Abstract
ILs, as essential innate immune modulators, are involved in an array of biological processes. In the European rabbit (Oryctolagus cuniculus) IL-1α, IL-1β, IL-2, IL-4, IL-8, IL-10, IL-12A, IL-12B, IL-15 and IL-18 have been implicated in inflammatory processes and in the immune response against rabbit hemorrhagic disease virus and myxoma virus infections. In this study we characterized these ILs in six Lagomorpha species (European rabbit, pygmy rabbit, two cottontail rabbit species, European brown hare and American pika). Overall, these ILs are conserved between lagomorphs, including in their exon/intron structure. Most differences were observed between leporids and American pika. Indeed, when comparing both, some relevant differences were observed in American pika, such as the location of the stop codon in IL-1α and IL-2, the existence of a different transcript in IL8 and the number of cysteine residues in IL-1β. Changes at N-glycosylation motifs were also detected in IL-1, IL-10, IL-12B and IL-15. IL-1α is the protein that presents the highest evolutionary distances, which is in contrast to IL-12A where the distances between lagomorphs are the lowest. For all these ILs, sequences of human and European rabbit are more closely related than between human and mouse or European rabbit and mouse.
Keywords: Immune system, interleukins, lagomorphs
Introduction
ILs are polypeptides of low molecular mass involved in several biological activities, including immunity, inflammation, inflammatory diseases, hematopoiesis, oncogenesis and fertility, among others.1–3 In vertebrates, many of these proteins participate in host defense with complementary and conflicting roles in induction, regulation and functioning of the immune system by regulating growth, differentiation, effector functions and survival of cells.4,5 ILs are crucial for the immune response. They are produced as an integral part of the innate immune response, and have the ability to influence the result and nature of adaptive immune response.4–8 Despite these functions, ILs are also considered important therapeutic targets;9–11 therefore, any changes in their sequence or structure may lead to alteration in their normal functioning. In mammals, ILs encoding genes are among the ones with faster evolution,4,6 and 28 of the 46 known ILs present signatures of positive selection.12
Although being well characterized in most mammalian groups, little is known about ILs in lagomorphs, with the exception of the European rabbit (Oryctolagus cuniculus). The order Lagomorpha comprises two families, Ochotonidae and Leporidae, that diverged about 35 million years ago (mya).13 While the family Ochotonidae comprise only one genus (Ochotona or pikas), the family Leporidae comprises 11 genera of hares and rabbits that are widely distributed: pygmy rabbit (genus Brachylagus), riverine rabbit (genus Bunolagus), striped rabbit (genus Nesolagus), European rabbit (genus Oryctolagus), Amami rabbit (genus Pentalagus), Bunyoro rabbit (genus Poelagus), red rock rabbit (genus Pronolagus), volcano rabbit (genus Romerolagus), cottontail rabbits (Sylvilagus sp.), hispid hare (genus Caprolagus) and true hares and jackrabbits (Lepus sp.).13–15 The ones that are more closely related to the Oryctolagus genus are Bunolagus, Caprolagus and Pentalagus with divergence times of ∼7, ∼8 and ∼9 mya, respectively. In contrast, Nesolagus, Poelagus and Pronolagus are less related to Oryctolagus, with divergence times of ∼15 mya.13 The other two leporidae genera with some information for ILs, Sylvilagus and Lepus, diverged from Oryctolagus at ∼12 mya.13,16–18
The European rabbit is one of the most used laboratory animal models for immunologic research, including the study of atherosclerosis,19 intestinal immunity,20 arthritis,21 cancer,22 Alzheimer’s disease23 and several viral infections.24–26 In addition, the European rabbit is a key species in the Mediterranean ecosystem, where it is strongly affected by two viral diseases, rabbit hemorrhagic disease (RHD) and myxomatosis.27–32 RHD is caused by a single-stranded RNA virus, the rabbit hemorrhagic disease virus (RHDV), while myxomatosis is caused by the myxoma virus (MYXV), a double-stranded DNA virus.
Rabbit resistance to both viral diseases is highly dependent on the immune response of the host in order to develop an effective adaptive immune response to control these viruses.25,33 ILs are not only important in the European rabbit immune response against RHDV and MYXV infections,24,25 but also in inflammatory processes.34 For RHDV, IL-1, IL-2, IL-6, IL-8 and IL-10 are among the most important ILs. Indeed, young European rabbits infected with RHDV showed an increase of IL-1 during an early stage of infection until 18 h post-infection, with a decrease to normal values at 24 h of infection, while IL-8 is particularly increased at 24 h after infection, probably owing to an increase of leukocyte migration to the site of infection.25 In contrast, in infected adult rabbits, IL-10 is significantly increased in the course of the disease.25,26
For MYXV, IL-12, IL-15 and IL-18 have an important anti-viral activity. Indeed, recombinants of myxoma virus expressing human IL-12, despite being similar to the wild type, do not induce myxomatosis in rabbits.35 IL-15 prevents lethal myxomatosis in the New Zealand rabbit breed through the stimulation of an immune response that leads to the elimination of the viral infection.36 Poxviruses developed the ability to block IL-18 by interfering with proteins crucial for IL-18 activation or function highlighting the importance of its anti-viral activity.37–39 In contrast, expression of European rabbit IL-4 by recombinant myxoma virus strains increases virus virulence and overcomes genetic resistance in wild rabbits.40
Considering the important biological role in the European rabbit immune response, we performed a genetic characterization of IL-1, which includes two biologically similar antagonist proteins IL-1α and IL-1β,41–45 IL-2, IL-4, IL-8, IL-10, IL-12A, IL-12B, IL-15 and IL-18 in five Lagomorpha genera (Oryctolagus, Brachylagus, Sylvilagus, Lepus and Ochotona).
Materials and methods
Samples of European rabbit (O. c. cuniculus and O. c. algirus), pygmy rabbit (Brachylagus idahoensis), cottontail rabbits (brush rabbit, Sylvilagus bachmani, and eastern cottontail, Sylvilagus floridanus), European brown hare (Lepus europaeus) and American pika (Ochotona princeps) were provided by the CIBIO Lagomorpha tissue collection. Genomic DNA (gDNA) was extracted using the EasySpin Genomic DNA Minipreps Tissue Kit (Citomed, Torun, Poland) according to the manufacturer’s instructions. Total RNA was extracted by using the RNeasy Mini Kit according to the manufacturer’s instructions (Qiagen, Hilden, Germany) from one specimen of European rabbit, European brown hare, eastern cottontail and American pika. Complementary DNA (cDNA) was synthesized using oligo(dT) as primers and SuperScript III reverse transcriptase (Invitrogen, Carlsbad, CA, USA). The European rabbit and American pika IL sequences were retrieved from public databases (accession numbers are given in bold in Figure 1). PCR amplification was performed with the Multiplex PCR Kit (Qiagen) using several pairs of primers designed according to the retrieved sequences (Supplementary Material Table 1). Sequencing was performed on an ABI PRISM 310 Genetic Analyzer (PE Applied Biosystems, Foster City, CA, USA) and PCR products were sequenced in both directions. The sequences obtained were submitted to GenBank (accession numbers: KT216045-KT216070; KT273911-273919; KT279631-KT279692).
Figure 1.
Alignment of the studied ILs for the different lagomorphs: (a) IL-1α; (b) IL-1β; (c) IL-2; (d) IL-4; (e) IL-8; (f) detail of the splicing region of exon 2 of IL-8, with the splicing regions underlined; (g) IL-10; (h) IL-12 A; (i) IL-12B; (j) IL-15; (k) IL-18. GenBank and Ensembl accession numbers are indicated in bold for the retrieved sequences. Positively selected aa are boxed according to Neves et al.,12 N-Glycosylation sites are shaded in light gray and cysteine residues are shaded in dark gray. (*) Represent stop codons; (–) represent deletions; *1 and *2 represent alleles. The numbering is according to the European rabbit sequences. The signal peptide and indels, indicated as (–), were included in the numbering.
The PHASE program, built into the software DnaSP,46 was used to reconstruct the haplotype phases of the obtained sequences that were aligned using MUltiple Sequence Comparison by Log-Expectation (MUSCLE) available at http://www.ebi.ac.uk/.47
Putative N-glycosylation sites were predicted using the NetNGlyc 1.0 server available at http://www.cbs.dtu.dk/services/NetNGlyc/.48
Predicted splicing sites were determined using the NetGene2 server available at http://www.cbs.dtu.dk/services/NetGene2/.49,50
The number of amino acid (aa) differences per site between sequences was estimated in MEGA6 with the following options:51 bootstrap method (1000 replicates), P-distance as model and pairwise deletion for gaps/missing data treatment. A maximum likelihood (ML) approach was used to estimate, for each gene, the phylogenetic relationships between the aa sequences. The ML trees were estimated in MEGA6 by using the best-fit nucleotide substitution model predicted by the same software and bootstrap 1000 replicates. In order to simplify the outputs and avoid duplicated sequences, only one sequence from each species was used.
The secondary structure of each IL was predicted using PsiPred (http://bioinf.cs.ucl.ac.uk/psipred/)52,53 and DiAminoacid Neural Network Application (DiANNA) (http://clavius.bc.edu/∼clotelab/DiANNA/).54 These methods predict protein cysteines that create permanent structural disulfide bonds. PsiPred uses Position Specific Iterated–BLAST (PSI-BLAST) searches of the non-redundant protein sequence database to obtain evolutionary information used to predict the secondary structure of the query protein. DiANNA is a neural network trained to recognize cysteines in an oxidized state (sulfur covalently bonded) telling them apart from those in a reduced state.
Results
IL-1α, IL-1β, IL-2, IL-4, IL-8, IL-10, IL-12A, IL-12B, IL-15 and IL-18 were amplified from gDNA for the two European rabbit subspecies, European brown hare, pygmy rabbit, brush rabbit and American pika. These ILs were also successfully amplified from cDNA of European rabbit (O. c. cuniculus), European brown hare, eastern cottontail and American pika. The allelic forms identified in DnaSP are shown in Figure 1. The results obtained are summarized in Table 1.
Table 1.
Summary of the alterations observed between lagomorph species for the ILs studied.
| Insertion/deletion | Cysteine residues | N-Glycosylation (N-X-T) | |
|---|---|---|---|
| IL-1α | American pika: deletion aa 52; insertion aa 179, aa 272–275 | – | Cottontails: N-X-T missing (Asn64) America pika: two different N-X-T (Asn55 and Asn92) |
| IL-1β | American pika: deletion aa 39–43; deletion aa 102 and 108 Pygmy rabbit and cottontail rabbits: deletion aa254 | American pika: extra Cys aa 135 | European rabbit and European brown hare: extra N-X-T (Asn254) |
| IL-2 | Pygmy rabbit: insertion aa 7 and 8 American pika: insertion aa 139–141 and aa 174; deletion of aa 146 | – | – |
| IL-4 | – | – | – |
| IL-8 | American pika: insertion aa 50–53 | – | – |
| IL-10 | – | – | European brown hare: extra N-X-T (Asn67) |
| IL-12A | – | Brush rabbit: extra Cys aa 54 | – |
| IL-12B | European rabbit, Brush rabbit and European brown hare: insertion aa 174–177 Brush rabbit, European brown hare and American pika: insertion aa 189 | – | Brush rabbit: extra N-X-T (Asn147 and Asn 280) |
| IL-15 | – | American pika: extra Cys aa 140 | Pygmy rabbit and European brown hare: N-X-T missing (Asn152) American pika: N-X-T missing (Asn168) |
| IL-18 | European rabbit: deletion aa 14 | – | – |
IL-1
IL-1 is composed by two proteins, IL-1α and IL-1β. According to the public databases NCBI, Ensembl and Uniprot, in the European rabbit these proteins are located in tandem in the reverse strand of chromosome 2, while in the American pika they are located in the forward strand.
IL-1α
In lagomorphs, IL-1α is organized into seven exons, with six of them being coding exons. The coding sequence comprises 801 base pairs (bp) that translate into a protein with 267 amino acids (aa). However, for American pika, and like the human IL-1α, the coding sequence has 813 bp translating into a protein with 271 aa (Figure 1a). Some differences existed between lagomorphs: American pika presents an aa deletion at position 52 of the European rabbit and an insertion of five aa at positions 179 and from 272 to 275. All cysteine residues are conserved between lagomorphs (Cys14 and Cys47). Three potential N-glycosylation sites were detected for all leporids (Asn64, Asn103 and Asn144), but Asn64 is absent in the cottontail rabbits studied. For American pika, there are two other potential N-glycosylation sites (Asn55 and Asn92).
IL-1β
In lagomorphs, IL-1β comprises six coding exons, translating into proteins with different aa lengths: 268 aa in the European rabbit and 261 aa in American pika (Figure 1b). In the American pika sequence we observed three deletions: five aa between positions 39 and 43, including the deletion of two cysteine residues that are conserved in all leporids; the two other deletions are located at aa 102 and 108. Between leporids there are nine conserved cysteines at sites 35, 42, 43, 54, 81, 105, 117, 126 and 189. In the American pika sequence, only four cysteines are conserved (Cys81, Cys117, Cys126 and Cys189), and an extra cysteine is found at position 135. All lagomorphs have a putative N-glycosylation site at Asn58; European rabbit and European brown hare have an additional putative N-glycosylation site at Asn254 that had been lost in both the brush rabbit and in the eastern cottontail rabbit due to a deletion of this aa.
IL-2
Regarding the IL-2 sequence, this protein is highly conserved between European rabbit, human, cat and horse.34 In the European rabbit, IL-2 is located in the reverse strand of chromosome 15, with four coding exons that translate into a protein with 153 aa. In American pika IL-2 also has four exons that translate into a 156aa protein (Figure 1c). Differences between lagomorphs include the insertion of two aa in pygmy rabbit (Met7 and Phe8) and four aa in American pika (Thr139, Ser140, Arg141 and Asn174). In American pika we also observed a deletion of one aa that corresponds to position 146 of European rabbit. All cysteine residues are conserved between lagomorphs (Cys11, Cys94, Cys144 and Cys164). The search of putative N-glycosylation sites identified one site (Asn127) common to all lagomorphs.
IL-4
In the European rabbit, IL-4 is located in chromosome 3 (forward strand). In the European rabbit and American pika, IL-4 is organized into four coding exons that translate into a protein of 147 aa (Figure 1d). No alterations, including insertions/deletions, were observed between IL-4 of lagomorphs. Also, cysteine residues were conserved (Cys13, Cys48, Cys70, Cys89, Cys115 and Cys123) and the same potential N-glycosylation site was detected (Asn62).
IL8
IL-8 from European rabbit is located in the reverse strand of chromosome 15 and it is encoded by four exons that translate into a 101-aa protein. For American pika, the coding sequence available in Ensembl.org (ENSOPRT00000002239) is only known by projection and it is incomplete by not starting with an initiation codon and by missing the stop codon. Nevertheless, we successfully amplified IL-8 for American pika from both gDNA and cDNA (Figure 1e, f). Our results showed that American pika has an insertion of four aa (Gly50, Phe51, Pro52 and Thr53). The cysteine residues are conserved among lagomorph species (Cys48, Cys61, Cys63, Cys93 and Cys109) and there are no putative N-glycosylation sites.
IL-10
IL-10 from European rabbit is located in the forward strand of chromosome 2 and, in both the European rabbit and American pika, it is composed of five coding exons that translate into a protein with 178 aa (Figure 1g). The cysteine residues are conserved between species (Cys8, Cys9 Cys30, Cys80, Cys126 and Cys132). The search for N-glycosylation sites revealed two potential sites in leporids (Asn100 and Asn134). Ans67 was also predicted for European brown hare and American pika has only one of the sites predicted for N-glycosylation (Asn134).
IL-12
IL-12 is a heterodimeric protein composed by two proteins, IL-12A and IL-12B. For European rabbit, and according to the public databases NCBI and Ensembl, the encoding genes are located in the forward strand of chromosome 14 and in the reverse strand of chromosome 3, respectively.
IL-12A
IL-12A from European rabbit is composed of seven coding exons that translate into a protein with 219 aa. For American pika, there are two different coding sequences available in NCBI (XM_012930663.1) and Ensembl (ENSOPRT00000006274). When compared with the European rabbit sequence, the American pika sequence available in NCBI has a start codon 109 aa upstream, while the sequence available in Ensembl is incomplete, as it lacks an initiation codon and aa information in the middle of the sequence. We were not able to amplify IL12A from gDNA or cDNA of American pika (Figure 1h). All cysteine residues are conserved between leporids (Cys2, Cys71, Cys98, Cys119, Cys130, Cys144, Cys157 and Cys230). In addition, brush rabbit presents an extra cysteine at position 54. Four potential N-glycosylation sites were detected for all leporids (Asn73, Asn127, Asn141 and Asn214).
IL-12B
IL-12B from European rabbit and American pika is composed by six coding exons that translate into a protein with 324 and 329 aa, respectively. Despite all the attempts, we were unable to amplify IL-12B from cDNA of any of the lagomorphs studied. Successful amplification from gDNA for all leporids showed that in the middle of exon 4 there is an insertion of four aa (Thr174 to Ala177) that is absent in the predicted European rabbit sequence available in NCBI (XM_002710347.2). However, this insertion is present in human, mouse and rat cDNA (Figure 1i). All cysteine residues (Cys2, Cys50, Cys90, Cys131, Cys142, Cys170, Cys194 and Cys200, Cys275, Cys291, Cys309 and Cys336) and the two potential N-glycosylation sites (Asn135, Asn223) are conserved between lagomorphs. In addition, brush rabbit evidenced two other potential N-glycosylation sites (Asn147 and Asn280).
IL-15
In the European rabbit, IL-15 is located in chromosome 15 (reverse strand) and it is organized into six coding exons that translate into a protein with 162 aa. There are two sequences described for the European rabbit in NCBI, with the accession numbers NM_001082216.2 and XM_008267272.1, being the latter a variant with an insertion of one aa (Gln127) (Figure 1j). For American pika the information available in NCBI and Ensembl is different for the 5’ coding region. Amplification was only successful using the primers constructed according to the American pika IL15 sequence available in Ensembl (ENSOPRT00000008393). All cysteine residues are conserved in lagomorphs (Cys39, Cys42, Cys62, Cys107, Cys114, Cys158 and Cys161), although American pika has an extra cysteine (Cys40). Four potential N-glycosylation sites were detected in lagomorphs, Asn132, Asn143, Asn152 and Asn168, but Asn152 is absent in pygmy rabbit and in European brown hare, and Asn168 is absent in American pika.
IL-18
The European rabbit’s IL-18 is located in the forward strand of chromosome 1 and is organized into five coding exons that translate into a protein with 192 aa. For American pika there are two different predicted sequences available, one from NCBI (XM_004585222.1) and the other from Ensembl (ENSOPRG00000012703). When compared with the other lagomorphs, the American pika sequence from NCBI had an insertion of six aa in the first exon, while exon 5 presented low aa identity and a smaller size. The sequence from Ensembl is incomplete as it lacks an initiation codon and exon 5 (Figure 1k). For this reason, and despite all attempts, we were unable to amplify IL-18 from gDNA or cDNA of American pika. The European rabbit has a deletion at aa 14, whereas all the other leporids have a Gln residue. All cysteine residues are conserved between leporids (Cys16, Cys80, Cys110, Cys118 and Cys169). Our search revealed that Asn135 is the only putative N-glycosylation site.
Regarding the number of aa differences, brush rabbit and pygmy rabbit are among the species more related for IL-1, IL-2, IL-4 and IL-18, while cottontail rabbits and European rabbit evidenced a closer relationship for IL-4, IL-8, IL-10, IL-12 and IL-15. However, American pika was the most distantly related when compared with other lagomorphs (Supplementary Material Table 2). IL-1α presented the higher aa distances between lagomorphs (0.038–0.308), while IL-12A showed the lowest (0.018–0.096). When comparing the aa distances between human, mouse and European rabbit ILs sequences (Table 2), we observed that for all the ILs studied the aa distances are significantly lower between human–European rabbit than between human–mouse or European rabbit–mouse. These results were further confirmed with the phylogenetic analysis (Figure 2). Overall, we also observed that for the studied ILs, lagomorphs cluster according to the accepted evolutionary topology. However, for IL-1, IL-10, IL-12A, IL-15 and IL-18, these proteins present low bootstrap values to support these clusters.
Table 2.
Aa distances between rabbit, mouse and human for the different ILs studied (the lowest values for each IL are in bold).
| Rabbit | Mouse | ||
|---|---|---|---|
| IL-1α | Human | 0.350 | 0.383 |
| Rabbit | – | 0.383 | |
| IL-1β | Human | 0.254 | 0.315 |
| Rabbit | – | 0.271 | |
| IL-2 | Human | 0.199 | 0.373 |
| Rabbit | – | 0.399 | |
| IL-4 | Human | 0.463 | 0.564 |
| Rabbit | – | 0.593 | |
| IL-8 | Human | 0.192 | – |
| Rabbit | – | – | |
| IL-10 | Human | 0.202 | 0.270 |
| Rabbit | – | 0.281 | |
| IL-12A | Human | 0.183 | 0.400 |
| Rabbit | – | 0.391 | |
| IL-12B | Human | 0.177 | 0.310 |
| Rabbit | – | 0.341 | |
| IL-15 | Human | 0.167 | 0.272 |
| Rabbit | – | 0.259 | |
| IL-18 | Human | 0.266 | 0.358 |
| Rabbit | – | 0.389 |
Figure 2.
ML trees of the ILs studied: (a) IL-1α; (b) IL-1β; (c) IL-2; (d) IL-4; (e) IL-8; (f) IL-10; (g) IL-12A; (h) IL-12B; (i) IL-15; (j) IL-18. Only bootstrap values ≥ 75% are shown. In order to facilitate visualization, only one sequence/allele of each species was used. GenBank and Ensembl accession numbers are indicated for human, rodents and, in some cases, American pika sequences.
Discussion
ILs are important proteins for the innate and adaptive immune responses. In the European rabbit IL-1α, IL-1β, IL-2, IL-4, IL-8, IL-10, IL-12A, IL-12B, IL-15 and IL-18 have been implicated in the immune response against two highly fatal viral diseases, RHD and myxomatosis, and in inflammatory processes. Recently, we described a mutation in the European rabbit and Amami rabbit (Pentalagus furnessi) IL-6 stop codon that leads to a larger protein with a considerable increase of the number of cysteines.55 For the ILs included in this study, no significant alterations were observed in any of the stop codons. Interestingly, American pika IL-1α and IL-2 sequences presented an insertion of four and one aa, respectively, in the site where the others lagomorphs have a stop codon.
Some differences were observed between the European rabbit and the American pika sequences reported here and those already available in public databases (Supplementary Material Table 3). The major differences are in American pika IL-8 and in the European rabbit IL-12B, where both sequences have an insertion of four aa. However, the sequences available in public databases are predicted, which might explain the differences observed. The remaining punctual differences are due to single polymorphisms arising from species diversity.
Alternative splicing is an important step for the production of mature mRNA that leads to protein diversity and may occur by exon skipping (38%); alternative 5’ or 3’ spliced sites (26%); intron retention (3%); mutually exclusive exons, alternative promoters or multiple polyadenylation sites (33%).56,57 Exon skipping occurs by cleavage of specific exon–intron motifs located in the 5’ and 3’ termini regions of the intron. These motifs include, in decreasing order of relevance, GT–AG, CG–AG, AC–AT or AT–AC in the 5’ and 3’ regions, respectively.58–60 Studies in the European rabbit reported alternative splicing in several ILs that leads to novel and functional proteins. Indeed, in IL-2 and IL-7 alternative splicing occurs with exclusion of exon 2;34,61 for IL-4 the presence of two variants (δ2, δ3) results from splicing out of exons 2 and 3, respectively;34 for IL-10 a spliced variant was described with the spliced region occurring between exon 5 and the 3’ UTR region.34 Beside these variants, abnormal transcripts were also characterized for European rabbit IL-4 and IL-10 (IL-4 int2A, IL-4 int3A, IL-4 int3B and IL-10/C).34 Although, more recently Mage and Mage suggested that there is no frameshit in exon 2 of IL-4.62 When comparing American pika IL-8 sequence with the corresponding European rabbit sequence, we observed an insertion of four aa in the 5’ region of exon 2, which is derived from the 3’ region of the preceding intron. This suggests that in American pika the splicing occurred in a CA motif located 12 bp upstream of the motif found in the European rabbit (Figure 1f). These results were further confirmed by analyzing the human and American pika sequences using NetGene2.49,50 Indeed, for human IL-8 the splicing is predicted to occur by NetGene2 at nucleotide position 18 (Figure 1f) (CCCCCAACA|GGTGCAGTTTT; with a 0.77 confidence value), while for American pika the confidence value is of 0.17, with the most likely splicing motif (confidence value of 0.65) being located at nucleotide position 6 (TAATTTTCA|GGTTTCCCCA). Thus, IL-8 from American pika has a different transcript.
Disulfide bonds and glycosylation are important for protein protection against denaturation and proteolytic degradation.63 Disulfide bonds are formed between the thiol groups of cysteine residues and ensure important roles in folding, stability, function and in the regulation of protein activity, being well conserved between species.63–66 In the studied ILs, and with the exception of IL-1β, the cysteine residues are well conserved in lagomorphs. For American pika IL-1β, and when comparing with other lagomorphs, the number and location of five out of nine cysteines residues is different. The analysis of protein secondary structure using PsiPred52,53 and DiANNA54 indicated that the loss of cysteine residues reduces the predicted bonds from four in the European rabbit (Cys35-Cys115; Cys42-Cys186; Cys43-Cys123 and Cys54-Cys81) to two in American pika (Cys108-Cys125 and Cys116-Cys179). These differences lead to the formation of an extra helix (Gly84 to Phe89) in American pika (Figure 3).
Figure 3.
IL-1β PsiPred sequence analysis results for the (a) European rabbit and (b) American pika (the extra helix is marked with a black circle).
The glycosylation process is also important for several functions that include protein folding and interaction with cell surface receptors.67,68 Thus, glycosylation can lead to protein diversity and modulation of protein properties.67,69–71 Putative N-glycosylation sites prediction showed some differences between lagomorphs, in particular for IL-1α, IL-1β, IL-10, IL-12B and IL-15; however, the consequences of these alterations should be considered in future studies.
Animal models are important tools for the study of human diseases. Historically, the European rabbit was the first animal model for immunological studies. Indeed, some of the foundations of molecular immunology were laid with the use of the European rabbit reviewed in Pinheiro et al.72 Nevertheless, over the last decade the European rabbit as a research animal model has been replaced by the mouse model due to its smaller size, lower cost, ease of breeding, etc.73 However, the choice of mouse as the best model to study human diseases has been controversial due to the high variability of results observed between mouse and human, especially when the focus is inflammatory diseases.73–78 A previous study in ILs showed that European rabbit–human sequences are more related than mouse–human or European rabbit–mouse sequences.34 Overall, the results of the phylogenetic analysis are in agreement with the evolutionary topology described for lagomorphs,13 where the American pika is the most divergent species. Nevertheless, the ML trees depicted in Figure 2 correspond to gene trees that do not necessarily reflect the true species tree.79,80 Overall, the human IL sequences appear more closely related to those from European rabbit than with mouse and rat. This pattern had been already observed for other molecular markers and might suggest that the European rabbit might be a more appropriate animal model for the study of immunity in humans.34,81–85
Conclusions
In this study we sequenced and characterized ten ILs for six lagomorph species and in the two European rabbit subspecies. As expected, most differences were observed between leporids and American pika. While these differences may lead to alterations in the biological roles of these proteins, the overall genomic organization, the location of the cysteine residues and the presence of N-glycosylation sites is well conserved. In addition, and according to divergence between the European rabbit and human for the ILs studied, the European rabbit might be a more suitable animal model for studies in the human innate immunity.
Acknowledgements
We would like to thank Jeff Wilcox and Dr. Michael Hamilton from Blue Oak Ranch Reserve, University of California, Berkeley, USA, for providing brush rabbit (Sylvilagus bachmani) tissue samples; and Dr. Jay Storz from School of Biological Sciences, University of Nebraska, USA, for the liver sample of Ochotona princeps. We are grateful to Dr. Janet Rachlow, Dr. Lisette Waits and Dr. Caren Goldberg from Department of Fish and Wildlife Sciences, University of Idaho, USA, for providing pygmy rabbit (Brachylagus idahoensis) tissue samples.
Funding
This work is funded by FEDER funds through the Operational Programme for Competitiveness Factors—COMPETE, and by National Funds through Foundation for Science and Technology (FCT) under projects PTDC/BIA-ANM/3963/2012 and FCOMP-01-0124-FEDER-028286. FCT also supported the doctoral grants of Fabiana Neves (ref.: SFRH/BD/81916/2011) and the FCT Investigator grant of Joana Abrantes (ref.: IF/01396/2013). ‘Genomics Applied To Genetic Resources’, co-financed by North Portugal Regional Operational Programme 2007/2013 (ON.2 – O Novo Norte), under the National Strategic Reference Framework (NSRF), through the European Regional Development Fund (ERDF), also supported this work.
Conflict of interest
The authors declared no potential conflicts of interest with respect to the research, authorship, and/or publication of this article.
References
- 1.Afzal N, Tahir R, Jahan S. Cytokines: an ever expanding area. Biol Biomed Rep 2012; 2: 37–43. [Google Scholar]
- 2.Heinrich PC, Behrmann I, Haan S, et al. Principles of interleukin (IL)-6-type cytokine signalling and its regulation. Biochem J 2003; 374: 1–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Ishihara K, Hirano T. Molecular basis of the cell specificity of cytokine action. Biochim Biophys Acta 2002; 1592: 281–296. [DOI] [PubMed] [Google Scholar]
- 4.Brocker C, Thompson D, Matsumoto A, et al. Evolutionary divergence and functions of the human interleukin (IL) gene family. Hum Genomics 2010; 5: 30–55. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Kaiser P, Rothwell L, Avery S, Balu S. Evolution of the interleukins. Develop Comp Immunol 2004; 28: 375–394. [DOI] [PubMed] [Google Scholar]
- 6.O'Connell MJ, McInerney JO. Gamma chain receptor interleukins: evidence for positive selection driving the evolution of cell-to-cell communicators in the mammalian immune system. J Mol Evol 2005; 61: 608–619. [DOI] [PubMed] [Google Scholar]
- 7.Zelus D, Robinson-Rechavi M, Delacre M, et al. Fast evolution of interleukin-2 in mammals and positive selection in ruminants. J Mol Evol 2000; 51: 234–244. [DOI] [PubMed] [Google Scholar]
- 8.Zhang J, Nei M. Positive selection in the evolution of mammalian interleukin-2 genes. Mol Biol Evol 2000; 17: 1413–1416. [DOI] [PubMed] [Google Scholar]
- 9.Rossi J-F, Lu Z-Y, Jourdan M, Klein B. Interleukin-6 as a therapeutic target. Clin Cancer Res 2015; 21: 1248–1257. [DOI] [PubMed] [Google Scholar]
- 10.He Y, Huang C, Zhang L, et al. Interleukin-21, a potential diagnostic and therapeutic target for systemic lupus erythematosus. Rheumatol Int 2014; 34: 1027–1028. [DOI] [PubMed] [Google Scholar]
- 11.Bessis N, Boissier MC. Novel pro-inflammatory interleukins: potential therapeutic targets in rheumatoid arthritis. Joint Bone Spine 2001; 68: 477–481. [DOI] [PubMed] [Google Scholar]
- 12.Neves F, Abrantes J, Steinke JW, Esteves PJ. Maximum-likelihood approaches reveal signatures of positive selection in IL genes in mammals. Innate Immun 2014; 20: 184–191. [DOI] [PubMed] [Google Scholar]
- 13.Matthee CA, van Vuuren BJ, Bell D, Robinson TJ. A molecular supermatrix of the rabbits and hares (Leporidae) allows for the identification of five intercontinental exchanges during the Miocene. Syst Biol 2004; 53: 433–447. [DOI] [PubMed] [Google Scholar]
- 14.Ge D, Wen Z, Xia L, et al. Evolutionary history of lagomorphs in response to global environmental change. PloS One 2013; 8: e59668–e59668. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Chapman JA and John EC Flux. Introduction to the Lagomorpha. In: C AP, Nuno F and Klaus H (eds) Lagomorph biology: evolution, ecology, and conservation. Berlin: Springer-Verlag, 2008, pp. 1–9.
- 16.Esteves PJ, Lanning D, Ferrand N, et al. The evolution of the immunoglobulin heavy chain variable region (IgVH) in Leporids: an unusual case of transspecies polymorphism. Immunogenetics 2005; 57: 874–882. [DOI] [PubMed] [Google Scholar]
- 17.Pinheiro A, de Mera IG, Alves PC, et al. Sequencing of modern Lepus VDJ genes shows that the usage of VHn genes has been retained in both Oryctolagus and Lepus that diverged 12 million years ago. Immunogenetics 2013; 65: 777–784. [DOI] [PubMed] [Google Scholar]
- 18.van der Loo W, Abrantes J, Esteves PJ. Sharing of endogenous lentiviral gene fragments among leporid lineages separated for more than 12 million years. J Virol 2009; 83: 2386–2388. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Tian J, Hu S, Sun Y, et al. A novel model of atherosclerosis in rabbits using injury to arterial walls induced by ferric chloride as evaluated by optical coherence tomography as well as intravascular ultrasound and histology. J Biomed Biotechnol 2012; 2012: 121867–121867. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Jimenez-Garcia A, Balongo-Garcia R, Alconero FF, et al. Intestinal wall damage in simple ileus in rabbits: immune-modulator role of somatostatin. Hepato-gastroenterology 2004; 51: 1030–1036. [PubMed] [Google Scholar]
- 21.Desando G, Cavallo C, Sartoni F, et al. Intra-articular delivery of adipose derived stromal cells attenuates osteoarthritis progression in an experimental rabbit model. Arthritis Res Ther 2013; 15: R22–R22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Kang SJ, Grossniklaus HE. Rabbit model of retinoblastoma. J Biomed Biotechnol 2011; 2011: 394730–394730. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Woodruff-Pak DS, Agelan A, Del Valle L. A rabbit model of Alzheimer's disease: valid at neuropathological, cognitive, and therapeutic levels. J Alzheimer Dis 2007; 11: 371–383. [DOI] [PubMed] [Google Scholar]
- 24.Marchandeau S, Pontier D, Guitton JS, et al. Early infections by myxoma virus of young rabbits (Oryctolagus cuniculus) protected by maternal antibodies activate their immune system and enhance herd immunity in wild populations. Vet Res 2014; 45: 26–26. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Marques RM, Costa ESA, Aguas AP, et al. Early inflammatory response of young rabbits attending natural resistance to calicivirus (RHDV) infection. Vet Immunol Immunopathol 2012; 150: 181–188. [DOI] [PubMed] [Google Scholar]
- 26.Teixeira L, Marques RM, Aguas AP, Ferreira PG. Regulatory T cells are decreased in acute RHDV lethal infection of adult rabbits. Vet Immunol Immunopathol 2012; 148: 343–347. [DOI] [PubMed] [Google Scholar]
- 27.Muller A, Freitas J, Silva E, et al. Evolution of rabbit haemorrhagic disease virus (RHDV) in the European rabbit (Oryctolagus cuniculus) from the Iberian Peninsula. Vet Microbiol 2009; 135: 368–373. [DOI] [PubMed] [Google Scholar]
- 28.Lopes AM, Correia J, Abrantes J, et al. Is the new variant RHDV replacing genogroup 1 in Portuguese wild rabbit populations? Viruses 2015; 7: 27–36. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Dalton KP, Nicieza I, Abrantes J, et al. Spread of new variant RHDV in domestic rabbits on the Iberian Peninsula. Vet Microbiol 2014; 169: 67–73. [DOI] [PubMed] [Google Scholar]
- 30.Garcia-Bocanegra I, Astorga RJ, Napp S, et al. Myxomatosis in wild rabbit: design of control programs in Mediterranean ecosystems. Prevent Vet Med 2010; 93: 42–50. [DOI] [PubMed] [Google Scholar]
- 31.Delibes-Mateos M, Delibes M, Ferreras P, Villafuerte R. Key role of European rabbits in the conservation of the Western Mediterranean basin hotspot. Conserv Biol 2008; 22: 1106–1117. [DOI] [PubMed] [Google Scholar]
- 32.Abrantes J, Lopes AM, Dalton KP, et al. New variant of rabbit hemorrhagic disease virus, Portugal, 2012–2013. Emerg Infect Dis 2013; 19: 1900–1902. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Kerr P, McFadden G. Immune responses to myxoma virus. Viral Immunol 2002; 15: 229–246. [DOI] [PubMed] [Google Scholar]
- 34.Perkins HD, van Leeuwen BH, Hardy CM, Kerr PJ. The complete cDNA sequences of IL-2, IL-4, IL-6 AND IL-10 from the European rabbit (Oryctolagus cuniculus). Cytokine 2000; 12: 555–565. [DOI] [PubMed] [Google Scholar]
- 35.Stanford MM, Werden SJ, McFadden G. Myxoma virus in the European rabbit: interactions between the virus and its susceptible host. Vet Res 2007; 38: 299–318. [DOI] [PubMed] [Google Scholar]
- 36.Liu J, Wennier S, Reinhard M, et al. Myxoma virus expressing interleukin-15 fails to cause lethal myxomatosis in European rabbits. J Virol 2009; 83: 5933–5938. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Vande Walle L, Lamkanfi M. Inflammasomes: caspase-1-activating platforms with critical roles in host defense. Front Microbiol 2011; 2: 3–3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Johnston JB, McFadden G. Poxvirus immunomodulatory strategies: current perspectives. J Virol 2003; 77: 6093–6100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Johnston JB, McFadden G. Technical knockout: understanding poxvirus pathogenesis by selectively deleting viral immunomodulatory genes. Cell Microbiol 2004; 6: 695–705. [DOI] [PubMed] [Google Scholar]
- 40.Kerr PJ, Perkins HD, Inglis B, et al. Expression of rabbit IL-4 by recombinant myxoma viruses enhances virulence and overcomes genetic resistance to myxomatosis. Virology 2004; 324: 117–128. [DOI] [PubMed] [Google Scholar]
- 41.Boraschi D, Bossu P, Macchia G, et al. Structure-function relationship in the IL-1 family. Front Biosci 1996; 1: d270–d308. [DOI] [PubMed] [Google Scholar]
- 42.Borish LC, Steinke JW. 2. Cytokines and chemokines. J Allergy Clin Immunol 2003; 111: S460–S475. [DOI] [PubMed] [Google Scholar]
- 43.Dinarello CA. Biology of interleukin 1. FASEB J 1988; 2: 108–115. [PubMed] [Google Scholar]
- 44.Schrader JW. Interleukin is as interleukin does. J Immunol Methods 2003; 276: 1–3. [DOI] [PubMed] [Google Scholar]
- 45.Dinarello CA. Interleukin 1 and interleukin 18 as mediators of inflammation and the aging process. Am J Clin Nutr 2006; 83: 447S–455S. [DOI] [PubMed] [Google Scholar]
- 46.Librado P, Rozas J. DnaSP v5: a software for comprehensive analysis of DNA polymorphism data. Bioinformatics 2009; 25: 1451–1452. [DOI] [PubMed] [Google Scholar]
- 47.Edgar RC. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res 2004; 32: 1792–1797. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Blom N, Sicheritz-Pontèn T, Gupta R, et al. Prediction of post-translational glycosylation and phosphorylation of proteins from the amino acid sequence. Proteomics 2004; 4(6): 1633–1649. [DOI] [PubMed]
- 49.Hebsgaard SM, Korning PG, Tolstrup N, et al. Splice site prediction in Arabidopsis thaliana pre-mRNA by combining local and global sequence information. Nucleic Acids Res 1996; 24: 3439–3452. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Brunak S, Engelbrecht J, Knudsen S. Prediction of human mRNA donor and acceptor sites from the DNA sequence. J Mol Biol 1991; 220: 49–65. [DOI] [PubMed] [Google Scholar]
- 51.Tamura K, Stecher G, Peterson D, et al. MEGA6: Molecular Evolutionary Genetics Analysis version 6.0. Mol Biol Evol 2013; 30: 2725–2729. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Buchan DW, Minneci F, Nugent TC, et al. Scalable web services for the PSIPRED Protein Analysis Workbench. Nucleic Acids Research 2013; 41: W349–W357. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Jones DT. Protein secondary structure prediction based on position-specific scoring matrices. J Mol Biol 1999; 292: 195–202. [DOI] [PubMed] [Google Scholar]
- 54.Ferre F, Clote P. DiANNA 1.1: an extension of the DiANNA web server for ternary cysteine classification. Nucleic Acids Res 2006; 34: W182–W185. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Neves F, Abrantes J, Pinheiro A, et al. Convergent evolution of IL-6 in two leporids (Oryctolagus and Pentalagus) originated an extended protein. Immunogenetics 2014; 66: 589–595. [DOI] [PubMed] [Google Scholar]
- 56.Sahoo A, Im SH. Interleukin and interleukin receptor diversity: role of alternative splicing. Int Rev Immunol 2010; 29: 77–109. [DOI] [PubMed] [Google Scholar]
- 57.Keren H, Lev-Maor G, Ast G. Alternative splicing and evolution: diversification, exon definition and function. Nat Rev Genetics 2010; 11: 345–355. [DOI] [PubMed] [Google Scholar]
- 58.Holste D, Ohler U. Strategies for identifying RNA splicing regulatory motifs and predicting alternative splicing events. PLoS Comput Biol 2008; 4: e21–e21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Wu Q, Krainer AR. AT-AC pre-mRNA splicing mechanisms and conservation of minor introns in voltage-gated ion channel genes. Mol Cell Biol 1999; 19: 3225–3236. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Shimada MK, Hayakawa Y, Takeda J, et al. A comprehensive survey of human polymorphisms at conserved splice dinucleotides and its evolutionary relationship with alternative splicing. BMC Evol Biol 2010; 10: 122–122. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Siewe BT, Kalis SL, Esteves PJ, et al. A novel functional rabbit IL-7 isoform. Develop Comp Immunol 2010; 34: 828–836. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Mage RG, Mage MG. Sequence of rabbit (Oryctolagus cuniculus) DNA from the OryCun2.0 donor does not confirm a frameshift in exon 2 of IL4. Immunol Immunogenet Insights 2012; 2012: 1–5. [Google Scholar]
- 63.Bulaj G. Formation of disulfide bonds in proteins and peptides. Biotechnol Adv 2005; 23: 87–92. [DOI] [PubMed] [Google Scholar]
- 64.Fass D. Disulfide bonding in protein biophysics. Annu Rev Biophys 2012; 41: 63–79. [DOI] [PubMed] [Google Scholar]
- 65.Li XQ, Zhang T, Donnelly D. Selective loss of cysteine residues and disulphide bonds in a potato proteinase inhibitor II family. PloS One 2011; 6: e18615–e18615. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Wong JW, Ho SY, Hogg PJ. Disulfide bond acquisition through eukaryotic protein evolution. Mol Biol Evol 2011; 28: 327–334. [DOI] [PubMed] [Google Scholar]
- 67.Chamorey AL, Magne N, Pivot X, Milano G. Impact of glycosylation on the effect of cytokines. A special focus on oncology. Eur Cytokine Netw 2002; 13: 154–160. [PubMed] [Google Scholar]
- 68.Schwarz F, Aebi M. Mechanisms and principles of N-linked protein glycosylation. Curr Opin Struct Biol 2011; 21: 576–582. [DOI] [PubMed] [Google Scholar]
- 69.Helenius A, Aebi M. Roles of N-linked glycans in the endoplasmic reticulum. Annu Rev Biochem 2004; 73: 1019–1049. [DOI] [PubMed] [Google Scholar]
- 70.Rudd PM, Elliott T, Cresswell P, et al. Glycosylation and the immune system. Science 2001; 291: 2370–2376. [DOI] [PubMed] [Google Scholar]
- 71.Shental-Bechor D, Levy Y. Effect of glycosylation on protein folding: a close look at thermodynamic stabilization. Proc Natl Acad Sci USA 2008; 105: 8256–8261. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Pinheiro A, Lanning D, Alves PC, et al. Molecular bases of genetic diversity and evolution of the immunoglobulin heavy chain variable region (IGHV) gene locus in leporids. Immunogenetics 2011; 63: 397–408. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Webb DR. Animal models of human disease: inflammation. Biochem Pharmacol 2014; 87: 121–130. [DOI] [PubMed] [Google Scholar]
- 74.Burkhardt AM, Zlotnik A. Translating translational research: mouse models of human disease. Cell Mol Immunol 2013; 10: 373–374. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Mullane K, Williams M. Animal models of asthma: reprise or reboot? Biochem Pharmacol 2014; 87: 131–139. [DOI] [PubMed] [Google Scholar]
- 76.Seok J, Warren HS, Cuenca AG, et al. Genomic responses in mouse models poorly mimic human inflammatory diseases. Proc Natl Acad Sci USA 2013; 110: 3507–3512. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Takao K, Miyakawa T. Genomic responses in mouse models greatly mimic human inflammatory diseases. Proc Natl Acad Sci U S A 2015; 112: 1167–1172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Shay T, Lederer JA, Benoist C. Genomic responses to inflammation in mouse models mimic humans: We concur, apples to oranges comparisons won't do. Proc Natl Acad Sci USA 2015; 112: E346–E346. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Nichols R. Gene trees and species trees are not the same. Trends Ecol Evol 2001; 16: 358–364. [DOI] [PubMed] [Google Scholar]
- 80.Nakhleh L, Ruths D and Innan H. Gene trees, species trees, and species networks. In: Guerra R and Goldstein D (eds) Meta-analysis and Combining Information in Genetics and Genomics. Boca Raton, FL: CRC Press, 2009, pp. 275–293.
- 81.Vaure C, Liu Y. A comparative review of toll-like receptor 4 expression and functionality in different animal species. Front Immunol 2014; 5: 316–316. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Koch E, Hue-Beauvais C, Galio L, et al. Leptin gene in rabbit: cloning and expression in mammary epithelial cells during pregnancy and lactation. Physiol Genomics 2013; 45: 645–652. [DOI] [PubMed] [Google Scholar]
- 83.Fischer B, Chavatte-Palmer P, Viebahn C, et al. Rabbit as a reproductive model for human health. Reproduction 2012; 144: 1–10. [DOI] [PubMed] [Google Scholar]
- 84.Vuillaumier S, Kaltenboeck B, Lecointre G, et al. Phylogenetic analysis of cystic fibrosis transmembrane conductance regulator gene in mammalian species argues for the development of a rabbit model for cystic fibrosis. Mol Biol Evol 1997; 14: 372–380. [DOI] [PubMed] [Google Scholar]
- 85.Wang Y, Fan N, Song J, et al. Generation of knockout rabbits using transcription activator-like effector nucleases. Cell Regen (Lond) 2014; 3: 3–3. [DOI] [PMC free article] [PubMed] [Google Scholar]






