interactions)
(Fig. 4a). Additional file 4provides information for all
gene interactions represented in Fig. 4a. Following visu-
alisation of the large GIN in Cytoscape, the jActivesMo-
dules Cytoscape plugin was used to detect statistically
significant differentially activated subnetworks (modules)
at the 24 hpi and 48 hpi time points. The top five sub-
networks at each time point were retained for down-
stream analyses and consisted of 198 genes in module 1
at 24 hpi (M124), 287 genes in M224, 272 genes in
M324, 53 genes in M424, 171 genes in M524, 381
genes in M148, 330 genes in M248, 403 genes in
M348, 371 genes in M448, and 399 genes in M548
(Additional file 4). As an example, Fig. 4b shows the
subnetwork of genes and gene interactions representing
module 5 at 24 hpi (M524).
The genes contained in the top five modules at 24 hpi
and 48 hpi were filtered to remove duplicates and con-
solidated into two separate gene sets for GWAS integra-
tion with gwinteR. The consolidated gene sets for the
top five modules at 24 hpi and 48 hpi contained 339 and
495 unique genes, respectively. There were 245 genes
that overlapped between the two subnetwork gene sets
for the two post-infection time points. The two differen-
tial expression network (DEN) gwinteR input gene sets
(DEN-24 and DEN-48 see Fig. 1) are also detailed in
Additional file 4.
GWAS integration and identification of additional SNP
trait associations
The six gene sets generated from the three separate
analyses of DE genes in bAM challenged with M. bovis
at 24 hpi and 48 hpi are summarised in Table 2and fur-
ther detailed in Additional files 1,2and 3. Also, S4Fig
and S5Fig show Venn diagrams with the overlaps for
the DEG, CON and DEN input gene sets at 24 hpi and
Hall et al. BMC Genomics (2021) 22:343 Page 6 of 20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Fig. 3 Differential co-expression correlation networks and submodules at 24 and 48 hpi. aThe complete correlation networks for M. bovis-
infected bAM at 24 and 48 hpi. bThe 22 subnetwork modules detected at 24 hpi and the 14 subnetwork modules detected at 48 hpi with
individual example modules highlighted in yellow. cIndividual example subnetwork modules at 24 and 48 hpi
Hall et al. BMC Genomics (2021) 22:343 Page 7 of 20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Fig. 4 A large gene interaction network (GIN) with superimposed differentially expressed genes. aCytoscape tuberculosis gene interaction
network with superimposed differentially expressed genes at 24 and 48 hpi. bExample subnetwork showing module number 5 detected with
the 24 hpi gene expression data
Hall et al. BMC Genomics (2021) 22:343 Page 8 of 20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
48 hpi, respectively. In addition to these six putative
functionally relevant gene sets, one hundred sets of 250
genes randomly sampled from the bovine genome were
used for statistical context and comparison. These
random gene sets (RAN) are detailed in Additional file 5
(see also Fig. 1). The results from the integrative analyses
using gwinteR with the DEG-24, DEG-48, CON-24,
CON-48, DEN-24, DEN-48, and RAN gene sets are
summarised graphically in Fig. 5and detailed in Add-
itional files 6and 7. We also used DE gene sets at 2 and
6 hpi with the default statistical threshold and fold-
change cut-offs (FDR P
adj.
< 0.05; |log
2
FC| > 1) for
GWAS integration using gwinteR (i.e., DEG-2 and DEG-
6). However, no significant SNP enrichment were ob-
served for these input gene sets, most likely because of
the low numbers of input genes (3 and 97, respectively).
Figure 5a shows circular Manhattan plots with
GWAS results (P
adj.
values) for each of the three
breeds prior to data integration using gwinteR.
Figure 5bshowsthegwinteR permuted P-values
(P
perm.
) for each of the 10 genomic intervals used
and for each of the six input gene sets plus the
RAN gene set with P
adj.
< 0.10. Figure 5cshowscir-
cular Manhattan plots with GWAS results post data
integration using gwinteR. Inspection of Fig. 5b
shows that, in terms of SNP enrichment (P
perm.
<
0.05), the integrative analyses using gwinteR were
most effective for the HOF breed group where the
CON-24, CON-48, and DEG-48 input gene sets pro-
duced enriched SNPs across all 10 genomic ranges.
In addition, the DEG-24 and DEN-24 input gene sets
were effective for the HOF breed across the ±20 to
100 kb and ± 30 to 50 kb genomic ranges, respect-
ively. In the case of the LIM breed, the DEG-48 in-
put gene set produced enriched SNPs across all 10
genomic ranges, the CON-48 between ±10 to 70 kb
and the CON-24 at ±10 kb. For the CHA breed, SNP
enrichment using gwinteR was only observed for the
CON-24 input gene set for the genomic interval be-
tween ±10 to 40 kb.
Figure 6summarises the numbers of statistically
significant SNPs pre- and post-data integration, again with
SNP enrichment being most evident for the HOF breed
with a 24-fold post-gwinteR SNP enrichment at P
adj.
<
0.10 from 40 to 961 SNPs. The SNP enrichments for the
CHA and LIM breed were more modest, although there
was a 2.3-fold enrichment at P
adj.
<0.10from80to182
SNPs for the LIM breed. Inspection of Additional file 7re-
veals notable gene loci associated with enriched GWAS
SNPs for the HOF breed including the allograft inflamma-
tory factor 1 gene (AIF1), which encodes a protein that
promotes macrophage activation and proinflammatory ac-
tivity [63]; the ciliogenesis associated kinase 1 gene (CILK1
aka ICK); IL17A and IL7F, which encode proinflammatory
cytokines and contain polymorphisms in the human
orthologs that have been associated with lower human TB
disease incidence [64,65]; the integrin subunit beta 3 gene
(ITGB3), which encodes a protein that has been shown to
regulate matrix metalloproteinase secretion in pulmonary
human TB [66]; the neuraminidase 1 gene (NEU1), which
encodes a protein that regulates phagocytosis in macro-
phages [67]; and the TNF gene that encodes TNF (aka
TNF-α), a key cytokine for generation and maintenance of
the granuloma, and where gene polymorphisms have been
linked to resistance to M. bovis infection [68]. Gene loci
associated with enriched GWAS SNPs for the CHA breed
group also included the CILK1 gene (aka ICK) and the
Kelch repeat and BTB domain containing 3 gene
(KCNJ15), which has been detected as an expression bio-
marker for human TB [69]; the T cell immune regulator
1, ATPase H+ transporting V0 subunit a3 gene (TCIRG1),
a known antimycobacterial host defence gene that has
been shown to be a key hub gene associated with IFN-γ
stimulation of human macrophages [70]; and the Von
Willebrand factor gene (VWF). Notable gene loci associ-
ated with enriched GWAS SNPs for the Limousin breed
group included the cellular repressor of E1A stimulated
genes 1 gene (CREG1), which encodes a regulator of core
macrophage differentiation genes [71]; the desmoplakin
gene (DSP), which increases in expression during M. tu-
berculosis-derived ESAT6-regulated transition of bone
marrow-derived macrophages (BMDMs) into epithelioid
macrophages [72]; and the SP110 nuclear body protein
gene (SP110), which encodes a protein that modulates
growth of MTBC pathogens in macrophages and has been
successfully exploited for genome editing of cattle to en-
hance resistance to M. bovis infection [73].
Discussion
During the last decade, integrative genomics, multi-
omics analyses and network biology have come to the
fore as powerful strategies for exploring, dissecting
Table 2 Six different input gene sets used for GWAS integration
Post-
infection
time
point
Filtered differentially
expressed genes (DEG)
Correlation networks
analysis (CON)
Differential expression
network analysis (DEN)
Input gene set code No. of genes Input gene set code No. of genes Input gene set code No. of genes
24 hpi DEG-24 378 CON-24 460 DEN-24 338
48 hpi DEG-48 390 CON-48 416 DEN-48 493
Hall et al. BMC Genomics (2021) 22:343 Page 9 of 20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
and unpicking the complexities of the vertebrate im-
mune system and immune responses to specific mi-
crobial pathogens [7477]. In the present study we
have used these approaches to integrate transcripto-
mics data from a pivotal immune effector cell in M.
bovis infection with high-resolution GWAS data for a
bTB resistance trait. We developed a new computa-
tional tool, gwinteR, to enhance detection of QTLs by
leveraging nominal SNP P-values from large GWAS
data sets for resistance to infection by M. bovis in
cattle. Three different integration strategies were
employed with transcriptomics data from bAM
infected with M. bovis across a 48-h time course. The
first and most straightforward method was based on
DE gene sets at 24 and 48 hpi with stringent fold-
change and P-value thresholds as filtering criteria
(DEG-24 and DEG-48). For the second method, a
correlation network approach was used to identify
subnetworks (modules) of co-expressed functional
gene clusters from the bAM transcriptomics data at
the two post-infection time points (CON-24 and
CON-48). The third method was also network-based
but took advantage of the extensive scientific litera-
ture and curated biomolecular data for mycobacterial
Fig. 5 Integration of bAM functional genomics and GWAS data for resistance to M. bovis infection in three cattle breeds. a Circular Manhattan
plots showing GWAS results pre-integration with blue and red data points indicating binned SNP clusters with FDR P
adj.
< 0.10 and < 0.05,
respectively. b Line plots of permuted P-values across different genomic intervals for SNPs from six different input gene sets. c Circular Manhattan
plots showing GWAS results post-integration using gwinteR with blue and red data points indicating binned SNP clusters with FDR P
adj.
< 0.10
and < 0.05, respectively
Hall et al. BMC Genomics (2021) 22:343 Page 10 of 20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
infections and tuberculosis disease. For this approach,
a base GIN was constructed and functional modules
containing overlaid differentially expressed genes were
identified to provide two post-infection input gene
sets (DEN-24 and DEN-48) for downstream data
integration.
Integration of these six input gene sets with bTB
GWAS data from three different cattle breeds (CHA,
LIM and HOF) revealed substantial differences among
the three methods in their capacity to detect additional
QTLs for a M. bovis infection resistance trait in these
particular GWAS data sets. For example, the correlation
network approach was the only method that enriched
SNPs for the CHA breed and that worked for at least
one bAM post-infection time point across all three
breeds (see Fig. 5). Surprisingly, perhaps, the functional
modules obtained using the GIN differential network
(DEN) approachthe most complex method to imple-
mentproduced the least effective input gene sets for
prioritising additional SNPs from the GWAS data set.
This method proved effective only for the DEN-24 input
gene set with the HOF breed and then only for the ±30,
±40, and ± 50 kb genomic intervals. Conversely, the sim-
plest method based on DE genes at 24 and 48 hpi
enriched SNPs for both the LIM and HOF breed groups,
with the DEG-48 input gene set being the most effective
(Fig. 5). The relatively poor performance of the DEN ap-
proach for integration of functional genomics and
GWAS data may be a consequence of the human-
focused GeneCards and InnateDB resources we used to
generate the base GIN [62,78]. Therefore, it would be
instructive to conduct similar integrative analyses for
Fig. 6 Histogram showing numbers of significant GWAS SNPs for the bTB resistance trait, pre- and post-enrichment for the three cattle breeds
Hall et al. BMC Genomics (2021) 22:343 Page 11 of 20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
human TB using appropriate functional genomics and
GWAS data sets.
In summary, across the two different bAM infection
time points and the three breeds, the CON method
enriched 970 SNPs, the DEG method enriched 163 SNPs
and the DEN method enriched only 11 SNPs (Additional
file 7). Although other factors such as linkage disequilib-
rium need to be considered in interpreting these differ-
ences, it is reasonable to hypothesise that GWAS
integration using the correlation network approach is
more sensitive to regulatory genomic variants that alter
expression of co-ordinately regulated protein compo-
nents of the alveolar macrophage pathways and pro-
cesses underpinning host-pathogen interaction for the
early stages of intracellular MTBC infection [27,28,30].
Figure 7illustrates this using the example of SNPs
within and proximal to gene loci associated with PI3K/
AKT signalling, which based on IPA data mining for
bAM DE genes was a highly activated pathway, particu-
larly at 24 hpi. In this regard, previous work has shown
that macrophage PI3K/AKT signalling is key to a range
of cellular processes associated with host-pathogen
interaction in MTBC infections, including modulation of
cell death pathways, manipulation of signalling down-
stream of TLRs, and initiation of granuloma formation
[7984]. We have also recently demonstrated that genes
encoding protein products embedded in the PI3K/AKT
pathway are primary targets for chromatin modifications
that substantially alter bAM gene expression in response
to M. bovis infection [46].
There were also notable breed differences in the effect-
iveness of the three methods used for multi-omics data in-
tegration and enrichment of GWAS SNPs. The original
GWAS data set for the HOF breed was the least powered
of the three breeds in terms of both sample size (n=1502
sires) and genetic markers (12,740,315 genome-wide
SNPs). This is reflected in the relatively small number of
GWAS SNPs that were detected pre-integration for the
HOF breed: 40 SNPs at P
adj.
<0.10 compared to 475 for
the CHA breed and 80 for the LIM breed (see Fig. 6).
However, the SNP enrichment post-integration was mark-
edly more effective for the HOF breed with a 24-fold in-
crease from 40 to 961 SNPs (P
adj.
<0.10) compared to
1.08-fold (475 to 511 SNPs; P
adj.
< 0.10) and 2.28-fold (80
to 182 SNPs; P
adj.
< 0.10) for the CHA and LIM breeds, re-
spectively (Fig. 6). In addition, a total of 48 genes were
captured by the enriched GWAS SNPs for the HOF breed
compared to 20 and 16 genes for the CHA and LIM
breeds, respectively (Additional file 7).
Regarding the enhanced SNP detection observed for
the HOF breed GWAS after multi-omic integration, it is
noteworthy that the M. bovis infection transcriptomics
data set was generated using bAM from this population
[43]. Most of the SNPs (n= 829, 88%) and genes (n= 36,
67%) obtained post-integration in the HOF breed were
detected using the correlation co-expression network
method, which, again, is likely caused by enrichment of
genomic regulatory variants that modulate expression of
genes associated with alveolar macrophage pathways and
processes critical to early host-pathogen interaction.
This pattern of variation may also reflect the polygenic
architecture of M. bovis infection resistance in domestic
cattle [52,54,55,85].
It is well established that intracellular mycobacterial
pathogens use a host of evolved mechanisms to survive
within, and ultimately disseminate from host macro-
phage cells. Several of these mechanisms are ultimately
linked to regulators of TNF and cognate downstream
signalling pathways [29,30,39]. It is noteworthy, there-
fore, that our analysis prioritised SNPs at the TNF gene
locus, a key nexus for host-pathogen interaction in
mycobacterial infections. In addition, another important
component of host-pathogen interaction and target for
immunoevasion by intracellular mycobacterial infections
is the PI3K/AKT pathway [33,79,83] and our integra-
tive genomics approach was able to prioritise genomic
variation at genes encoding ligands and receptors within
this signalling network, specifically TNXB,GNG11,
ITGB3,andVWF (see Fig. 7).
Conclusions
Elucidation of the mechanisms used by M. bovis to estab-
lish infection in cattle, and ultimately cause disease, re-
quires an intimate knowledge of host-pathogen
interactions, especially at the interface between the bAM
and the invading bacilli. Using a systems biology approach,
we identified and catalogued patterns of gene expression
and gene-gene interactions that occur in bAM during the
initial stages of M. bovis infection, highlighting key re-
sponse pathways and hub genes. Additionally, a new R
package, gwinteR, facilitated integration of these functional
genomics outputs with three large breed based GWAS
data sets for resistance/susceptibility to M. bovis infection.
This revealed genomic variants associated with resistance
in key innate immune genes and supported the hypothesis
that the response to M. bovis infection at the level of the
alveolar macrophage may be more breed-specific than
previously assumed.
Integrative multi-omics approaches to data integra-
tion are now widely used to explore and dissect the
genomic architecture and physiological basis of com-
plex traits in domestic livestock, including network-
based methods to integrate functional genomics and
GWAS data [8691]. However, to the best of our
knowledge, this study is the first that uses network
biology to systematically combine transcriptomics
data from M. bovis-infected macrophages with
GWAS data for M. bovis infection resistance in
Hall et al. BMC Genomics (2021) 22:343 Page 12 of 20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Fig. 7 The PI3K/AKT signalling pathway and genomic variants within associated genes. aThe IPA-predicted activation state of the PI3K/AKT
signalling pathway with red and blue colours indicating increased and decreased activity of pathway components, respectively. bSchematic of
enriched bTB disease resistance GWAS SNPs in four genes associated with the PI3K/AKT signalling pathway
Hall et al. BMC Genomics (2021) 22:343 Page 13 of 20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
cattle. Therefore, it provides a novel framework for
integrative genomics studies of complex infectious
disease resistance traits in livestock, particularly
those involving other intracellular bacterial patho-
gens such as Brucella abortus,Mycobacterium avium
subsp. paratuberculosis and Salmonella enterica.The
work is also relevant to development of methods for
integrative analyses of outputs from the Functional
Annotation of Animal Genomics (FAANG) initiative
[92] and for identification and prioritization of tar-
gets for genome editing to enhance resistance to in-
fection in domestic livestock species [93]. The
results from this study may also inform genome-
enabled breeding programmes for resistance to M.
bovis infection in production cattle populations [85,
94]. Finally, this integrative multi-omics approach
could also be used to combine relevant functional
genomics and GWAS data sets to improve know-
ledge of innate immune responses and establishment
of infection in human TB caused by M. tuberculosis.
Methods
Genomics data acquisition and computational and
bioinformatics workflow
Genome-wide RNA-seq transcriptomics data from a
48-h bAM time course challenge experiment using
the sequenced M. bovis AF2122/97 strain have been
previously generated by our group (GEO accession:
GSE62506). The complete laboratory methods used to
isolate, culture and infect bAM with M. bovis
AF2122/9 and generate strand-specific RNA-seq li-
braries using RNA harvested from these cells are de-
scribed in detail elsewhere [42,43,45]. Briefly, these
RNA-seq data were generated using bAM obtained by
lung lavage of ten unrelated age-matched 7 to 12-
week-old male Holstein-Friesian calves. Bovine AM
were either infected in vitro with M. bovis AF2122/97
or incubated with media only. Following total RNA
extraction from M. bovis-infected and control non-
infected alveolar macrophages, 78 strand-specific
RNA-seq libraries were prepared (paired-end 2 × 90
nucleotide reads). These comprised M. bovis-and
non-infected samples from each post-infection time
point (2, 6, 24 and 48 hpi) across 10 animals with the
exception of one animal that did not yield sufficient
alveolar macrophages for in vitro infection at 48 hpi.
GWAS data sets for the present study were obtained
from intra-breed imputed WGS-based GWAS analyses
that used estimated breeding values (EBVs) derived from
a bTB infection phenotype, which were generated for
2039 Charolais, 1964 Limousin and 1502 Holstein-
Friesian sires [55]. The bTB phenotype, the WGS-based
imputed SNP data, and the quantitative genetics
methods are described in detail elsewhere [55]; however,
the following provides a brief summary. The bTB infec-
tion phenotype was defined for every animal present
during each herd-level bTB breakdown when a bTB re-
actor or a slaughterhouse case was identified. Cattle that
yielded a positive single intradermal comparative tuber-
culin test (SICTT), and/or post-mortem lymph node le-
sion, or laboratory culture result/s were coded as bTB =
1 and all other cattle present in the herd during the
bTB-breakdown were coded as bTB = 0; potential expos-
ure of cattle within the bTB breakdown was also consid-
ered in this study [55]. After phenotype data edits, bTB
resistance EBVs were generated for 781,270 phenotyped
cattle (plus their recorded ancestors). After within-breed
SNP filtering using thresholds for minor allele frequency
(MAF 0.002) and deviation from Hardy-Weinberg
equilibrium (HWE; P<1×10
6
), there were 17,250,600,
17,267,260 and 15,017,692 autosomal SNPs for the 2039
Charolais (CHA), 1964 Limousin (LIM) and 1502
Holstein-Friesian (HOF) sire analyses. A single-SNP re-
gression analyses was performed for each breed separ-
ately using weighted (i.e., by an effective record
contribution) sire EBVs for M. bovis infection resistance/
susceptibility and the nominal P-values were used for
downstream integrative genomics analyses.
All data-intensive computational procedures were
performed on a 36-core/72-thread compute server (2×
Intel
®
Xeon
®
CPU E52697 v4 processors, 2.30 GHz
with 18 cores each), with 512 GB of RAM, 96 TB SAS
storage (12 × 8 TB at 7200 rpm), 480 GB SSD storage,
and with Ubuntu Linux OS (version 18.04 LTS). The
complete computational and bioinformatics workflow
is available with additional information as a public
GitHub repository (github.com/ThomasHall1688/
Bovine_multi-omic_integration). The individual com-
ponents of the experimental and computational work-
flows are shown in Fig. 1and described in more
detail below.
Differential gene expression analysis of RNA-seq data
A custom Perl script was used to deconvolute barcoded
RNA-seq reads into individual libraries, filter out adapter
sequence reads, and remove poor quality reads [43]. At
each stage of the process, a quality check was performed
on the FASTQ files with FastQC (version 0.11.8) [95].
Paired-end sequence reads were then aligned to the Bos
taurus reference genome (ARS-UCD1.2, GenBank as-
sembly accession: GCA_002263795.2) [96] using the
STAR aligner (version 2.7) [97]. Read counts for each
gene were calculated using featureCounts (version 1.6.4)
[98], set to unambiguously assign uniquely aligned
paired-end reads in a stranded manner to gene exon an-
notation. Using the R statistical programming language
(version 3.5.3) [99], gene annotation was derived from
the NCBI database via a GFF annotation file GCF_
Hall et al. BMC Genomics (2021) 22:343 Page 14 of 20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
002263805.1 with additional descriptions and chromo-
somal locations annotated using GO.db (version 3.8.2)
[100] and biomaRt packages (version 2.40.0) [101]. Dif-
ferential gene expression analysis was performed using
the DESeq2 package (version 1.24.0) [102] with a longi-
tudinal time series design that accounted for time (hpi)
and experimental treatment (M. bovis-infected versus
control). Lowly expressed reads were removed using the
mean of normalized counts as a filter statistic; individual
genes with very low read counts would typically not ex-
hibit significant differential expression due to high dis-
persion [102]. In addition, extreme count outliers were
removed within DESeq2 using the Cooks distance [103]
as previously described [102]. Multiple testing correction
was performed on each time point using the Benjamini-
Hochberg (B-H) false discovery rate (FDR) method
[104]. The default criteria for differentially expressed
(DE) genes were an FDR-adjusted P-value less than 0.05
(P
adj.
< 0.05).
Ingenuity
®
pathway analysis (IPA) of differentially
expressed genes
Ingenuity
®
Pathway AnalysisIPA
®
(version 1.1, summer
2020 release; Qiagen, Redwood City, CA, USA) was used
to perform a statistical enrichment analysis of DE gene
sets and expression data [58]. This enabled identification
of canonical pathways and functional processes of bio-
logical importance in alveolar macrophages challenged
with M. bovis across the longitudinal infection time
course. The target species selected was Homo sapiens
and the cell type used was Macrophage with the Experi-
mentally Observed and High Predicted confidence set-
tings. Following best practice, the default background
gene set for pathway and functional process enrichment
testing was the set of detectable genes across all RNA-
seq libraries for each time point contrast and not the
complete bovine transcriptome [105].
Identification of functional gene modules using
differential co-expression network analysis
An integrated computational pipeline for differential
co-expression network analysis was implemented
using: 1) the Differential Gene Correlation Analysis
(DGCA) R package (version 1.0.2) [61]; 2) the Cytos-
cape open source Java platform for visualisation and
integration of biomolecular interaction networks (ver-
sion 3.7.0) [106]; and 3) the Multiscale Embedded
Gene Co-expression Network Analysis (MEGENA) R
package (version 1.3.7) [107].
Duetothelowvarianceingeneexpressionob-
served among samples at 2 and 6 hpi (Fig. 2aandb),
only data from the 24 and 48 hpi time points were
used for the differential correlation analysis. The
DGCA R package was used to filter normalised gene
counts such that genes in the lower 30th percentile
of median expression values were removed. Pearson
correlation coefficients were then calculated for each
gene pair between the control non-infected and M.
bovis-infected samples at 24 and 48 hpi. Following
this, for each time point, the infected and control
samples were randomly shuffled, and the analysis was
repeated for a total of ten iterations. Additionally,
using the permutation testing and reference pool dis-
tribution approaches implemented in DGCA, an em-
pirical P-value was calculated for each observed
correlation coefficient and q-values were calculated
based on empirical P-values and the estimated pro-
portion of null hypotheses; gene pairs were then con-
sidered to be differentially correlated with a q-value
threshold of 0.10 [61].
The correlation networks and network parameters
generated by DGCA were initially visualised, examined
and evaluated using the Cytoscape platform. In correl-
ation networks, based on gene co-expression, each gene
acts as a node and each correlation acts as a weighted
edge, depending on the strength of the correlation coef-
ficient [108,109]. The DGCA network data for the 24
and 48 hpi time points were then imported into MEGE
NA for identification of functional subnetworks (mod-
ules) using differentially correlated (q< 0.10) gene pairs
ranked by q-value to a maximum of 3500 gene pairs; this
gave 3500 and 1085 gene pairs for the 24 and 48 hpi
correlation networks, respectively. Each gene pair was
assigned to a class that described the change in correl-
ation depending on infection status, and only those
genes that exhibited a change in expression pattern were
included in the MEGENA analysis to identify functional
modules.
Functional modules in the 24 and 48 hpi correlation
networks were detected as locally coherent subnetwork
clusters with a minimum of 20 unique genes that MEGE
NA classified as statistically significant (P-value < 0.05)
based on analyses of shortest path indices, local path
index, weight of the correlation and overall modularity.
The resulting MEGENA functional modules were
visualised using the Cytoscape platform and genes
embedded in functional modules at 24 and 48 hpi
were combined and annotated for downstream GWAS
integration as described below. The genes contained
in these functional modules were also subject to gene
ontology (GO) term enrichment analyses within the
MEGENA package [107].
Detection of active gene subnetworks using a
tuberculosis and mycobacterial infection gene interaction
network
The GeneCards
®
gene compendium and knowledgebase
(www.genecards.org; version 4.9), which integrates
Hall et al. BMC Genomics (2021) 22:343 Page 15 of 20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
multiple sources of biological information on all anno-
tated and predicted human genes [78], was used to iden-
tify a set of genes that are functionally associated with
the host response to TB and other diseases caused by in-
fection with mycobacteria. The GeneCards search query
generated a total of 2291 gene hits (Additional file 4)
using the search terms: tuberculosis OR mycobac-
terium OR mycobacteria OR mycobacterial.
Genes were ranked by a GeneCards statisticthe Rele-
vance Scorebased on the Elasticsearch algorithm [110],
which determines the strength of the relationships be-
tween genes and keyword terms. Gene IDs were con-
verted to human Ensembl gene IDs [111] and retained
for downstream analysis using the InnateDB knowledge-
base and analysis platform for systems level analysis of
the innate immune response (www.innatedb.com; ver-
sion 5.4) [62].
A gene interaction network (GIN) was generated
with the gene list output from GeneCards using Inna-
teDB with default settings and this network was
visualised using Cytoscape. The jActivesModules
Cytoscape plugin (version 3.12.1) [112] was then used
to superimpose the bAM RNA-seq gene expression
data and detect, through a greedy search algorithm,
differentially active subnetworks (modules) of genes at
the 24 and 48 hpi time points. Locally coherent clus-
ters that contain genes that are differentially
expressed were identified using the log
2
FC and P
adj.
values of each differentially expressed gene; the over-
all connectivity of those genes with their immediate
module co-members; and the comparison of that con-
nectivity with a background comprised of randomly
drawn networks using the same genes, but independ-
ent of the base network. Genes embedded in active
modules that were detected as statistically significant
at 24 and 48 hpi were combined and annotated for
downstream GWAS integration as described below.
Integration of M. bovis-infected bovine AM gene
expression data with bTB GWAS data
To facilitate integration of GWAS data with gene sets
generated from functional genomics data analyses, an R
software package was developedgwinteR (github.com/
ThomasHall1688/gwinteR), which can be used to test
the hypothesis that a specific set of genes is enriched for
signal in a GWAS data set relative to the genomic back-
ground. This gene set, for example, could be an output
from an active gene module network analysis of tran-
scriptomics data from a cell type or tissue relevant to
the GWAS phenotype. To formally test the primary hy-
pothesis, the gwinteR tool was designed to determine if
genomic regions containing GWAS SNPs that are prox-
imal to genes within a gene set are enriched for statis-
tical associations with the trait/s analysed in the GWAS.
The gwinteR tool works as follows: 1) a set of signifi-
cant and non-significant SNPs (named the target SNP
set) is collated across all genes in a specific gene set at
increasing genomic intervals upstream and downstream
from each gene inclusive of the coding sequence (e.g., ±
10 kb, ±20 kb, ±30 kb ……±100 kb); 2) for each gen-
omic region, a null distribution of 1000 SNP sets, each
of which contains the same number of total significant
and non-significant combined SNPs as the target SNP
set, is generated by resampling with replacement from
the search space of the total population of SNPs in the
GWAS data set; 3) the nominal (uncorrected) GWAS P-
values for the target SNP set and the null distribution
SNP sets are converted to local FDR-adjusted P-values
(P
adj.
) using the fdrtool R package (current version
1.2.15) [113]; 4) a permuted P-value (P
perm.
) to the test
the primary hypothesis for each observed genomic inter-
val target SNP set is generated based on the proportion
of permuted random SNP sets where the same or a lar-
ger number of SNPs exhibiting significant q-values (e.g.
q< 0.05 or q< 0.10) are observed; 5) gwinteR generates
data to plot P
perm.
results by genomic interval class and
obtain a graphical representation of the GWAS signal
surrounding genes within the target gene set; and 6) a
summary output file of all SNPs in the observed target
SNP set with genomic locations and q-values is gener-
ated for subsequent investigation.
In the original bTB GWAS data set used for the
present study [55], the WGS-imputed SNPs were
mapped to the UMD3.1 bovine genome assembly [114].
Consequently, prior to GWAS data integration, the im-
puted and previously filtered SNPs for each of the three
breed groups were mapped, using a custom R pipeline
(github.com/ThomasHall1688/Bovine_multi-omic_
integration), to the most recent ARS-UCD1.2 cattle ref-
erence genome assembly [96]. After this step, there were
14,583,567, 14,586,972 and 12,740,315 autosomal SNPs
with nominal GWAS P-values that could be used for in-
tegrative genomics analyses of the CHA, LIM and HOF
breeds, respectively.
For the integrative analyses of bAM functional
genomics outputs with the bTB GWAS data, three
different subsets of genes were used: 1) basic DE
gene sets that were filtered to ensure manageable
computational loads using stringent expression
threshold criteria of |log
2
FC| > 2 and P
adj.
<0.01 and
P
adj.
< 0.000001 for 24 and 48 hpi, respectively; 2)
genes embedded in functional modules at 24 and 48
hpi that were detected using the MEGENA package
in the differential co-expression network analyses;
and 3) genes embedded in active modules at 24 and
48 hpi that were identified using jActiveModules
within the tuberculosis and mycobacterial infection
interaction network.
Hall et al. BMC Genomics (2021) 22:343 Page 16 of 20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Abbreviations
AM: Alveolar macrophage; bAM: Bovine alveolar macrophage; B-
H: Benjamini-Hochberg; bTB: Bovine tuberculosis; CHA: Charolais; ChIP-
seq: Chromatin immunoprecipitation sequencing; CON: Correlation network;
DEG: Differentially expressed gene; DEN: Differential expression network;
DGCA: Differential Gene Correlation Analysis; EBV: Estimated breeding value;
GCRS: GeneCards Relevance Score; FDR: False discovery rate; GIN: Gene
interaction network; GO: Gene ontology; GWAS: Genome-wide association
study; HOF: Holstein-Friesian; HWE: Hardy-Weinberg equilibrium;
IFN: Interferon; IPA: Ingenuity Pathway Analysis; LIM: Limousin; MEGE
NA: Multiscale Embedded Gene Co-expression Network Analysis; MAF: Minor
allele frequency; MTBC: Mycobacterium tuberculosis complex;
QTL: Quantitative trait locus; RNA-seq: RNA sequencing; RNI: Reactive
nitrogen intermediate; ROS: Reactive oxygen species; SICTT: Single
intradermal comparative tuberculin test; SNP: Single nucleotide
polymorphism; TB: Tuberculosis; WGS: Whole-genome sequence;
zTB: Zoonotic tuberculosis
Supplementary Information
The online version contains supplementary material available at https://doi.
org/10.1186/s12864-021-07643-w.
Additional file 1. Differentially expressed genes across the infection
time course and DEG-24 and DEG-48 input gene sets for GWAS
integration.
Additional file 2. Enriched IPA canonical pathways at 24 and 48 hpi.
Additional file 3. Outputs from correlation network analyses using
DGCA with MEGENA and CON-24 and CON-48 input gene sets for GWAS
integration.
Additional file 4. Outputs from differential network analyses using
GeneCards
®
, InnateDB with Cytoscape/jActiveModules and DEN-24 and
DEN-48 input gene sets for GWAS integration.
Additional file 5. 100 random (RAN) input gene sets used for GWAS
integration.
Additional file 6. Functional genomics and GWAS integration results for
Charolais (CHA), Limousin (LIM) and Holstein-Friesian (HOF).
Additional file 7. GWAS prioritised SNP results for Charolais (CHA),
Limousin (LIM) and Holstein-Friesian (HOF).
Additional file 8: S1 Fig. Principal component analysis (PCA) plots for
individual animal bAM gene expression data at a2 hpi, b6 hpi, c24 hpi,
and d48 hpi. S2 Fig. Gene ontology (GO) enrichment for functional
modules identified from the differential co-expression correlation network
generated from M. bovis-infected bAM gene expression at 24 hpi. S3 Fig.
Gene ontology (GO) enrichment for functional modules identified from
the differential co-expression correlation network generated from M.
bovis-infected bAM gene expression at 48 hpi. S4 Fig. Venn diagram il-
lustrating the overlaps for the three 24 hpi input gene sets used for inte-
gration with cattle GWAS data sets. S5 Fig. Venn diagram illustrating the
overlaps for the three 48 hpi input gene sets used for integration with
cattle GWAS data sets. RNA-seq statistics and results for this study are
available at the Dryad Digital Repository: https://doi.org/10.5061/dryad.
83bk3j9q6.
Acknowledgements
This work was presented at the 37th International Society for Animal
Genetics (ISAG) Conference held in Lleida, Spain; 712th July 2019 (www.
isag.us/Docs/Proceedings/ISAG2019_Proceedings.pdf).
Authorscontributions
TJH, MPM, KEK, DPB, SVG and DEM conceived and designed the study. JAB
performed experimental work. SCR and DPB provided cattle genotype and
other genomic data. TJH, MPM, GPM, KEK, SCR and CNC performed
bioinformatics and computational analyses. TJH, CNC and DEM wrote and
prepared the manuscript and figures. All authors read and approved the final
manuscript.
Funding
This study was supported by Science Foundation Ireland (SFI) Investigator
Programme Awards to D.E.M. and S.V.G. (grant nos. SFI/08/IN.1/B2038 and
SFI/15/IA/3154); a Department of Agriculture, Food and the Marine (DAFM)
project award to D.E.M (TARGET-TB; grant no. 17/RD/US-ROI/52); and a
European Union Framework 7 project grant to D.E.M. (no: KBBE-211602-
MACROSYS). The funding agencies had no role in the study design, collec-
tion, analysis, and interpretation of data, and no role in writing the
manuscript.
Availability of data and materials
The RNA-seq data set was generated by the authors and can be obtained
from the NCBI Gene Expression Omnibus (GEO): accession number
GSE62506. GWAS summary statistics data were obtained from a published
study that provides additional information about sequence and genotype
data availability [55]. The source code to reproduce the results of the ana-
lyses is available from a GitHub repository (https://github.com/ThomasHall16
88/Bovine_multi-omic_integration). The source code used to integrate the
outputs from functional genomics and GWAS data sets is available from a
GitHub repository (https://github.com/ThomasHall1688/gwinteR) Detailed
RNA-seq filtering and mapping statistics are provided in a data file available
from the Dryad Digital Repository: https://doi.org/10.5061/dryad.83bk3j9q6.
The computational pipeline and software used for this study is detailed
below.
Complete pipeline: https://github.com/ThomasHall1688/Bovine_multi-omic_
integration
gwinteR: https://github.com/ThomasHall1688/gwinteR
DESeq2: https://github.com/mikelove/DESeq2
DGCA: https://github.com/andymckenzie/DGCA
MEGENA: https://github.com/cran/MEGENA
Cattle genome assembly and annotation: www.ncbi.nlm.nih.gov/datasets/
genomes/?txid=9913
Declarations
Ethics approval and consent to participate
All animal procedures were performed in accordance with EU Directive
2010/63/EU and Irish Statutory Instrument 543/2012, with ethical approval
from the University College Dublin (UCD) Animal Research Ethics Committee
(AREC-1314-Gordon).
Consent for publication
Not applicable.
Competing interests
The authors declare that they have no competing interests.
Author details
1
Animal Genomics Laboratory, UCD School of Agriculture and Food Science,
UCD College of Health and Agricultural Sciences, University College Dublin,
Belfield, Dublin D04 V1W8, Ireland.
2
Bioscience Research Institute, Athlone
Institute of Technology, Dublin Road, Athlone, Westmeath N37 HD68, Ireland.
3
Present address: Genuity Science, Cherrywood Business Park. Loughlinstown,
Dublin D18 K7W4, Ireland.
4
Irish Cattle Breeding Federation, Highfield House,
Shinagh, Bandon, Cork P72 X050, Ireland.
5
Teagasc, Animal and Grassland
Research and Innovation Centre, Moorepark, Fermoy, Cork P61 C996, Ireland.
6
UCD School of Veterinary Medicine, UCD College of Health and Agricultural
Sciences, University College Dublin, Belfield, Dublin D04 V1W8, Ireland.
7
UCD
Conway Institute of Biomolecular and Biomedical Research, University
College Dublin, Belfield, Dublin D04 V1W8, Ireland.
Received: 19 November 2020 Accepted: 22 April 2021
References
1. Steele JH. Introduction (Part 2 Regional and Country Status Reports). In:
Thoen CO, Steele JH, editors. Mycobacterium bovis infection in animals and
humans. Ames: Iowa State University Press; 1995. p. 16972.
2. Waters WR, Palmer MV, Buddle BM, Vordermeier HM. Bovine tuberculosis
vaccine research: historical perspectives and recent advances. Vaccine. 2012;
30(16):261122. https://doi.org/10.1016/j.vaccine.2012.02.018.
Hall et al. BMC Genomics (2021) 22:343 Page 17 of 20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
3. Fitzgerald SD, Kaneene JB. Wildlife reservoirs of bovine tuberculosis
worldwide: hosts, pathology, surveillance, and control. Vet Pathol. 2013;
50(3):48899. https://doi.org/10.1177/0300985812467472.
4. Gormley E, Corner LAL. Wild animal tuberculosis: stakeholder value systems
and management of disease. Front Vet Sci. 2018;5:327. https://doi.org/10.33
89/fvets.2018.00327.
5. Malone KM, Gordon SV. Mycobacterium tuberculosis complex members
adapted to wild and domestic animals. Adv Exp Med Biol. 2017;1019:135
54. https://doi.org/10.1007/978-3-319-64371-7_7.
6. Palmer MV. Mycobacterium bovis: characteristics of wildlife reservoir hosts.
Transbound Emerg Dis. 2013;60(Suppl 1):113. https://doi.org/10.1111/
tbed.12115.
7. Garnier T, Eiglmeier K, Camus JC, Medina N, Mansoor H, Pryor M, et al. The
complete genome sequence of Mycobacterium bovis. Proc Natl Acad Sci U S
A. 2003;100(13):787782. https://doi.org/10.1073/pnas.1130426100.
8. World Health Organization. Global Tuberculosis Report 2019. Geneva: World
Health Organization; 2019.
9. Luciano SA, Roess A. Human zoonotic tuberculosis and livestock exposure
in low- and middle-income countries: a systematic review identifying
challenges in laboratory diagnosis. Zoonoses Public Health. 2020;67(2):97
111. https://doi.org/10.1111/zph.12684.
10. Olea-Popelka F, Muwonge A, Perera A, Dean AS, Mumford E, Erlacher-Vindel
E, et al. Zoonotic tuberculosis in human beings caused by Mycobacterium
bovis a call for action. Lancet Infect Dis. 2017;17(1):e215. https://doi.org/1
0.1016/S1473-3099(16)30139-6.
11. Thoen CO, Kaplan B, Thoen TC, Gilsdorf MJ, Shere JA. Zoonotic tuberculosis.
A comprehensive ONE HEALTH approach. Medicina (B Aires). 2016;76(3):
15965.
12. Vayr F, Martin-Blondel G, Savall F, Soulat JM, Deffontaines G, Herin F.
Occupational exposure to human Mycobacterium bovis infection: a
systematic review. PLoS Negl Trop Dis. 2018;12(1):e0006208. https://doi.
org/10.1371/journal.pntd.0006208.
13. Daniel TM. The history of tuberculosis. Respir Med. 2006;100(11):186270.
https://doi.org/10.1016/j.rmed.2006.08.006.
14. Cambau E, Drancourt M. Steps towards the discovery of Mycobacterium
tuberculosis by Robert Koch, 1882. Clin Microbiol Infect. 2014;20(3):196201.
https://doi.org/10.1111/1469-0691.12555.
15. Neill SD, Bryson DG, Pollock JM. Pathogenesis of tuberculosis in cattle.
Tuberculosis (Edinb). 2001;81(12):7986. https://doi.org/10.1054/tube.2000.
0279.
16. Russell DG. Highlighting the parallels between human and bovine
tuberculosis. J Vet Med Educ. 2003;30(2):1402. https://doi.org/10.3138/
jvme.30.2.140.
17. Cassidy JP. The pathogenesis and pathology of bovine tuberculosis with
insights from studies of tuberculosis in humans and laboratory animal
models. Vet Microbiol. 2006;112(24):15161. https://doi.org/10.1016/j.
vetmic.2005.11.031.
18. Pollock JM, Rodgers JD, Welsh MD, McNair J. Pathogenesis of bovine
tuberculosis: the role of experimental models of infection. Vet Microbiol.
2006;112(24):14150. https://doi.org/10.1016/j.vetmic.2005.11.032.
19. WatersWR,MaggioliMF,McGillJL,LyashchenkoKP,PalmerMV.Relevanceof
bovine tuberculosis research to the understanding of human disease: historical
perspectives, approaches, and immunologic mechanisms. Vet Immunol
Immunopathol. 2014;159(34):11332. https://doi.org/10.1016/j.vetimm.2014.02.009.
20. Hein WR, Griebel PJ. A road less travelled: large animal models in
immunological research. Nat Rev Immunol. 2003;3(1):7984. https://doi.
org/10.1038/nri977.
21. Van Rhijn I, Godfroid J, Michel A, Rutten V. Bovine tuberculosis as a model
for human tuberculosis: advantages over small animal models. Microbes
Infect. 2008;10(7):7115. https://doi.org/10.1016/j.micinf.2008.04.005.
22. Waters WR, Palmer MV, Thacker TC, Davis WC, Sreevatsan S, Coussens P,
et al. Tuberculosis immunity: opportunities from studies with cattle. Clin
Dev Immunol. 2011;2011:768542.
23. Williams A, Orme IM. Animal models of tuberculosis: an overview. Microbiol
Spectr. 2016;4(4). https://doi.org/10.1128/microbiolspec.TBTB2-0004-2015.
24. Gong W, Liang Y, Wu X. Animal models of tuberculosis vaccine research: an
important component in the fight against tuberculosis. Biomed Res Int.
2020;2020:4263079. https://doi.org/10.1155/2020/4263079.
25. Palmer MV, Waters WR, Whipple DL. Aerosol delivery of virulent
Mycobacterium bovis to cattle. Tuberculosis (Edinb). 2002;82(6):27582.
https://doi.org/10.1054/tube.2002.0341.
26. Palmer MV, Wiarda J, Kanipe C, Thacker TC. Early pulmonary lesions in cattle
infected via aerosolized Mycobacterium bovis. Vet Pathol. 2019;56(4):54454.
https://doi.org/10.1177/0300985819833454.
27. Kaufmann SHE, Dorhoi A. Molecular determinants in phagocyte-bacteria
interactions. Immunity. 2016;44(3):47691. https://doi.org/10.1016/j.immuni.2
016.02.014.
28. Weiss G, Schaible UE. Macrophage defense mechanisms against intracellular
bacteria. Immunol Rev. 2015;264(1):182203. https://doi.org/10.1111/imr.12266.
29. Awuh JA, Flo TH. Molecular basis of mycobacterial survival in macrophages. Cell Mol
Life Sci. 2017;74(9):162548. https://doi.org/10.1007/s00018-016-2422-8.
30. Schorey JS, Schlesinger LS. Innate immune responses to tuberculosis.
Microbiol Spectr. 2016;4(6). https://doi.org/10.1128/microbiolspec.TBTB2-001
0-2016.
31. Cambier CJ, Falkow S, Ramakrishnan L. Host evasion and exploitation
schemes of Mycobacterium tuberculosis. Cell. 2014;159(7):1497509. https://
doi.org/10.1016/j.cell.2014.11.024.
32. de Chastellier C. The many niches and strategies used by pathogenic
mycobacteria for survival within host macrophages. Immunobiology. 2009;
214(7):52642. https://doi.org/10.1016/j.imbio.2008.12.005.
33. BoseDasgupta S, Pieters J. Macrophage-microbe interaction: lessons learned
from the pathogen Mycobacterium tuberculosis. Semin Immunopathol. 2018;
40(6):57791. https://doi.org/10.1007/s00281-018-0710-0.
34. Chaurasiya SK. Tuberculosis: smart manipulation of a lethal host. Microbiol
Immunol. 2018;62(6):36179. https://doi.org/10.1111/1348-0421.12593.
35. Hussain Bhat K, Mukhopadhyay S. Macrophage takeover and the host-bacilli
interplay during tuberculosis. Future Microbiol. 2015;10(5):85372. https://
doi.org/10.2217/fmb.15.11.
36. Leopold Wager CM, Arnett E, Schlesinger LS. Mycobacterium tuberculosis and
macrophage nuclear receptors: What we do and don't know. Tuberculosis
(Edinb). 2019;116s:S98s106. https://doi.org/10.1016/j.tube.2019.04.016.
37. Queval CJ, Brosch R, Simeone R. The macrophage: a disputed fortress in the
battle against Mycobacterium tuberculosis. Front Microbiol. 2017;8:2284.
https://doi.org/10.3389/fmicb.2017.02284.
38. Russell DG, Huang L, VanderVen BC. Immunometabolism at the interface
between macrophages and pathogens. Nat Rev Immunol. 2019;19(5):291
304. https://doi.org/10.1038/s41577-019-0124-9.
39. Stutz MD, Clark MP, Doerflinger M, Pellegrini M. Mycobacterium tuberculosis:
rewiring host cell signaling to promote infection. J Leukoc Biol. 2018;103(2):
25968. https://doi.org/10.1002/JLB.4MR0717-277R.
40. Widdison S, Watson M, Piercy J, Howard C, Coffey TJ. Granulocyte
chemotactic properties of M. tuberculosis versus M. bovis-infected bovine
alveolar macrophages. Mol Immunol. 2008;45(3):7409. https://doi.org/10.1
016/j.molimm.2007.06.357.
41. Widdison S, Watson M, Coffey TJ. Early response of bovine alveolar
macrophages to infection with live and heat-killed Mycobacterium bovis.Dev
Comp Immunol. 2011;35(5):58091. https://doi.org/10.1016/j.dci.2011.01.001.
42. Magee DA, Conlon KM, Nalpas NC, Browne JA, Pirson C, Healy C, et al.
Innate cytokine profiling of bovine alveolar macrophages reveals
commonalities and divergence in the response to Mycobacterium bovis and
Mycobacterium tuberculosis infection. Tuberculosis (Edinb). 2014;94(4):441
50. https://doi.org/10.1016/j.tube.2014.04.004.
43. Nalpas NC, Magee DA, Conlon KM, Browne JA, Healy C, McLoughlin KE,
et al. RNA sequencing provides exquisite insight into the manipulation of
the alveolar macrophage by tubercle bacilli. Sci Rep. 2015;5(1):13629.
https://doi.org/10.1038/srep13629.
44. Vegh P, Magee DA, Nalpas NC, Bryan K, McCabe MS, Browne JA, et al.
MicroRNA profiling of the bovine alveolar macrophage response to
Mycobacterium bovis infection suggests pathogen survival is enhanced by
microRNA regulation of endocytosis and lysosome trafficking. Tuberculosis.
2015;95(1):607. https://doi.org/10.1016/j.tube.2014.10.011.
45. Malone KM, Rue-Albrecht K, Magee DA, Conlon K, Schubert OT, Nalpas NC,
et al. Comparative omics analyses differentiate Mycobacterium tuberculosis
and Mycobacterium bovis and reveal distinct macrophage responses to
infection with the human and bovine tubercle bacilli. Microb Genom. 2018;
4(3):e000163. https://doi.org/10.1099/mgen.0.000163.
46. Hall TJ, Vernimmen D, Browne JA, Mullen MP, Gordon SV, MacHugh DE,
et al. Alveolar macrophage chromatin is modified to orchestrate host
response to Mycobacterium bovis infection. Front Genet. 2020;10:1386.
https://doi.org/10.3389/fgene.2019.01386.
47. Falkow S. I never met a microbe I didn't like. Nat Med. 2008;14(10):10537.
https://doi.org/10.1038/nm1008-1053.
Hall et al. BMC Genomics (2021) 22:343 Page 18 of 20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
48. O'Doherty AM, Rue-Albrecht KC, Magee DA, Ahting S, Irwin RE, Hall TJ, et al.
The bovine alveolar macrophage DNA methylome is resilient to infection
with Mycobacterium bovis. Sci Rep. 2019;9(1):1510. https://doi.org/10.1038/
s41598-018-37618-z.
49. Finlay EK, Berry DP, Wickham B, Gormley EP, Bradley DG. A genome wide
association scan of bovine tuberculosis susceptibility in Holstein-Friesian dairy
cattle. PLoS One. 2012;7(2):e30545. https://doi.org/10.1371/journal.pone.0030545.
50. Bermingham ML, Bishop SC, Woolliams JA, Pong-Wong R, Allen AR, McBride
SH, et al. Genome-wide association study identifies novel loci associated
with resistance to bovine tuberculosis. Heredity (Edinb). 2014;112(5):54351.
https://doi.org/10.1038/hdy.2013.137.
51. Richardson IW, Berry DP, Wiencko HL, Higgins IM, More SJ, McClure J, et al.
A genome-wide association study for genetic susceptibility to
Mycobacterium bovis infection in dairy cattle identifies a susceptibility QTL
on chromosome 23. Genet Sel Evol. 2016;48(1):19. https://doi.org/10.1186/
s12711-016-0197-x.
52. Raphaka K, Matika O, Sanchez-Molano E, Mrode R, Coffey MP, Riggio V, et al.
Genomic regions underlying susceptibility to bovine tuberculosis in
Holstein-Friesian cattle. BMC Genet. 2017;18(1):27. https://doi.org/10.1186/
s12863-017-0493-7.
53. Wilkinson S, Bishop SC, Allen AR, McBride SH, Skuce RA, Bermingham M,
et al. Fine-mapping host genetic variation underlying outcomes to
Mycobacterium bovis infection in dairy cows. BMC Genomics. 2017;18(1):477.
https://doi.org/10.1186/s12864-017-3836-x.
54. Tsairidou S, Allen AR, Pong-Wong R, McBride SH, Wright DM, Matika O, et al.
An analysis of effects of heterozygosity in dairy cattle for bovine
tuberculosis resistance. Anim Genet. 2018;49(2):1039. https://doi.org/1
0.1111/age.12637.
55. Ring SC, Purfield DC, Good M, Breslin P, Ryan E, Blom A, et al. Variance
components for bovine tuberculosis infection and multi-breed genome-wide
association analysis using imputed whole genome sequence data. PLoS One.
2019;14(2):e0212067. https://doi.org/10.1371/journal.pone.0212067.
56. Karczewski KJ, Snyder MP. Integrative omics for health and disease. Nat Rev
Genet. 2018;19(5):299310. https://doi.org/10.1038/nrg.2018.4.
57. Cano-Gamez E, Trynka G. From GWAS to function: using functional
genomics to identify the mechanisms underlying complex diseases. Front
Genet. 2020;11:424. https://doi.org/10.3389/fgene.2020.00424.
58. Kramer A, Green J, Pollard J Jr, Tugendreich S. Causal analysis approaches in
ingenuity pathway analysis. Bioinformatics. 2014;30(4):52330. https://doi.
org/10.1093/bioinformatics/btt703.
59. Qiagen: Qiagen Ingenuity Pathway Analysis Online Manual. 2019.
60. Foroushani AB, Brinkman FS, Lynn DJ. Pathway-GPS and SIGORA: identifying
relevant pathways based on the over-representation of their gene-pair
signatures. PeerJ. 2013;1:e229. https://doi.org/10.7717/peerj.229.
61. McKenzie AT, Katsyv I, Song WM, Wang M, Zhang B. DGCA: a
comprehensive R package for differential gene correlation analysis. BMC
Syst Biol. 2016;10(1):106. https://doi.org/10.1186/s12918-016-0349-1.
62. Breuer K, Foroushani AK, Laird MR, Chen C, Sribnaia A, Lo R, et al. InnateDB:
systems biology of innate immunity and beyondrecent updates and
continuing curation. Nucleic Acids Res. 2013;41(Database issue):D122833.
https://doi.org/10.1093/nar/gks1147.
63. Liu G, Ma H, Jiang L, Zhao Y. Allograft inflammatory factor-1 and its
immune regulation. Autoimmunity. 2007;40(2):95102. https://doi.org/10.1
080/08916930601083946.
64. Eskandari-Nasab E, Moghadampour M, Tahmasebi A, Asadi-Saghandi A.
Interleukin-17 a and F gene polymorphisms affect the risk of tuberculosis:
an updated meta-analysis. Indian J Tuberc. 2018;65(3):2007. https://doi.
org/10.1016/j.ijtb.2017.08.027.
65. Wang M, Xu G, Lü L, Xu K, Chen Y, Pan H, et al. Genetic polymorphisms of
IL-17A, IL-17F, TLR4 and miR-146a in association with the risk of pulmonary
tuberculosis. Sci Rep. 2016;6(1):28586. https://doi.org/10.1038/srep28586.
66. Brilha S, Wysoczanski R, Whittington AM, Friedland JS, Porter JC. Monocyte
adhesion, migration, and extracellular matrix breakdown are regulated by
integrin αVβ3inMycobacterium tuberculosis infection. J Immunol. 2017;
199(3):98291. https://doi.org/10.4049/jimmunol.1700128.
67. Seyrantepe V, Iannello A, Liang F, Kanshin E, Jayanth P, Samarani S, et al.
Regulation of phagocytosis in macrophages by neuraminidase 1. J Biol
Chem. 2010;285(1):20615. https://doi.org/10.1074/jbc.M109.055475.
68. Cheng Y, Huang C, Tsai HJ. Relationship of bovine TNF-αgene
polymorphisms with the risk of bovine tuberculosis in Holstein cattle. J Vet
Med Sci. 2016;78(5):72732. https://doi.org/10.1292/jvms.15-0506.
69. Satproedprai N, Wichukchinda N, Suphankong S, Inunchot W, Kuntima T,
Kumpeerasart S, et al. Diagnostic value of blood gene expression signatures
in active tuberculosis in Thais: a pilot study. Genes Immun. 2015;16(4):253
60. https://doi.org/10.1038/gene.2015.4.
70. Steiger J, Stephan A, Inkeles MS, Realegeno S, Bruns H, Kröll P, et al. Imatinib
triggers phagolysosome acidification and antimicrobial activity against
Mycobacterium bovis Bacille Calmette-Guérin in glucocorticoid-treated
human macrophages. J Immunol. 2016;197(1):22232. https://doi.org/10.404
9/jimmunol.1502407.
71. Gautier EL, Shay T, Miller J, Greter M, Jakubzick C, Ivanov S, et al. Gene-
expression profiles and transcriptional regulatory pathways that underlie the
identity and diversity of mouse tissue macrophages. Nat Immunol. 2012;
13(11):111828. https://doi.org/10.1038/ni.2419.
72. Lin J, Jiang Y, Liu D, Dai X, Wang M, Dai Y. Early secreted antigenic target of
6-kDa of Mycobacterium tuberculosis induces transition of macrophages into
epithelioid macrophages by downregulating iNOS / NO-mediated H3K27
trimethylation in macrophages. Mol Immunol. 2020;117:189200. https://doi.
org/10.1016/j.molimm.2019.11.013.
73. Wu HB, Wang YS, Zhang Y, Yang MQ, Lv JX, Liu J. TALE nickase-mediated
SP110 knockin endows cattle with increased resistance to tuberculosis. Proc
Natl Acad Sci U S A. 2015;112:E15309. https://doi.org/10.1073/pnas.1421
587112.
74. Schubert C. Systems immunology: complexity captured. Nature. 2011;
473(7345):1134. https://doi.org/10.1038/nj7345-113a.
75. Kidd BA, Peters LA, Schadt EE, Dudley JT. Unifying immunology with
informatics and multiscale biology. Nat Immunol. 2014;15(2):11827. https://
doi.org/10.1038/ni.2787.
76. Vodovotz Y, Xia A, Read EL, Bassaganya-Riera J, Hafler DA, Sontag E, et al.
Solving immunology? Trends Immunol. 2017;38(2):11627. https://doi.org/1
0.1016/j.it.2016.11.006.
77. Hao S, Yan KK, Ding L, Qian C, Chi H, Yu J. Network approaches for
dissecting the immune system. iScience. 2020;23(8):101354. https://doi.org/1
0.1016/j.isci.2020.101354.
78. Stelzer G, Rosen N, Plaschkes I, Zimmerman S, Twik M, Fishilevich S, et al. The
GeneCards suite: from gene data mining to disease genome sequence analyses.
Curr Protoc Bioinformatics. 2016;54:1.30.3131.30.33. https://doi.org/10.1002/cpbi.5.
79. Brace PT, Tezera LB, Bielecka MK, Mellows T, Garay D, Tian S, et al.
Mycobacterium tuberculosis subverts negative regulatory pathways in human
macrophages to drive immunopathology. PLoS Pathog. 2017;13(6):
e1006367. https://doi.org/10.1371/journal.ppat.1006367.
80. Cho JE, Park S, Lee H, Cho SN, Kim YS. Mycobacterium tuberculosis-induced
expression of granulocyte-macrophage colony stimulating factor is
mediated by PI3-K/MEK1/p38 MAPK signaling pathway. BMB Rep. 2013;
46(4):2138. https://doi.org/10.5483/BMBRep.2013.46.4.200.
81. Maiti D, Bhattacharyya A, Basu J. Lipoarabinomannan from Mycobacterium
tuberculosis promotes macrophage survival by phosphorylating bad
through a phosphatidylinositol 3-kinase/Akt pathway. J Biol Chem. 2001;
276(1):32933. https://doi.org/10.1074/jbc.M002650200.
82. Senthil Kumar A, Bansal K, Holla S, Verma-Kumar S, Sharma P, Balaji KN.
ESAT-6 induced COX-2 expression involves coordinated interplay between
PI3K and MAPK signaling. Mol Immunol. 2012;49(4):65563. https://doi.org/1
0.1016/j.molimm.2011.11.011.
83. Yang Y, Sun Y, Xu J, Bao K, Luo M, Liu X, et al. Epithelial cells attenuate toll-
like receptor-mediated inflammatory responses in monocyte-derived
macrophage-like cells to Mycobacterium tuberculosis by modulating the
PI3K/Akt/mTOR signaling pathway. Mediat Inflamm. 2018;2018:3685948.
https://doi.org/10.1155/2018/3685948.
84. Liu Y, Li JY, Chen ST, Huang HR, Cai H. The rLrp of Mycobacterium
tuberculosis inhibits proinflammatory cytokine production and
downregulates APC function in mouse macrophages via a TLR2-mediated
PI3K/Akt pathway activation-dependent mechanism. Cell Mol Immunol.
2016;13(6):72946. https://doi.org/10.1038/cmi.2015.58.
85. Tsairidou S, Allen A, Banos G, Coffey M, Anacleto O, Byrne AW, et al. Can we
breed cattle for lower bovine TB infectivity? Front Vet Sci. 2018;5:310.
https://doi.org/10.3389/fvets.2018.00310.
86. Cai Z, Guldbrandtsen B, Lund MS, Sahana G. Prioritizing candidate genes post-
GWAS using multiple sources of data for mastitis resistance in dairy cattle. BMC
Genomics. 2 018;19(1):656. https://doi.org/10.1186/s12864-018-5050-x.
87. Canovas A, Reverter A, DeAtley KL, Ashley RL, Colgrave ML, Fortes MR, et al.
Multi-tissue omics analyses reveal molecular regulatory networks for
Hall et al. BMC Genomics (2021) 22:343 Page 19 of 20
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
puberty in composite beef cattle. PLoS One. 2014;9(7):e102551. https://doi.
org/10.1371/journal.pone.0102551.
88. Deng T, Liang A, Liang S, Ma X, Lu X, Duan A, et al. Integrative analysis of
transcriptome and GWAS data to identify the hub genes associated with
milk yield trait in buffalo. Front Genet. 2019;10:36. https://doi.org/10.3389/
fgene.2019.00036.
89. Fang L, Sørensen P, Sahana G, Panitz F, Su G, Zhang S, et al. MicroRNA-
guided prioritization of genome-wide association signals reveals the
importance of microRNA-target gene networks for complex traits in cattle.
Sci Rep. 2018;8(1):9345. https://doi.org/10.1038/s41598-018-27729-y.
90. Yan Z, Huang H, Freebern E, Santos DJA, Dai D, Si J, et al. Integrating RNA-
Seq with GWAS reveals novel insights into the molecular mechanism
underpinning ketosis in cattle. BMC Genomics. 2020;21(1):489. https://doi.
org/10.1186/s12864-020-06909-z.
91. Fang L, Sahana G, Su G, Yu Y, Zhang S, Lund MS, et al. Integrating
sequence-based GWAS and RNA-seq provides novel insights into the
genetic basis of mastitis and milk production in dairy cattle. Sci Rep. 2017;
7(1):45560. https://doi.org/10.1038/srep45560.
92. Giuffra E, Tuggle CK. The FAANG consortium: functional annotation of
animal genomes (FAANG): current achievements and roadmap. Annu Rev
Anim Biosci. 2019;7(1):6588. https://doi.org/10.1146/annurev-animal-02051
8-114913.
93. Bishop TF, Van Eenennaam AL. Genome editing approaches to augment
livestock breeding programs. J Exp Biol. 2020;223(Pt Suppl 1):jeb207159.
https://doi.org/10.1242/jeb.207159.
94. Banos G, Winters M, Mrode R, Mitchell AP, Bishop SC, Woolliams JA, et al.
Genetic evaluation for bovine tuberculosis resistance in dairy cattle. J Dairy
Sci. 2017;100(2):127281. https://doi.org/10.3168/jds.2016-11897.
95. Andrews S. FastQC: a quality control tool for high throughput sequence
data. Cambridge: Babraham Research Campus: Bioinformatics Group,
Babraham Institute; 2019.
96. Rosen BD, Bickhart DM, Schnabel RD, Koren S, Elsik CG, Tseng E, et al. De novo
assembly of the cattle reference genome with single-molecule sequencing.
Gigascience. 2020;9(3):giaa021. https://doi.org/10.1093/gigascience/giaa021.
97. Dobin A, Davis CA, Schlesinger F, Drenkow J, Zaleski C, Jha S, et al. STAR:
ultrafast universal RNA-seq aligner. Bioinformatics. 2013;29(1):1521. https://
doi.org/10.1093/bioinformatics/bts635.
98. Liao Y, Smyth GK, Shi W. featureCounts: an efficient general purpose
program for assigning sequence reads to genomic features. Bioinformatics.
2014;30(7):92330. https://doi.org/10.1093/bioinformatics/btt656.
99. R Core Team. R: a language and environment for statistical computing.
Vienna: R Foundation for statistical Computing; 2019.
100. Carlson M. GO.db: A set of annotation maps describing the entire Gene
Ontology. R package version 3.8.2; 2019.
101. Durinck S, Spellman PT, Birney E, Huber W. Mapping identifiers for the
integration of genomic datasets with the R/bioconductor package biomaRt.
Nat Protoc. 2009;4(8):118491. https://doi.org/10.1038/nprot.2009.97.
102. Love MI, Huber W, Anders S. Moderated estimation of fold change and
dispersion for RNA-seq data with DESeq2. Genome Biol. 2014;15(12):550.
https://doi.org/10.1186/s13059-014-0550-8.
103. Cook RD. Detection of influential observation in linear-regression.
Technometrics. 1977;19(1):158. https://doi.org/10.2307/1268249.
104. Benjamini Y, Hochberg Y. Controlling the false discovery rate - a practical
and powerful approach to multiple testing. J R Stat Soc Ser B Method. 1995;
57(1):289300. https://doi.org/10.1111/j.2517-6161.1995.tb02031.x.
105. Timmons JA, Szkop KJ, Gallagher IJ. Multiple sources of bias confound
functional enrichment analysis of global -omics data. Genome Biol. 2015;
16(1):186. https://doi.org/10.1186/s13059-015-0761-7.
106. Shannon P, Markiel A, Ozier O, Baliga NS, Wang JT, Ramage D, et al.
Cytoscape: a software environment for integrated models of biomolecular
interaction networks. Genome Res. 2003;13(11):2498504. https://doi.org/1
0.1101/gr.1239303.
107. Song WM, Zhang B. Multiscale embedded gene co-expression network
analysis. PLoS Comput Biol. 2015;11(11):e1004574. https://doi.org/10.1371/
journal.pcbi.1004574.
108. van Dam S, Vosa U, van der Graaf A, Franke L, de Magalhaes JP. Gene co-
expression analysis for functional classification and gene-disease predictions.
Brief Bioinform. 2018;19(4):57592. https://doi.org/10.1093/bib/bbw139.
109. Schaefer RJ, Michno JM, Myers CL. Unraveling gene function in agricultural
species using gene co-expression networks. Biochimica et biophysica acta
Gene regulatory mechanisms. 2017;1860(1):5363. https://doi.org/10.1016/j.
bbagrm.2016.07.016.
110. Gormley C, Tong Z. Elasticsearch: the definitive guide. 1st ed. Sebastopol:
OReilly Media, Inc; 2015.
111. Yates AD, Achuthan P, Akanni W, Allen J, Allen J, Alvarez-Jarreta J, et al.
Ensembl 2020. Nucleic Acids Res. 2020;48(D1):D682d688. https://doi.org/1
0.1093/nar/gkz966.
112. Ideker T, Ozier O, Schwikowski B, Siegel AF. Discovering regulatory and
signalling circuits in molecular interaction networks. Bioinformatics. 2002;
18(Suppl 1):S23340. https://doi.org/10.1093/bioinformatics/18.suppl_1.S233.
113. Strimmer K. Fdrtool: a versatile R package for estimating local and tail area-
based false discovery rates. Bioinformatics. 2008;24(12):14612. https://doi.
org/10.1093/bioinformatics/btn209.
114. Zimin AV, Delcher AL, Florea L, Kelley DR, Schatz MC, Puiu D, et al. A whole-
genome assembly of the domestic cow, Bos taurus. Genome Biol. 2009;
10(4):R42. https://doi.org/10.1186/gb-2009-10-4-r42.
PublishersNote
Springer Nature remains neutral with regard to jurisdictional claims in
published maps and institutional affiliations.
Hall et al. BMC Genomics (2021) 22:343 Page 20 of 20
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
... Previous studies, including those performed by our research group, have demonstrated that the bovine and human alveolar macrophage (bAM and hAM) transcriptomes are extensively reprogrammed in response to infection with M. bovis and M. tuberculosis [36][37][38][39][40][41][42][43]. These studies have also highlighted a complex system of gene expression regulation that drives host-pathogen interactions and innate immune response pathway execution, which are functionally associated with many macrophage processes that control or eliminate intracellular microbes. ...
... All data-intensive computational procedures were performed on a 36-core/72-thread compute server (2 × Intel® Xeon® CPU E5-2697 v4 processors, 2.30 GHz with 18 cores each), with 512 GB of RAM, 96 TB SAS storage (12 × 8 TB at 7200 rpm), 480 GB SSD storage, and with Ubuntu Linux OS (version 18.04 LTS).The complete computational and bioinformatics workflow is available with additional information as a public GitHub repository (github.com/ThomasHall1688/Bovine_multi-omic_integration). The individual components of the experimental and computational workflows are described below and these methodologies and procedures used are modified from those detailed previously by us [41]. ...
... To facilitate integration of GWAS data with gene sets generated from functional genomics data analyses, the gwinteR software package was used (github.com/ThomasHall1688/gwinteR) [41]. The gwinteR tool can be used to test the hypothesis that a specific set of genes is enriched for signal in a GWAS data set relative to the genomic background. ...
Article
Full-text available
Mycobacterium tuberculosis, the causative agent of human tuberculosis (hTB), is a close evolutionary relative of Mycobacterium bovis, which causes bovine tuberculosis (bTB), one of the most damaging infectious diseases to livestock agriculture. Previous studies have shown that the pathogenesis of bTB disease is comparable to hTB disease, and that the bovine and human alveolar macrophage (bAM and hAM, respectively) transcriptomes are extensively reprogrammed in response to infection with these intracellular mycobacterial pathogens. In this study, a multi-omics integrative approach was applied with functional genomics and GWAS data sets across the two primary hosts (Bos taurus and Homo sapiens) and both pathogens (M. bovis and M. tuberculosis). Four different experimental infection groups were used: 1) bAM infected with M. bovis, 2) bAM infected with M. tuberculosis, 3) hAM infected with M. tuberculosis, and 4) human monocyte-derived macrophages (hMDM) infected with M. tuberculosis. RNA-seq data from these experiments 24 h post-infection (24 hpi) was analysed using three computational pipelines: 1) differentially expressed genes, 2) differential gene expression interaction networks, and 3) combined pathway analysis. The results were integrated with high-resolution bovine and human GWAS data sets to detect novel quantitative trait loci (QTLs) for resistance to mycobacterial infection and resilience to disease. This revealed common and unique response macrophage pathways for both pathogens and identified 32 genes (12 bovine and 20 human) significantly enriched for SNPs associated with disease resistance, the majority of which encode key components of the NF-κB signalling pathway and that also drive formation of the granuloma.
... Li CJ has been involved in studies in different perspectives ranging from transcriptomic profiling of organs in the digestive system such as rumen (47), duodenum (48) and butyrate-treated cells (49) in cattle, fat tissue development (50), epigenetic studies in which DNA methylation profiles (51) are monitored according to tissues and studies in which regulatory variants are examined in multiple tissue atlases (52). Brown JA, on the other hand, has been involved in reproductive studies and fertility genomics studies (53,54) as well as some studies examining the responses in peripheral blood cells of cattle affected with Mycobacterium avium and bovis (11,(55)(56)(57). It was determined that the identified authors have been working and publishing in the relevant research field for many years. ...
Article
Full-text available
The aim of this study is to examine the development of research articles on "gene expression and RNA-seq" in cattle species in the Web of Science (WOS) database between 2010 and 2023 using bibliometric mapping methods. Initially, 500 articles were screened using relevant keywords, and 353 articles suitable for analysis were analyzed using the Bibliometrix R package's shiny web application and some analyses were conducted using the VOSviewer application. The conducted analyses included subheadings such as main data information, annual scientific production, countries, and institution analysis, bibliographic coupling with sources, Bradford analysis, Lotka’s law analyasis, highly cited articles, and most influential authors. According to the results obtained after the analysis, it was determined that the institution conducting the most studies in the relevant field is "Universidade De Sao Paulo", the country with the most publications is the USA, and the most published journal is "BMC Genomics". Keyword analysis revealed that the trending topics in recent years are mastitis, dairy cattle farming, and heat stress. The studies were categorized into different clusters related to the reproductive system, immune system and diseases, meat and dairy cattle production. It is recommended that researchers planning to work in this research area on cattle species should design their research, determine the journal to be published, or establish institutional connections by examining the reported study and planning accordingly.
... Adapter sequence reads were removed and quality trimming was carried out using the fastP adapter removal software (28). RNA seq data was processed as described previously (29). Briefly, 50 million, 2 × 150 bp paired-end sequence reads were generated from each RNA sample. ...
Article
Full-text available
Bovine tuberculosis (bTB), caused by infection with Mycobacterium bovis, continues to cause significant issues for the global agriculture industry as well as for human health. An incomplete understanding of the host immune response contributes to the challenges of control and eradication of this zoonotic disease. In this study, high-throughput bulk RNA sequencing (RNA-seq) was used to characterise differential gene expression in γδ T cells – a subgroup of T cells that bridge innate and adaptive immunity and have known anti-mycobacterial response mechanisms. γδ T cell subsets are classified based on expression of a pathogen-recognition receptor known as Workshop Cluster 1 (WC1) and we hypothesised that bTB disease may alter the phenotype and function of specific γδ T cell subsets. Peripheral blood was collected from naturally M. bovis-infected (positive for single intradermal comparative tuberculin test (SICTT) and IFN-γ ELISA) and age- and sex-matched, non-infected control Holstein-Friesian cattle. γδ T subsets were isolated using fluorescence activated cell sorting (n = 10–12 per group) and high-quality RNA extracted from each purified lymphocyte subset (WC1.1⁺, WC1.2⁺, WC1⁻ and γδ⁻) was used to generate transcriptomes using bulk RNA-seq (n = 6 per group, representing a total of 48 RNA-seq libraries). Relatively low numbers of differentially expressed genes (DEGs) were observed between most cell subsets; however, 189 genes were significantly differentially expressed in the M. bovis-infected compared to the control groups for the WC1.1⁺ γδ T cell compartment (absolute log2 FC ≥ 1.5 and FDR P adj. ≤ 0.1). The majority of these DEGs (168) were significantly increased in expression in cells from the bTB+ cattle and included genes encoding transcription factors (TBX21 and EOMES), chemokine receptors (CCR5 and CCR7), granzymes (GZMA, GZMM, and GZMH) and multiple killer cell immunoglobulin-like receptor (KIR) proteins indicating cytotoxic functions. Biological pathway overrepresentation analysis revealed enrichment of genes with multiple immune functions including cell activation, proliferation, chemotaxis, and cytotoxicity of lymphocytes. In conclusion, γδ T cells have important inflammatory and regulatory functions in cattle, and we provide evidence for preferential differential activation of the WC1.1⁺ specific subset in cattle naturally infected with M. bovis.
... To search for the terms "clozapine" and "olanzapine", the GeneCards Database (https://www.genecards.org/, Version 5.13) was used with a Relevance score[?]1 as the standard screening target [28]. ...
Preprint
Erchen Decoction(ECD) has been clinically verified to be effective for treating metabolic syndrome(MetS) induced by second-generation antipsychotics(SGAs). Network pharmacology analysis results exhibited that 129 active components and 221 potential targets of ECD, 1027 targets related to MetS, and 361 targets of clozapine and olanzapine were gained after screening. Then, 79 intersection targets of ECD and MetS were obtained by Venn analysis to construct a PPI network. However, it could not explain the potential target and mechanism of ECD to improve the metabolic disorder caused by SGAs. GO function and KEGG pathway analyses were performed on 23 common targets of ECD, clozapine, olanzapine, and metabolic syndrome, and visualization networks were constructed. The results revealed the potential signaling pathway of ECD for treating drug-induced MetS, especially the AMPK signaling pathway. The visualization network reflects that ADRA1A, AHR, NR3C1, and SLC6A4 may be the core targets. Ultimately, molecular docking results displayed that active components of ECD naringenin, baicalein, and quercetin had a good binding activity with NR3C1 and SLC6A4 targets.
... An r software package, gwinteR (Hall et al., 2021), was used to integrate the SNP data arising from the CSS analyses with HEG sets generated from RNA-seq data. The gwinteR software tool works as follows: (1) a set of significant and non-significant SNPs (named the | 5 GENES FOR EQUINE BEHAVIOUR target SNP set) is collated across all genes in a specific gene set at increasing genomic intervals upstream and downstream from each gene inclusive of the coding sequence (e.g, ±10 kb, ±20 kb, ±30 kb, …, ± 100 kb); (2) for each genomic region, a null distribution of 1000 SNP sets, each of which contains the same number of total significant and non-significant combined SNPs as the target SNP set, is generated by resampling with replacement from the search space of the total population of SNPs in the CSS data set; (3) the nominal (uncorrected) CSS p-values for the target SNP set and the null distribution SNP sets are converted to local false discovery rate-adjusted p-values (p adj. ) using the fdrtool r package (version 1.2.15) (Strimmer, 2008); (4) a permuted p-value (P perm. ) to test the primary hypothesis for each observed genomic interval target SNP set is generated based on the proportion of permuted random SNP sets where the same or a larger number of SNPs exhibiting significant q-values (e.g. ...
Article
Full-text available
Behavioural plasticity enables horses entering an exercise training programme to adapt with reduced stress. We characterised SNPs associated with behaviour in yearling Thoroughbred horses using genomics analyses for two phenotypes: (1) handler‐assessed coping with early training events [coping] (n = 96); and (2) variation in salivary cortisol concentration at the first backing event [cortisol] (n = 34). Using RNA‐seq derived gene expression data for amygdala and hippocampus tissues from n = 2 Thoroughbred stallions, we refined the SNPs to those with functional relevance to behaviour by cross‐referencing to the 500 most highly expressed genes in each tissue. The SNPs of high significance (q < 0.01) were in proximity to genes (coping – GABARAP, NDM, OAZ1, RPS15A, SPARCL1, VAMP2; cortisol – CEBPA, COA3, DUSP1, HNRNPH1, RACK1) with biological functions in social behaviour, autism spectrum disorder, suicide, stress‐induced anxiety and depression, Alzheimer's disease, neurodevelopmental disorders, neuroinflammatory disease, fear‐induced behaviours and alcohol and cocaine addiction. The strongest association (q = 0.0002) was with NDN, a gene previously associated with temperament in cattle. This approach highlights functionally relevant genes in the behavioural adaptation of Thoroughbred horses that will contribute to the development of genetic markers to improve racehorse welfare.
Preprint
Full-text available
Background Erchen Decoction (ECD) has garnered clinical recognition for its efficacy in managing metabolic syndrome (MetS) induced by second-generation antipsychotics (SGAs). Despite its therapeutic success, the intricate pharmacological mechanisms underpinning ECD's action remain to be elucidated. Methods To predict protein interactions within the pharmacological framework of ECD, we constructed a protein-protein interaction (PPI) network using the shared targets between ECD components and MetS. Subsequently, we conducted Gene Ontology (GO) function and Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway enrichment analyses on the common targets of ECD, SGAs, and MetS. A component-core target visualization network was developed for clearer representation. Molecular docking simulations were performed using Autodock Vina 1.2.0, and corroborative animal experiments were undertaken to validate ECD's mechanisms of action. Results Our research identified 221 potential targets of ECD, 1027 MetS-related targets, and 361 targets associated with clozapine and olanzapine. A PPI network was established from 79 intersecting targets of ECD and MetS. Analyses of 23 shared targets among ECD, SGAs, and MetS highlighted the AMPK pathway as potentially pivotal in the treatment of SGAs-induced MetS. The visualization network suggested ADRA1A, AHR, NR3C1, and SLC6A4 as core targets. In silico molecular docking revealed strong binding affinities of naringenin, baicalein, and quercetin in ECD with the NR3C1 and SLC6A4 targets. In vivo, ECD mitigated olanzapine-induced MetS in rats, accompanied by reduced expression of AMPK and SREBP1 in the liver. Conclusions Our findings propose that ECD may exert its therapeutic effects by targeting NR3C1 and SLC6A4 and modulating the AMPK pathway in the treatment of MetS induced by SGAs. These insights are in congruence with the results obtained from molecular docking and animal model studies.
Preprint
Full-text available
Bovine tuberculosis (bTB), caused by infection with Mycobacterium bovis, continues to cause significant issues for the global agriculture industry as well as for human health. An incomplete understanding of the host immune response contributes to the challenges of control and eradication of this zoonotic disease. In this study, high-throughput bulk RNA sequencing (RNA-seq) was used to characterize differential gene expression in GD T cells- a subgroup of T cells that bridge innate and adaptive immunity and have known anti-mycobacterial response mechanisms. GD T cell subsets are classified based on expression of a pathogen-recognition receptor known as Workshop Cluster 1 (WC1) and we hypothesised that bTB disease may alter the phenotype and function of specific GD T cell subsets. Peripheral blood was collected from naturally M. bovis-infected (positive for single intradermal comparative tuberculin test (SICTT) and IFN-Gamma ELISA) and age- and sex-matched, non-infected control Holstein-Friesian cattle. GD T subsets were isolated using fluorescence activated cell sorting (n = 10-12 per group) and high-quality RNA extracted from each purified lymphocyte subset (WC1.1+, WC1.2+, WC1- and GD TCR-) was used to generate transcriptomes using bulk RNA-seq (n = 6 per group, representing a total of 48 RNA-seq libraries). Relatively low numbers of differentially expressed genes (DEGs) were observed between most cell subsets; however, 189 genes were significantly differentially expressed in the M. bovis-infected compared to the control groups for the WC1.1+ GD T cell compartment (absolute log2 FC ≥ 1.5 and FDR Padj. ≤ 0.1). The majority of these DEGs (168) were significantly increased in expression in cells from the bTB+ cattle and included genes encoding transcription factors (TBX21 and EOMES), chemokine receptors (CCR5 and CCR7), granzymes (GZMA, GZMM, and GZMH) and multiple killer cell immunoglobulin-like receptor (KIR) proteins indicating cytotoxic functions. Biological pathway overrepresentation analysis revealed enrichment of genes with multiple immune functions including cell activation, proliferation, chemotaxis, and cytotoxicity of lymphocytes. In conclusion, WC1.1+ GD T cells have been proposed as major regulatory cell subset in cattle, and we provide evidence for preferential differential activation of this specific subset in cattle naturally infected with M. bovis.
Preprint
Full-text available
Mycobacterium tuberculosis, the causative agent of human tuberculosis (hTB), is currently classed as the thirteenth leading cause of death worldwide. Mycobacterium bovis, a close evolutionary relative of M. tuberculosis, causes bovine tuberculosis (bTB) and is one of the most damaging infectious diseases to livestock agriculture. Previous studies have shown that the pathogenesis of bTB disease is comparable to hTB disease, and that the bovine and human alveolar macrophage (bAM and hAM, respectively) transcriptomes are extensively reprogrammed in response to infection with these intracellular mycobacterial pathogens. However, although M. bovis and M. tuberculosis share over 99% identity at the genome level, the innate immune responses to these pathogens have been shown to be different in human or cattle hosts. In this study, a multi-omics integrative approach was applied to encompass functional genomics and GWAS data sets across the two primary hosts (Bos taurus and Homo sapiens) and both pathogens (M. bovis and M. tuberculosis). Four different experimental infection groups were used, each with parallel non-infected control cells: 1) bAM infected with M. bovis, 2) bAM infected with M. tuberculosis, 3) hAM infected with M. tuberculosis, and 4) human monocyte-derived macrophages (hMDM) infected with M. tuberculosis. RNA-seq data from these experiments 24 hours post-infection (24 hpi) was analysed using three separate computational pipelines: 1) differentially expressed genes, 2) differential gene expression interaction networks, and 3) combined pathway analysis. The results of these analyses were then integrated with high-resolution bovine and human GWAS data sets to detect novel quantitative trait loci (QTLs) for resistance to mycobacterial infection and resilience to disease. Results from this study revealed common and unique response macrophage pathways for both pathogens and identified 32 genes (12 bovine and 20 human) significantly enriched for SNPs associated with disease resistance, the majority of which encode key components of the NF-κB signalling pathway and that also drive formation of the granuloma.
Article
Full-text available
During mycobacterial infections, pathogenic mycobacteria manipulate both host immune and stromal cells to establish and maintain a productive infection. In humans, non-human primates, and zebrafish models of infection, pathogenic mycobacteria produce and modify the specialized lipid trehalose 6,6′-dimycolate (TDM) in the bacterial cell envelope to drive host angiogenesis toward the site of forming granulomas, leading to enhanced bacterial growth. Here, we use the zebrafish-Mycobacterium marinum infection model to define the signaling basis of the host angiogenic response. Through intravital imaging and cell-restricted peptide-mediated inhibition, we identify macrophage-specific activation of NFAT signaling as essential to TDM-mediated angiogenesis in vivo. Exposure of cultured human cells to Mycobacterium tuberculosis results in robust induction of VEGFA, which is dependent on a signaling pathway downstream of host TDM detection and culminates in NFATC2 activation. As granuloma-associated angiogenesis is known to serve bacterial-beneficial roles, these findings identify potential host targets to improve tuberculosis disease outcomes.
Article
Full-text available
Selection for system-wide morphological, physiological, and metabolic adaptations has led to extreme athletic phenotypes among geographically diverse horse breeds. Here, we identify genes contributing to exercise adaptation in racehorses by applying genomics approaches for racing performance, an end-point athletic phenotype. Using an integrative genomics strategy to first combine population genomics results with skeletal muscle exercise and training transcriptomic data, followed by whole-genome resequencing of Asian horses, we identify protein-coding variants in genes of interest in galloping racehorse breeds (Arabian, Mongolian and Thoroughbred). A core set of genes, G6PC2, HDAC9, KTN1, MYLK2, NTM, SLC16A1 and SYNDIG1 , with central roles in muscle, metabolism, and neurobiology, are key drivers of the racing phenotype. Although racing potential is a multifactorial trait, the genomic architecture shaping the common athletic phenotype in horse populations bred for racing provides evidence for the influence of protein-coding variants in fundamental exercise-relevant genes. Variation in these genes may therefore be exploited for genetic improvement of horse populations towards specific types of racing.
Article
Full-text available
Background: Ketosis is a common metabolic disease during the transition period in dairy cattle, resulting in long-term economic loss to the dairy industry worldwide. While genetic selection of resistance to ketosis has been adopted by many countries, the genetic and biological basis underlying ketosis is poorly understood. Results: We collected a total of 24 blood samples from 12 Holstein cows, including 4 healthy and 8 ketosis-diagnosed ones, before (2 weeks) and after (5 days) calving, respectively. We then generated RNA-Sequencing (RNA-Seq) data and seven blood biochemical indicators (bio-indicators) from leukocytes and plasma in each of these samples, respectively. By employing a weighted gene co-expression network analysis (WGCNA), we detected that 4 out of 16 gene-modules, which were significantly engaged in lipid metabolism and immune responses, were transcriptionally (FDR < 0.05) correlated with postpartum ketosis and several bio-indicators (e.g., high-density lipoprotein and low-density lipoprotein). By conducting genome-wide association signal (GWAS) enrichment analysis among six common health traits (ketosis, mastitis, displaced abomasum, metritis, hypocalcemia and livability), we found that 4 out of 16 modules were genetically (FDR < 0.05) associated with ketosis, among which three were correlated with postpartum ketosis based on WGCNA. We further identified five candidate genes for ketosis, including GRINA, MAF1, MAFA, C14H8orf82 and RECQL4. Our phenome-wide association analysis (Phe-WAS) demonstrated that human orthologues of these candidate genes were also significantly associated with many metabolic, endocrine, and immune traits in humans. For instance, MAFA, which is involved in insulin secretion, glucose response, and transcriptional regulation, showed a significantly higher association with metabolic and endocrine traits compared to other types of traits in humans. Conclusions: In summary, our study provides novel insights into the molecular mechanism underlying ketosis in cattle, and highlights that an integrative analysis of omics data and cross-species mapping are promising for illustrating the genetic architecture underpinning complex traits.
Article
Full-text available
The immune system is a complex biological network composed of hierarchically organized genes, proteins, and cellular components that combat external pathogens and monitor the onset of internal disease. To meet and ultimately defeat these challenges, the immune system orchestrates an exquisitely complex interplay of numerous cells, often with highly specialized functions, in a tissue-specific manner. One of the major methodologies of systems immunology is to measure quantitatively the components and interaction levels in the immunologic networks to construct a computational network and predict the response of the components to perturbations. The recent advances in high-throughput sequencing techniques have provided us with a powerful approach to dissecting the complexity of the immune system. Here we summarize the latest progress in integrating omics data and network approaches to construct networks and to infer the underlying signaling and transcriptional landscape, as well as cell–cell communication, in the immune system, with a focus on hematopoiesis, adaptive immunity, and tumor immunology. Understanding the network regulation of immune cells has provided new insights into immune homeostasis and disease, with important therapeutic implications for inflammation, cancer, and other immune-mediated disorders.
Article
Full-text available
Genome-wide association studies (GWAS) have successfully mapped thousands of loci associated with complex traits. These associations could reveal the molecular mechanisms altered in common complex diseases and result in the identification of novel drug targets. However, GWAS have also left a number of outstanding questions. In particular, the majority of disease-associated loci lie in non-coding regions of the genome and, even though they are thought to play a role in gene expression regulation, it is unclear which genes they regulate and in which cell types or physiological contexts this regulation occurs. This has hindered the translation of GWAS findings into clinical interventions. In this review we summarize how these challenges have been addressed over the last decade, with a particular focus on the integration of GWAS results with functional genomics datasets. Firstly, we investigate how the tissues and cell types involved in diseases can be identified using methods that test for enrichment of GWAS variants in genomic annotations. Secondly, we explore how to find the genes regulated by GWAS loci using methods that test for colocalization of GWAS signals with molecular phenotypes such as quantitative trait loci (QTLs). Finally, we highlight potential future research avenues such as integrating GWAS results with single-cell sequencing read-outs, designing functionally informed polygenic risk scores (PRS), and validating disease associated genes using genetic engineering. These tools will be crucial to identify new drug targets for common complex diseases.
Article
Full-text available
Background Major advances in selection progress for cattle have been made following the introduction of genomic tools over the past 10–12 years. These tools depend upon the Bos taurus reference genome (UMD3.1.1), which was created using now-outdated technologies and is hindered by a variety of deficiencies and inaccuracies. Results We present the new reference genome for cattle, ARS-UCD1.2, based on the same animal as the original to facilitate transfer and interpretation of results obtained from the earlier version, but applying a combination of modern technologies in a de novo assembly to increase continuity, accuracy, and completeness. The assembly includes 2.7 Gb and is >250× more continuous than the original assembly, with contig N50 >25 Mb and L50 of 32. We also greatly expanded supporting RNA-based data for annotation that identifies 30,396 total genes (21,039 protein coding). The new reference assembly is accessible in annotated form for public use. Conclusions We demonstrate that improved continuity of assembled sequence warrants the adoption of ARS-UCD1.2 as the new cattle reference genome and that increased assembly accuracy will benefit future research on this species.
Article
Full-text available
The prospect of genome editing offers a number of promising opportunities for livestock breeders. Firstly, these tools can be used in functional genomics to elucidate gene function, and identify causal variants underlying monogenic traits. Secondly, they can be used to precisely introduce useful genetic variation into structured livestock breeding programs. Such variation may include repair of genetic defects, the inactivation of undesired genes, and the moving of useful alleles and haplotypes between breeds in the absence of linkage drag. Editing could also be used to accelerate the rate of genetic progress by enabling the replacement of the germ cell lineage of commercial breeding animals with cells derived from genetically elite lines. In the future, editing may also provide a useful complement to evolving approaches to decrease the length of the generation interval through in vitro generation of gametes. For editing to be adopted, it will need to seamlessly integrate with livestock breeding schemes. This will likely involve introducing edits into multiple elite animals to avoid genetic bottlenecks. It will also require editing of different breeds and lines to maintain genetic diversity, and enable structured cross-breeding. This requirement is at odds with the process-based trigger and event-based regulatory approach that has been proposed for the products of genome editing by several countries. In the absence of regulatory harmony, researchers in some countries will have the ability to use genome editing in food animals, while others will not, resulting in disparate access to these tools, and ultimately the potential for global trade disruptions.
Article
Full-text available
Bovine tuberculosis is caused by infection with Mycobacterium bovis, which can also cause disease in a range of other mammals, including humans. Alveolar macrophages are the key immune effector cells that first encounter M. bovis and how the macrophage epigenome responds to mycobacterial pathogens is currently not well understood. Here, we have used chromatin immunoprecipitation sequencing (ChIP-seq), RNA-seq and miRNA-seq to examine the effect of M. bovis infection on the bovine alveolar macrophage (bAM) epigenome. We show that H3K4me3 is more prevalent, at a genome-wide level, in chromatin from M. bovis-infected bAM compared to control non-infected bAM; this was particularly evident at the transcriptional start sites of genes that determine programmed macrophage responses to mycobacterial infection (e.g. M1/M2 macrophage polarisation). This pattern was also supported by the distribution of RNA Polymerase II (Pol II) ChIP-seq results, which highlighted significantly increased transcriptional activity at genes demarcated by permissive chromatin. Identification of these genes enabled integration of high-density genome-wide association study (GWAS) data, which revealed genomic regions associated with resilience to infection with M. bovis in cattle. Through integration of these data, we show that bAM transcriptional reprogramming occurs through differential distribution of H3K4me3 and Pol II at key immune genes. Furthermore, this subset of genes can be used to prioritise genomic variants from a relevant GWAS data set.
Article
Full-text available
Background: Zoonotic tuberculosis (zTB) accounts for 1.4% of the global tuberculosis burden, with the largest disease burden in low- and middle-income countries (LMICs). These populations have increased exposure to zTB due to livestock rearing practices and raw dairy consumption. This qualitative systematic literature review evaluates the quality of the literature that examines the association between human zTB in LMICs and frequent exposure to livestock and livestock products and summarizes current gaps in laboratory detection methods. Methods: The Navigation Guide, a systematic review framework utilized to assess environmental health exposures, was used to conduct this literature review. Peer-reviewed research articles were selected and evaluated for risk of bias and quality of evidence. Only studies conducted in LMICs that mentioned livestock or livestock product exposure and had a confirmed diagnosis were eligible. Results: Fifteen studies met the inclusion criteria. Reported prevalence of human zTB ranged from 0% to 28%, with cattle and raw dairy the primary exposures. All confirmed zTB cases were Mycobacterium bovis. Eight of the 15 studies included livestock sampling, predominantly cattle and reported prevalence of zTB between 0% and 23%. Laboratory methods used included nearly a dozen different culture methods and a variety of molecular methods, some of which are not appropriate for zTB. Conclusions: This review revealed the need for appropriate and standardized laboratory diagnostic methods, and large prospective studies of at-risk populations to determine exposures that lead to an increased risk of tuberculosis conversion/infection to better understand the true burden of disease. Standardized, easy to implement laboratory diagnostics is an imperative focus for this scientific field to better identify the burden of zTB. Future studies pairing livestock and human subjects will allow better characterization of the high zTB transmission areas for targeted control and prevention programmes.
Article
Full-text available
Tuberculosis (TB), an infectious disease caused by Mycobacterium tuberculosis , is one of the top ten infectious diseases worldwide, and is the leading cause of morbidity from a single infectious agent. M. tuberculosis can cause infection in several species of animals in addition to humans as the natural hosts. Although animal models of TB disease cannot completely simulate the occurrence and development of human TB, they play an important role in studying the pathogenesis, immune responses, and pathological changes as well as for vaccine research. This review summarizes the commonly employed animal models, including mouse, guinea pig, rabbit, rat, goat, cattle, and nonhuman primates, and their characteristics as used in TB vaccine research, and provides a basis for selecting appropriate animal models according to specific research needs. Furthermore, some of the newest animal models used for TB vaccine research (such as humanized animal models, zebrafish, Drosophila , and amoeba) are introduced, and their characteristics and research progress are discussed.
Article
Full-text available
The Ensembl (https://www.ensembl.org) is a system for generating and distributing genome annotation such as genes, variation, regulation and comparative genomics across the vertebrate subphylum and key model organisms. The Ensembl annotation pipeline is capable of integrating experimental and reference data from multiple providers into a single integrated resource. Here, we present 94 newly annotated and re-annotated genomes, bringing the total number of genomes offered by Ensembl to 227. This represents the single largest expansion of the resource since its inception. We also detail our continued efforts to improve human annotation, developments in our epigenome analysis and display, a new tool for imputing causal genes from genome-wide association studies and visualisation of variation within a 3D protein model. Finally, we present information on our new website. Both software and data are made available without restriction via our website, online tools platform and programmatic interfaces (available under an Apache 2.0 license) and data updates made available four times a year.
Article
Background: Tuberculosis (TB) is a chronic infectious disease caused by Mycobacterium tuberculosis (Mtb). Granuloma is a pathological feature of tuberculosis and is a tight immune cell aggregation caused by Mtb. The main constituent cells are macrophages and their derivative cells including epithelioid macrophages. However, the molecular mechanism of the transition has not been reported. The purpose of this study was to investigate whether early secreted antigenic target of 6-kDa (ESAT6) can induce the transition of bone marrow-derived macrophages (BMDMs) into epithelioid macrophages and its possible molecular mechanism. Methods: The recombinant ESAT6 protein was obtained from E.coli carrying esat6 gene after isopropyl β-d-thiogalactopyranoside (IPTG) induction. BMDMs were isolated from bone marrow of mice hind legs. Cells viability was detected by Cell Counting Kit 8 (CCK8) assays. The expression levels of mRNA and proteins were detected by qPCR and Western blot, or evaluated by flow cytometry. The expression level of nitric oxide (NO) was measured with a nitric oxide indicator. Results: ESAT6 could significantly induce mRNA and protein expression levels of a group of epithelioid macrophages marker molecules (EMMMs), including E-cadherin, junction plakoglobin, ZO1, desmoplakin, desmoglein3 and catenin porteins, in BMDMs. These events could be abrogated in macrophage from TLR2 deficiency mice. ESAT6 could also markedly induce iNOS/NO production that could significantly inhibit trimethylation of H3K27 in the cells. ESAT6-induced expressions of epithelioid macrophages marker molecules were significantly inhibited in the presence of H3K27 histone demethylase inhibitor GSK J1. Furthermore, ROS scavenging agent N,N'-Dimethylthiourea (DMTU) could markedly inhibit the transition induced by ESAT6 in macrophages. Conclusion: This study demonstrates that ESAT6 bound with TLR2 can activate iNOS/NO and ROS signalings to reduce the trimethylation of H3K27 resulting in the increment of EMMMs expression that is beneficial to the transition of macrophages into epithelioid macrophages. However, hypoxia can inhibit this transition event. This study has provided new evidence of pathogenesis of granuloma caused by Mtb and also proposed new ideas for the treatment of TB.