ArticlePDF Available

Differential immune response modulation in early Leishmania amazonensis infection of BALB/c and C57BL/6 macrophages based on transcriptome profiles

Authors:

Abstract and Figures

The fate of Leishmania infection can be strongly influenced by the host genetic background. In this work, we describe gene expression modulation of the immune system based on dual global transcriptome profiles of bone marrow-derived macrophages (BMDMs) from BALB/c and C57BL/6 mice infected with Leishmania amazonensis. A total of 12,641 host transcripts were identified according to the alignment to the Mus musculus genome. Differentially expressed genes (DEGs) profiling revealed a differential modulation of the basal genetic background between the two hosts independent of L. amazonensis infection. In addition, in response to early L. amazonensis infection, 10 genes were modulated in infected BALB/c vs. non-infected BALB/c macrophages; and 127 genes were modulated in infected C57BL/6 vs. non-infected C57BL/6 macrophages. These modulated genes appeared to be related to the main immune response processes, such as recognition, antigen presentation, costimulation and proliferation. The distinct gene expression was correlated with the susceptibility and resistance to infection of each host. Furthermore, upon comparing the DEGs in BMDMs vs. peritoneal macrophages, we observed no differences in the gene expression patterns of Jun, Fcgr1 and Il1b, suggesting a similar activation trends of transcription factor binding, recognition and phagocytosis, as well as the proinflammatory cytokine production in response to early L. amazonensis infection. Analysis of the DEG profile of the parasite revealed only one DEG among the 8,282 transcripts, indicating that parasite gene expression in early infection does not depend on the host genetic background.
This content is subject to copyright. Terms and conditions apply.
1
Scientific RepoRtS | (2019) 9:19841 | https://doi.org/10.1038/s41598-019-56305-1
www.nature.com/scientificreports
Dierential immune response
modulation in early Leishmania
amazonensis infection of BALB/c
and C57BL/6 macrophages based
on transcriptome proles
Juliana Ide Aoki
1*, Sandra Marcia Muxel1, Ricardo Andrade Zampieri1, Karl Erik Müller2,3,
Audun Helge Nerland
2 & Lucile Maria Floeter-Winter1*
The fate of Leishmania infection can be strongly inuenced by the host genetic background. In
this work, we describe gene expression modulation of the immune system based on dual global
transcriptome proles of bone marrow-derived macrophages (BMDMs) from BALB/c and C57BL/6 mice
infected with Leishmania amazonensis. A total of 12,641 host transcripts were identied according to
the alignment to the Mus musculus genome. Dierentially expressed genes (DEGs) proling revealed
a dierential modulation of the basal genetic background between the two hosts independent of
L. amazonensis infection. In addition, in response to early L. amazonensis infection, 10 genes were
modulated in infected BALB/c vs. non-infected BALB/c macrophages; and 127 genes were modulated
in infected C57BL/6 vs. non-infected C57BL/6 macrophages. These modulated genes appeared to
be related to the main immune response processes, such as recognition, antigen presentation,
costimulation and proliferation. The distinct gene expression was correlated with the susceptibility and
resistance to infection of each host. Furthermore, upon comparing the DEGs in BMDMs vs. peritoneal
macrophages, we observed no dierences in the gene expression patterns of Jun, Fcgr1 and Il1b,
suggesting a similar activation trends of transcription factor binding, recognition and phagocytosis, as
well as the proinammatory cytokine production in response to early L. amazonensis infection. Analysis
of the DEG prole of the parasite revealed only one DEG among the 8,282 transcripts, indicating that
parasite gene expression in early infection does not depend on the host genetic background.
Leishmania is a protozoan parasite and the causative agent of several clinical infections, generically known as
leishmaniases. In general, these infections are characterized by cutaneous, mucosal or visceral manifestations1,2.
Leishmaniases are considered neglected tropical diseases by the World Health Organization. ere is no vaccine
available to prevent the disease due to a range of factors, such as diversity among Leishmania species and the
interaction of these parasites with the host immune system36. Treatment can be complicated since most of the
drugs available are expensive and toxic and may require long treatment regimens7,8. Furthermore, resistance to
several commonly used drugs has been reported9. In humans, L. amazonensis infection can cause chronic cutane-
ous lesions, although diuse cutaneous and visceral manifestations have been reported1,7.
e immune response to Leishmania involves a complex range of cells. Neutrophils and monocytes are rst
recruited to the site of the insect bite, which leads to the dierentiation of macrophages; this dierentiation is fol-
lowed by the recognition and phagocytosis of the parasite, as well as the induction of a range of inammatory sig-
nals. Other phagocytes, such as dendritic cells, also play important roles since they induce the response in other
inammatory tissues. However, macrophages that play a critical roles in the establishment of infection, as they are
the main host cells for Leishmania replication inside the phagolysosome1013. e infection is characterized by 1
cell-mediated production of interferon gamma (IFN-γ), tumor necrosis factor alpha (TNF-α) and granulocyte
1Department of Physiology, Institute of Bioscience, University of São Paulo, São Paulo, Brazil. 2Department of
Clinical Science, University of Bergen, Bergen, Norway. 3Department of Internal Medicine, Drammen Hospital,
Drammen, Norway. *email: juaoki@usp.br; lucile@ib.usp.br
open
Corrected: Author Correction
Content courtesy of Springer Nature, terms of use apply. Rights reserved
2
Scientific RepoRtS | (2019) 9:19841 | https://doi.org/10.1038/s41598-019-56305-1
www.nature.com/scientificreports
www.nature.com/scientificreports/
macrophage colony-stimulating factor (GM-CSF), which polarizes macrophages to the proinammatory M1
phenotype and increases nitric oxide synthase 2 (NOS2) and nitric oxide (NO) levels, resulting in parasite control,
or by 2 cell-mediated production of interleukin (IL) 4 (IL4), IL13, IL10, tumor growth factor beta (TGFβ) and
macrophage colony-stimulating factor (M-CSF), which polarizes macrophages to an anti-inammatory M2 phe-
notype and increases arginase 1 and polyamine production, resulting in parasite replication3,1417. However, the
parasite is able to subvert macrophage killing mechanisms through the modication of host cytokine expression,
preventing antigen display by MHC class II molecules and reducing NO production with consequent amastigote
dierentiation and proliferation11,12.
Leishmania infection in murine models has been extensively characterized and varies according to the parasite
species and host genetic background3,1822. Progressive disease occurs due to impaired cellular immunity, with
dysfunction of T cells, macrophages, or both23. Regulation of the host immune response to Leishmania has been
well dened in L. major model in which the BALB/c mouse strain is susceptible to infection due to early bursts of
IL4 that lead to disease progression. On the other hand, the C57BL/6 mouse strain is resistant to infection due to
a dominant 1-type response leading to infection control13,1820,24. Experimental murine infections with L. ama-
zonensis have demonstrated distinct susceptibilities compared to those for L. major25,26. L. amazonensis induces
severe lesions upon cutaneous inoculation in susceptible BALB/c mice, while the same parasite causes only mod-
erate lesions in resistant C57BL/6 mice21,27. Such variations in infection have been observed as dierences in the
lesion size, parasite burden, cellular activation and 1/2 ratio between the dierent infected strains21,25,28.
Furthermore, studies involving knockout mouse strains have revealed interesting data concerning the
response of the host to Leishmania infection. Targeted deletion of the Il4 and Il10 genes results in a minimal
eects on the development of L. amazonensis29 and L. major infections30, due to reduced IL12 receptor expression,
which leads to reduced IL12 responsiveness and, consequently, to impairment of the 1 response31. In Tlr4- and
MyD88- decient mice, L. amazonensis shows increased in vitro infectivity; in contrast Tlr2-decient mice exhib-
ited a decreased parasite loads, indicating that this receptor is required for disease progression32.
Based on these ndings, we analyzed the modulation of the early immune responses dened by the dual
transcriptome proles of BMDMs from the BALB/c and C57BL/6 mouse strains aer infection with L. ama-
zonensis for 4 h. Previous transcriptomic data have revealed novel information about the coordinated response
of Leishmania-infected macrophages3336 and about parasite biology, physiology and gene expression modula-
tion3742. In this work, we identied a total of 12,641 total mouse transcripts, and analyses of the DEGs prole
involved in immune response modulation conrmed the existence of dierences between these two hosts that can
regulate susceptibility and resistance to L. amazonensis infection. Interestingly, the parasite transcriptome prole
showed only one DEG, a noncoding RNA, indicating that the parasite presents no modulation of gene expression
in early infection regardless of the host genetic background.
Results
BMDMs from BALB/c mice exhibited a lower infection index than those from C57BL/6 mice at
4 h after infection. BMDMs from the BALB/c and C57BL/6 mouse strains were infected with L. amazon-
ensis (MOI 5:1), and the infection index was analyzed at 4 h aer infection. First, no signicant dierences were
observed in the infection rate or the number of intracellular parasites per infected macrophage (Fig.S1A,B).
However, the infection index was signicantly lower in infected BALB/c than in infected C57BL/6 macrophages
(Fig.S1C).
Host transcriptome proling revealed greater gene expression modulation in BMDMs from
C57BL/6 mice than in BALB/c mice in response to L. amazonensis infection. Transcriptomic anal-
yses were performed on ve independent biological replicates per analysis of BMDMs from BALB/c and C57BL/6
mice infected or not infected with L. amazonensis for 4 h, using Illumina NovaSeq. 6000 sequencing, which
generated millions of reads. e sequencing data are available in the NCBI BioProject database (https://www.
ncbi.nlm.nih.gov/bioproject/) under accession numbers PRJNA481041 and PRJNA481042 and in the Sequence
Read Archive (SRA) database (https://www.ncbi.nlm.nih.gov/sra) under accession numbers SRP156183 and
SRP156466. e RNA-seq data were aligned to the M. musculus reference genome, and 12,641 transcripts were
identied (TableS1).
Analysis of DEGs with a statistical signicance threshold of a fold change 2 and a p-value < 0.05 revealed
dierential basal backgrounds in non-infected BALB/c vs. non-infected C57BL/6 macrophages; specically,
313 genes were upregulated, and 254 genes were downregulated. Comparison of BALB/c_La vs. BALB/c mac-
rophages revealed only 20 upregulated genes and 2 downregulated genes. In contrast, comparison of C57BL/6_La
vs. C57BL/6 macrophages revealed 358 upregulated genes and 139 downregulated genes, and comparison of
BALB/c_La vs. C57BL/6_La macrophages revealed 318 upregulated genes and 434 downregulated genes (Fig.1).
In addition, we generated volcano plots comparing the fold changes in expression (log2) with the correspond-
ing adjusted p-values (-log10) (Fig.S2A) and volume plots comparing the fold changes in expression (log2) with
the volumes (Fig.S2B). Based on these results, we identied the ve most highly modulated transcripts among
the comparisons (TableS2). Functional annotation and gene enrichment analyses were performed using the Gene
Ontology (GO) and Kyoto Encyclopedia of Genes and Genomes (KEGG) databases. KEGG enrichment analysis
showed the 20 most dierentially regulated pathways among the samples (Fig.S2C).
RNA-seq generates a large amount of information that can be analyzed from various perspectives. According
to GO enrichment analysis of the DEGs, the most modulated subcategories were associated with biological pro-
cesses, molecular functions and cellular components (Fig.2). In this work, we focused on the immune system
process term, comprising 361 modulated transcripts (TableS3), to elucidate how the host genetic background
dierences can dene the fate of L. amazonensis infection.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
3
Scientific RepoRtS | (2019) 9:19841 | https://doi.org/10.1038/s41598-019-56305-1
www.nature.com/scientificreports
www.nature.com/scientificreports/
The gene expression modulation patterns revealed higher immune response activation in
BMDMs from C57BL/6 mice than in BALB/c mice in response to L. amazonensis infection.
Among the 361 modulated transcripts related to immune system processes, 150 of them appeared to dier in
expression in non-infected BALB/c vs. C57BL/6 macrophages, indicating the existence of dierential basal gene
expression in these two host backgrounds, independent of L. amazonensis infection. Aer L. amazonensis infec-
tion, we observed only 10 modulated genes in BALB/c_La vs. BALB/c macrophages; 127 modulated genes in
C57BL/6_La vs. C57BL/6 macrophages; and 221 modulated genes in C57BL/6_La vs. BALB/c_La macrophages
(Table1).
Furthermore, we categorized the identied molecules according to the main types of immune system pro-
cesses in response to L. amazonensis infection (Fig.3). In the BALB/c_La vs. BALB/c comparison, we found that
most of the modulated transcripts were immunomodulatory (Il1b, Irg1 and Tnfrsf26) and chemokine signaling
molecules (Cxcl1, Cxcl2 and Cxcl3) (Fig.3 and Table1). In the C57BL/6_La vs. C57BL/6 comparison, most of the
modulated transcripts were immunomodulatory molecules (Clec4d, Clec4e, Clec5a, Il16, Il17ra, Il1rn, Il27, Irak2,
Figure 1. Transcriptome proles of BMDMs from BALB/c and C57BL/6 mice infected with L. amazonensis.
Dierential gene expression proles of BMDMs from BALB/c and C57BL/6 mice infected with L. amazonensis,
presented as the numbers of upregulated (light gray) and downregulated (dark gray) transcripts in the following
comparisons: non-infected BALB/c vs. non-infected C57BL/6 macrophages; infected BALB/c vs. non-infected
BALB/c macrophages; infected C57BL/6 vs. non-infected C57BL/6 macrophages; and infected BALB/c vs.
infected C57BL/6 macrophages. e data are from ve independent biological replicates, considering a fold
change 2 and a p-value < 0.05. La, L. amazonensis.
Figure 2. GO enrichment analysis of DEGs in BMDMs from BALB/c and C57BL/6 mice in response to L.
amazonensis infection. e GO enrichment analysis results are presented as numbers of transcripts distributed
in three main categories: biological process, molecular function and cellular component. Immune system
processes were the focus of this work.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
4
Scientific RepoRtS | (2019) 9:19841 | https://doi.org/10.1038/s41598-019-56305-1
www.nature.com/scientificreports
www.nature.com/scientificreports/
Irg1, Lcp2 Mefv, emis2, Tnf, Tnfaip3, Tnfaip8l2, Tnfrsf26, Tnfsf9, Tnip1 and Tnip3), transcription factors (Batf,
Bcl3, Cebpb, Foxo3, Hhex, Id2, Irf1, Jun, Lyl1, Ma, Nb1, Nb2, Nbia, Tiparp, Trim13, Trim14 and Tsc22d3),
adaptor proteins (Malt1, Mef2c, Nck1, Nfe2l2, Nr1h3, Pik3cd, Procr, Ptafr, Rbpj, Rgcc, Sh2b2, Smad6 and Src) and
members of recognition pathways (Birc3, Jag1, Lilrb4a, Mapk14, Mb21d1, Nlrp3, Nod2, Ripk2, Tlr2, Tlr8 and
Traf3) (Fig.3 and Table1).
The exclusive dierential gene expression patterns in BMDMs from C57BL/6 mice appeared to
be mostly related to proliferation signaling and transcription factor molecules. Venn diagram
analysis was performed, and based on the results, we grouped the exclusively and commonly modulated genes
involved in the immune response. We identied 28 exclusively modulated genes in the comparison of the two
host backgrounds (in non-infected macrophages). Additionally, we identied only one exclusively modulated
gene in BALB/c_La vs. BALB/c macrophages, 39 exclusively modulated genes in C57BL/6_La vs. C57BL/6 mac-
rophages, and 47 exclusively modulated genes in C57BL/6_La vs. BALB/c_La macrophages. Interestingly, only
one gene, Smad6, was common among all comparisons (Fig.4A).
Examination of the immune response modulation associated with L. amazonensis infection and the pattern
of exclusively expressed genes in the comparison of BALB/c_La vs. BALB/c macrophages revealed the downreg-
ulation of Il1b as unique (Fig.4B). In contrast, comparison of C57BL/6_La vs. C57BL/6 macrophages revealed a
set of 39 modulated genes, among which 22 were upregulated genes and 17 were downregulated. Most of these
genes appeared to be involved in the proliferation pathway, such as the downregulated Pik3cd and Mtus1 genes
comparison downregulated genes upregulated genes p-value FDR
BALB/c vs.
C57BL/6
Adgre1, AF251705, Ang, Apobec3, Asb2, Batf3,
Blnk, Bst2, C1qa, C1qb, C1qc, C5ar2, Camp,
Ccl5, Ccr2, Cd300a, Cd4, Cd40, Cd79b, Clec1b,
Clec2d, Clec4a2, Ctsh, Cxcl10, Cxcl9, Emr1,
Erbb2ip, Fcgr4, Fcna, Gbp2, Gbp3, Gbp5, Gbp7,
Hfe, It1bl1, Iigp1, Il15, Il18bp, Irf1, Irgm2,
Itga4, Itgad, Itgal, Lcn2, Lgals1, Ly86, Marco,
Mertk, Mill2, Pde4b, Pnp, Prdm1, S100a8,
Samhd1, Skil, Slamf7, Slc11a1, Slc40a1, Smad6,
Tfrc, Tgtp1, Tlr8, Tmem176a, Tmem176b,
Trim34a, Vcam1, Veg fa, Vsig4, Wwp 1
Ada, Ahcy, Alpk1, Armc6, Batf, Bcl2a1a, Bcl2a1d,
Bst1, Ccl2, Ccl24, Ccl3, Ccl4, Ccl7, Ccnb2, Ccr1,
Cd109, Cd14, Cd24a, Cd300lf, Cd86, Cdk6, Clec4n,
Clec5a, Col3a1, Colec12, Csf1, Ctse, Cx3cr1, Cxcl14,
Fam20c, Glo1, Gm8909, Gpr183, H2-Ab1, H2-
DMb1, H2-K1, H2-L, H2-Q1, H2-Q2, H2-Q4, H2-
Q6, H2-Q8, H2-Q9, H2-T22, H2-T24, Hist1h2bf,
Hist1h2bk, Hist1h2bl, Hist2h3c2, Hist1h3a,
Hist1h3b, Hist1h3d, Hist1h3g, Hist1h3h, Hist1h3i,
Hist1h4a, Hist1h4f, Hist1h4i, Hist4h4, Itm1,
Itm3, Il1rn, Irf7, Kdr, Lat2, Malt1, Mmp14, Myc,
Ndrg1, Npy, Oasl1, Pla2g7, Ripk3, Serpine1, Slpi,
Spn, Spp1, Stap1, Tnfsf13, Tnfsf8, Top2a
2.88e79 1.00e76
BALB/c_La
vs. BALB/c Il1b, Mef2c Cxcl1, Cxcl2, Cxcl3, Hilpda, Id2, Irg1, Smad6,
Tnfrsf26 2.43e82.24e6
C57BL/6_La
vs. C57BL/6
Ccr2, Ccr5, Fcgr1, Foxo3, Gcnt1, Gpr183,
Hhex, Hist1h2bc, Hist1h2be, Hist1h2bg,
Hist1h3e, Hist1h4c, Hist1h4d, Hist1h4h,
Hist1h4m, Hist2h3b, Hist2h4, Il16, Lyl1, Ma,
Mapk14, Mef2c, Mertk, Mtus1, Pik3cd, Rassf2,
emis2, Tlr8, Tnfaip8l2, Trim14, Tsc22d3,
Zfp36l1, Zfp36l2
Adora2b, Ampd3, Batf, Bcl2a1a, Bcl2a1d, Bcl3,
Birc3, Ccl3, Ccl4, Cd24a, Cd274, Cd40, Cd83,
Cd86, Cdkn1a, Cdkn2b, Cebpb, Clec4d, Clec4e,
Clec5a, Cxcl1, Cxcl2, Cxcl3, Ednrb, Ezr, Fam20c,
Fas, Gadd45g, Gbp5, Gch1, Gpr68, H2-M2, Hcar2,
Hilpda, Hmox1, Hsp90aa1, Hyal2, Icam1, Icosl,
Id2, Il17ra, Il1rn, Il27, Irak2, Irf1, Irg1, Jag1, Jun,
Lcp2, Lilrb4a, Malt1, Mb21d1, Mefv, Mmp14, Nck1,
Nfe2l2, Nb1, Nb2, Nbia, Nlrp3, Nod2, Nr1h3,
Olr1, Osm, Pde4b, Pmaip1, Ppp4r2, Prdx1, Procr,
Ptafr, Rbpj, Rgcc, Ripk2, Rnf19b, Samsn1, Serpine1,
Sh2b2, Slamf7, Slc11a2, Smad6, Sod2, Sqstm1,Src,
Stx11, Tiparp, Tlr2, Tnf , Tnfaip3, Tnfrsf26, Tnfsf9,
Tnip1, Tnip3, Traf3, Trim13
1.40e65 2.37e63
C57BL/6_La
vs. BALB/
c_La
Adgre1, Ampd3, Ang, Apobec3, Axl, Bcl3, Bcl6,
Birc3, Blnk, Bloc1s6, C1qa, C1qb, C1qc, C5ar1,
Camp, Casp1, Ccl3, Ccl4, Ccl5, Ccl9, Ccr2,
Ccr3, Cd274, Cd38, Cd4, Cd40, Cd79b, Cd83,
Cdkn1a, Cebpb, Clec1b, Clec2d, Clec2i, Clec4e,
Cnr2, Ctsh, Cxcl1, Cxcl10, Cxcl2, Cxcl3, Cxcl9,
Ednrb, Fas, Fcgr4, Fcna, Fyb, Fzd7, Gbp2,
Gbp3, Gbp5, Gbp6, Gbp7, Gpr68, H2-Ab1, H2-
DMb1, H2-K1, H2-L, H2-M2, Hmox1, Icam1,
Icosl, Iig p1, Il10, Il17ra, Il18bp, Il1a, Il1f9,
Il27, Irak2, Irak3, Irf1, Irg1, Itgad, Itgal, Jag1,
Kdr, Lcn2, Lcp2, Malt1, Mapkapk2, Marco,
Mefv, Mertk, Mill2, Nb1, Nb2, Nbia,
Nbid, Nlrc4, Nlrp3, Nod1, Nr1h3, Olr1,
Pde4b, Pmaip1, Pnp, Ppp4r2, Prdm1, Procr,
Ptafr, Ptprj, Rab32, Rela, Relb, Rgcc, Ripk2,
Rnf19b, S100a8, Sh2b2, Skil, Slamf7, Slc40a1,
Smad6, Snx10, Sod2, Sqstm1, Stx11, Tapbpl,
Tbk1, Tgtp1, bs1, Tlr1, Tlr2, Tmem176a,
Tmem176b, Tnf, Tnfaip3, Tnfrsf1b, Tnfsf9,
Tnip3, Traf3, Treml4, Trib1, Trim13, Vcam1,
Veg fa , Vsig4
Ada, Ahcy,Aim2, Alpk1, Armc6, Bst1, Ccl2,
Ccl24, Ccl7, Ccnb2, Cd109, Cd300lf, C, Clec4n,
Csf1, Colec12, Col3a1, Ctse, Cxcl14, Gcnt1, Glo1,
Gm8909, Gpr183, H2-Q1, H2-Q2, H2-Q6, H2-
Q8, H2-Q9, H2-T22, H2-T24, Hhex, Hist1h2ba,
Hist1h2be, Hist1h2bf, Hist1h2bg, Hist1h2bk,
Hist1h2bl, Hist1h3a, Hist1h3b, Hist1h3d, Hist1h3e,
Hist1h3h, Hist1h3i, Hist1h3g, Hist1h4a, Hist1h4b,
Hist1h4d, Hist1h4f, Hist1h4h, Hist1h4i, Hist1h4j,
Hist1h4k, Hist1h4m, Hist1h4n, Hist2h3b,
Hist2h3c2, Hist2h4, Hist4h4, Itm1, Itm3, Il16,
Il1rn, Irf4, Irf7, Junb, Lgals1, Mmp9, Ndrg1, Npy,
Pdgfrb, Pla2g7, Ripk3, Serpine1, Slfn1, Slpi, Spn,
Spp1, Tacc3, Tnfsf13, Tnfsf8 , Top2a,
2.02e125 9.61e123
Table 1. Prole of DEGs involved in immune system processes in BMDMs from BALB/c and C57BL/6
in response to L. amazonensis infection. Gene Ontology (GO) enrichment analysis and the proling of
dierentially expressed genes (DEGs) involved in the immune system processes in bone marrow-derived
macrophages (BMDMs) from BALB/c and C57BL/6 in response to L. amazonensis (La) infection. e analysis
was based on p-values and false discovery rates (FDRs).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
5
Scientific RepoRtS | (2019) 9:19841 | https://doi.org/10.1038/s41598-019-56305-1
www.nature.com/scientificreports
www.nature.com/scientificreports/
and the upregulated Samsn1, Prdx1, Osm and Hsp90aa1 genes. Another group of genes contained transcription
factors, including the downregulated Tsc22d3, Trim14, Ma and Lyl1 genes and the upregulated Tiparp, Rbpj,
Nfe2l2, Jun and Foxo3 genes. We also identied genes involved in recognition and costimulation pathways, as
well as genes encoding adaptor molecules: Fcgr1, Mapk14 and emis2 were downregulated, while Mb21d1/cGas,
Nod2, Clec4d, Lilrb4a, Cdkn2b, Nck1, Src and Tnip1 were upregulated. Among apoptosis-related molecules, Rassf2
and Gadd45g appeared downregulated. e immunomodulatory molecules Tnfaip8l2 and Ccr5 were downregu-
lated. e metal transporter Slc11a2 (formerly Nramp2) was upregulated. e histones Hist1h2bc and Hist1h4c,
as well as the RNA-binding proteins Zfp36l1 and Zfp36l2, were downregulated. Poorly studied molecules, such
as Adora2b, Ezr, Gch1 and Hyal2 were upregulated and were classied as belonging to other pathways (Fig.5).
RT-qPCR validation assays were performed on some of the most modulated molecules from the RNA-seq
data: Il1b, Fcgr1, Ccr5, Smad6, Jun and Mapk14. Comparative analyses showed concordance between the RNA-seq
and RT-qPCR data with no statistically signicant dierences, thus validating the RNA-seq results (Fig.6).
Similar to BMDMs, peritoneal macrophages were collected and infected with L. amazonensis, and the gene
expression modulation of selected genes was analyzed by RT-qPCR to evaluate whether a similar trend occurred
in another macrophage subtype. e infection index appeared signicantly lower in C57BL/6_La macrophages
than in BALB/c_La macrophages (Fig.S3A), indicating a distinct phenotypic dierence between BMDMs and
peritoneal macrophages in response to L. amazonensis infection. Comparison of the gene expression in BMDMs
and peritoneal macrophage subtypes from BALB/c_La mice revealed lower expression of Smad6 and Mapk14. No
modulation of Il1b, Ccr5, Fcgr1 or Jun was observed. On the other hand, we observed lower expression of Smad6,
higher expression of Ccr5 and no modulation of Il1b, Mapk14, Fcgr1 and Jun expression in BMDMs compared
with peritoneal macrophage subtype from C57BL/6_La (Fig.S3B).
e transcriptomic data presented here corroborate the ndings of previous studies on how dierential genetic
backgrounds from dierent hosts dene susceptibility or resistance to Leishmania infection. e DEGs proles
described in this work represents new knowledge obtained from transcriptome analyses of immune responses
between two dierent host genetic backgrounds. e analyses identied molecular markers that could be linked to
susceptibility and resistance to L. amazonensis infection, as illustrated by the schematic representation of the exclu-
sively and DEGs in BMDMs from BALB/c and C57BL/6 mice in response to early L. amazonensis infection (Fig.7).
Parasite transcriptome proling revealed only one DEG between L. amazonensis infecting
BALB/c and L. amazonensis infecting C57BL/6 macrophages. We also analyzed the gene expression
of L. amazonensis via alignment to the L. mexicana genome database (TableS1). e sequencing data are available
in the NCBI BioProject and SRA databases, as previously described.
Aer initial assembly, 8,282 parasite transcripts were identied. Analysis of DEGs with signicant threshold
of a fold change 2 and a p-value < 0.05, as statistically signicant, revealed only one DEG, a noncoding RNA
(ncRNA) (LmxM.32.ncRNA:rfamscan:912871–912976), which showed higher expression in infected BALB/c
than in infected C57BL/6 macrophages.
Additionally, we performed RT-qPCR validation assays of our RNA-seq data for Amastin-like gene (LmxM.33.0960).
Similar to the case for the host comparative analyses, we observed concordance between the RNA-seq and RT-qPCR
data (Fig.6), thus validating the RNA-seq results.
Finally, we observed lower expression of the Amastin-like gene (LmxM.33.0960) in peritoneal macrophages
than in BMDMs from BALB/c and C57BL/6 mice (Fig.S3B).
Figure 3. Immune response analysis of DEGs in BMDMs from BALB/c and C57BL/6 mice in response to
L. amazonensis infection. Pie chart of the modulated molecules involved in the immune response processes
grouped into main immune signaling pathways.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
6
Scientific RepoRtS | (2019) 9:19841 | https://doi.org/10.1038/s41598-019-56305-1
www.nature.com/scientificreports
www.nature.com/scientificreports/
Discussion
e 1/2 paradigm correlating resistance/susceptibility to Leishmania infection has been extensively stud-
ied3,14,1821,23,35,41. Identication of potential biomarkers for leishmaniases can be useful for dierent approaches,
such as diagnosis, prognosis, disease progression monitoring, clinical intervention and host immune response
characterization33,34,41,4345. e host-parasite interaction depends on both host genetic backgrounds33,35,36,41 and
the genetic complexity of Leishmania species39,40,46.
L. amazonensis infection elicits dierent immune responses than those previously described for L. major
infection25,26,28,29,31,35. In this work, we present the global transcriptome proles of BMDMs from BALB/c and
Figure 4. Venn diagram analysis of DEGs in BMDMs from BALB/c and C57BL/6 mice in response to L.
amazonensis infection. (A) Venn diagram of the 361 DEGs involved in the immune response processes, showing
the numbers of exclusively and common genes for each comparison. (B) List of exclusively and common genes
according for each comparison in the Venn diagram.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
7
Scientific RepoRtS | (2019) 9:19841 | https://doi.org/10.1038/s41598-019-56305-1
www.nature.com/scientificreports
www.nature.com/scientificreports/
C57BL/6 mice non-infected and infected with L. amazonensis, focusing on the modulation of the immune
response. In the absence of L. amazonensis infection, we identied signicantly dierent basal gene expression
patterns between the two hosts, corroborating with previous ndings47. Analysis of the immune response in early
L. amazonensis infection revealed 361 modulated genes among the comparisons. Comparison of infected BALB/c
to non-infected BALB/c BMDMs revealed low levels of gene expression modulation; this pattern could be related
to limited immune response activation, leading to susceptibility of this host to L. amazonensis infection, as previ-
ously described21,48. e DEGs involved in immune response modulation comprised mostly immunomodulatory
and chemokine signaling molecules, suggesting a link to the inammation process. In contrast, we observed high
levels of gene expression modulation in infected C57BL/6 compared to non-infected C57BL/6 BMDMs. is
pattern could be related to increased immune response activation via augmentation of recognition processes
and, consequently, activation of signaling cascades, leading to moderate resistance of this host to L. amazonensis
infection. Dierent proles associated with dierent host genetic backgrounds have previously been described as
being due to dierent parasite burdens, inammatory cell populations and cytokine production21,48.
e infection index of BMDMs from BALB/c mice appeared smaller than that of BMDMs from C57BL/6
mice aer 4 h of infection. As the infection index represents the number of intracellular parasites multiplied by
the percentage of infected macrophages, the biological impact of this dierence indicates that at an early stage of
infection, C57BL/6 macrophages exhibit greater phagocytosis, which in subsequent times of infections may ena-
ble control of parasite replication. Previous studies by our group have demonstrated increased infection index val-
ues in BALB/c macrophages aer 24 and 48 h of infection; in contrast, the index values of C57BL/6 macrophages
appeared to remain stable49,50. However, most gene expression modulation has been described to occur during
early Leishmania infection33,38,41,50,51.
e fact that Il1b appeared to be downregulated and was an exclusively modulated gene involved in the
immune response in infected BALB/c compared to non-infected BALB/c BMDMs corroborates the important
role of this molecule in Leishmania infection. IL1β has previously been identied as an important signaling fac-
tor for host resistance to C57BL/6 infection, since this cytokine signals through IL1R and MyD88 to induce
NOS2-mediated NO production, which is a major host defense mechanism against Leishmania52. Furthermore,
polymorphisms in the Il1b gene are associated with the severity of the disease in patients infected with L. mex-
icana53. Given these ndings, we reinforce the importance of this molecule in Leishmania infection in both
hosts5254.
e 39 exclusively modulated immune response-related genes in infected C57BL/6 compared to non-infected
C57BL/6 BMDMs were associated with important signaling pathways, suggesting enhancement of immune
response activation resulting in moderated resistance against L. amazonensis infection. e recognition signa-
ling cascade included a large number of modulated molecules, highlighting the importance of the host genetic
background in the initial steps of macrophage activation55. Among these molecules, NOD-like receptors play
protective roles during Leishmania infection52,56,57. e upregulation of Nod2 in infected C57BL/6 compared
Figure 5. DEGs prole of the exclusively modulated genes involved in the immune response processes in
infected C57BL/6 vs. non-infected C57BL/6 BMDMs. e proles of DEGs are presented as the log2-fold
changes in the expression of the 39 exclusively modulated genes involved in the immune response processes
in BMDMs from C57BL/6 infected with L. amazonensis vs. non-infected C57BL/6 BMDMs. e genes were
classied by their involvement in main immune response signaling pathways or by their identities as regulatory
molecules of the immune response pathways. La, L. amazonensis.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
8
Scientific RepoRtS | (2019) 9:19841 | https://doi.org/10.1038/s41598-019-56305-1
www.nature.com/scientificreports
www.nature.com/scientificreports/
Figure 6. RT-qPCR validation of some modulated genes in BALB/c and C57BL/6 BMDMs in response to L.
amazonensis infection. Comparative analysis of the relative expression levels of selected genes determined by
RNA-seq and validated by RT-qPCR. e bars represent the mean ± SD values of the fold changes in Il1b, Fcgr1,
Ccr5, Smad6, Jun and Mapk14 expression determined with ve independent biological replicates analyzed
in duplicate. e fold changes were calculated through relative quantication using the ΔΔCt method. e
data were normalized to Gapdh expression and the relative gene expression was set to 1 for the control (non-
infected) samples. Statistical analysis was performed using the t-tests, and no signicant dierences were
observed (p-value < 0.05) between the RT-qPCR and RNA-seq results for the BALB/c_La and C57BL/6_La
groups. e bars for Amastin-like (LmxM.33.0960) show the mean aer normalization to Gapdh in L.
amazonensis infecting BALB/c and L. amazonensis infecting C57BL/6 macrophages. La, L. amazonensis.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
9
Scientific RepoRtS | (2019) 9:19841 | https://doi.org/10.1038/s41598-019-56305-1
www.nature.com/scientificreports
www.nature.com/scientificreports/
to non-infected BMDMs indicates greater macrophage activation in C57BL/6 mice. NOD2 mediates the
parasite-induced production of cytokines, such as IL-17 and IFN-γ production, in L. infantum and L. amazonen-
sis infections, whereas NOD1 is not relevant to these infections56,57. Recognition signaling also involves MAPKs,
which play important roles against parasitic infections58, driving the switch in macrophage activation from proin-
ammatory IL12 to anti-inammatory IL10 cytokines59. Previous studies have demonstrated that signaling dur-
ing L. amazonensis infection leads to the activation of MAPK1 and MAPK358. MAPK14 has been poorly studied
in the context of Leishmania infection, although downregulation of Mapk14 has previously been described to
occur in L. braziliensis and L. major infections41,60. Most molecules from the recognition pathway were upregu-
lated, indicating the activation of the downstream steps in recognition signaling cascades.
Immunomodulatory molecules play important roles in macrophage activation and the induction of adaptive
immune responses via cytokine production in response to Leishmania infection13,20,61. Among the main cytokines
studied, TNF is a multipotent cytokine implicated in a wide range of immune responses occuring in response
to many infections62,63. In particular, the TNF-related molecules Tnip1 and Tnfaip8l2 appeared upregulated and
downregulated, respectively, indicating signaling cascade activation and repression to maintain immune home-
ostasis. Leishmania infection can also induce the expression of numerous chemokines26,51,64,65. is event could
potentially benet the parasite due the ability to repress the induction of proinammatory cytokine expression66.
e downregulation of Ccr5 in infected C57BL/6 vs. non-infected C57BL/6 BMDMs could be correlated with the
fact that this receptor directs the 1 immune response and is thus associated with inammation, cell inltra-
tion and the development of infectious disease67. Previous studies have demonstrated that CCR5 knockout mice
exhibit increased resistance to L. major infection68.
Similarly, human macrophage infections with L. amazonensis, L. major and L. panamensis have been shown to
elicit immune response modulation of TNF, NF-kB and NOD-like receptor signaling pathways, oxidative stress
pathways and proliferation signaling pathways41,69.
e expression of proliferation signals and transcription factor-related molecules was highly modulated
according to our data. ere are limited descriptions of these molecules; however, they are known to control the
expression of many genes required for the eective activation of the immune responses, such as transcriptional
activators or repressors, as well as for FOXO transcriptional activity, NF-kB recruitment and Notch signaling7073.
e release and activation of histones occur in response to stress, leading to Toll-like receptor binding and
triggering the activation of multiple signaling pathways74. e downregulation of Hist1h4c and Hist1h2bc could
be related to the negative modulation of transcription factors listed above that are involved in macrophage
activation.
e metal transporter natural resistance-associated macrophage protein (Nramp) has been associated with
resistance to intracellular pathogens due to enhanced NOS2 expression and NO production75,76. Point mutations
in Nramp1 promote susceptibility to Leishmania infection by modulating iron acquisition from intracellular com-
partments76,77, starving pathogens of this essential nutrient and impacting parasite survival and replication78.
Although Nramp2 shares a conserved structure and iron transport functions with Nramp1, its role in Leishmania
infection has been poorly studied. e upregulation of Slc11a2 (formerly Nramp2) was upregulated in infected
C57BL/6 macrophages could be correlated with increased NO production and resistance to L. amazonensis
infection.
Apoptosis induced by Leishmania may permit successful infection through modulation of host immunity79.
RASSF2 and GADD45G are involved in the regulation of growth and apoptotic processes. Consistent with these
ndings, we identied downregulation of Rassf2 and the upregulation of Gadd45g as important factors in the
modulation of the host immunity in response to L. amazonensis infection.
Molecules not acting in any of the described pathways were classied as “other” due to their limited descrip-
tions in Leishmania infection. Further studies and functional validation could implicate the role of these
Figure 7. Schematic representation of the exclusive genes and DEGs in BALB/c and C57BL/6 BMDMs in
response to L. amazonensis infection. Summary of the data of the exclusive genes and DEGs in BMDMs derived
from BALB/c and C657BL/6 mice in response to early L. amazonensis infection.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
10
Scientific RepoRtS | (2019) 9:19841 | https://doi.org/10.1038/s41598-019-56305-1
www.nature.com/scientificreports
www.nature.com/scientificreports/
molecules in host immune modulation in response to L. amazonensis infection, but this study provides only a
global transcriptomic view based on the prole of the DEGs involved in immune response modulation in the two
dierent host genetic backgrounds.
Macrophages form a vast and diverse population with considerable plasticity to adapt to dierent tissues
and change in response to environmental variations8083. e dierences between peritoneal macrophages and
BMDMs are believed to arise from dierential physiological conditions and organ specicity along with the het-
erogeneity of macrophages83,84. us, we compared BMDMs and peritoneal macrophages with regard to some
of the modulated genes to reinforce our ndings and provide a representation of the in vivo scenario. According
to our results, the infection index appeared lower in peritoneal macrophages from C57BL/6 mice than in those
from BALB/c mice, indicating a distinct phenotypic dierence between the macrophage subtypes in response
to early L. amazonensis infection and suggesting that BALB/c mice are more susceptible models than C57BL/6
mice, as previously described21,48. Analyses of gene expression have shown a similar gene expression proles in
the comparison of BMDMs and pre-existing populations, although some dierences have also been reported,
suggesting that tissue environments dictate the macrophage phenotype required to trigger an eective immune
response80,81. In our comparisons we observed nondierential and dierential modulation patterns, indicating
that some of the analyzed genes were involved in distinct signaling cascades that lead to a distinct network activ-
ity. Smad6 showed a lower gene expression pattern in peritoneal macrophages than in BMDMs in both BALB/c
and C57BL/6 mice. Since Smad6 is a regulator of myeloid dierentiation85, this expression pattern conrms the
dierences between the macrophage subtypes. ere were no dierences in the gene expression patterns of Jun,
Fcgr1 and Il1b between macrophage subtypes, suggesting similar trends of activation of transcription factor bind-
ing, recognition, phagocytosis and proinammatory cytokines production. Ccr5 showed high modulation only
in peritoneal macrophages from infected C57BL/6 mice, indicating upregulation of this chemokine receptor in
this macrophage subtype. Mapk14 showed low modulation only in peritoneal macrophages from infected BALB/c
mice, indicating low activation of the cellular response cascade in this macrophage subtype.
Altogether, our ndings indicate the need to be cautious in extrapolating ndings to in vivo scenarios that may
or may not dier from those observed in the present study, especially considering that other immune cells, such
as monocytes, neutrophils and lymphocytes, migrates to local cutaneous lesions with Leishmania8688. Both host
and parasite genetic backgrounds also need to be considered in translational approaches to identify biomarkers
for the prognosis determination and treatment of the leishmaniases.
Finally, the transcriptome proling of the parasite revealed only one DEG between L. amazonensis infecting
BALB/c macrophages and L. amazonensis infecting C57BL/6 macrophages, a noncoding RNA (LmxM.32.ncR-
NA:rfamscan:912871–912976). ncRNAs have several functions; for example, they mediate transcription by RNA
polymerase II, polyadenylate 3´-ends, regulate transcript expression and are potentially associated with small
ribonucleoprotein complexes89. Our observations indicate that during early infection, the parasite exhibits the
same gene expression pattern regardless of the host genetic background.
Methods
Animals. Female BALB/c and C57BL/6 mice (6–8 weeks old) were obtained from the Animal Center of the
Medical School of the University of São Paulo and were maintained at the Animal Center of the Department of
Physiology of the Institute of Bioscience of the University of São Paulo with access to food and water ad libitum.
Leishmania culture. L. amazonensis (MHOM/BR/1973/M2269) was grown at 25 °C in M199 medium
(Gibco, Grand Island NY, USA), pH 7.0, supplemented with L-glutamine, 10% heat-inactivated fetal bovine
serum, 0.25% hemin, 40 mM NaHCO3, 100 μM adenine, 40 mM HEPES, 100 U/mL penicillin and 100 μg/mL
streptomycin, as previously described3739. e parasites were counted in a Neubauer chamber.
In vitro macrophage infections. BMDMs were obtained from the femurs of BALB/c and C57BL/6 mice
through PBS washing, and the cells were collected by centrifugation at 500 x g for 10 min at 4 °C. Lysis of eryth-
rocytes was performed with NH4Cl (145 mM) and Tris-base (200 mM), pH 7.0, followed by incubation on ice for
20 min. Aer lysis, the cells were washed with cold PBS, centrifuged at 500 x g for 10 min at 4 °C and incubated in
RPMI 1640 medium supplemented with penicillin (100 U/mL), streptomycin (100 µg/mL), 2-mercaptoethanol
(50 µM), L-glutamine (2 mM), sodium pyruvate (1 mM), 10% fetal bovine serum and 10% L929 conditioned
medium as a macrophage stimulating factor source. e cells were dierentiated for 7 days at 34 °C in 5% CO2. e
BMDMs were used aer phenotypic analysis by ow cytometry showed at least 95% F4/80 and CD11b-positive
cells, as previously described50. Aer macrophage dierentiation, cellular viability was evaluated with Trypan
blue staining (1:1 (v:v)), and the cells were counted in a Neubauer chamber. Approximately 5 × 106 BMDMs from
BALB/c and C57BL/6 mice were incubated in sterile 6-well plates (SPL Life Sciences, Korea) overnight at 34 °C in
5% CO2. Non-adherent cells were removed by washing with PBS, and infection was performed with L. amazonen-
sis promastigotes in the stationary growth phase (MOI 5:1). Aer 4 h of infection, the cultures were washed with
PBS; then, RNA was extracted, or the infection index was determinated. Non-infected macrophages maintained
in culture under the same conditions were used as the controls. e infections were evaluated by determining
the percentage of infected cells aer counting 400 panoptic-stained (Laborclin, Parana, Brazil) macrophages.
e infection index was determined by multiplying the percentage of infected macrophages by the mean num-
ber of intracellular parasites per infected cell90,91. Statistical analyses were performed using Student´s t-test and
p-value < 0.05 was considered to indicate a signicant dierence between infected C57BL/6 macrophages or
infected BALB/c macrophages and the corresponding non-infected macrophages.
Peritoneal macrophages were collected from BALB/c and C57BL/6 mice by injection and recovery of 5 mL of
RPMI 1640 medium supplemented, as previously described. e cells were recovered by centrifugation at 500 × g
Content courtesy of Springer Nature, terms of use apply. Rights reserved
11
Scientific RepoRtS | (2019) 9:19841 | https://doi.org/10.1038/s41598-019-56305-1
www.nature.com/scientificreports
www.nature.com/scientificreports/
for 10 min at 4 °C. Cellular viability was evaluated with Trypan blue staining (1:1 (v:v)), and the cells were counted
in a Neubauer chamber. Approximately, 1 × 106 peritoneal macrophages were incubated in sterile 6-well plates
(SPL Life Sciences, Korea) overnight at 34 °C in 5% CO2. Non-adherent cells were removed by washing with PBS,
and infection was performed with L. amazonensis promastigotes in the stationary growth phase (MOI 5:1). Aer
4 h of infection, cultures were washed with PBS; then, RNA was extracted or the infection index was determined.
Non-infected macrophages maintained in culture under the same conditions were used as the controls. e infec-
tions were evaluated as previously described for BMDMs.
Total RNA isolation and library construction. Total RNA was isolated from ve independent biological
replicates of each infected and non-infected group using TRIzol reagent (Life Technologies, Carlsbad, CA, USA)
according to the manufacturer’s instructions and as previously described39. e RNA samples were treated with
DNase I (1 U per µg of RNA) (ermo Scientic, Lithuania, EU) at 37 °C for 1 h, and the RNA concentration was
determined from the A260/A280 ratio using a NanoDrop ND1000 (ermo Scientic, USA). In addition, RNA
integrity was evaluated using an Agilent 2100 Bioanalyzer and a Pico Agilent RNA 6000 kit (Agilent Technologies,
Santa Clara, CA, USA) according to the manufacturer’s instructions. rRNA depletion was performed using a
poly(A) magnetic bead capture protocol and a TrueSeq Stranded Total RNA Sample Prep kit (Illumina) according
to the manufacturer´s instructions. Libraries were prepared using a TrueSeq Stranded RNA-seq Library Prep Kit
(Illumina), according to the manufacturer’s instructions.
RNA-seq and data analysis. Paired end reads (100 bp) were obtained using an Illumina NovaSeq. 6000
platform at Macrogen Inc. (Seoul, South Korea). Quality control was performed on the sequenced raw reads
based on the read quality, total bases, total reads, GC content (%) and basic statistics. e quality of the reads
was analyzed using FastQC according to the Phred quality score92. Reads with Phred quality scores lower than
20 were discarded. To reduce bias in the analysis and artifacts, such as low-quality reads and adaptor sequences,
Trimmomatic was used93. e trimmed reads were mapped to the reference genome L. mexicana reference
genome (MHOMGT2001U1103) with genomic data obtained from TriTrypDB version 36 (www.tritryp.org) and
to the M. musculus genome using the TopHat splice-aware aligner94,95. A maximum of two mismatches were
allowed. e transcripts were assembled in Cuinks through read alignment, providing information on the
known transcripts. e expression proles of the assembled transcripts and the abundance estimates for each
sample were generated by Cuinks96. e expression values were calculated as fragments per kilobase of tran-
script per million mapped reads (FPKM) and are represented as normalized values based on the transcript length
and coverage depth97. Gene expression level values were calculated from the transcript counts. DEG analysis was
performed for the following comparisons: (1) C57BL/6 vs. BALB/c, (2) BALB/c infected with L. amazonensis vs.
BALB/c, (3) C57BL/6 infected with L. amazonensis vs. C57BL/6, and (4) C57BL/6 infected with L. amazonensis
vs. BALB/c infected with L. amazonensis. Genes with FPKM values of zero were excluded. Groups under dierent
conditions or with dierent DEGs were ltered out through statistical hypothesis tests. e false discovery rate
(FDR) was controlled by adjusting the p-value using the Benjamini-Hochberg algorithm98. Functional annota-
tion was performed using GO and KEGG analyses. All analyses were performed by Macrogen Inc. (Seoul, South
Korea).
RT-qPCR validation. RT-qPCR validation assays were performed using total RNA isolated as previ-
ously described above from ve biological replicates. Reverse transcription was performed using 2 µg of total
RNA as a template, reverse transcriptase and random primers (RevertAid H Minus Reverse Transcriptase Kit,
ermo-Scientic, Canada), according to the manufacturer’s instructions. Equal amounts of cDNA were assessed
in total volumes of 25 μL containing Maxima SYBR Green qPCR Master Mix (ermo Scientic, Lithuania, EU)
and primers (200 nM) (TableS4). e mixtures were incubated at 94 °C for 5 min, followed by 40 cycles at 94 °C
for 30 s, 60 °C for 30 s and 72 °C for 30 s. A negative control in the absence of reverse transcriptase was included
in the RT-qPCR assays to detect DNA contamination in the RNA samples. e reactions were carried out using
a PikoReal Real-time PCR System (ermo Scientic, Finland). e reactions were performed in duplicate, and
analyses were performed using PikoReal Soware 2.2 (ermo Scientic). e fold changes were calculated by
relative quantication using the ΔΔCt method99. e data were normalized by Gapdh expression, and the relative
gene expression was set to 1 for the control (non-infected) samples. e normalized absolute copy number of the
amastin-like gene (LmxM.33.0960) was calculated based on the normalization to a reference, considering the
molar mass concentration, according to a standard curve generated from a ten-fold dilution of a quantied PCR
product. e normalized Amastin/Gapdh ratio of the absolute number of molecules was used as an expression
parameter according to a standard curve generated from a ten-fold serial dilution of a quantied and linearized
plasmid containing the target fragment.
Statistical analysis. e experiments were performed with ve biological replicates per group and the
results are presented as the means ± SDs. DEGs were considered statistically significant considering fold
changes 2, p-value < 0.05 and FDR analysis. RT-qPCR validation assays were performed with ve biological
replicates, and the results are presented the means ± SDs. Statistical analysis was based on Student´s t-test with
p-value < 0.05 indicating statistical signicance.
Ethics statement. e experimental protocols for animals were approved by the Animal Care and Use Committee
at the Institute of Bioscience of the University of São Paulo (CEUA 233/2015). is study was carried out in strict
accordance with the recommended guidelines and the policies for the care and use of laboratory animals of São Paulo
State (Lei Estadual 11.977, de 25/08/2005) and the Brazilian government (Lei Federal 11.794, de 08/10/2008).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
12
Scientific RepoRtS | (2019) 9:19841 | https://doi.org/10.1038/s41598-019-56305-1
www.nature.com/scientificreports
www.nature.com/scientificreports/
Received: 1 March 2019; Accepted: 10 December 2019;
Published online: 27 December 2019
References
1. Ashford, . W. e leishmaniases as emerging and reemerging zoonoses. Int J Parasitol 30, 1269–1281 (2000).
2. Burza, S., Cro, S. L. & Boelaert, M. Leishmaniasis. Lancet 392, 951–970, https://doi.org/10.1016/S0140-6736(18)31204-2 (2018).
3. Muxel, S. M. et al. Arginine and Polyamines Fate in Leishmania Infection. Front Microbiol 8, 2682, https://doi.org/10.3389/
fmicb.2017.02682 (2017).
4. hamesipour, A., afati, S., Davoudi, N., Maboudi, F. & Modabber, F. Leishmaniasis vaccine candidates for development: a global
overview. Indian J Med es 123, 423–438 (2006).
5. Müller, . E., Solberg, C. T., Aoi, J. I., Floeter-Winter, L. M. & Nerland, A. H. Developing a vaccine for leishmaniasis: how biology
shapes policy. Tidssr Nor Laegeforen 137, https://doi.org/10.4045/tidssr.17.0620 (2018).
6. umar, . & Engwerda, C. Vaccines to prevent leishmaniasis. Clin Transl Immunology 3, e13, https://doi.org/10.1038/cti.2014.4 (2014).
7. Murray, H. W., Berman, J. D., Davies, C. . & Saravia, N. G. Advances in leishmaniasis. Lancet 366, 1561–1577, https://doi.
org/10.1016/S0140-6736(05)67629-5 (2005).
8. Hotez, P. J., Bottazzi, M. E., Franco-Paredes, C., Ault, S. . & Periago, M. . e neglected tropical diseases of Latin America and the
Caribbean: a review of disease burden and distribution and a roadmap for control and elimination. PLoS Negl Trop Dis 2, e300,
https://doi.org/10.1371/journal.pntd.0000300 (2008).
9. Cro, S. L., Sundar, S. & Fairlamb, A. H. Drug resistance in leishmaniasis. Clin Microbiol ev 19, 111–126, https://doi.org/10.1128/
CM.19.1.111-126.2006 (2006).
10. Nathan, C. & Shiloh, M. U. eactive oxygen and nitrogen intermediates in the relationship between mammalian hosts and microbial
pathogens. Proc Natl Acad Sci USA 97, 8841–8848 (2000).
11. Gregory, D. J. & Olivier, M. Subversion of host cell signalling by the protozoan parasite Leishmania. Parasitology 130(Suppl), S27–35,
https://doi.org/10.1017/S0031182005008139 (2005).
12. ossi, M. & Fasel, N. How to master the host immune system? Leishmania parasites have the solutions! Int Immunol 30, 103–111,
https://doi.org/10.1093/intimm/dxx075 (2018).
13. Bogdan, C. & öllingho, M. e immune response to Leishmania: mechanisms of parasite control and evasion. Int J Parasitol 28,
121–134 (1998).
14. Mantovani, A. et al. e chemoine system in diverse forms of macrophage activation and polarization. Trends in immunology 25,
677–686, https://doi.org/10.1016/j.it.2004.09.015 (2004).
15. Wang, N., Liang, H. & Zen, . Molecular mechanisms that inuence the macrophage m1-m2 polarization balance. Frontiers in
immunology 5, 614, https://doi.org/10.3389/mmu.2014.00614 (2014).
16. Gordon, S. & Taylor, P. . Monocyte and macrophage heterogeneity. Nat ev Immunol 5, 953–964, https://doi.org/10.1038/nri1733 (2005).
17. Lawrence, T. & Natoli, G. Transcriptional regulation of macrophage polarization: enabling diversity with identity. Nat ev Immunol
11, 750–761, https://doi.org/10.1038/nri3088 (2011).
18. Von Stebut, E. et al. Interleuin 1alpha promotes Th1 differentiation and inhibits disease progression in Leishmania major-
susceptible BALB/c mice. J Exp Med 198, 191–199, https://doi.org/10.1084/jem.20030159 (2003).
19. von Stebut, E. & Udey, M. C. equirements for 1-dependent immunity against infection with Leishmania major. Microbes Infect
6, 1102–1109, https://doi.org/10.1016/j.micinf.2004.05.024 (2004).
20. Alexander, J. & Brombacher, F. T helper1/t helper2 cells and resistance/susceptibility to leishmania infection: is this paradigm still
relevant? Front Immunol 3, 80, https://doi.org/10.3389/mmu.2012.00080 (2012).
21. Velasquez, L. G. et al. Distinct courses of infection with Leishmania (L.) amazonensis are observed in BALB/c, BALB/c nude and
C57BL/6 mice. Parasitology 143, 692–703, https://doi.org/10.1017/S003118201600024X (2016).
22. osas, L. E. et al. Genetic bacground inuences immune responses and disease outcome of cutaneous L. mexicana infection in
mice. Int Immun ol 17, 1347–1357, https://doi.org/10.1093/intimm/dxh313 (2005).
23. ong, F. e t al. Transcriptional Proling in Experimental Visceral Leishmaniasis eveals a Broad Splenic Inammatory Environment
that Conditions Macrophages toward a Disease-Promoting Phenotype. PLoS Pathog 13, e1006165, https://doi.org/10.1371/journal.
ppat.1006165 (2017).
24. Himmelrich, H., Parra-Lopez, C., Tacchini-Cottier, F., Louis, J. A. & Launois, P. e IL-4 rapidly produced in BALB/c mice aer
infection with Leishmania major down-regulates IL-12 receptor beta 2-chain expression on CD4+ T cells resulting in a state of
unresponsiveness to IL-12. J Immunol 161, 6156–6163 (1998).
25. Ji, J., Sun, J., Qi, H. & Soong, L. Analysis of T helper cell responses during infection with Leishmania amazonensis. Am J Trop Med
Hyg 66, 338–345 (2002).
26. Ji, J., Sun, J. & Soong, L. Impaired expression of inammatory cytoines and chemoines at early stages of infection with Leishmania
amazonensis. Infect Immun 71, 4278–4288 (2003).
27. Felizardo, T. C., Toma, L. S., Borges, N. B., Lima, G. M. & Abrahamsohn, I. A. Leishmania (Leishmania) amazonensis infection and
dissemination in mice inoculated with stationary-phase or with puried metacyclic promastigotes. Parasitology 134, 1699–1707,
https://doi.org/10.1017/S0031182007003186 (2007).
28. Soong, L. et al.  ole of CD4+ T cells in pathogenesis associated with Leishmania amazonensis infection. J Immunol 158, 5374–5383 (1997).
29. Jones, D. E., Acermann, M. ., Wille, U., Hunter, C. A. & Scott, P. Early enhanced 1 response aer Leishmania amazonensis
infection of C57BL/6 interleuin-10-decient mice does not lead to resolution of infection. Infe ct Immun 70, 2151–2158 (2002).
30. ane, M. M. & Mosser, D. M. e role of IL-10 in promoting disease progression in leishmaniasis. J Immunol 166, 1141–1147 (2001).
31. Jones, D. E., Buxbaum, L. U. & Scott, P. IL-4-independent inhibition of IL-12 responsiveness during Leishmania amazonensis
infection. J Immunol 165, 364–372 (2000).
32. Guerra, C. S. et al. Histopathological analysis of initial cellular response in TL-2 deficient mice experimentally infected by
Leishmania (L.) amazonensis. Int J Exp Pathol 91, 451–459, https://doi.org/10.1111/j.1365-2613.2010.00717.x (2010).
33. Dillon, L. A. et al. Simultaneous transcriptional proling of Leishmania major and its murine macrophage host cell reveals insights
into host-pathogen interactions. BMC Genomics 16, 1108, https://doi.org/10.1186/s12864-015-2237-2 (2015).
34. Christensen, S. M. et al. Meta-transcriptome Proling of the Human-Leishmania braziliensis Cutaneous Lesion. PLoS Negl Trop Dis
10, e0004992, https://doi.org/10.1371/journal.pntd.0004992 (2016).
35. Probst, C. M. et al. A comparison of two distinct murine macrophage gene expression profiles in response to Leishmania
amazonensis infection. BMC Microbiol 12, 22, https://doi.org/10.1186/1471-2180-12-22 (2012).
36. Ontoria, E. et al. Transcriptional Proling of Immune-elated Genes In. Front Cell Infect Microbiol 8, 197, https://doi.org/10.3389/
fcimb.2018.00197 (2018).
37. Acuña, S. M. et al. Arginase expression modulates nitric oxide production in Leishmania (Leishmania) amazonensis. PLoS One 12,
e0187186, https://doi.org/10.1371/journal.pone.0187186 (2017).
38. Aoi, J. I. et al. L-arginine availability and arginase activity: Characterization of amino acid permease 3 in Leishmania amazonensis.
PLoS Negl Trop Dis 11, e0006025, https://doi.org/10.1371/journal.pntd.0006025 (2017).
39. Aoi, J. I. et al. NA-seq transcriptional proling of Leishmania amazonensis reveals an arginase-dependent gene expression
regulation. PLoS Negl Trop Dis 11, e0006026, https://doi.org/10.1371/journal.pntd.0006026 (2017).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
13
Scientific RepoRtS | (2019) 9:19841 | https://doi.org/10.1038/s41598-019-56305-1
www.nature.com/scientificreports
www.nature.com/scientificreports/
40. astrojo, A . et al. e transcriptome of Leishmania major in the axenic promastigote stage: transcript annotation and relative
expression levels by NA-seq. BMC Genomics 14, 223, https://doi.org/10.1186/1471-2164-14-223 (2013).
41. Fernandes, M. C. et al. Dual Transcriptome Proling of Leishmania-Infected Human Macrophages eveals Distinct eprogramming
Signatures. MBio 7, https://doi.org/10.1128/mBio.00027-16 (2016).
42. Fiebig, M., elly, S. & Gluenz, E. Comparative Life Cycle Transcriptomics evises Leishmania mexicana Genome Annotation and
Lins a Chromosome Duplication with Parasitism of Vertebrates. PLoS Pathog 11, e1005186, https://doi.org/10.1371/journal.
ppat.1005186 (2015).
43. Veras, P. S. T., amos, P. I. P. & de Menezes, J. P. B. In Search of Biomarers for Pathogenesis and Control of Leishmaniasis by Global
Analyses of. Front Cell Infect Microbiol 8, 326, https://doi.org/10.3389/fcimb.2018.00326 (2018).
44. ip, A. E. et al. Systematic review of biomarers to monitor therapeutic response in leishmaniasis. Antimicrob Agents Chemother 59,
1–14, https://doi.org/10.1128/AAC.04298-14 (2015).
45. Bahrami, F., Harandi, A. M. & afati, S. Biomarers of Cutaneous Leishmaniasis. Front Cell Infect Microbiol 8, 222, https://doi.
org/10.3389/fcimb.2018.00222 (2018).
46. Dillon, L. A. et al. Transcriptomic proling of gene expression and NA processing during Leishmania major dierentiation. Nucleic
Acids es 43, 6799–6813, https://doi.org/10.1093/nar/gv656 (2015).
47. Jenner, . G. & Young, . A. Insights into host responses against pathogens from transcriptional proling. Nat ev Microbiol 3,
281–294, https://doi.org/10.1038/nrmicro1126 (2005).
48. Paladi, C. S. et al. Treatment of Leishmania (Leishmania) amazonensis-infected mice with a combination of a Palladacycle complex
and heat-illed Propionibacterium acnes triggers protective cellular immune response. Front Microbiol 8, 333, https://doi.
org/10.3389/fmicb.2017.00333 (2017).
49. Fernandes, J. C. . et al. Melatonin and Leishmania amazonensis infection altered mi-294, mi-30e, and mi-302d i mpacting on
Tnf, Mcp-1, and Nos2 e xpression. Front Cell Infect Microbiol 9, 60, https://doi.org/10.3389/fcimb.2019.00060 (2019).
50. Muxel, S. M., Laranjeira-Silva, M. F., Zampieri, . A. & Floeter-Winter, L. M. Leishmania (Leishmania) amazonensis induces
macrophage mi-294 and mi-721 expression and modulates infection by targeting NOS2 and L-arginine metabolism. Sci ep 7,
44141, https://doi.org/10.1038/srep44141 (2017).
51. Matte, C. & Olivier, M. Leishmania-induced cellular recruitment during the early inflammatory response: modulation of
proinammatory mediators. J Infect Dis 185, 673–681, https://doi.org/10.1086/339260 (2002).
52. Lima-Junior, D. S. et al. Inammasome-derived IL-1beta production induces nitric oxide-mediated resistance to Leishmania. Nature
medicine 19, 909–915, https://doi.org/10.1038/nm.3221 (2013).
53. Fernández-Figueroa, E. A. et al. Disease severity in patients infected with Leishmania mexicana relates to IL-1β. PLoS Negl Trop Dis
6, e1533, https://doi.org/10.1371/journal.pntd.0001533 (2012).
54. Charmoy, M. et al. e Nlrp3 inammasome, IL-1β, and neutrophil recruitment are required for susceptibility to a nonhealing
strain of Leishmania major in C57BL/6 mice. Eur J Immunol 46, 897–911, https://doi.org/10.1002/eji.201546015 (2016).
55. Liu, D. & Uzonna, J. E. e early interaction of Leishmania with macrophages and dendritic cells and its inuence on the host
immune response. Front Cell Infect Microbiol 2, 83, https://doi.org/10.3389/fcimb.2012.00083 (2012).
56. Nascimento, M. S. et al. NOD2-IP2-Mediated Signaling Helps Shape Adaptive Immunity in Visceral Leishmaniasis. e Journal of
infectious diseases 214, 1647–1657, https://doi.org/10.1093/infdis/jiw446 (2016).
57. Dos Santos, J. C. et al. e NOD2 receptor is crucial for immune responses towards New World Leishmania species. Sci ep 7, 15219,
https://doi.org/10.1038/s41598-017-15412-7 (2017).
58. Yang, Z., Mosser, D. M. & Zhang, X. Activation of the MAP, E, following Leishmania amazonensis infection of macrophages. J
Immunol 178, 1077–1085 (2007).
59. Zambrano-Villa, S., osales-Borjas, D., Carrero, J. C. & Ortiz-Ortiz, L. How protozoan parasites evade the immune response. Trend s
in parasitology 18, 272–278 (2002).
60. Sousa, . et al. Early Suppression of Macrophage Gene Expression by. Front Microbiol 9, 2464, https://doi.org/10.3389/
fmicb.2018.02464 (2018).
61. Cunningham, A. C. Parasitic adaptive mechanisms in infection by leishmania. Exp Mol Pathol 72, 132–141, https://doi.org/10.1006/
exmp.2002.2418 (2002).
62. Murray, H. W., Jungbluth, A., itter, E., Montelibano, C. & Marino, M. W. Visceral leishmaniasis in mice devoid of tumor necrosis
factor and response to treatment. Infect Immun 68, 6289–6293 (2000).
63. oach, D. . et al. TNF regulates chemoine induction essential for cell recruitment, granuloma formation, and clearance of
mycobacterial infection. J Immunol 168, 4620–4627 (2002).
64. acoosin, E. L. & Beverley, S. M. Leishmania major: promastigotes induce expression of a subset of chemoine genes in murine
macrophages. Exp Parasitol 85, 283–295, https://doi.org/10.1006/expr.1996.4139 (1997).
65. itter, U. et al. Dierential expression of chemoines in patients with localized and diuse cutaneous American leishmaniasis. J
Infect Dis 173, 699–709 (1996).
66. Teixeira, M. J., Teixeira, C. ., Andrade, B. B., Barral-Netto, M. & Barral, A. Chemoines in host-parasite interactions in
leishmaniasis. Trends Parasitol 22, 32–40, https://doi.org/10.1016/j.pt.2005.11.010 (2006).
67. Oghumu, S., Lezama-Dávila, C. M., Isaac-Márquez, A. P. & Satosar, A. . ole of chemoines in regulation of immunity against
leishmaniasis. Exp Parasitol 126, 389–396, https://doi.org/10.1016/j.exppara.2010.02.010 (2010).
68. Yurcheno, E. et al. CC5-dependent homing of naturally occurring CD4+ regulatory T cells to sites of Leishmania major infection
favors pathogen persistence. J Exp Med 203, 2451–2460, https://doi.org/10.1084/jem.20060956 (2006).
69. amírez, C. et al. Human macrophage response to L. (Viannia) panamensis: microarray evidence for an early inflammatory
response. PLoS Negl Trop Dis 6, e1866, https://doi.org/10.1371/journal.pntd.0001866 (2012).
70. Wang, Y., Zhou, Y. & Graves, D. T. FOXO transcription factors: their clinical signicance and regulation. Biomed es Int 2014,
925350, https://doi.org/10.1155/2014/925350 (2014).
71. Gupta, P., Srivastav, S., Saha, S., Das, P. . & Uil, A. Leishmania donovani inhibits macrophage apoptosis and pro-inammatory
response through AT-mediated regulation of β-catenin and FOXO-1. Cell Death Dier 23, 1815–1826, https://doi.org/10.1038/
cdd.2016.101 (2016).
72. Tu, L. et al. Notch signaling is an important regulator of type 2 immunity. J Exp Med 202, 1037–1042, https://doi.org/10.1084/
jem.20050923 (2005).
73. Auderset, F. et al. edundant Notch1 and Notch2 signaling is necessary for IFNγ secretion by T helper 1 cells during infection with
Leishmania major. PLoS Pathog 8, e1002560, https://doi.org/10.1371/journal.ppat.1002560 (2012).
74. Chen, ., ang, ., Fan, X. G. & Tang, D. elease and activity of histone in diseases. Cell Death Dis 5, e1370, https://doi.org/10.1038/
cddis.2014.337 (2014).
75. Canonne-Hergaux, F., Gruenheid, S., Govoni, G. & Gros, P. The Nramp1 protein and its role in resistance to infection and
macrophage function. Proc Assoc Am Physicians 111, 283–289 (1999).
76. Blacwell, J. M. et al. SLC11A1 (formerly NAMP1) and disease resistance. Cell Microbiol 3, 773–784 (2001).
77. Mittra, B. et al. Iron uptae controls the generation of Leishmania infective forms through regulation of OS levels. J Exp Med 210,
401–416, https://doi.org/10.1084/jem.20121368 (2013).
78. Huynh, C., Sacs, D. L. & Andrews, N. W. A Leishmania amazonensis ZIP family iron transporter is essential for parasite replication
within macrophage phagolysosomes. J Exp Med 203, 2363–2375, https://doi.org/10.1084/jem.20060559 (2006).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
14
Scientific RepoRtS | (2019) 9:19841 | https://doi.org/10.1038/s41598-019-56305-1
www.nature.com/scientificreports
www.nature.com/scientificreports/
79. Lüder, C. G., Campos-Salinas, J., Gonzalez-ey, E. & van Zandbergen, G. Impact of protozoan cell death on parasite-host
interactions and pathogenesis. Parasit Vectors 3, 116, https://doi.org/10.1186/1756-3305-3-116 (2010).
80. Ginhoux, F., Schultze, J. L., Murray, P. J., Ochando, J. & Biswas, S. . New insights into the multidimensional concept of macrophage
ontogeny, activation and function. Nat Immunol 17, 34–40, https://doi.org/10.1038/ni.3324 (2016).
81. Ginhoux, F. & Jung, S. Monocytes and macrophages: developmental pathways and tissue homeostasis. Nat ev Immunol 14,
392–404, https://doi.org/10.1038/nri3671 (2014).
82. Oabe, Y. & Medzhitov, . Tissue-specific signals control reversible program of localization and functional polarization of
macrophages. Cell 157, 832–844, https://doi.org/10.1016/j.cell.2014.04.016 (2014).
83. Lavin, Y. et al. Tissue-resident macrophage enhancer landscapes are shaped by the local microenvironment. Cell 159, 1312–1326,
https://doi.org/10.1016/j.cell.2014.11.018 (2014).
84. Wang, C. et al. Characterization of murine macrophages from bone marrow, spleen and peritoneum. BMC Immunol 14, 6, https://
doi.org/10.1186/1471-2172-14-6 (2013).
85. Tran, D. D. et al. Transcriptional regulation of immediate-early gene response by THOC5, a member of mNA export complex,
contributes to the M-CSF-induced macrophage dierentiation. Cell Death Dis 4, e879, https://doi.org/10.1038/cddis.2013.409 (2013).
86. Tubo, N. J. & Jenins, M. . CD4+ T Cells: guardians of the phagosome. Clin Microbiol ev 27, 200–213, https://doi.org/10.1128/
CM.00097-13 (2014).
87. omano, A. et al. Divergent roles for Ly6C+CC2+CX3C1+ inammatory monocytes during primary or secondary infection of
the sin with the intra-phagosomal pathogen Leishmania major. PLoS Pathog 13, e1006479, https://doi.org/10.1371/journal.
ppat.1006479 (2017).
88. ibeiro-Gomes, F. L. et al. Site-dependent recruitment of inammatory cells determines the eective dose of Leishmania major.
Infec t Immun 82, 2713–2727, https://doi.org/10.1128/IAI.01600-13 (2014).
89. Dumas, C., Chow, C., Müller, M. & Papadopoulou, B. A novel class of developmentally regulated noncoding NAs in Leishmania.
Euaryot C ell 5, 2033–2046, https://doi.org/10.1128/EC.00147-06 (2006).
90. Aoi, J. I., Yamashiro-anashiro, E. H., amos, D. C. C. & Cotrim, P. C. Ecacy of the tubercidin antileishmania action associated with
an inhibitor of the nucleoside transport. Parasitology  esearch 104, 223–228, https://doi.org/10.1007/s00436-008-1177-z (2009).
91. do Socorro S osa, Mo. S. et al. Antileishmanial activity of a linalool-rich essential oil from Croton cajucara. Antimicrob Agents
Chemother 47, 1895–1901 (2003).
92. Van der Auwera, G. A. et al. From FastQ data to high condence variant calls: the Genome Analysis Toolit best practices pipeline.
Curr Protoc Bioinformatics 43, 11.10.11–33, https://doi.org/10.1002/0471250953.bi1110s43 (2013).
93. Bolger, A. M., Lohse, M. & Usadel, B. Trimmomatic: a exible trimmer for Illumina sequence data. Bioinformatics 30, 2114–2120,
https://doi.org/10.1093/bioinformatics/btu170 (2014).
94. im, D. et al . TopHat2: accurate alignment of transcriptomes in the presence of insertions, deletions and gene fusions. Genome Biol
14, 36, https://doi.org/10.1186/gb-2013-14-4-r36 (2013).
95. Trapnell, C., Pachter, L. & Salzberg, S. L. TopHat: discovering splice junctions with NA-Seq. Bioinformatics 25, 1105–1111, https://
doi.org/10.1093/bioinformatics/btp120 (2009).
96. Trapnell, C. et al. Dierential gene and transcript expression analysis of NA-seq experiments with TopHat and Cuins. Nat
Protoc 7, 562–578, https://doi.org/10.1038/nprot.2012.016 (2012).
97. Mortazavi, A., Williams, B. A., McCue, ., Schaeer, L. & Wold, B. Mapping and quantifying mammalian transcriptomes by NA-
Seq. Nat Methods 5, 621–628, https://doi.org/10.1038/nmeth.1226 (2008).
98. Benjamini, Y., Drai, D., Elmer, G., afa, N. & Golani, I. Controlling the false discovery rate in behavior genetics research. Behav
Brain es 125, 279–284 (2001).
99. Liva, . J. & Schmittgen, T. D. Analysis of relative gene expression data using real-time quantitative PC and the 2(-Delta Delta
C(T)) Method. Methods 25, 402–408, https://doi.org/10.1006/meth.2001.1262 (2001).
Acknowledgements
We would like to thank Juliane Cristina Ribeiro Fernandes and Stephanie Maia Acuña for their comments and
suggestions.
Author contributions
Conceived and designed the experiments: J.I.A., S.M.M., L.M.F.W. Performed the experiments: J.I.A., S.M.M.,
R.A.Z. Analyzed the data: J.I.A., S.M.M., L.M.F.W. Contributed reagents/materials/analysis tools: J.I.A., K.E.M.,
A.H.N., L.M.F.W. Wrote the dra of the manuscript: J.I.A., S.M.M. and L.M.F.W. Revised the manuscript: J.I.A.,
S.M.M., R.A.Z., K.E.M., A.H.N., L.M.F.W.
Competing interests
e authors declare no competing interests.
Additional information
Supplementary information is available for this paper at https://doi.org/10.1038/s41598-019-56305-1.
Correspondence and requests for materials should be addressed to J.I.A. or L.M.F.-W.
Reprints and permissions information is available at www.nature.com/reprints.
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional aliations.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Cre-
ative Commons license, and indicate if changes were made. e images or other third party material in this
article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons license and your intended use is not per-
mitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the
copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.
© e Author(s) 2019
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... Previous studies have compared the differential responses of macrophages from resistant and susceptible mice induced by Leishmania (6,7). However, to the best of our knowledge, no study has focused on the impact of the basal transcriptomic profile of macrophages, from different host backgrounds, on their response to Leishmania infection. ...
... Global transcriptomic events associated with M-CSFdependent monocyte-to-macrophage differentiation have been analyzed and associated with major changes in the global transcriptomic profile (17). Moreover, differential basal backgrounds in non-infected BALB/c vs. non-infected C57BL/6 L929 differentiated macrophages have also been previously reported (6). Among the DEGs between the two mouse strains, our analysis showed that Cathepsin E (Supplementary File 1) is one of the most strongly modulated genes in the BALB/c derived macrophages. ...
... In particular, these profiles could account for the limited amplitude of the inflammatory response observed in Leishmaniainfected BALB/c BMdMs, while the conditions that prevail in uninfected C57BL/6 BMdMs could explain the greater magnitude of the immune response induced by Leishmania infection in resistant BMDMs (Figure 2). Similarly, L. amazoniensis has also been reported to induce a stronger immune response activation in BMDMs from C57BL/6 mice as compared to BALB/c mice (6). In vivo, these different amounts of cytokines/chemokines produced by the macrophages in different backgrounds could affect leukocyte recruitment, especially at the site of infection. ...
Article
Full-text available
Leishmaniases are a group of diseases with different clinical manifestations. Macrophage-Leishmania interactions are central to the course of the infection. The outcome of the disease depends not only on the pathogenicity and virulence of the parasite, but also on the activation state, the genetic background, and the underlying complex interaction networks operative in the host macrophages. Mouse models, with mice strains having contrasting behavior in response to parasite infection, have been very helpful in exploring the mechanisms underlying differences in disease progression. We here analyzed previously generated dynamic transcriptome data obtained from Leishmania major (L. major) infected bone marrow derived macrophages (BMdMs) from resistant and susceptible mouse. We first identified differentially expressed genes (DEGs) between the M-CSF differentiated macrophages derived from the two hosts, and found a differential basal transcriptome profile independent of Leishmania infection. These host signatures, in which 75% of the genes are directly or indirectly related to the immune system, may account for the differences in the immune response to infection between the two strains. To gain further insights into the underlying biological processes induced by L. major infection driven by the M-CSF DEGs, we mapped the time-resolved expression profiles onto a large protein-protein interaction (PPI) network and performed network propagation to identify modules of interacting proteins that agglomerate infection response signals for each strain. This analysis revealed profound differences in the resulting responses networks related to immune signaling and metabolism that were validated by qRT-PCR time series experiments leading to plausible and provable hypotheses for the differences in disease pathophysiology. In summary, we demonstrate that the host’s gene expression background determines to a large degree its response to L. major infection, and that the gene expression analysis combined with network propagation is an effective approach to help identifying dynamically altered mouse strain-specific networks that hold mechanistic information about these contrasting responses to infection.
... The complete lists of genes are provided in S1 Table (24h) and S2 Table (48h). A direct comparison of deregulated genes at each timepoint revealed that the macrophage response to infection at 24 h was more pronounced compared to the response at 48 h, confirming previous findings in murine macrophage infection models, which evidenced a greater gene expression modulation during early infection and a progressive reduction at later timepoints until 72 h post-infection [4,6,28,29]. A fraction of these genes was consistently up-or downregulated both at 24 h and 48h, while another fraction (19% to 69%) was deregulated uniquely at 48h post-infection (Tables 2 and S3). ...
Article
Full-text available
Background In the Mediterranean basin, three Leishmania species have been identified: L . infantum , L . major and L . tropica , causing zoonotic visceral leishmaniasis (VL), zoonotic cutaneous leishmaniasis (CL) and anthroponotic CL, respectively. Despite animal models and genomic/transcriptomic studies provided important insights, the pathogenic determinants modulating the development of VL and CL are still poorly understood. This work aimed to identify host transcriptional signatures shared by cells infected with L . infantum , L . major , and L . tropica , as well as specific transcriptional signatures elicited by parasites causing VL (i.e., L . infantum ) and parasites involved in CL (i.e., L . major , L . tropica ). Methodology/Principal findings U937 cells differentiated into macrophage-like cells were infected with L . infantum , L . major and L . tropica for 24h and 48h, and total RNA was extracted. RNA sequencing, performed on an Illumina NovaSeq 6000 platform, was used to evaluate the transcriptional signatures of infected cells with respect to non-infected cells at both time points. The EdgeR package was used to identify differentially expressed genes (fold change > 2 and FDR-adjusted p-values < 0.05). Then, functional enrichment analysis was employed to identify the enriched ontology terms in which these genes are involved. At 24h post-infection, a common signature of 463 dysregulated genes shared among all infection conditions was recognized, while at 48h post-infection the common signature was reduced to 120 genes. Aside from a common transcriptional response, we evidenced different upregulated functional pathways characterizing L . infantum -infected cells, such as VEGFA-VEGFR2 and NFE2L2-related pathways, indicating vascular remodeling and reduction of oxidative stress as potentially important factors for visceralization. Conclusions The identification of pathways elicited by parasites causing VL or CL could lead to new therapeutic strategies for leishmaniasis, combining the canonical anti-leishmania compounds with host-directed therapy.
... Rab15 is involved in positive regulation of regulated secretory pathway-associated immunodeficiency [61]. The Fcgr1 gene encodes a protein that plays an important role in the immune response, acting as a high-affinity Fc-gamma receptor [62]. Upon examining the KEGG pathway using these genes, the cAMP signaling pathway, anti-inflammatory response, inflammatory mediator regulation of TRP channels, and FcγR-mediated phagocytosis pathways in cancer were included. ...
Article
Full-text available
This study investigates the immunomodulatory potential of Galium aparine L. (GAE) in immunodeficient animals. In this study, animals were categorized into five groups: the normal group, CYP group (cyclophosphamide intraperitoneal injection), GA5 group (cyclophosphamide + 5 μg GAE), GA50 group (cyclophosphamide + 50 μg GAE), and GA500 group (cyclophosphamide + 500 μg GAE). The CYP group exhibited significantly reduced spleen weights compared to the normal group, while the groups obtaining GAE displayed a dose-dependent increase in spleen weight. Furthermore, the GAE demonstrated dose-dependent enhancement of splenocyte proliferating activity, with significant increases observed in both LPS and ConA-induced assays. NK cell activity significantly increased in the GA50 and GA500 groups compared to the CYP group. Cytokine analysis revealed a significant increase in IL-6, TNF-α, and IFN-γ levels in ConA-induced splenocytes treated with GAE. Gene expression analysis identified 2434 DEG genes in the extract groups. Notable genes, such as Entpd1, Pgf, Thdb, Syt7, Sqor, and Rsc1al, displayed substantial differences in individual gene expression levels, suggesting their potential as target genes for immune enhancement. In conclusion, Galium aparine L. extract exhibits immunomodulatory properties. The observed gene expression changes further support the potential of Galium aparine L. extract as a natural agent for immune augmentation.
... The CEBPB (ENSG00000172216) TF, a member of the bZIP family, is related to the main immune response processes. The expression of CEBPB gene up-regulated in response to Leishmania infection, as previously reported [61]. The CTCF (ENSG00000102974) (a C2H2 ZF TF) mediates the macrophage function of mice infected with Leishmania major [62]. ...
Article
Full-text available
Background Leishmaniasis is a parasitic disease caused by the Leishmania protozoan affecting millions of people worldwide, especially in tropical and subtropical regions. The immune response involves the activation of various cells to eliminate the infection. Understanding the complex interplay between Leishmania and the host immune system is crucial for developing effective treatments against this disease. Methods This study collected extensive transcriptomic data from macrophages, dendritic, and NK cells exposed to Leishmania spp. Our objective was to determine the Leishmania -responsive genes in immune system cells by applying meta-analysis and feature selection algorithms, followed by co-expression analysis. Results As a result of meta-analysis, we discovered 703 differentially expressed genes (DEGs), primarily associated with the immune system and cellular metabolic processes. In addition, we have substantiated the significance of transcription factor families, such as bZIP and C2H2 ZF, in response to Leishmania infection. Furthermore, the feature selection techniques revealed the potential of two genes, namely G0S2 and CXCL8 , as biomarkers and therapeutic targets for Leishmania infection. Lastly, our co-expression analysis has unveiled seven hub genes, including PFKFB3 , DIAPH1 , BSG , BIRC3 , GOT2 , EIF3H , and ATF3 , chiefly related to signaling pathways. Conclusions These findings provide valuable insights into the molecular mechanisms underlying the response of immune system cells to Leishmania infection and offer novel potential targets for the therapeutic goals.
... The role of IL-1β and IL-18 in Leishmania infection has been the subject of numerous studies [39,46,48]. It was reported that IL-1β can modulate the immune responses, while IL-18 shifts the T-cell activation pathway towards Th2, however, both cytokines contribute to the progression of the disease [50,51]. Notably, IL-1β has been identified as a significant signaling factor for host resistance against infection, as this cytokine transmits signals through IL-1R and myeloid differentiation primary response protein (MyD) 88, leading to the induction of NOS2-mediated nitric oxide (NO) production. ...
Article
Full-text available
Background Cutaneous Leishmaniasis (CL) is a parasitic disease with diverse outcomes. Clinical diversity is influenced by various factors such as Leishmania species and host genetic background. The role of Leishmania RNA virus (LRV), as an endosymbiont, is suggested to not only affect the pathogenesis of Leishmania, but also impact host immune responses. This study aimed to investigate the influence of LRV2 on the expression of a number of virulence factors (VFs) of Leishmania and pro-inflammatory biomarkers. Materials and methods Sample were obtained from CL patients from Golestan province. Leishmania species were identified by PCR (LIN 4, 17), and the presence of LRV2 was checked using the semi-nested PCR (RdRp gene). Human monocyte cell line (THP-1) was treated with three isolates of L. major with LRV2 and one isolate of L. major without LRV2. The treatments with four isolates were administered for the time points: zero, 12, 24, 36, and 48 h after co-infection. The expression levels of Leishmania VFs genes including GP63, HSP83, and MPI, as well as pro-inflammatory biomarkers genes including NLRP3, IL18, and IL1β, were measured using quantitative real-time PCR. Results The expression of GP63, HSP83, and MPI revealed up-regulation in LRV2 + isolates compared to LRV2- isolates. The expression of the pro-inflammatory biomarkers including NLRP3, IL1β, and IL18 genes in LRV2- were higher than LRV2 + isolates. Conclusion This finding suggests that LRV2 + may have a probable effect on the Leishmania VFs and pro-inflammatory biomarkers in the human macrophage model.
... Balb/c mice have strong Th2 immune response, a common response in allergic reactions and infectious diseases, high production of IgG1 and IgA, high complement activity, and low interferon production. Additionally, Balb/c mice produce stronger humoral response against antigens in comparison with C57BL/6 strain (Bleul et al., 2021;Aoki et al., 2019;Zhang et al., 2023;Laurenti et al., 2004;Kuroda et al., 2002;Koo and Gan, 2006;Watanabe et al., 2004). ...
Article
Full-text available
Inflammation is a general term for a wide variety of both physiological and pathophysiological processes in the body which primarily prevents the body from diseases and helps to remove dead tissues. It has a crucial part in the body immune system. Tissue damage can recruit inflammatory cells and cytokines and induce inflammation. Inflammation can be classified as acute, sub-acute, and chronic. If it remained unresolved and lasted for prolonged periods, it would be considered as chronic inflammation (CI), which consequently exacerbates tissue damage in different organs. CI is the main pathophysiological cause of many disorders such as obesity, diabetes, arthritis, myocardial infarction, and cancer. Thus, it is critical to investigate different mechanisms involved in CI to understand its processes and to find proper anti-inflammatory therapeutic approaches for it. Animal models are one of the most useful tools for study about different diseases and mechanisms in the body, and are important in pharmacological studies to find proper treatments. In this study, we discussed the various experimental animal models that have been used to recreate CI which can help us to enhance the understanding of CI mechanisms in human and contribute to the development of potent new therapies.
... Intriguingly, the functional enrichment analysis carried out using DAVID revealed that many pathways related to parasitic diseases such as amoebiasis, trypanosomiasis, toxoplasmosis, and leishmaniasis were also enriched in the up-regulated dataset of the infected animals (O'Gorman et al., 2006;Aoki et al., 2019;Oliveira et al., 2020;Kalavi et al., 2021;Salloum et al., 2021;Sun et al., 2021;Wang et al., 2022). The constituent genes associated with these infections and their expression statistics observed in our study are shown in Table 2. ...
Article
Bovine anaplasmosis caused by Anaplasma marginale is a tick-borne disease of livestock with widespread prevalence and huge economic implications. In order to get new insights into modulation of host gene expression in response to natural infections of anaplasmosis, this study is the first attempt that compared the transcriptome profiles of peripheral blood mononuclear cells (PBMCs) of A. marginale infected and healthy crossbred cattle. Transcriptome analysis identified shared as well as unique functional pathways in the two groups. Translation and structural constituent of ribosome were the important terms for the genes abundantly expressed in the infected as well as healthy animals. Gene Ontology (GO) and Kyoto Encyclopedia of Genes and Genomes (KEGG) enrichment analysis of the differentially expressed genes revealed that immunity and signal transduction related terms were enriched for the up-regulated genes in the infected animals. The over-represented pathways were cytokine-cytokine receptor interaction and signaling pathways involving chemokines, Interleukin 17 (IL17), Tumour Necrosis Factor (TNF), Nuclear Factor Kappa B (NFKB) etc. Interestingly, many genes previously associated with parasite-borne diseases such as amoebiasis, trypanosomiasis, toxoplasmosis, and leishmaniasis were profusely expressed in the dataset of the diseased animals. High expression was also evident for the genes for acute phase response proteins, anti-microbial peptides and many inflammatory cytokines. Role of cytokines in mediating communication between immune cells was the most conspicuous gene network identified through the Ingenuity Pathway Analysis. This study provides comprehensive information about the crosstalk of genes involved in host defense as well as parasite persistence in the host upon infection with A. marginale.
... Spermidine reduces the secretion of TNF-α and IL-1β in LPS-stimulated RAW 264.7 macrophages [67] and MCP-1 secretion in THP-1-macrophages treated with IFN-γ [90]. IL-1β induces NOS2 and NO production and resistance to infection in C57BL/6 BMDM infected with L. amazonensis [91], and transcriptome data showed downregulation of Il1b in L. amazonensis infected BALB/c-BMDM [92]. In L. amazonensis skin lesions on C57BL/6 mice, the lack of CCR2 (receptor for MCP-1) CD11b +cells showed lower ARG1 and NOS2 and a reduction in parasite load [70]. ...
Article
Full-text available
Leishmania is a protozoan that causes leishmaniasis, a neglected tropical disease with clinical manifestations classified as cutaneous, mucocutaneous, and visceral leishmaniasis. In the infection context, the parasite can modulate macrophage gene expression affecting the microbicidal activity and immune response. The metabolism of L-arginine into polyamines putrescine, spermidine, and spermine reduces nitric oxide (NO) production, favoring Leish-mania survival. Here, we investigate the effect of supplementation with L-arginine and poly-amines in infection of murine BALB/c macrophages by L. amazonensis and in the transcriptional regulation of genes involved in arginine metabolism and proinflammatory response. We showed a reduction in the percentage of infected macrophages upon putres-cine supplementation compared to L-arginine, spermidine, and spermine supplementation. Unexpectedly, deprivation of L-arginine increased nitric oxide synthase (Nos2) gene expression without changes in NO production. Putrescine supplementation increased transcript levels of polyamine metabolism-related genes Arg2, ornithine decarboxylase (Odc1), Sper-midine synthase (SpdS), and Spermine synthase (SpmS), but reduced Arg1 in L. amazo-nensis infected macrophages, while spermidine and spermine promoted opposite effects. Putrescine increased Nos2 expression without leading to NO production, while L-arginine plus spermine led to NO production in uninfected macrophages, suggesting that polyamines can induce NO production. Besides, L-arginine supplementation reduced Il-1b during infection , and L-arginine or L-arginine plus putrescine increased Mcp1 at 24h of infection, suggesting that polyamines availability can interfere with cytokine/chemokine production. Our data showed that putrescine shifts L-arginine-metabolism related-genes on BALB/c macro-phages and affects infection by L. amazonensis.
... Previous studies determined the transcriptome of Li infection in THP-1 derived macrophages (Gatto et al., 2020), La and L. major infection in primary human macrophages (hMDM) (Fernandes et al., 2016) and Lb infection in patient's lesions (Maretti-Mira et al., 2012). In murine macrophages, the transcriptome of BALB/c and C57BL/6 macrophages infected with La indicated an inflammatory response different from the spectrum extremes M1 and M2 polarized macrophages (Osorio y Fortéa et al., 2009;Aoki et al., 2019) observed in L. major (Sacks & Noben-Trauth, 2002). All of the above-mentioned results of transcriptome-wide experiments exhibit some contrasting results with the literature because macrophage response to Leishmania is highly dependent on the parasite species and strain and host cell type, thus the importance of investigating different models (Salloum et al., 2021). ...
Article
Full-text available
It is well established that infection with Leishmania alters the host cell’s transcriptome. Since mammalian cells have multiple mechanisms to control gene expression, different molecules, such as noncoding RNAs, can be involved in this process. MicroRNAs have been extensively studied upon Leishmania infection, but whether long noncoding RNAs (lncRNAs) are also altered in macrophages is still unexplored. We performed RNA-seq from THP-1-derived macrophages infected with Leishmania amazonensis (La), L. braziliensis (Lb), and L. infantum (Li), investigating a previously unappreciated fraction of macrophage transcriptome. We found that more than 24% of the total annotated transcripts and 30% of differentially expressed (DE) RNAs in Leishmania-infected macrophage correspond to lncRNAs. LncRNAs and protein coding RNAs with altered expression are similar among macrophages infected with the Leishmania species. Still, some species-specific alterations could occur due to distinct pathophysiology in which Li infection led to a more significant number of exclusively DE RNAs. The most represented classes among DE lncRNAs were intergenic and antisense lncRNAs. We also found enrichment for immune response-related pathways in the DE protein coding RNAs, as well as putative targets of the lncRNAs. We performed a coexpression analysis to explore potential cis regulation of coding and antisense noncoding transcripts. We identified that antisense lncRNAs are similarly regulated as its neighbor protein coding genes, such as the BAALC/BAALC-AS1, BAALC/BAALC-AS2, HIF1A/HIF1A-AS1, HIF1A/HIF1A-AS3 and IRF1/IRF1-AS1 pairs, which can occur as a species-specific modulation. These findings are a novelty in the field because, to date, no study has focused on analyzing lncRNAs in Leishmania-infected macrophage. Our results suggest that lncRNAs may account for a novel mechanism by which Leishmania can control macrophage function. Further research must validate putative lncRNA targets and provide additional prospects in lncRNA function during Leishmania infection.
Article
Full-text available
Leishmaniases are neglected diseases that cause a large spectrum of clinical manifestations, from cutaneous to visceral lesions. The initial steps of the inflammatory response involve the phagocytosis of Leishmania and the parasite replication inside the macrophage phagolysosome. Melatonin, the darkness-signaling hormone, is involved in modulation of macrophage activation during infectious diseases, controlling the inflammatory response against parasites. In this work, we showed that exogenous melatonin treatment of BALB/c macrophages reduced Leishmania amazonensis infection and modulated host microRNA (miRNA) expression profile, as well as cytokine production such as IL-6, MCP-1/CCL2, and, RANTES/CCL9. The role of one of the regulated miRNA (miR-294-3p) in L. amazonensis BALB/c infection was confirmed with miRNA inhibition assays, which led to increased expression levels of Tnf and Mcp-1/Ccl2 and diminished infectivity. Additionally, melatonin treatment or miR-30e-5p and miR-302d-3p inhibition increased nitric oxide synthase 2 (Nos2) mRNA expression levels and nitric oxide (NO) production, altering the macrophage activation state and reducing infection. Altogether, these data demonstrated the impact of melatonin treatment on the miRNA profile of BALB/c macrophage infected with L. amazonensis defining the infection outcome.
Article
Full-text available
Leishmania braziliensis is an intracellular parasite that resides mostly in macrophages. Both the parasite genome and the clinical disease manifestations show considerable polymorphism. Clinical syndromes caused by L. braziliensis include localized cutaneous (CL), mucosal (ML), and disseminated leishmaniasis (DL). Our prior studies showed that genetically distinct L. braziliensis clades associate with different clinical types. Herein, we hypothesized that: (1) L. braziliensis induces changes in macrophage gene expression that facilitates infection; (2) infection of macrophages with strains associated with CL (clade B), ML (clade C), or DL (clade A) will differentially affect host cell gene expression, reflecting their different pathogenic mechanisms; and (3) differences between the strains will be reflected by differences in macrophage gene expression after initial exposure to the parasite. Human monocyte derived macrophages were infected with L. braziliensis isolates from clades A, B, or C. Patterns of gene expression were compared using Affymetrix DNA microarrays. Many transcripts were significantly decreased by infection with all isolates. The most dramatically decreased transcripts encoded proteins involved in signaling pathways, apoptosis, or mitochondrial oxidative phosphorylation. Some transcripts encoding stress response proteins were up-regulated. Differences between L. braziliensis clades were observed in the magnitude of change, rather than the identity of transcripts. Isolates from subjects with metastatic disease (ML and DL) induced a greater magnitude of change than isolates from CL. We conclude that L. braziliensis enhances its intracellular survival by inhibiting macrophage pathways leading to microbicidal activity. Parasite strains destined for dissemination may exert a more profound suppression than less invasive L. braziliensis strains that remain near the cutaneous site of inoculation.
Article
Full-text available
Leishmaniasis is a vector-borne, neglected tropical disease with a worldwide distribution that can present in a variety of clinical forms, depending on the parasite species and host genetic background. The pathogenesis of this disease remains far from being elucidated because the involvement of a complex immune response orchestrated by host cells significantly affects the clinical outcome. Among these cells, macrophages are the main host cells, produce cytokines and chemokines, thereby triggering events that contribute to the mediation of the host immune response and, subsequently, to the establishment of infection or, alternatively, disease control. There has been relatively limited commercial interest in developing new pharmaceutical compounds to treat leishmaniasis. Moreover, advances in the understanding of the underlying biology of Leishmania spp. have not translated into the development of effective new chemotherapeutic compounds. As a result, biomarkers as surrogate disease endpoints present several potential advantages to be used in the identification of targets capable of facilitating therapeutic interventions considered to ameliorate disease outcome. More recently, large-scale genomic and proteomic analyses have allowed the identification and characterization of the pathways involved in the infection process in both parasites and the host, and these analyses have been shown to be more effective than studying individual molecules to elucidate disease pathogenesis. RNA-seq and proteomics are large-scale approaches that characterize genes or proteins in a given cell line, tissue, or organism to provide a global and more integrated view of the myriad biological processes that occur within a cell than focusing on an individual gene or protein. Bioinformatics provides us with the means to computationally analyze and integrate the large volumes of data generated by high-throughput sequencing approaches. The integration of genomic expression and proteomic data offers a rich multi-dimensional analysis, despite the inherent technical and statistical challenges. We propose that these types of global analyses facilitate the identification, among a large number of genes and proteins, those that hold potential as biomarkers. The present review focuses on large-scale studies that have identified and evaluated relevant biomarkers in macrophages in response to Leishmania infection.
Article
Full-text available
This study was conducted to investigate body mass index (BMI), levels of cholesterol and triglycerides in prison inmates at the Institution for Reform and Rehabilitation in Southern Libya to be considered as an indication about their health and the provided foods. The results of this study showed that 26.5% of BMI of the prison inmates were found to be higher than the normal levels. Generally, the average level of cholesterol and triglycerides concentrations were found to be within normal range 142.6 mg/dl and 135.4 mg/dl, respectively. The findings also established that there were a significant relationship and direct correlation between BMI levels and age and concentration of cholesterol and triglycerides levels. The results of this showed that the served foods for these prison inmates are well balanced as indicated by their cholesterol and triglycerides levels.
Article
Full-text available
Cutaneous leishmaniasis (CL) is an immune-mediated skin pathology caused mainly by Leishmania (L.) major, Leishmania tropica, Leishmania braziliensis, L. mexicana, and L. amazonensis. The burden of CL in terms of morbidity and social stigmas are concentrated on certain developing countries in Asia, Africa, and South America. People with asymptomatic CL represent a large proportion of the infected individuals in the endemic areas who exhibit no lesion and can control the infection by as yet not fully understood mechanisms. Currently, there is no approved prophylactic control measure for CL. Discovery of biomarkers of CL infection and immunity can inform the development of more precise diagnostics tools as well as curative or preventive strategies to control CL. Herein, we provide a brief overview of the state-of-the-art for the biomarkers of CL with a special emphasis on the asymptomatic CL biomarkers. Among the identified CL biomarkers so far, direct biomarkers which indicate the actual presence of the infection as well as indirect biomarkers which reflect the host's reaction to the infection, such as alterations in delayed type hypersensitivity, T-cell subpopulations and cytokines, adenosine deaminase, and antibodies against the sand fly saliva proteins are discussed in detail. The future avenues such as the use of systems analysis to identify and characterize novel CL biomarkers are also discussed.
Article
Full-text available
Leishmania spp. is a protozoan parasite that affects millions of people around the world. At present, there is no effective vaccine to prevent leishmaniases in humans. A major limitation in vaccine development is the lack of precise understanding of the particular immunological mechanisms that allow parasite survival in the host. The parasite-host cell interaction induces dramatic changes in transcriptome patterns in both organisms, therefore, a detailed analysis of gene expression in infected tissues will contribute to the evaluation of drug and vaccine candidates, the identification of potential biomarkers, and the understanding of the immunological pathways that lead to protection or progression of disease. In this large-scale analysis, differential expression of 112 immune-related genes has been analyzed using high-throughput qPCR in spleens of infected and naïve Balb/c mice at four different time points. This analysis revealed that early response against Leishmania infection is characterized by the upregulation of Th1 markers and M1-macrophage activation molecules such as Ifng, Stat1, Cxcl9, Cxcl10, Ccr5, Cxcr3, Xcl1, and Ccl3. This activation doesn't protect spleen from infection, since parasitic burden rises along time. This marked difference in gene expression between infected and control mice disappears during intermediate stages of infection, probably related to the strong anti-inflammatory and immunosuppresory signals that are activated early upon infection (Ctla4) or remain activated throughout the experiment (Il18bp). The overexpression of these Th1/M1 markers is restored later in the chronic phase (8 wpi), suggesting the generation of a classical “protective response” against leishmaniasis. Nonetheless, the parasitic burden rockets at this timepoint. This apparent contradiction can be explained by the generation of a regulatory immune response characterized by overexpression of Ifng, Tnfa, Il10, and downregulation Il4 that counteracts the Th1/M1 response. This large pool of data was also used to identify potential biomarkers of infection and parasitic burden in spleen, on the bases of two different regression models. Given the results, gene expression signature analysis appears as a useful tool to identify mechanisms involved in disease outcome and to establish a rational approach for the identification of potential biomarkers useful for monitoring disease progression, new therapies or vaccine development.
Article
Full-text available
Leishmania is a protozoan parasite that alternates its life cycle between the sand fly and the mammalian host macrophages, involving several environmental changes. The parasite responds to these changes by promoting a rapid metabolic adaptation through cellular signaling modifications that lead to transcriptional and post-transcriptional gene expression regulation and morphological modifications. Molecular approaches such as gene expression regulation, next-generation sequencing (NGS), microRNA (miRNA) expression profiling, in cell Western blot analyses and enzymatic activity profiling, have been used to characterize the infection of murine BALB/c and C57BL/6 macrophages, as well as the human monocytic cell-lineage THP-1, with Leishmania amazonensis wild type (La-WT) or arginase knockout (La-arg - ). These models are being used to elucidate physiological roles of arginine and polyamines pathways and the importance of arginase for the establishment of the infection. In this review, we will describe the main aspects of Leishmania-host interaction, focusing on the arginine and polyamines pathways and pointing to possible targets to be used for prognosis and/or in the control of the infection. The parasite enzymes, arginase and nitric oxide synthase-like, have essential roles in the parasite survival and in the maintenance of infection. On the other hand, in mammalian macrophages, defense mechanisms are activated inducing alterations in the mRNA, miRNA and enzymatic profiles that lead to the control of infection. Furthermore, the genetic background of both parasite and host are also important to define the fate of infection.
Article
Full-text available
One aim of the Sustainable Development Goals is to end the epidemics of neglected tropical diseases by 2030. A multifaceted approach is needed to tackle leishmaniasis, keeping in mind the parasite, as well as its vector, host and reservoir. Attention should be focused on sustainable and achievable solutions rather than fairytale goals, and biology should play a more prominent role.
Article
Full-text available
Infection by protozoan parasites of the genus Leishmania results in the development of leishmaniases, an increasingly prevalent group of diseases affecting over 12 million people worldwide. Leishmaniasis can have very different outcomes ranging from cutaneous lesions, mucosal lesions to visceralization depending on the species of the infecting parasite and on the immune response developed by the host. As an obligate intracellular parasite, residing within macrophages, Leishmania evolved in strict contact with the host immune system, developing different mechanisms to evade or modulate the immune response. Various types of immune responses are observed during different Leishmania spp. infections, resulting in parasite clearance but also contributing to the pathogenesis, thus increasing the complexity of the course of the disease. Interestingly, depending on the type of leishmaniasis developed, opposite treatment strategies, which either boost or inhibit the inflammatory response, have shown efficacy. In this review, we summarize the contribution of different immune cell types to the development of the anti-leishmanial immune response and the parasite strategies to evade and modulate host immunity. Further, we discuss the involvement of co-infecting pathogens in the determination of the outcome of leishmaniasis and on the effectiveness of treatment and the implication of the immune response for treatment and vaccine development.
Article
Leishmaniasis is a poverty-related disease with two main clinical forms: visceral leishmaniasis and cutaneous leishmaniasis. An estimated 0·7-1 million new cases of leishmaniasis per year are reported from nearly 100 endemic countries. The number of reported visceral leishmaniasis cases has decreased substantially in the past decade as a result of better access to diagnosis and treatment and more intense vector control within an elimination initiative in Asia, although natural cycles in transmission intensity might play a role. In east Africa however, the case numbers of this fatal disease continue to be sustained. Increased conflict in endemic areas of cutaneous leishmaniasis and forced displacement has resulted in a surge in these endemic areas as well as clinics across the world. WHO lists leishmaniasis as one of the neglected tropical diseases for which the development of new treatments is a priority. Major evidence gaps remain, and new tools are needed before leishmaniasis can be definitively controlled.