Since the publication of the first chicken genome sequence, we have encountered genes playing key roles in mammalian immunology, but being seemingly absent in birds. One of those was, until recently, Foxp3, the master transcription factor of regulatory T cells in mammals. Therefore, avian regulatory T cell research is still poorly standardized. In this study we identify a chicken ortholog of Foxp3. We prove sequence homology with known mammalian and sauropsid sequences, but also reveal differences in major domains. Expression profiling shows an association of Foxp3 and CD25 expression levels in CD4+CD25+ peripheral T cells and identifies a CD4CD25+Foxp3high subset of thymic lymphocytes that likely represents yet undescribed avian regulatory T precursor cells. We conclude that Foxp3 is existent in chickens and that it shares certain functional characteristics with its mammalian ortholog. Nevertheless, pathways for regulatory T cell development and Foxp3 function are likely to differ between mammals and birds. The identification and characterization of chicken Foxp3 will help to define avian regulatory T cells and to analyze their functional properties and thereby advance the field of avian immunology.

Ever since the publication of the first chicken genome in 2004 (1), the avian research community has been struggling with the apparent absence of avian orthologs of genes that are considered vitally relevant to mammalian physiology (24). As evolutionary success requires adaptation to environmental changes, speciation events go hand in hand with gain or loss of genes. Therefore, it is likely that some, or even most, of these genes are indeed not present in birds (2, 4). However, some presumably missing genes have recently been described in chickens (2, 5, 6). As the original publication on the chicken genome had indicated—and others emphasized later—avian genomes feature a high overall GC content, which impairs sequencing (1, 2). This characteristic has been reported to be displayed by certain mammalian genes as well: due to GC-biased gene conversion, genes that have been the subject of vast evolutionary changes are especially GC-rich and usually underrepresented in sequencing datasets (7).

One of the genes that was assumed to be absent in birds until recently is Foxp3. Foxp3 plays a central role in the maintenance of immune homeostasis, representing the master transcription factor of regulatory T cells (Tregs) (8). All attempts to find traces of avian Foxp3 yielded no results (9, 10). In 2013, however, the highest quality avian genome so far, namely that of the Tibetan ground tit (Pseudopodoces humilis), was published (11) and, surprisingly, P. humilis Foxp3 was automatically annotated. Subsequently, Denyer et al. (12) used the ground tit sequence as a template for further database searches and identified single Foxp3 exons in two falcon species (Falco peregrinus and Falco cherrug). However, the function of avian Foxp3 and whether it possesses any relevance for Tregs have not been evaluated in any bird species yet.

Tregs were first described as suppressors of T cell help in mice by Gershon et al. (13) in 1972. Even though a lot of knowledge remains to be gathered, they represent a well-established cell population today. Tregs in mice are defined as a CD4+CD25+Foxp3+ T lymphocyte subset (14), but it has been shown that their phenotype is highly variable depending on species and condition (15). Generation of Tregs can either occur by thymic selection and priming (naive Tregs [nTregs]) or by peripheral induction of Foxp3 in CD4+CD25+ naive T cells (iTregs) (16). Both pathways are promoted by IL-2/IL-2R ligation in the presence of TGF-β in a non-inflammatory environment. As TCR stimulation together with coreceptor ligation is reported to be crucial for effector T cell induction, Treg development can be triggered earliest during the double-positive (DP) stage (i.e., CD4+CD8+) of T cell development in the thymus or in mature single-positive (SP) peripheral T cells (i.e., CD4+ or CD8+). The expression of Foxp3 then indicates lineage determination toward the regulatory branch (8). Tregs express a distinct set of effector cytokines, among which are IL-10 and TGF-β. Along with other suppressor functions, these cytokines diminish proinflammatory immune response mechanisms (17, 18). This is achieved mainly by the IL-10–mediated downregulation of IL-2, which is not critical for the maintenance of Tregs or their induction outside the thymus, but for the activation of naive Th cells (19).

Due to the presumed absence of avian Foxp3, the body of literature on Tregs in birds is comparably small and there is low consensus on the cells’ phenotype. Nevertheless, only a few years after the initial description of mouse suppressor T cells (13), suppression of Ab production in response to infection by a subset of T cells was described in the chicken (20). Various groups have since worked on avian suppressor T cells: the group who discovered the suppressive T cell–mediated effect on Ab production later characterized a subset of these cells as CD8+ γδ T cells (21). A number of studies of another group concluded that most CD4+CD25+ avian T cells display regulatory properties (9, 2224), and another study recently suggested TGF-β as a marker for Tregs during pathology (25). This heterogeneity of data calls for a universal definition of avian Tregs.

Therefore, the discovery of avian Foxp3 finally paves the way for standardized Treg research in birds. This study, aiming to characterize chicken Foxp3, is hence the crucially required next step toward a new and fundamentally more precise description of avian Tregs.

To assemble the chicken Foxp3 and chCcdc22 sequences, dataset PRJEB4677, obtained from the National Center for Biotechnology Information (NCBI) (https://www.ncbi.nlm.nih.gov) Sequence Read Archive (SRA) database, was used. Sequences of non-avian vertebrate Foxp3 genes and P. humilis Foxp3 (Table I) were used as probes in basic local alignment search tool (BLAST) searches of the SRA dataset. The sequences obtained were downloaded and assembled manually either with CLC genomics workbench 8.0.1 (www.clcbio.com) or with Lasergene 10.0.0 (DNASTAR, Madison, WI). The resulting short contigs were used as probes in subsequent rounds of BLAST searches against SRA datasets, until the full open reading frame was completed. The assemblies of chicken Foxp3 and chicken Ccdc22 have been published in the GenBank database (https://www.ncbi.nlm.nih.gov/genbank/) with the respective accession numbers MT133687 and MT133688.

Table I.

NCBI transcript and protein identifiers of Foxp family member sequences used throughout this study

Foxp1Foxp2Foxp3Foxp4
Human (Homo sapiensNM_001349337.1:259–2295
NP_001336266.1 
XM_017012801.2:981–3203
XP_016868290.1 
NM_014009.3:189–1484
NP_054728.2 
XM_017010233.1:1468–3531
XP_016865722.1 
Mouse (Mus musculusNM_001347345.1:705–2822
NP_001334274.1 
NM_212435.1:323–2467
NP_997600.1 
NM_001199348.1:201–1490
NP_001186277.1 
NM_001110824.1:445–2502
NP_001104294.1 
Lizard (Anolis carolinensisXM_016991767.1:1228–3339
XP_016847256.1 
XM_008111156.1:345–2666
XP_008109363.1 
XM_016991205.1:150–1256
XP_016846694.1 
XM_008109692.2:335–2425
XP_008107899.1 
Alligator (Alligator sinensisXM_025197991.1:1425–3470
XP_025053776.1 
XM_025208808.1:330–2411
XP_025064593.1 
XM_025213651.1:197–1267
XP_025069436.1 
XM_025205820.1:423–2447
XP_025061605.1 
Ground tit (Pseudopodoces humilisXM_014251617.1:322–2409
XP_014107092 
XM_005519253.2:318–2504
XP_005519310 
NM_001323961.1
NP_001310890.1 
XM_014258084.1:12–1721
XP_014113559 FOXP4 
Chicken (Gallus gallusXM_025154490.1:152–2212
XP_025010258.1 
XM_025151552.1:11–2137
XP_025007320.1 
MT133687
Translation of MT133687 
XM_004948413.3:287–2287
XP_004948470.1 
Foxp1Foxp2Foxp3Foxp4
Human (Homo sapiensNM_001349337.1:259–2295
NP_001336266.1 
XM_017012801.2:981–3203
XP_016868290.1 
NM_014009.3:189–1484
NP_054728.2 
XM_017010233.1:1468–3531
XP_016865722.1 
Mouse (Mus musculusNM_001347345.1:705–2822
NP_001334274.1 
NM_212435.1:323–2467
NP_997600.1 
NM_001199348.1:201–1490
NP_001186277.1 
NM_001110824.1:445–2502
NP_001104294.1 
Lizard (Anolis carolinensisXM_016991767.1:1228–3339
XP_016847256.1 
XM_008111156.1:345–2666
XP_008109363.1 
XM_016991205.1:150–1256
XP_016846694.1 
XM_008109692.2:335–2425
XP_008107899.1 
Alligator (Alligator sinensisXM_025197991.1:1425–3470
XP_025053776.1 
XM_025208808.1:330–2411
XP_025064593.1 
XM_025213651.1:197–1267
XP_025069436.1 
XM_025205820.1:423–2447
XP_025061605.1 
Ground tit (Pseudopodoces humilisXM_014251617.1:322–2409
XP_014107092 
XM_005519253.2:318–2504
XP_005519310 
NM_001323961.1
NP_001310890.1 
XM_014258084.1:12–1721
XP_014113559 FOXP4 
Chicken (Gallus gallusXM_025154490.1:152–2212
XP_025010258.1 
XM_025151552.1:11–2137
XP_025007320.1 
MT133687
Translation of MT133687 
XM_004948413.3:287–2287
XP_004948470.1 

The transcript ID is shown on the top line for entry; the protein ID is shown on the bottom line.

The protein sequences listed in Table I were aligned using the ClustalX 2.1 algorithm (http://www.clustal.org). The similarity between sequence stretches of chicken and human Foxp3 was calculated using the EMBOSS needle pairwise alignment tool (https://www.ebi.ac.uk/Tools/psa/emboss_needle/).

The maximum likelihood method based on the JTT matrix–based model (26) was used by Mega6 (27) to create 500 individual trees using all sites of a previously made MUSCLE alignment of the protein sequences listed in Table I. The bootstrap method was used to assess the likelihood of the displayed taxa association.

To obtain long-read sequences from the chicken genome, we used genomic DNA samples from a female inbred line C White Leghorn chicken (28). Genomic DNA was quantified on the Qubit fluorometer (Invitrogen, Carlsbad, CA) and used for sequencing library preparation (SQK-LSK109; Oxford Nanopore Technologies, Oxford, U.K.). The sequencing was performed on the GridION X5 platform using R9.4.1 chemistry, and base calling was performed in real time using the GridION sequencing software version 19.12.2 (all from Oxford Nanopore Technologies). To select reads from the Foxp3 locus, a BLAST library was generated using CLC Genomics Workbench, and searched with Foxp3 coding sequences. Two reads were selected that covered the entire Foxp3 coding exons and neighboring genomic sequence, including the Ccdc22 gene. Individual exons were mapped to these Nanopore reads by pairwise alignment. The Nanopore reads used to assess Foxp3 synteny have been assigned the NCBI SRA accession number PRJNA610550 (https://www.ncbi.nlm.nih.gov/sra).

White Leghorn line M11 chickens (Federal Research Institute for Animal Health, Neustadt, Germany) were hatched, vaccinated only against Marek’s disease, and conventionally housed. Water and a commercial diet were provided ad libitum. At the age of 3–6 mo, the chickens were euthanized for tissue collection.

Organs were collected in sterile PBS for immediate lymphocyte preparation. Organs were mashed to create a cell suspension as described elsewhere (29). Cell suspensions were subsequently density centrifuged over 1.077 g/ml Biocoll separating solution (Merck, Burlington, MA) and the lymphocyte layer was extracted and washed.

Lymphocytes from respective organs were incubated for 20 min with a mix of anti-CD25, anti-CD4, and anti-Bu1 mAbs, followed by 20-min staining with the respective isotype-matched and fluorochrome-labeled secondary Abs (Table II). Cell sorting for viable single cells was performed with a FACSAria IIIu (Becton Dickinson, Franklin Lakes, NJ) to a purity of >91%. Purified cells were pelleted and stored at −80°C until RNA extraction.

Table II.

Abs and staining conditions used throughout this study

ApplicationAntigenCloneConcentrationSecondary AbConcentration
Cell sort chBu-1 AV20 (SouthernBiotech) 2 µg/ml Anti-mouse IgG1-PE (Jackson ImmunoResearch) 1.25 µg/ml 
Cell sort chCD4 CT4 (SouthernBiotech) 0.5 µg/ml Anti-mouse IgG2b-FITC (Invitrogen) 10 µg/ml 
Cell sort chCD25 28-4 (supernatant) (611:100 Anti-mouse IgG3-AF647 (SouthernBiotech) 2.5 µg/ml 
Flow cytometry, intracellular Myc-Tag 9E10 (MBL International) 0.25 µg/ml Anti-mouse IgG1-allophycocyanin (Jackson ImmunoResearch) 1.67 µg/ml 
Flow cytometry, intracellular muFoxp3 FJK-16s-PE (eBioscience) 1.25 µg/ml — — 
Flow cytometry, intracellular muFoxp3 MF-14-PE (BioLegend) 4 µg/ml — — 
Flow cytometry, surface chCD25 28-4 (purified) (611:200 Anti-mouse IgG3-AF647 (SouthernBiotech) 2.5 µg/ml 
Flow cytometry, surface chCD8 3-298-allophycocyanin-Cy7 (621:100 — — 
Flow cytometry, surface chCD4 CT4-FITC (SouthernBiotech) 0.5 µg/ml — — 
Flow cytometry, surface chCD3 CT3-PE (SouthernBiotech) 2 µg/ml — — 
Flow cytometry, surface Isotype control B10 (eBioscience) 5 µg/ml Anti-mouse IgG3-AF647 (SouthernBiotech) 2.5 µg/ml 
ApplicationAntigenCloneConcentrationSecondary AbConcentration
Cell sort chBu-1 AV20 (SouthernBiotech) 2 µg/ml Anti-mouse IgG1-PE (Jackson ImmunoResearch) 1.25 µg/ml 
Cell sort chCD4 CT4 (SouthernBiotech) 0.5 µg/ml Anti-mouse IgG2b-FITC (Invitrogen) 10 µg/ml 
Cell sort chCD25 28-4 (supernatant) (611:100 Anti-mouse IgG3-AF647 (SouthernBiotech) 2.5 µg/ml 
Flow cytometry, intracellular Myc-Tag 9E10 (MBL International) 0.25 µg/ml Anti-mouse IgG1-allophycocyanin (Jackson ImmunoResearch) 1.67 µg/ml 
Flow cytometry, intracellular muFoxp3 FJK-16s-PE (eBioscience) 1.25 µg/ml — — 
Flow cytometry, intracellular muFoxp3 MF-14-PE (BioLegend) 4 µg/ml — — 
Flow cytometry, surface chCD25 28-4 (purified) (611:200 Anti-mouse IgG3-AF647 (SouthernBiotech) 2.5 µg/ml 
Flow cytometry, surface chCD8 3-298-allophycocyanin-Cy7 (621:100 — — 
Flow cytometry, surface chCD4 CT4-FITC (SouthernBiotech) 0.5 µg/ml — — 
Flow cytometry, surface chCD3 CT3-PE (SouthernBiotech) 2 µg/ml — — 
Flow cytometry, surface Isotype control B10 (eBioscience) 5 µg/ml Anti-mouse IgG3-AF647 (SouthernBiotech) 2.5 µg/ml 

–, not applicable (probe not used in qPCR, or PCR efficiency ND for regular PCR); ch, chicken; mu, murine.

Unless otherwise indicated, total RNA was isolated from sort-purified lymphocyte preparations using the ReliaPrep cell RNA isolation system (Promega, Madison, WI). RNA quality was determined with a Bioanalyzer (Agilent Technologies, Santa Clara, CA); only RNA with a minimum RNA integrity number of 7.5 was used for further analyses. Reverse transcription of 400 ng of RNA was carried out using GoScript reverse transcription master mix (Promega).

Total RNA for RT-PCR was isolated from chicken spleen tissue using TRI reagent (Sigma-Aldrich, St. Louis, MO). Reverse transcription was performed using the SMART RACE (Takara Bio, Kusatsu, Japan) protocol with 1 µg of RNA as a template. For the amplification of the highly GC-rich Foxp3 gene, previously described conditions (5, 30) were used: a 1:200 mixture of Deep Vent and Taq polymerases (both New England Biolabs, Ipswich, MA), and PCR primers with high melting temperature and long amplification times (Table III). The PCR program was set to 30 cycles of 95°C for 30 s and 65°C for 10 min. The PCR product was extracted from a 0.6% agarose gel using a Qiaex II gel extraction kit (Qiagen, Hilden, Germany) and directly sequenced.

Table III.

Primers and PCR conditions used throughout this study

GeneAppl.NCBI accession no.Forward Primer SequenceReverse Primer SequenceProbe SequenceTA, Elong.qPCR Eff. (%)
chFoxp3 Full-length PCR MT133687 5′-CACCCGGTGTCACCCAAAG-3′ 5′-CCTGCTTTGTGGTTTTGGGGT CCTC-3′ — 65°C, 10 min — 
chFoxp3 qPCR MT133687 5′-AGTACGCCACAACCTGAGCCT-3′ 5′-TTGGGGTCCTCTCAGCTCCGT-3′ 5′-TGCGGGTGGAGAA CGTACGTGGG-3′ 60°C, 30 s 80 
ch18S rRNA qPCR AF173612.1 5′-CATGTCTAAGTACACACGGGCGGTA-3′ 5′-GGCGCTCGTCGGCATGTATTA-3′ — 59°C, 30 s 102 
chIL-2 qPCR AJ224516.1 5′-GTCTTACGGGTCTAAATCACACC-3′ 5′-AAAGTTGGTCAGTTCATGGAGAA-3′ — 59°C, 30 s 103 
chIL-10 qPCR EF554720.1 5′-CGGGAGCTGAGGGTGAAGT-3′ 5′-CAGCCAAAGGTCCCCTTAAAC-3′ — 59°C, 30 s 113 
chCD8 qPCR NM_205235.1 QuantiTect primer assay (Qiagen)  — 60°C, 30 s 80 
chTGF-β qPCR NM_001318456.1 5′-CCACAGCATCTTCTTCGTGT-3′ 5′-GCATTGCCGTAACCCTGG-′3 — 59°C, 30 s 85 
chFoxp3 Cloning MT133687 5′-CCCCAAGCTGGCCTCTGACCATGGAACAA AAACTCATCTCAGAAGAGGATCTAGCAGG CGCCAGAGATC-3′ 5′-CCCAAGCTTGGCCTCAGCTTCTTGG AGGCAG-3′ — 59°C, 30 s — 
muFoxp3 Cloning DQ387959.1 5′-CCCCAAGCTGGCCTCTGACCATGGAACAAA AACTCATCTCAGAAGAGGATCTACCCAACCC TAGGCCAG-3′ 5′-CCCAAGCTTGGCCTCAAGGGCAGGG ATTGGA-3′ — 59°C, 30 s — 
GeneAppl.NCBI accession no.Forward Primer SequenceReverse Primer SequenceProbe SequenceTA, Elong.qPCR Eff. (%)
chFoxp3 Full-length PCR MT133687 5′-CACCCGGTGTCACCCAAAG-3′ 5′-CCTGCTTTGTGGTTTTGGGGT CCTC-3′ — 65°C, 10 min — 
chFoxp3 qPCR MT133687 5′-AGTACGCCACAACCTGAGCCT-3′ 5′-TTGGGGTCCTCTCAGCTCCGT-3′ 5′-TGCGGGTGGAGAA CGTACGTGGG-3′ 60°C, 30 s 80 
ch18S rRNA qPCR AF173612.1 5′-CATGTCTAAGTACACACGGGCGGTA-3′ 5′-GGCGCTCGTCGGCATGTATTA-3′ — 59°C, 30 s 102 
chIL-2 qPCR AJ224516.1 5′-GTCTTACGGGTCTAAATCACACC-3′ 5′-AAAGTTGGTCAGTTCATGGAGAA-3′ — 59°C, 30 s 103 
chIL-10 qPCR EF554720.1 5′-CGGGAGCTGAGGGTGAAGT-3′ 5′-CAGCCAAAGGTCCCCTTAAAC-3′ — 59°C, 30 s 113 
chCD8 qPCR NM_205235.1 QuantiTect primer assay (Qiagen)  — 60°C, 30 s 80 
chTGF-β qPCR NM_001318456.1 5′-CCACAGCATCTTCTTCGTGT-3′ 5′-GCATTGCCGTAACCCTGG-′3 — 59°C, 30 s 85 
chFoxp3 Cloning MT133687 5′-CCCCAAGCTGGCCTCTGACCATGGAACAA AAACTCATCTCAGAAGAGGATCTAGCAGG CGCCAGAGATC-3′ 5′-CCCAAGCTTGGCCTCAGCTTCTTGG AGGCAG-3′ — 59°C, 30 s — 
muFoxp3 Cloning DQ387959.1 5′-CCCCAAGCTGGCCTCTGACCATGGAACAAA AACTCATCTCAGAAGAGGATCTACCCAACCC TAGGCCAG-3′ 5′-CCCAAGCTTGGCCTCAAGGGCAGGG ATTGGA-3′ — 59°C, 30 s — 

—, not applicable (probe not used in qPCR, or PCR efficiency ND for regular PCR); Appl., application; ch, chicken; Eff., efficiency; Elong., elongation time; mu, murine; Ta, annealing temperature.

SYBR Green–based real-time RT-PCR (subsequently called quantitative PCR [qPCR]) was performed with cDNA in a dilution ensuring linear amplification of each target gene, using GoTaq probe qPCR master mix or GoTaq qPCR master mix (Promega) to amplify Foxp3 or 18S rRNA, IL-10, IL-2, and CD8 (Table III), respectively. For the amplification of CD8, the QuantiTect primer assay for chicken CD8α (Qiagen) was used according to the manufacturer’s instructions. All other primers and the Foxp3 probe were synthesized by Eurofins Genomics (Ebersberg, Germany) and used at a final concentration of 300 nM (primers) or 150 nM (probe) each. Primer efficiency and linear amplification range was determined using serial dilutions of a cDNA mix. In addition to the primers, each 25-µl reaction contained 12.5 µl of the GoTaq qPCR master mix, 0.25 µl of 100× CXR reference dye, 4.25 µl of nuclease-free water, and 5 µl of diluted cDNA. The cycling conditions were set to 95°C for 2 min, followed by 40 cycles of 95°C for 15 s, 59 (Eurofins primers without probe) or 60°C (Foxp3 and CD8) for 30 s, and 72°C for 30 s. Each assay included relevant non-template controls. A melting curve performed following each run with the qPCR master mix confirmed a single, specific PCR product. Raw data created by a 7300 real-time PCR system was analyzed using the SDS 7300 software (both from Applied Biosystems, Foster City, CA). Relative expression fold changes of Foxp3, IL-10, IL-2, and CD8 were assessed using normalized cycle threshold values (ΔCt). Normalization was carried out against 18S rRNA expression for each sample. ΔCt values were subtracted from the total number of cycles (40 − ΔCt) and used as an exponent of 2 to calculate fold changes (31). The mean of all ΔCt values measured for each gene and organ was set to 1.

Chicken Foxp3 was codon optimized for expression in human cells using the GeneArt Thermo Fisher Scientific web tool. The resulting sequence was synthesized as gBlocks Gene Fragment including a 3xFLAG tag on the 5′ end. The gBlocks Gene Fragment was cloned into pcDNA3.1 using the NEBuilder kit (New England Biolabs) according to the manufacturer’s instructions. Obtained clones were checked by Sanger sequencing. One correct clone was expressed in HEK293 cells, and cDNA of these cells was subsequently used as a template. The cloning template for full-length murine Foxp3 was synthesized without any modifications (GeneArt, Regensburg, Germany). Assembly primers were designed to overlap the junction of the pSBbi(GP)rev vector backbone and the respective Foxp3 sequence; the forward primers contain the Myc-tag sequence (see Table III). The NEBuilder HiFi DNA assembly cloning kit (New England Biolabs) was used to assemble the constructs according to the manufacturer’s instructions. The resulting plasmids were expressed in Escherichia coli and purified with the PureYield plasmid miniprep system (Promega).

293T cells (105) were cultured overnight and transfected with 250 ng of the mouse or chicken Foxp3 construct, respectively, and 12.5 ng of a transposase-containing plasmid using X-tremeGENE 9 DNA transfection reagent (Sigma-Aldrich) according to the manufacturer’s instructions. GFP expression was confirmed microscopically and the cells were harvested for flow cytometry with a 0.02% EDTA solution in PBS 48 h after transfection.

For intracellular staining, 293T cells transfected with either chicken or mouse Myc-Foxp3 plasmid were incubated with intracellular fixation buffer for 30 min in the dark at room temperature and permeabilized using 1× permeabilization buffer in every subsequent washing step. Viability staining was performed with fixable viability dye (all from eBioscience, San Diego, CA). Abs against the Myc epitope, mouse Foxp3, respective isotype-matched negative controls, and, where necessary, fluorochrome-labeled isotype-specific secondary Abs (see Table II) were diluted in 1× permeabilization buffer. The staining protocol used is described elsewhere (32). For cell surface staining, splenic lymphocytes were stained with anti-CD25, anti-CD8, anti-CD4, and anti-CD3 mAbs or respective control Abs (see Table II) according to standard procedures. Viability was assessed using 7-aminoactinomycin D (Invitrogen). Analyses were performed with a FACSCanto II, with sort purification of cell populations performed with a FACSAria IIIu using FACSDiva and FlowJo software (all from Becton Dickinson).

Using previously published avian Foxp3 sequences (12) for homology searches in the NCBI SRA, we assembled the entire coding sequence of putative chicken Foxp3 from the database short reads. Subsequently, we confirmed the correctness of the assembly by RT-PCR amplification and sequencing of full-length Foxp3 from chicken spleen RNA. Previously missing avian genes that have recently been assembled manually are similar in respect to their comparably high overall GC content and long GC stretches (2). To assess whether such features might have impeded the annotation of the newly identified chicken sequence as Foxp3 in the past, we plotted them for each Foxp coding sequence (Foxp1–Foxp4) of six vertebrate species (Supplemental Fig. 1). We found that avian Foxp sequences do not generally differ from mammalian and reptilian sequences concerning GC characteristics. Foxp1 and Foxp2 sequences display no significant species differences (Supplemental Fig. 1A), whereas Foxp4 is slightly more variable and displays a higher GC content (Supplemental Fig. 1C). Foxp3, however, shows a clear increase in both, that is, length of GC stretches and overall GC content, that peaks in the two avian species (Supplemental Fig. 1B). The hypothesis that chicken Foxp3 could not be annotated before the present due to an extraordinary GC content is thereby further supported.

To verify the identity of the newly assembled Foxp3 sequence, we aligned it with all four previously described members of the Foxp family from a total of six mammalian and sauropsid species. Phylogenetic analysis (Fig. 1) generated from this alignment reveals that the protein sequences of Foxp1 and Foxp2 are highly conserved among species. A slightly elevated substitution rate separates mammalian and sauropsid Foxp4 sequences. Foxp3, however, displays considerable variation between species, whereas mammalian sequences, on the one hand, and sauropsid sequences, on the other hand, cluster together. The non-archosaur sauropsid lizard Anolis carolinensis represents a separate branch within the latter clade. Consistent with previous analyses (33), it is also shown that Foxp3 sequences are generally less conserved than those of the other Foxp family members.

FIGURE 1.

Phylogenetic relationship of Foxp1, Foxp2, Foxp3, and Foxp4 proteins of mammals, reptiles, and birds. The tree represents the most likely clade organization (likelihood −12,305.6245). The percentage of trees in which the associated taxa clustered together is shown next to the branches. The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. Mouse FoxN1 was used as an outlier to root the tree.

FIGURE 1.

Phylogenetic relationship of Foxp1, Foxp2, Foxp3, and Foxp4 proteins of mammals, reptiles, and birds. The tree represents the most likely clade organization (likelihood −12,305.6245). The percentage of trees in which the associated taxa clustered together is shown next to the branches. The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. Mouse FoxN1 was used as an outlier to root the tree.

Close modal

We also aimed to assess whether the chicken Foxp3 locus shows synteny. To this end, we assembled the sequence of the closest neighboring gene, chicken Ccdc22, using a similar approach as for chicken Foxp3. Next, we mapped the position of both genes in long-read Nanopore sequences from the chicken genome. Indeed, both chicken Foxp3 and Ccdc22 lie on a single contiguous Nanopore read (Supplemental Fig. 2), suggesting strongly that the two-gene synteny is conserved in the chicken genome, as it is in the Tibetan ground tit genome (12). Additionally, this analysis indicates that chicken Foxp3 is divided into 11 coding exons, similar to the human ortholog. Thus, both, phylogenetic analysis and synteny, indicate that the newly assembled sequence is a true chicken Foxp3 ortholog.

As a next step, we analyzed the Foxp3 protein sequences of the aforementioned six species for conservation of functionally relevant domains (Fig. 2). The N terminus contains the repressor domain, which is represented by a proline-rich region and conveys the repressive function of Foxp3 (34). This domain is poorly conserved in chickens. It is only ∼100 aa long, instead of 170 aa in human Foxp3, and displays low similarity values. Nevertheless, with roughly 30% proline residues compared with 17% in the human sequence, it is highly proline rich. The zinc finger domain, located downstream of the proline-rich region at aa 198–221 of human Foxp3, is absent in non-mammalian species. However, the leucine zipper domain (aa 240–261) (35), which in non-mammals connects the N- and C-terminal domains, is highly conserved, displaying 67% similarity. It contains a conserved glutamic acid residue essential for dimerization of the protein (E251) (34, 36). The C-terminal forkhead (FKH) domain is 91% similar to the mammalian domain. It includes the sites of the nuclear localization signals essential for the translocation of Foxp3 to the nucleus (R347 and aa 414–417) (34), which are 100% similar to the respective positions in the mammalian protein.

FIGURE 2.

ClustalX protein alignment of Foxp3 sequences of mammals, reptiles, and birds. Shading and the histogram below the amino acid sequences indicate conservation. The residues are numbered according to the human sequence. Domain annotations according to the human protein are as follows: repressor domain, aa 1–193; zinc finger domain, aa 200–223; leucine zipper, aa 240–261, containing glutamic acid residue E251, which is essential for dimerization (35); forkhead domain, aa 336–421, containing nuclear localization signals (NLSs) essential for translocation of Foxp3 to nucleus (34, 35).

FIGURE 2.

ClustalX protein alignment of Foxp3 sequences of mammals, reptiles, and birds. Shading and the histogram below the amino acid sequences indicate conservation. The residues are numbered according to the human sequence. Domain annotations according to the human protein are as follows: repressor domain, aa 1–193; zinc finger domain, aa 200–223; leucine zipper, aa 240–261, containing glutamic acid residue E251, which is essential for dimerization (35); forkhead domain, aa 336–421, containing nuclear localization signals (NLSs) essential for translocation of Foxp3 to nucleus (34, 35).

Close modal

These findings prove structural homology of the newly assembled chicken Foxp3 gene and its previously described orthologs. In addition, they hint on possibly conserved functions despite considerable differences in the sequence and in domain organization.

Our next approach aimed to determine the expression pattern of chicken Foxp3 in relevant lymphocyte populations. We sort purified cells from spleen, cecal tonsil, and thymus and analyzed them using real-time RT-PCR (Fig. 3). We first distinguished chicken T cells according to the human Treg phenotype (CD4+CD25+Foxp3+) (14) by analyzing CD4+CD25+ and CD4+CD25 lymphocyte subsets. Chicken Foxp3 mRNA was highly abundant in CD25+ but not in CD25 subsets of CD4+ lymphocytes from secondary lymphatic organs (Fig. 3A, left and left-middle panels, dark gray bars).

FIGURE 3.

Relative quantification of Foxp3 and Treg-associated cytokines and CD8 expression in sort-purified lymphocyte populations. (AC) Lymphocytes were sort purified according to their CD4 and CD25 (A and B) or CD4 and AV20 expression (C). Gating strategies for spleen, caecal tonsil, and thymus are shown in the top panels. The sorted lymphocyte populations were analyzed for Foxp3 and cytokine (A and C) or CD8 (B) mRNA expression by quantitative RT-PCR. B cells (C) do not express Foxp3; CD8 is only expressed in CD25 subsets of thymic lymphocytes. Therefore, Foxp3 is expressed only in CD4+CD25+ T cells from secondary lymphatic organs and in DN CD25+ thymocytes (A–C). (D) Hypothesis on chicken Th cell and Treg lineages found in this experiment. Font size indicates high or low expression of the respective marker. Shown are representative staining patterns or mean values of three independent experiments; n = 3.

FIGURE 3.

Relative quantification of Foxp3 and Treg-associated cytokines and CD8 expression in sort-purified lymphocyte populations. (AC) Lymphocytes were sort purified according to their CD4 and CD25 (A and B) or CD4 and AV20 expression (C). Gating strategies for spleen, caecal tonsil, and thymus are shown in the top panels. The sorted lymphocyte populations were analyzed for Foxp3 and cytokine (A and C) or CD8 (B) mRNA expression by quantitative RT-PCR. B cells (C) do not express Foxp3; CD8 is only expressed in CD25 subsets of thymic lymphocytes. Therefore, Foxp3 is expressed only in CD4+CD25+ T cells from secondary lymphatic organs and in DN CD25+ thymocytes (A–C). (D) Hypothesis on chicken Th cell and Treg lineages found in this experiment. Font size indicates high or low expression of the respective marker. Shown are representative staining patterns or mean values of three independent experiments; n = 3.

Close modal

As CD25 expression in CD4+ splenocytes was quite variable, we performed a separate analysis of Foxp3 expression for CD25, CD25low, and CD25high subsets of chicken CD4+ splenocytes (Supplemental Fig. 4). We found that CD25low cells express 66-fold more Foxp3 than does the CD25 population. Furthermore, in CD25high cells, Foxp3 expression is 130-fold increased as compared with CD25 cells (Supplemental Fig. 4).

Thymic CD4+CD25+ cells expressed very low levels of Foxp3 (Fig. 3A, right-middle panel, dark gray bars). Taken together, chicken Foxp3 is expressed considerably by CD4+CD25+ cells in secondary lymphatic organs, but not in the thymus.

To investigate possibly deviating pathways of Foxp3 induction, we next investigated the CD4 thymocyte population. Much to our surprise, we found that, in association with CD25, CD4 thymocytes expressed very high levels of Foxp3 (Fig. 3A, right panel, dark gray bars). To exclude cytotoxic T cells as the source of the detected Foxp3 in this CD4 subset, analysis of splenic CD8+ lymphocytes was carried out and revealed very low Foxp3 mRNA levels (data not shown). Due to very low cell numbers, a sort of cytotoxic T cells for the activation marker, that is, of CD8+CD25+ lymphocytes, was not possible from any of the examined lymphocyte sources. However, real-time RT-PCR analysis of the CD4CD25+ thymocyte subset showed CD8 expression levels equal to those of any CD4+ peripheral lymphocyte subsets and way below those of CD25 thymocytes (Fig. 3B). Furthermore, B cells from secondary lymphatic organs were also excluded as Foxp3-expressing cells in chickens by analysis of AV20+ lymphocyte populations (Fig. 3C, dark gray bars). Thus, Foxp3 expression seems to be restricted mainly to CD25+ subsets of CD4+ peripheral and CD4CD8 thymic T cells (Fig. 3D). Further examination of the Foxp3-expressing thymocytes revealed that only 6.64% of these double-negative (DN) CD25+ cells were CD3+ (Supplemental Fig. 3).

IL-2 is expressed in effector T cells and conveys their activation via the IL-2 receptor complex, of which CD25 is a part. Besides other mechanisms, Tregs suppress IL-2 expression. IL-10 and TGF-β are the major suppressive cytokines used by mammalian Tregs to achieve regulatory function. To investigate the cytokine pattern of putative chicken Tregs, we measured IL-2, IL-10, and TGF-β mRNA levels in the sort-purified lymphocyte subsets mentioned above. IL-10 expression was mostly limited to CD4+CD25+ subsets in secondary lymphatic organs and to CD4CD25+ thymocytes (light gray bars in (Fig. 3A). As described above, this pattern was in line with the expression of Foxp3. IL-2, however, showed no signs of association with Foxp3, but was expressed only by CD4+CD25+ cells from thymus, and, to a much lesser extent, from cecal tonsil (white bars in (Fig. 3A). B cells expressed neither IL-2 nor IL-10 (Fig. 3C). Surprisingly, TGF-β was not differentially expressed between any of the analyzed subsets (data not shown). These results show that IL-10 and IL-2 expression levels in chicken T cell subsets generally resemble their mammalian counterparts, whereas TGF-β expression is not associated with Foxp3 or with CD25. These data suggest the existence of several distinct T cell phenotypes associated with regulatory pathways in chickens.

Future analyses of avian Tregs must extend to the protein level. Therefore, specific mAbs are indispensable. To assess cross-reactivity with chicken Foxp3, we hence transiently transfected HEK293T cells with full-length mouse or chicken Foxp3, respectively, and tested binding of selected anti-mouse Foxp3 mAbs with proven mammalian interspecies cross-reactivity by cytoplasmic staining. Using the mouse Foxp3-transfected cells as a positive control, we confirmed that both Abs bound mouse Foxp3 in the experimental setup used in the present study. However, no cross-reactivity with the chicken protein could be observed (Fig. 4).

FIGURE 4.

Test for cross-reactivity of selected anti-murine Foxp3 mAbs with chicken Foxp3. 293T cells were transfected with plasmids containing GFP and either Myc and murine Foxp3 (right panel, gray outline) or Myc and chicken Foxp3 (right panel, black outline). Forty-eight hours after transfection, cells were stained with anti-Myc and either MF-14 (top) or FJK-16s anti-murine Foxp3 mAbs (bottom), respectively, and analyzed by flow cytometry. Cells were gated for viability and Myc versus GFP (left panel). Empty vector-transfected cells (right panel, white outline) were used as a negative control (gated on GFP+). Figure is representative of two independent experiments. chFoxp3, chicken Foxp3; muFoxp3, murine Foxp3.

FIGURE 4.

Test for cross-reactivity of selected anti-murine Foxp3 mAbs with chicken Foxp3. 293T cells were transfected with plasmids containing GFP and either Myc and murine Foxp3 (right panel, gray outline) or Myc and chicken Foxp3 (right panel, black outline). Forty-eight hours after transfection, cells were stained with anti-Myc and either MF-14 (top) or FJK-16s anti-murine Foxp3 mAbs (bottom), respectively, and analyzed by flow cytometry. Cells were gated for viability and Myc versus GFP (left panel). Empty vector-transfected cells (right panel, white outline) were used as a negative control (gated on GFP+). Figure is representative of two independent experiments. chFoxp3, chicken Foxp3; muFoxp3, murine Foxp3.

Close modal

The ongoing search for “hidden” genes in birds is rewarded with increasing success due to progression in fields of sequencing techniques and bioinformatics. Using a large chicken RNA sequencing dataset, we were able to manually assemble a gene sequence homologous to mammalian and reptilian Foxp3, as well as to the sequence automatically annotated as Foxp3 in the ground tit (P. humilis) genome published in 2013 (11). Phylogenetic analysis and sequence alignment revealed unique characteristics of Foxp3 among the Foxp family members. The various analyzed Foxp3 sequences cluster to one clade. Furthermore, the chicken sequence clusters closest with the automatically annotated sequence of the ground tit, followed by alligator Foxp3. This mirrors the proven common archosaur origin of the crocodilian reptiles and avian species (37) and thereby emphasizes the accuracy of the chicken Foxp3 assembly (Fig. 1). PCR amplification of the assembled Foxp3 sequence from chicken tissue mRNA subsequently proved the gene’s transcription in vivo. In addition, we succeeded in mapping Ccdc22, a gene within the syntenic locus of mammalian Foxp3, to the same long sequencing read as the assembled chicken Foxp3 sequence. We thereby prove conserved synteny (Supplemental Fig. 2). The Foxp3 clade, however, displayed the highest heterogeneity compared with the clades of Foxp1, Foxp2, or Foxp4 proteins, respectively: only the two included mammalian species, mice and humans, showed vast similarity for Foxp3. Nevertheless, these species’ Foxp3 sequences showed higher numbers of substitutions per site than did the other three members of the Foxp family between these species. In the course of evolution, Foxp3 therefore appears to have been the subject of fundamental structural changes. This assumption is underlined by the amount of guanine and cytosine residues present in Foxp3 transcripts (Supplemental Fig. 1). It is known that structural damage to DNA is more likely to be repaired using G or C rather than A or T bases (7). As the class of birds is phylogenetically older than mammals, their genomes have been subject to a higher accumulated mutation probability throughout evolution. Consequently, high GC content is observable especially in immune-related genes, which are liable to a high selection pressure (1). This led to delayed annotation of Foxp3 in birds, because the strong triple covalent bonds between G and C impair amplification. Thereby, high GC content, especially in combination with long stretches of only G and C, significantly reduces sequencing accuracy using classical methods (2). Such sequences are also predicted to form various secondary structures that might impede analysis (2, 38). However, in this study we present the first successful full-length assembly of a chicken homolog of mammalian Foxp3.

Due to the comprehensive remodeling of Foxp3 since the split of the avian and mammalian lineages from the common ancestor, Foxp3 proteins in modern species are likely to vary in domain organization. To investigate this hypothesis further, we aligned the translation of the newly assembled chicken Foxp3 transcript sequence with Foxp3 of five other vertebrates (Fig. 2). The highest consensus is present at the C-terminal FKH domain. It is, in mammals, responsible for binding of the transcription factor Foxp3 to DNA, particularly conveyed by the nuclear localization signals on both ends of the domain (34). Both sites are conserved in the chicken sequence, indicating Foxp3’s ancient function as a transcription factor. However, single residues with predicted crucial roles in NFAT interaction and DNA binding (33) are only partially conserved, suggesting distinct class-specific differences for functional pathways that Foxp3 is involved in. The FKH domain is connected to the N-terminal effector domain by a leucine zipper, which is also highly conserved among all species. It conveys dimerization. To influence the transcription of >700 genes in total (39), Foxp3 forms large multiprotein complexes. The initial dimerization, however, results in homodimers or in heterodimers with Foxp1 (40, 41). The glutamic acid residue at position 251, within the leucine zipper domain, is crucial for these dimerization events. We found E251 to be conserved in the chicken sequence, rendering a conservation of Foxp3’s overall quaternary structure in this species likely. The zinc finger domain, in contrast, does not seem to be present in any non-mammalian species we investigated, which supports findings by an earlier study on Foxp3 ontogeny (33). However, a nonredundant role of the Foxp3 zinc finger domain has been doubted by a study examining truncated versions of the protein that reported unrestricted Foxp3 functionality regardless of the presence or absence of the zinc finger (34). The repressive function itself is conveyed by the N-terminal proline-rich domain (34). This domain distinguishes Foxp3 from the other members of the Foxp family, which are characterized by a glutamic acid–rich N terminus. Several studies describe functional motifs within the N-terminal domain (42, 43), but the exact mechanisms by which the proline content affects repressor function remain to be elucidated (44). However, even though consensus of the six examined vertebrate species is poor in the N-terminal region, the chicken protein features an even higher proline content than for the mammalian sequences. This finding confirms results of a previous study on avian Foxp3, which assembled partial sequences for two falcon species (12). Taken together, the comparison of Foxp3 sequences among species implies that Foxp3 adapted its functional repertoire throughout evolution to meet newly arisen requirements of avian and mammalian physiology, respectively. Nevertheless, structural similarities indicate that Foxp3 fulfilled certain basic tasks, which are likely to be of repressive nature, ever since the common vertebrate ancestor (33). Therefore, chicken Foxp3 promises to be an excellent candidate for studying the protein’s nonredundant functions in the future.

The common development of all T cell subsets takes place in the thymus. It proceeds from a CD4/CD8 DN via a DP to finally an SP state. These mature, primed SP cells are (revocably) committed to a T cell lineage before they leave the thymus and enter peripheral lymphatic organs, for example, spleen or cecal tonsil. For murine Tregs, lineage determination is characterized by the expression of CD4, CD25, and Foxp3 (14). Hence, we examined Foxp3 expression in lymphocyte subsets that had been extracted from lymphatic organs and sort purified for the presence or absence of surface CD4 and CD25. We found that Foxp3 expression is associated with CD4+CD25+ subpopulations in peripheral lymphatic organs, that is, spleen and cecal tonsil, but not in the thymus. As the thymus represents the major organ for T cell development, this finding raised the question on the origin of Foxp3+ avian T cells. To further investigate this issue, we subsequently analyzed CD4 thymic lymphocytes. Within this subset, surprisingly, we detected high Foxp3 expression in the rare CD25+ cells (Fig. 3A, dark gray bars). We could not detect any CD8 expression in this population, whereas the CD25 subsets of CD4 as well as CD4+ thymocytes were highly CD8+ (Fig. 3B). This distribution of CD8 to the Foxp3 subsets only suggests a minor role of Foxp3-expressing CD8+ T cells in chickens. In line with what is known about mammalian thymic T cell development, the CD25 (and CD8+) subpopulations of CD4+ or CD4 lymphocytes are therefore likely to represent the CD4/CD8 DP and CD8 SP developmental stages, respectively. We also ruled out B cells as a source of Foxp3 in the CD4 population by analysis of AV20+ subsets. These cells did not express Foxp3 (Fig. 3C, dark gray bars). From these results, we conclude that Foxp3 is mainly expressed by DN CD25+ T cells in the chicken thymus and in a subset of CD4+CD25+ cells in secondary lymphatic organs.

In order to draw a more comprehensive picture of the phenotype of the T cell subsets examined, we next investigated the expression of the marker cytokines IL-10 and IL-2 (Fig. 3A, 3C, light gray and white bars, respectively). In mammals, IL-10 is the major repressive cytokine secreted by Tregs. It suppresses the transcription of IL-2 in Th cells, which is required for the generation of all T cell subsets and for maintenance of Th cells themselves (45, 46). IL-2 binds to naive CD4+ T cells expressing the IL-2 receptor complex, part of which is CD25, and induces determination toward either the helper or regulatory lineage, depending on the presence or absence of proinflammatory cytokines (47). We could show on the mRNA level that, also in chickens, IL-10 expression was strongly associated with Foxp3, whereas IL-2 was expressed mostly by the Foxp3 thymic CD4+CD25+ population and to a lesser extent by peripheral CD25+ lymphocytes. As it has been shown that CD25 can act as a naive T cell activation marker in mouse (8) and also in chicken γδ T cells (48), these findings appear to emphasize that mammalian and avian T cell populations share vastly similar phenotypes. The chicken lymphocyte populations from secondary lymphatic organs very likely display the overall cytokine pattern of activated Th cells or mature Tregs, in line with low or very high expression of Foxp3, respectively. These subsets could not be distinguished from each other in this study, which resulted in a mixed CD4+CD8CD25+Foxp3+IL-10+IL-2low population in peripheral lymphatic organs. We hypothesize that this mixed population consists, on the one hand, of one subset of CD4+CD25+ lymphocytes, which is low in IL-10 and intermediate in IL-2 expression (Fig. 3D). This phenotype resembles murine peripherally activated Th cells (8). The second subset, which presumably merged into the peripheral CD4+CD25+ sort population, expresses high levels of Foxp3 and IL-10 and therefore resembles mature Tregs. As we were able to show that Foxp3 is expressed most strongly in CD25high splenocytes (Supplemental Fig. 4), the Treg subset is likely to be found among this subpopulation.

Another subset of thymic CD25+ lymphocytes, on the other hand, indicates a developmental course of chicken Tregs that differs in one decisive point from that in mammals; that is, the vast majority of Foxp3+ thymocytes are CD4 (Fig. 3A). We hypothesize that these cells represent immature nTregs, which in other species do express CD4 (49). Apart from this substantial difference, we found evidence that the classical Treg developmental cascade in the thymus resembles that between the mouse and chicken: in immature Tregs, IL-2 interaction with CD25 seems to have triggered Foxp3 induction and, in line with that, facilitated IL-10 expression. Due to their immaturity, also chicken thymic Tregs (CD4CD25+Foxp3+) were yet comparably low in CD25. As opposed to that, Th cells in the thymus, that is, CD4+CD25+Foxp3 thymocytes, were CD25high, which indicates that they had been subject to strong activating signals. This confirms our expectations, as the thymus is the location of positive T cell selection via TCR affinity to self-antigens, as shown in mice (50). The chicken CD4+CD25high thymic lymphocyte population is therefore likely to represent naive Th cells in the act of priming. This subset expresses high amounts of IL-2, which is crucial for the thymic generation of nTregs (51). TGF-β is another important Treg marker cytokine in mammals (52). In addition, it has recently been discussed as a substitute marker for chicken Tregs in Marek’s disease, defining a Treg subset mostly distinct from CD4+CD25+ T cells (25). The study found CD4+TGF-β+ cells in in vitro–activated lymphocyte preparations from spleen, caecal tonsil, and blood of chickens susceptible to Marek’s disease, but not in the thymus. The authors concluded that TGF-β is a marker for peripherally induced chicken Tregs (25). Surprisingly, we could not detect any differential expression of TGF-β between the analyzed lymphocyte subsets on the mRNA level (data not shown). Nonetheless, we cannot exclude a specific role of TGF-β+ T cells with regulatory function in certain diseases. We therefore conclude that the importance of TGF-β for chicken nTregs is neglectable under physiological conditions.

However, the role of the thymic CD4CD25+Foxp3highIL-10highIL-2low subset remained unclear at this point. In mice, Foxp3 is induced through TCR signaling and with the help of the coreceptors CD4 or CD8, respectively. Therefore, Foxp3 induction cannot occur before the DP state (49). In a human study, however, a small portion of DN thymocytes was shown to express Foxp3, along with low amounts of CD25, questioning the overall TCR dependency of Foxp3 induction (53). To test whether association of Foxp3 with the expression of a functional TCR in chickens is obligatory, we analyzed the subset of potential pre-Tregs (DN CD25+Foxp3highIL-10highIL-2low) for surface expression of CD3 (Supplemental Fig. 3). CD3 serves as co-TCR and as such is part of any functional TCR complex. It is hence essential for activation of T cells; its absence on the cell surface excludes TCR signaling in the respective cell (54). We observed that the vast majority (93%) of assumed pre-Tregs did indeed not express CD3. However, a small proportion of the examined population is CD3+ and therefore possesses the tools for TCR signaling. Without appropriate tools on our side, namely a specific anti-chicken Foxp3 Ab, we have no means to determine whether Foxp3 is expressed in the small CD3+ or the large CD3 subpopulation. Thus, it yet remains unclear whether the presumed pre-Treg population has not yet undergone T cell maturation. In this case, the population would be negative for surface CD3 and thereby TCR, and the expression of Foxp3 is hence TCR-independent. A comparably low expression level of CD25 in this subset, which is in line with data for human cells expressing Foxp3 in a TCR-independent manner (53), argues for this hypothesis. In addition, Foxp3 would be expected to be stronger diluted and thus not show very high expression levels (compare with (Fig. 3) in the overall population if only a small fraction of the examined population (7%) would express it. Nevertheless, another possibility is that somewhere in the course of their differentiation, the DN CD25+ thymocytes lost CD4 and/or CD8 expression after going through the classical maturation pathway and TCR-dependent Foxp3 induction. In that case, Foxp3+ cells would be found in the small CD3+ subpopulation, whereas the remaining 97% of this population would represent DN thymocytes activated throughout the course of their maturation toward CD4+ or CD8+ cells. However, a recent comprehensive study showed that there are at least two distinct pathways for the development of mature Tregs in humans (55). We can therefore also not exclude the option that both of the above-mentioned hypotheses or additional ways that are not discussed in the present study represent the actual events. Hence, the developmental pathways of chicken Tregs remain to be elucidated.

To further investigate the development and phenotype of Foxp3-expressing cells, appropriate tools are required, first and foremost specific anti-chicken Foxp3 Abs. As we showed within this study, chicken Foxp3 contains highly conserved domains as well as such that are highly deviant from other species’ Foxp3. Likely due to the very high conservation of the FKH domain between members of the Fox protein family, and consequently low specificity of this domain for Foxp3, we did not find any commercially available Abs with proven epitopes within the C-terminal region. We therefore tested a pair of anti-mouse Foxp3 mAbs against epitopes within the N terminus of murine Foxp3 for cross-reactivity with the recombinant chicken protein (Fig. 4). The Abs we chose to investigate in this concern are the clones FJK-16s and MF-14. FJK-16s is a standard tool used to study Foxp3 in a number of mammalian species, for example, swine (56), dogs (57), mice (58), and cats (59). Besides FJK-16s, MF-14 is frequently used for mouse Foxp3 studies. Their epitopes (aa 75–125 or aa 72–107 of the mouse protein, for FJK-16s and MF-14, respectively) are located within an area encoded by exon 2 and partly by exon 3, a part of the highly variate repressor domain. As expected, none of the two Ab clones possesses affinity to the chicken protein. A previous study on cross-reactive mAbs showed that only 2 out of 182 mAbs raised against various human leukocyte proteins cross-reacted with chicken leukocytes (60). The mentioned study, however, concerned itself only with Abs against surface molecules. mAbs raised against partially highly conserved intracellular proteins might be more likely to cross-react. Therefore, we cannot exclude the existence of Abs with sufficient affinity against chicken Foxp3. However, the two most widely used anti-murine Foxp3 Ab clones, FJK-16s and MF-14, are not suitable for Foxp3 detection in chickens. The next step toward a comprehensive assessment of the chicken Treg phenotype will therefore be the generation of a species-specific anti-chicken Foxp3 mAb.

In this study we identify an ortholog of Foxp3 in the chicken genome. We show structural homology as well as a number of features deviating from mammalian sequences and conclude that chicken Foxp3 is a model protein to study nonredundant Foxp3 function. We furthermore examine the phenotype of Foxp3-expressing chicken lymphocytes in secondary lymphatic organs and the thymus. From the results of these experiments we deduct the hypothesis that, except for CD4 expression in immature Tregs, chicken and mammalian Tregs are vastly similar in phenotype and are likely to resemble each other also in function and development. By revealing this similarity, this study allows for, and demands, a new definition of avian Tregs on the background of what is known about Foxp3 in other species. It thus paves the way toward standardized avian Treg research.

We thank Prof. Thomas Brocker and Prof. Gerhild Wildner for providing FJK-16s anti-mouse Foxp3 Abs, and Prof. Stefan Endres for providing MF-14 anti-mouse Foxp3 Abs. In addition, we thank Marina Kohn for expert technical assistance and Prof. Dr. Steffen Weigend (Institute of Farm Animal Genetics, Friedrich-Loeffler-Institut Mariensee) for providing M11 birds.

This work was supported by the ERA-NET Animal Health and Welfare Initiative, project ANIHWA-MICHIC, Horizon 2020 (EU Framework Programme for Research and Innovation) Grant PTBLE2814ERA01D (to B.K.) and by Deutsche Forschungsgemeinschaft research unit “Immuno Chick,” Grant FOR5130 (to T.G., S.H., and B.K.).

N.B.B., D.E., S.H., B.S., and V.K. performed experiments. N.B.B., D.E., S.H., T.W.G., and B.K. planned the study and analyzed data. N.B.B., D.E., S.H., and B.K. wrote the manuscript. All authors contributed to, read, and approved the final manuscript.

The sequences presented in this article have been submitted to the National Center for Biotechnology Information’s GenBank database under accession numbers MT133687 and MT133688 and the Sequence Read Archive under accession number PRJNA610550.

The online version of this article contains supplemental material.

Abbreviations used in this article:

BLAST

basic local alignment search tool

DN

double-negative

DP

double-positive

FKH

forkhead

NCBI

National Center for Biotechnology Information

nTreg

naive Treg

qPCR

quantitative PCR

SP

single-positive

SRA

Sequence Read Archive

Treg

regulatory T cell

1.
International Chicken Genome Sequencing Consortium
.
2004
.
Sequence and comparative analysis of the chicken genome provide unique perspectives on vertebrate evolution. [Published erratum appears in 2005 Nature 433: 777.]
Nature
432
:
695
716
.
2.
Hron
T.
,
P.
Pajer
,
J.
Pačes
,
P.
Bartůněk
,
D.
Elleder
.
2015
.
Hidden genes in birds.
Genome Biol.
16
:
164
.
3.
Yin
Z. T.
,
F.
Zhu
,
F. B.
Lin
,
T.
Jia
,
Z.
Wang
,
D. T.
Sun
,
G. S.
Li
,
C. L.
Zhang
,
J.
Smith
,
N.
Yang
,
Z. C.
Hou
.
2019
.
Revisiting avian “missing” genes from de novo assembled transcripts.
BMC Genomics
20
:
4
.
4.
Lovell
P. V.
,
M.
Wirthlin
,
L.
Wilhelm
,
P.
Minx
,
N. H.
Lazar
,
L.
Carbone
,
W. C.
Warren
,
C. V.
Mello
.
2014
.
Conserved syntenic clusters of protein coding genes are missing in birds.
Genome Biol.
15
:
565
.
5.
Rohde
F.
,
B.
Schusser
,
T.
Hron
,
H.
Farkašová
,
J.
Plachý
,
S.
Härtle
,
J.
Hejnar
,
D.
Elleder
,
B.
Kaspers
.
2018
.
Characterization of chicken tumor necrosis factor-α, a long missed cytokine in birds.
Front. Immunol.
9
:
605
.
6.
Botero-Castro
F.
,
E.
Figuet
,
M. K.
Tilak
,
B.
Nabholz
,
N.
Galtier
.
2017
.
Avian genomes revisited: hidden genes uncovered and the rates versus traits paradox in birds.
Mol. Biol. Evol.
34
:
3123
3131
.
7.
Duret
L.
,
N.
Galtier
.
2009
.
Biased gene conversion and the evolution of mammalian genomic landscapes.
Annu. Rev. Genomics Hum. Genet.
10
:
285
311
.
8.
Fontenot
J. D.
,
M. A.
Gavin
,
A. Y.
Rudensky
.
2003
.
Foxp3 programs the development and function of CD4+CD25+ regulatory T cells.
Nat. Immunol.
4
:
330
336
.
9.
Selvaraj
R. K.
2013
.
Avian CD4+CD25+ regulatory T cells: properties and therapeutic applications.
Dev. Comp. Immunol.
41
:
397
402
.
10.
Shack
L. A.
,
J. J.
Buza
,
S. C.
Burgess
.
2008
.
The neoplastically transformed (CD30hi) Marek’s disease lymphoma cell phenotype most closely resembles T-regulatory cells.
Cancer Immunol. Immunother.
57
:
1253
1262
.
11.
Qu
Y.
,
H.
Zhao
,
N.
Han
,
G.
Zhou
,
G.
Song
,
B.
Gao
,
S.
Tian
,
J.
Zhang
,
R.
Zhang
,
X.
Meng
, et al
2013
.
Ground tit genome reveals avian adaptation to living at high altitudes in the Tibetan plateau.
Nat. Commun.
4
:
2071
.
12.
Denyer
M. P.
,
D. Y.
Pinheiro
,
O. A.
Garden
,
A. J.
Shepherd
.
2016
.
Missed, Not Missing: Phylogenomic Evidence for the Existence of Avian FoxP3.
PLoS One
11
:
e0150988
.
13.
Gershon
R. K.
,
P.
Cohen
,
R.
Hencin
,
S. A.
Liebhaber
.
1972
.
Suppressor T cells.
J. Immunol.
108
:
586
590
.
14.
Sakaguchi
S.
,
N.
Sakaguchi
,
M.
Asano
,
M.
Itoh
,
M.
Toda
.
1995
.
Immunologic self-tolerance maintained by activated T cells expressing IL-2 receptor alpha-chains (CD25). Breakdown of a single mechanism of self-tolerance causes various autoimmune diseases.
J. Immunol.
155
:
1151
1164
.
15.
Garden
O. A.
,
D.
Pinheiro
,
F.
Cunningham
.
2011
.
All creatures great and small: regulatory T cells in mice, humans, dogs and other domestic animal species.
Int. Immunopharmacol.
11
:
576
588
.
16.
Sahin
E.
,
M.
Sahin
.
2019
.
Epigenetical targeting of the FOXP3 gene by S-adenosylmethionine diminishes the suppressive capacity of regulatory T cells ex vivo and alters the expression profiles.
J. Immunother.
42
:
11
22
.
17.
Schmidt
A.
,
N.
Oberle
,
P. H.
Krammer
.
2012
.
Molecular mechanisms of Treg-mediated T cell suppression.
Front. Immunol.
3
:
51
.
18.
Veiga-Parga
T.
2016
.
Regulatory T cells and their role in animal disease.
Vet. Pathol.
53
:
737
745
.
19.
Bayer
A. L.
,
T. R.
Malek
.
2008
.
Requirements for IL-2 within the thymus and peripheral lymphoid tisues.
In
Regulatory T Cells and Clinical Application.
S.
Jiang
.
Springer
,
New York, NY
, p.
62
66
.
20.
Grebenau
M. D.
,
S. P.
Lerman
,
M. A.
Palladino
,
G. J.
Thorbecke
.
1976
.
Suppression of adoptive antibody responses by addition of spleen cells from agammaglobulinaemic chickens “immunised” with histocompatible bursa cells.
Nature
260
:
46
48
.
21.
Quere
P.
,
M. D.
Cooper
,
G. J.
Thorbecke
.
1990
.
Characterization of suppressor T cells for antibody production by chicken spleen cells. I. Antigen-induced suppressor cells are CT8+, TcR1+ (gamma delta) T cells.
Immunology
71
:
517
522
.
22.
Shanmugasundaram
R.
,
M. H.
Kogut
,
R. J.
Arsenault
,
C. L.
Swaggerty
,
K.
Cole
,
J. M.
Reddish
,
R. K.
Selvaraj
.
2015
.
Effect of Salmonella infection on cecal tonsil regulatory T cell properties in chickens.
Poult. Sci.
94
:
1828
1835
.
23.
Shanmugasundaram
R.
,
R. K.
Selvaraj
.
2011
.
Regulatory T cell properties of chicken CD4+CD25+ cells.
J. Immunol.
186
:
1997
2002
.
24.
Shanmugasundaram
R.
,
R. K.
Selvaraj
.
2012
.
Effects of in vivo injection of anti-chicken CD25 monoclonal antibody on regulatory T cell depletion and CD4+CD25 T cell properties in chickens.
Dev. Comp. Immunol.
36
:
578
583
.
25.
Gurung
A.
,
N.
Kamble
,
B. B.
Kaufer
,
A.
Pathan
,
S.
Behboudi
.
2017
.
Association of Marek’s disease induced immunosuppression with activation of a novel regulatory T cells in chickens.
PLoS Pathog.
13
:
e1006745
.
26.
Jones
D. T.
,
W. R.
Taylor
,
J. M.
Thornton
.
1992
.
The rapid generation of mutation data matrices from protein sequences.
Comput. Appl. Biosci.
8
:
275
282
.
27.
Tamura
K.
,
G.
Stecher
,
D.
Peterson
,
A.
Filipski
,
S.
Kumar
.
2013
.
MEGA6: Molecular Evolutionary Genetics Analysis version 6.0.
Mol. Biol. Evol.
30
:
2725
2729
.
28.
Plachý
J.
,
M.
Vilhelmová
,
I.
Karakoz
,
J.
Schulmannová
.
1989
.
Prague inbred lines of chickens: a biological model for MHC research.
Folia Biol. (Praha)
35
:
177
196
.
29.
Göbel
T. W.
,
K.
Schneider
,
B.
Schaerer
,
I.
Mejri
,
F.
Puehler
,
S.
Weigend
,
P.
Staeheli
,
B.
Kaspers
.
2003
.
IL-18 stimulates the proliferation and IFN-γ release of CD4+ T cells in the chicken: conservation of a Th1-like system in a nonmammalian species.
J. Immunol.
171
:
1809
1815
.
30.
Farkašová
H.
,
T.
Hron
,
J.
Pačes
,
P.
Pajer
,
D.
Elleder
.
2016
.
Identification of a GC-rich leptin gene in chicken.
Agri Gene
1
:
88
92
.
31.
Borowska
D.
,
L.
Rothwell
,
R. A.
Bailey
,
K.
Watson
,
P.
Kaiser
.
2016
.
Identification of stable reference genes for quantitative PCR in cells derived from chicken lymphoid organs.
Vet. Immunol. Immunopathol.
170
:
20
24
.
32.
Kothlow
S.
,
I.
Morgenroth
,
C. A.
Tregaskes
,
B.
Kaspers
,
J. R.
Young
.
2008
.
CD40 ligand supports the long-term maintenance and differentiation of chicken B cells in culture.
Dev. Comp. Immunol.
32
:
1015
1026
.
33.
Andersen
K. G.
,
J. K.
Nissen
,
A. G.
Betz
.
2012
.
Comparative genomics reveals key gain-of-function events in Foxp3 during regulatory T cell evolution.
Front. Immunol.
3
:
113
.
34.
Lopes
J. E.
,
T. R.
Torgerson
,
L. A.
Schubert
,
S. D.
Anover
,
E. L.
Ocheltree
,
H. D.
Ochs
,
S. F.
Ziegler
.
2006
.
Analysis of FOXP3 reveals multiple domains required for its function as a transcriptional repressor.
J. Immunol.
177
:
3133
3142
.
35.
Li
J.
,
L.
Jiang
,
X.
Liang
,
L.
Qu
,
D.
Wu
,
X.
Chen
,
M.
Guo
,
Z.
Chen
,
L.
Chen
,
Y.
Chen
.
2017
.
DNA-binding properties of FOXP3 transcription factor.
Acta Biochim. Biophys. Sin. (Shanghai)
49
:
792
799
.
36.
Dang
C. V.
,
M.
McGuire
,
M.
Buckmire
,
W. M.
Lee
.
1989
.
Involvement of the “leucine zipper” region in the oligomerization and transforming activity of human c-myc protein.
Nature
337
:
664
666
.
37.
Towers
M.
2018
.
Evolution of antero-posterior patterning of the limb: insights from the chick.
Genesis
56
:
e23047
.
38.
Beauclair
L.
,
C.
Ramé
,
P.
Arensburger
,
B.
Piégu
,
F.
Guillou
,
J.
Dupont
,
Y.
Bigot
.
2019
.
Sequence properties of certain GC rich avian genes, their origins and absence from genome assemblies: case studies.
BMC Genomics
20
:
734
.
39.
Marson
A.
,
K.
Kretschmer
,
G. M.
Frampton
,
E. S.
Jacobsen
,
J. K.
Polansky
,
K. D.
MacIsaac
,
S. S.
Levine
,
E.
Fraenkel
,
H.
von Boehmer
,
R. A.
Young
.
2007
.
Foxp3 occupancy and regulation of key target genes during T-cell stimulation.
Nature
445
:
931
935
.
40.
Konopacki
C.
,
Y.
Pritykin
,
Y.
Rubtsov
,
C. S.
Leslie
,
A. Y.
Rudensky
.
2019
.
Transcription factor Foxp1 regulates Foxp3 chromatin binding and coordinates regulatory T cell function.
Nat. Immunol.
20
:
232
242
.
41.
Rudra
D.
,
P.
deRoos
,
A.
Chaudhry
,
R. E.
Niec
,
A.
Arvey
,
R. M.
Samstein
,
C.
Leslie
,
S. A.
Shaffer
,
D. R.
Goodlett
,
A. Y.
Rudensky
.
2012
.
Transcription factor Foxp3 and its protein partners form a complex regulatory network.
Nat. Immunol.
13
:
1010
1019
.
42.
Deng
G.
,
Y.
Xiao
,
Z.
Zhou
,
Y.
Nagai
,
H.
Zhang
,
B.
Li
,
M. I.
Greene
.
2012
.
Molecular and biological role of the FOXP3 N-terminal domain in immune regulation by T regulatory/suppressor cells.
Exp. Mol. Pathol.
93
:
334
338
.
43.
Morawski
P. A.
,
P.
Mehra
,
C.
Chen
,
T.
Bhatti
,
A. D.
Wells
.
2013
.
Foxp3 protein stability is regulated by cyclin-dependent kinase 2.
J. Biol. Chem.
288
:
24494
24502
.
44.
Deng
G.
,
X.
Song
,
M. I.
Greene
.
2020
.
FoxP3 in Treg cell biology: a molecular and structural perspective.
Clin. Exp. Immunol.
199
:
255
262
.
45.
Sugimoto
N.
,
T.
Oida
,
K.
Hirota
,
K.
Nakamura
,
T.
Nomura
,
T.
Uchiyama
,
S.
Sakaguchi
.
2006
.
Foxp3-dependent and -independent molecules specific for CD25+CD4+ natural regulatory T cells revealed by DNA microarray analysis.
Int. Immunol.
18
:
1197
1209
.
46.
Bayer
A. L.
,
A.
Yu
,
D.
Adeegbe
,
T. R.
Malek
.
2005
.
Essential role for interleukin-2 for CD4+CD25+ T regulatory cell development during the neonatal period.
J. Exp. Med.
201
:
769
777
.
47.
Nelson
B. H.
2002
.
Interleukin-2 signaling and the maintenance of self-tolerance.
Curr. Dir. Autoimmun.
5
:
92
112
.
48.
Wattrang
E.
,
P.
Thebo
,
A.
Lundén
,
T. S.
Dalgaard
.
2016
.
Monitoring of local CD8β-expressing cell populations during Eimeria tenella infection of naïve and immune chickens.
Parasite Immunol.
38
:
453
467
.
49.
Juvet
S. C.
,
L.
Zhang
.
2012
.
Double negative regulatory T cells in transplantation and autoimmunity: recent progress and future directions.
J. Mol. Cell Biol.
4
:
48
58
.
50.
Hemmers
S.
,
M.
Schizas
,
E.
Azizi
,
S.
Dikiy
,
Y.
Zhong
,
Y.
Feng
,
G.
Altan-Bonnet
,
A. Y.
Rudensky
.
2019
.
IL-2 production by self-reactive CD4 thymocytes scales regulatory T cell generation in the thymus.
J. Exp. Med.
216
:
2466
2478
.
51.
Bayer
A. L.
,
A.
Yu
,
T. R.
Malek
.
2007
.
Function of the IL-2R for thymic and peripheral CD4+CD25+ Foxp3+ T regulatory cells.
J. Immunol.
178
:
4062
4071
.
52.
Konkel
J. E.
,
D.
Zhang
,
P.
Zanvit
,
C.
Chia
,
T.
Zangarle-Murray
,
W.
Jin
,
S.
Wang
,
W.
Chen
.
2017
.
Transforming growth factor-β signaling in regulatory t cells controls T helper-17 cells and tissue-specific immune responses.
Immunity
46
:
660
674
.
53.
Tuovinen
H.
,
E.
Kekäläinen
,
L. H.
Rossi
,
J.
Puntila
,
T. P.
Arstila
.
2008
.
Cutting edge: human CD4CD8 thymocytes express FOXP3 in the absence of a TCR.
J. Immunol.
180
:
3651
3654
.
54.
Wegener
A. M.
,
F.
Letourneur
,
A.
Hoeveler
,
T.
Brocker
,
F.
Luton
,
B.
Malissen
.
1992
.
The T cell receptor/CD3 complex is composed of at least two autonomous transduction modules.
Cell
68
:
83
95
.
55.
Owen
D. L.
,
S. A.
Mahmud
,
L. E.
Sjaastad
,
J. B.
Williams
,
J. A.
Spanier
,
D. R.
Simeonov
,
R.
Ruscher
,
W.
Huang
,
I.
Proekt
,
C. N.
Miller
, et al
2019
.
Thymic regulatory T cells arise via two distinct developmental programs.
Nat. Immunol.
20
:
195
205
.
56.
Käser
T.
,
W.
Gerner
,
A.
Saalmüller
.
2011
.
Porcine regulatory T cells: mechanisms and T-cell targets of suppression.
Dev. Comp. Immunol.
35
:
1166
1172
.
57.
Rabiger
F. V.
,
K.
Rothe
,
H.
von Buttlar
,
D.
Bismarck
,
M.
Büttner
,
P. F.
Moore
,
M.
Eschke
,
G.
Alber
.
2019
.
Distinct features of canine non-conventional CD4CD8α double-negative TCRαβ+ vs. TCRγδ+ T cells.
Front. Immunol.
10
:
2748
.
58.
Majumder
K.
,
Y.
Jin
,
H.
Shibata
,
Y.
Mine
.
2020
.
Oral intervention of Lactobacillus pentosus S-PT84 attenuates the allergenic responses in a BALB/C mouse model of egg allergy.
Mol. Immunol.
120
:
43
51
.
59.
Sparger
E. E.
,
B. G.
Murphy
,
F. M.
Kamal
,
B.
Arzi
,
D.
Naydan
,
C. T.
Skouritakis
,
D. P.
Cox
,
K.
Skorupski
.
2018
.
Investigation of immune cell markers in feline oral squamous cell carcinoma.
Vet. Immunol. Immunopathol.
202
:
52
62
.
60.
Saalmüller
A.
,
J. K.
Lunney
,
C.
Daubenberger
,
W.
Davis
,
U.
Fischer
,
T. W.
Göbel
,
P.
Griebel
,
E.
Hollemweguer
,
T.
Lasco
,
R.
Meister
, et al
2005
.
Summary of the animal homologue section of HLDA8.
Cell. Immunol.
236
:
51
58
.
61.
Göbel
T.W.
,
B.
Kaspers
,
M.
Stangassinger
.
2001
.
NK and T cells constitute two major, functionally distinct intestinal epithelial lymphocyte subsets in the chicken.
Int. Immunol.
13
:
757
762
.
62.
Luhtala
M.
,
O.
Lassila
,
P.
Toivanen
,
O.
Vainio
.
1997
.
A novel peripheral CD4+ CD8+ T cell population: inheritance of CD8α expression on CD4+ T cells.
Eur. J. Immunol.
27
:
189
193
.

The authors have no financial conflicts of interest.

Supplementary data