Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Phyletic Distribution of Fatty Acid-Binding Protein Genes

  • Yadong Zheng,

    Affiliations School of Biology, University of Nottingham, Nottingham, United Kingdom, State Key Laboratory of Veterinary Etiological Biology, Key Laboratory of Veterinary Parasitology of Gansu Province, Lanzhou Veterinary Research Institute, CAAS, Lanzhou, Gansu, China

  • David Blair,

    Affiliation James Cook University, Townsville, Queensland, Australia

  • Janette E. Bradley

    jan.bradley@nottingham.ac.uk

    Affiliation School of Biology, University of Nottingham, Nottingham, United Kingdom

Abstract

Fatty acid-binding proteins (FABPs) are a family of fatty acid-binding small proteins essential for lipid trafficking, energy storage and gene regulation. Although they have 20 to 70% amino acid sequence identity, these proteins share a conserved tertiary structure comprised of ten beta sheets and two alpha helixes. Availability of the complete genomes of 34 invertebrates, together with transcriptomes and ESTs, allowed us to systematically investigate the gene structure and alternative splicing of FABP genes over a wide range of phyla. Only in genomes of two cnidarian species could FABP genes not be identified. The genomic loci for FABP genes were diverse and their genomic structure varied. In particular, the intronless FABP genes, in most of which the key residues involved in fatty acid binding varied, were common in five phyla. Interestingly, several species including one trematode, one nematode and four arthropods generated FABP mRNA variants via alternative splicing. These results demonstrate that both gene duplication and post-transcriptional modifications are used to generate diverse FABPs in species studied.

Introduction

Lipids are a very important subclass of constituents in the maintenance of normal physiology in organisms and a delicate balance of these hydrophobic molecules is partially regulated by fatty acid-binding proteins (FABPs). These small proteins of approximately 15 kDa execute fatty acid transport and, together with intracellular retinol- and retinoic acid-binding proteins, comprise a subfamily of intracellular lipid binding proteins (iLBPs) that are extensively present in animals. Ancestral iLBP genes are supposed to have arisen after separation of animals from fungi and plants [1]. FABPs are absent from archaebacteria and yeast [2,3]. Multiple gene duplications have occurred in this subfamily, giving rise to 16 iLBPs including 12 FABPs in vertebrates [1,46]. More than 30 FABP genes have been found in a wide range of invertebrates [2,7,8].

Mammalian FABP genes generally consist of four exons and some are dispersed on a single chromosome in humans, rats and mice [1,9]. The few studies on invertebrates show considerable variation in genomic organization of FABP genes, in aspects of size, exon and intron numbers [1,2]. For example, Caenorhabditis elegans expresses nine FABPs, also known as lipid binding proteins (LBP), and these mostly reside on different chromosomes. However, LBP-5 and LBP-6 are comprised of two exons and one intron and are positioned on chromosome I, suggesting that they might have arisen from tandem gene duplication.

Although FABPs share 20 to 70% identity at the amino acid level across and within invertebrate species, their tertiary structures are highly conserved, characterized by a cavity, formed by ten anti-parallel sheets and two helixes, that accommodates lipophilic compound(s), including fatty acids [2]. With a few exceptions, the residues related to ligand binding appear to be conserved in both invertebrate and vertebrate FABPs [9]. In the β-barrel cavity, the bound fatty acid(s) interacts with some residues Arg…Arg-x-Tyr, the so-called P2 motif. Moreover, Phe residues on the first helix and Ala/Pro-Asp in the turn between βE and βF are also critical for binding affinity in FABPs [10].

The systematic and genome-wide investigation of invertebrate FABP genes remains in its infancy. With availability of the complete genomes and transcriptome data for an increasing number of species, it is feasible to explore their genomic organization and post-transcriptional splicing paradigms. We have investigated gene organization and post-transcriptional modification of FABPs across 34 invertebrate species from 8 phyla (including lower chordates). Additionally, we have shown that an increase in gene copy numbers followed by divergence, as well as alternative splicing, are likely to be the mechanisms responsible for functional expansion and diversity of FABPs in invertebrate species.

Materials and Methods

Identification and annotation of FABP genes

In this study, most of 34 invertebrates have annotated genomes and FABP genes were directly retrieved from the databases. For the species without an annotated genome including Echinococcus multilocularis, Echinococcus granulosus, Heterorhabditis bacteriophoraTrichinella spiralisStrongyloides rattiRhodnius prolixus, Haemonchus contortus and Ciona savignyi, we searched the databases using the following strategies. Candidate FABP genes were identified using TBlastN, with experimentally or putatively identified FABP gene(s) from a closely related species as a query sequence, to search various genome databases with a cut-off e-value of 1-e10 (Table 1). Otherwise, Schistosoma mansoni FABPs (Smp_095360 and Smp_046800) or C. elegans FABPs (NP_505016, NP_508558, NP_508557, NP_491928, NP_506440, NP_491926, NP_001041249, NP_506444 and NP_001033511) were used as queries. This strategy was used because FABP genes share 20 to 70% similarity at the amino acid level. We then applied two criteria to resulting “hits” to identify FABP genes. First, considering that most known FABPs are ~130 amino acids (aa) in length, we arbitrarily set the size range of FABPs from 80 to 180aa (130 ± 50aa). In addition, the sequences within the size limit were used for secondary structure prediction and those with the putatively typical structural elements were considered to be FABP genes.

Species for which genome databases were searchedNum. loci found in genome draftsLengthaEvidencebAlternative splicingData originc
Cnidaria
Nematostella vectensis////JGI
Hydra magnipapillata////Metazome
Placozoa
Trichoplax adhaerens5120~1781/5NoJGI NCBI
Annelida
Capitella teleta7135~1677/7NoJGI NCBI
Helobdella robusta3119~1433/3NoJGI NCBI
Mollusca
Lottia gigantea7132~1637/7NoJGI NCBI
Platyhelminthes
Schmidtea mediterranea3123~1682/3NoSmedGD NCBI
Schistosoma mansoni2132, 1332/2YesGeneDB NCBI
Schistosoma japonicum11301/1NoGeneDB NCBI
Echinococcus granulosus5124~1432/5NoNCBI Sanger
Echinococcus multilocularis5124~1434/4NoSanger
Nematoda
Caenorhabditis elegans9135~1659/9YesNCBI
Pristionchus pacificus4118~1634/4NoNCBI WormBase WUGSC
Heterorhabditis bacteriophora3133~1643/3NoNCBI WUGSC
Trichinella spiralis3133~1433/3NoNCBI WUGSC
Haemonchus contortus0d133~1644/4NoSanger NCBI
Strongyloides ratti4132~1654/4NoSanger WormBase
Brugia malayi3130~1803/3NoNCBI
Arthropod
Daphnia pulex2130, 1312/2NowFleaBase NCBI
Pediculus humanus corporis3132~1350/3NoNCBI VectorBase VectorBaseFlyBase
Bombyx mori595~1424/5NoSilkDB
Tribolium castaneum11361/1YesNCBI
Nasonia vitripennis21322/2NoNCBI
Acyrthosiphon pisum3135, 1363/3YesNCBI
Apis mellifera2132, 1332/2YesNCBI
Drosophila melanogaster11301/1YesNCBI FlyBase
Anopheles gambiae21311/1NoVectorBase NCBI
Aedes aegypti11321/1NoNCBI
Culex pipiens quinquefasciatus11321/1NoNCBI
Rhodnius prolixus11341/1NoNCBI VectorBase
Echinodermata
Strongylocentrotus purpuratus21302/2NoNCBI JGI
Chordata
Branchiostoma floridae15135~1517/14NoJGI NCBI
Ciona savignyi3127~1333/3NoBroad NCBI
Saccoglossus kowalevskii3132~1383/3NoBaylor NCBI Metazome

Table 1. Distribution and features of FABP genes in invertebrates.

aNumber of amino acid residues;
bThe number of putative FABP transcript variants (the numbers after ‘/’) and the number of the variants for which expression was validated by transcriptomic or/and EST data or/and cDNA cloning (the numbers before ‘/’);
cJGI: Joint Genome Institute; NCBI: National Centre for Biotechnology Information; SmedGD: Schmidtea mediterranea Genome Database; Sanger: Wellcome Trust Sanger Institute; WUGSC: Washington University Genome Sequencing Centre; wFleaBase: Daphnia Water Flea Genome Database; FlyBase: Drosophila database; SilkDB: silkworm database; Broad: Broad Institute; Baylor: Baylor College of Medicine;
dNo genomic loci for FBAPs were found using Blast with its EST sequences.
CSV
Download CSV

Two sequential approaches were utilized for determination of the exons and exon boundaries. Firstly the exons and their boundaries were determined from TBlastN outcomes as highly-scored segment pairs or gaps within the segment pairs as described previously [11]. FABP gene structural models were then verified and finely modified using transcriptome data or expression sequence tags (ESTs). Segment pairs that dispersed over two or more supercontigs were not considered to build gene models in this study. The intron-exon boundaries were manually checked based on consensus splicing acceptor and donor sites and they conformed to the GT/AG rule.

FABP genes, identified using the approach above, were used as query sequences to search transcriptome and EST databases for the relevant species. This provided a means of validating the findings from genomic data alone (Table 1).

Sequence alignments and secondary structure prediction

The FABP protein sequences were aligned using Clustal W algorithm (MEGA 4.0) with default parameters [12] and then manually checked (Figure S1). The secondary structures of FABPs were predicted using Psipred [13].

Construction of a phylogenetic tree

Besides all the FABP amino acid sequences identified in this study, ten human FABP sequences were also included for phylogenetic analysis. Prior to tree construction, a best model was selected using TOPALi v2.5 [14]. A Bayesian tree was built using the following settings: WAG model plus gamma, 2 runs, 500,000 generations, 10 of sample frequency and 25% burn in. To confirm the topology of the tree, a ML tree was also built using the following settings: LG model plus gamma with 100 bootstraps.

Results

Identification and annotation of FABP genes across invertebrates

During sequence searching we obtained high-scoring hits that encoded more than 180aa or fewer than 80aa, but all of which were excluded from further analyses in this study. For instance, a Branchiostoma floridae hypothetical protein (987aa, XP_002589099) contained a region at the C terminal that shared 96% identity with Branchiostoma belcheri FABP (136aa, ADD10136).

In total, 107 sequences falling within the specified size range and exhibiting appropriate secondary structure were collected from 32 invertebrate species including one placozoan, two annelids, one mollusc, five platyhelminths, seven nematodes, twelve arthropods, one echinoderm and three chordates (Table 1 and Supplementary text file). The identity of these putative FABP amino acid sequences ranged from 29.0% to 99.3% and they were predicted to have the typical tertiary structure (Figure S2). No homologues of FABPs were identified in two Cnidaria species, Hydra magnipapillata and Nematostella vectensis. Notably, four Haemonchus contortus FABP genes were identified by TBlastN searches against transcriptome (NCBI) but none of them was found in the genome, possibly due to incomplete genomic data (Sanger). One putative FABP transcript was derived from transcriptome or EST data, but its locus was not found in the genome of each of the following species: Helobdella robusta, Lottia gigantea, Schistosoma japonicum, Heterorhabditis bacteriophora and Saccoglossus kowalevskii. With the exception of the body louse, Pediculus humanus corporis, some or all FABP genes found in genomes were validated by EST or transcriptomic data.

Numbers of genomic loci for FABPs ranged from one (several arthropods and S. japonicum) to fifteen (the chordate, B. floridae) in invertebrate genomes (Table 1). Echinococcus multilocularis, Anopheles gambiae and B. floridae each had two distinct loci that encoded identical FABPs at the amino acid level. The introns of the two E. multilocularis FABPs were identical, whilst those of the A. gambiae and B. floridae FABPs were different in size and sequence. But there is not enough evidence to support that these FABP genes are transcribed into the same mRNAs.

Phylogenetic analysis of FABPs

As shown in the Bayesian tree (Figure 1), nematode FABPs formed two distant clades and with an exception of T. spiralis, each clade was comprised of all the nematode species, suggesting that the FABP genes in nematodes may have evolved from different origins. Except S. mediterranea, the phylogenetic relationship within Platyhelminth species was clearly resolved. The subclades comprised of E. multilocularis and E. granulosus demonstrate that both parasites have a similar gene set for FABP, possibly descendent from their common ancestor.

thumbnail
Figure 1. A Bayesian tree of FABPs.

Bayesian probabilities more than 0.8 were shown at nodes. Tadh: Trichoplax adhaerens; Ctel: Capitella teleta; Hrob: Helobdella robusta; Lgig: Lottia gigantean; Smed: Schmidtea mediterranea; Sman: Schistosoma mansoni; Sjap: Schistosoma japonicum; Egra: Echinococcus granulosus; Emul: Echinococcus multilocularis; Cele: Caenorhabditis elegans; Ppac: Pristionchus pacificus; Hbac: Heterorhabditis bacteriophora; Tspi: Trichinella spiralis; Srat: Strongyloides ratti; Bmal: Brugia malayi; Dpul: Daphnia pulex; Phum: Pediculus humanus corporis; Bmor: Bombyx mori; Tcas: Tribolium castaneum; Nvit: Nasonia vitripennis; Apis: Acyrthosiphon pisum; Amel: Apis mellifera; Dmel: Drosophila melanogaster; Agam: Anopheles gambiae; Aaeg: Aedes aegypti; Cpip: Culex pipiens quinquefasciatus; Rpro: Rhodnius prolixus; Spur: Strongylocentrotus purpuratus; Bflo: Branchiostoma floridae; Csav: Ciona savignyi; Skow: Saccoglossus kowalevskii; Has: Homo sapiens Note: to make it simpler, ‘FABP’ was omitted in every branch name. For example: Tadh1 refers to Tadh_FABP1, Tadh2 to Tadh_FABP2 and so forth.

https://doi.org/10.1371/journal.pone.0077636.g001

Extraordinary gene expansion was observed in amphioxus, B. floridae, via gene duplications. Moreover, the phylogenetic analysis revealed that the current gene set might have resulted from multiple rounds of duplications and divergence during evolution, especially Bflo11 paralogues, and that these duplication events might have occurred recently (Figure 1). Essentially, a ML tree showed a similar topology to the Bayesian tree (Figure S3).

Diversity of FABP gene structures across invertebrates

Although intronless FABP pseudogenes have been described in several species including humans [1517], all functional mammalian FABP genes exhibit similar genomic organization, containing four exons and three introns [1]. An analysis of FABP gene organization revealed diversity in invertebrates, especially in Platyhelminthes and Nematoda, although the canonical organization (four exons) predominated. A six-exon five-intron structure for FABP was only found in the early-branching invertebrate Trichoplax adhaerens. FABP genes comprised of five exons and four introns were found in placozoans, molluscs, platyhelminths and nematodes (Table 2).

SpeciesaNumber of exons
654321
Placozoa
Trichoplax adhaerens113
Annelida
Capitella teleta7
Helobdella robusta2
Mollusca
Lottia gigantea17
Platyhelminthes
Schmidtea mediterranea12
Schistosoma mansoni2
Schistosoma japonicum1
Echinococcus granulosus32
Echinococcus multilocularis3b2
Nematoda
Caenorhabditis elegans144
Pristionchus pacificus12
Heterorhabditis bacteriophora21
Trichinella spiralis3
Strongyloides ratti13
Brugia malayi21
Arthropod
Daphnia pulex11
Pediculus humanus corporis12
Bombyx mori41
Tribolium castaneum1
Nasonia vitripennis2
Acyrthosiphon pisum3
Apis mellifera11
Drosophila melanogaster1
Anopheles gambiae2b
Aedes aegypti1
Culex pipiens quinquefasciatus1
Rhodnius prolixus1
Echinodermata
Strongylocentrotus purpuratus11
Chordata
Branchiostoma floridae12b3
Ciona savignyi21
Saccoglossus kowalevskii13
Total1652221614

Table 2. FABP genomic structures in invertebrates.

aIn total thirty-four species from eight phyla were included in this study;
bEach of these species has two different loci that encode identical FABPs at the amino acid level.
CSV
Download CSV

Intronless FABP genes

Single-exon FABP genes were found in the following species: Echinococcus granulosus, E. multilocularis, Strongyloides ratti, P. humanus corporis, Rhodnius prolixus, Strongylocentrotus purpuratus and S. kowalevskii (Table 2). Expression of most of these intronless genes was confirmed either by transcriptome analysis or analysis of ESTs. With the exception of R. prolixus, the species encoding intronless FABPs also encoded other FABP genes with two or more exons. The intronless FABP gene architecture dominated in three other metazoans Strongyloides ratti, P. humanus corporis and Saccoglossus kowalevskii.

Alignment of the intronless FABPs has revealed that the R. prolixus FABP contained intact key residues which are important in defining fatty acid binding [10], while absence or alterations in these sites occurred in the others (Figure S4). This suggests that R. prolixus FABP has capacity to bind to lipids, but the others may no longer be able to do so. Alternatively, they may have different binding spectra for fatty acids in comparison with those that have been characterised.

Alternative splicing in FABP genes

Alternative splicing is an important post-transcriptional modification in eukaryotic pre-mRNAs, accounting for the complexity and variety of proteomes. FABP genes underwent alternative splicing to generate isoforms in several invertebrates including S. mansoni, C. elegans, T. castaneum, Acyrthosiphon pisum, Apis mellifera and Drosophila melanogaster. Furthermore, all the transcripts derived from alternative splicing were confirmed by transcriptomic data.

Comparative analysis of FABP gene structure and transcripts revealed that FABP pre-mRNAs were alternatively spliced in different patterns (Figure 2). In S. mansoni, four FABP variants were produced via exon skipping. Moreover, these variants were differentially expressed at different developmental stages (http://www.genedb.org/Homepage/Smansoni), suggesting they have distinct roles. The arthropods Acyrthosiphon pisum (3 FABP genes) and Apis mellifera (2 FABP genes) utilized the same approach to yield four and three different FABP transcripts, respectively. In contrast to the exon exclusion mechanism seen in S. mansoni, spliced leader trans-splicing (SL trans-splicing), a type of alternative splicing whereby a spliced leader serves as a mini-exon to be added onto 5’ pre-mRNA ends [18], was used in C. elegans FABP mRNA precursors (Figure 2).

thumbnail
Figure 2. Alternative splicing in invertebrate FABP genes.

Typical alternative splicing patterns in C. elegans, T. castaneum and D. melanogaster are represented. In C. elegans, LBP-9 pre-mRNA is spliced to generate two variants 9a and 9b by addition of a short spliced leader sequence (SL1: 5’-GGTTTAATTACCCAAGTTTGAG-3’) at the 5’ end. Blank or filled boxes and straight lines represent exons and introns, respectively, and a poly (A) stretch present in each FABP cDNA clone or EST sequence is directly shown. In each group, an annotated FABP gene is placed above the variants that are indicated by a, b, c or/and d. Numbers above the boxes and under the lines show the sizes of corresponding exons and introns, respectively. The sizes of exons, where these differ, are indicated above the corresponding exons in the spliced transcripts. The length of variants is also shown after each transcript and the number of ESTs is shown in the brackets.

https://doi.org/10.1371/journal.pone.0077636.g002

Both T. castaneum and D. melanogaster genomes contained only one FABP gene locus but variants were found in transcriptome or EST datasets. Mapping of these transcripts revealed that another alternative splicing mechanism, intron retention, was involved in the post-transcriptional splicing of FABP genes together with exon skipping (Figure 2). Noticeably, the last exons of the FABP genes were retained during alternative splicing in all the invertebrates studied except D. melanogaster that also used the partial sequence of the second intron as the last exon to generate FABP isoforms.

Discussion

In this study, the number of the FABP genomic loci identified was remarkably variable, from 1 in several invertebrates to 15 in B. floridae. Interestingly, no FABP loci were identified in genomes of the cnidarians H. magnipapillata and N. vectensis, yet they were present in the simplest known free-living metazoan, T. adhaerens, which is considered as a basal metazoan [19,20]. Consistent with the gene structure, T. adhaerens FABP genes seem to be prototypes of this family (most loci exhibit the “canonical” four-exon structure). The lack of FABP expression in cnidarians may be explained by gene loss but the possibility remains that these species may express extremely heterogeneous FABPs and investigations of more cnidarians are required.

Also of interest is the finding that each of three invertebrate genomes contained two loci to encode the same proteins, suggesting that they might have arisen from recent gene duplication. A phylogenetic analysis suggested that the current gene set in E. multilocularis may have been generated before the speciation of Echinococcus species (Figure 1), supporting the idea that FABP gene duplication may have occurred in their common ancestor. This finding does not fully support the previous assumption that E. granulosus FABPs 2 and 4 arose from a recent duplication event [21].

The ancestral FABP gene might have evolved from a lipocalin gene and have undergone the first duplication approximately 930 million years ago with subsequent duplications and divergence [1,22]. The FABP sequences annotated in this study were heterogeneous with regard to length, composition and identity, possibly driven by the need to transport numerous different fatty acids [1]. In contrast to fifteen FABP copies in the lower chordate B. floridae, arthropods often possessed only a single FABP locus. This was the situation in two species of mosquito, Aedes aegypti and C. pipiens quinquefasciatus. However, a malaria mosquito, Anopheles gambiae, had two copies of FABP genes that resided on the same scaffold. It is noteworthy that mosquito FABPs have been annotated as allergens (XP_001657349 for Aedes aegypti FABP; XP_001864031 for C. pipiens quinquefasciatus FABP) [23]. Although no evidence is to date available, it is possible that mosquito FABPs act as allergens. FABPs from mites [24,25] and other lipid-binding proteins from nematodes [2628] have been shown to be allergic. Secondary structural prediction with high confidence showed that all of these mosquito allergens had two alpha helixes and ten beta sheets typical of FABP structural elements (data not shown). In addition to the conserved secondary structures, they contained fatty acid binding-related key residues except Val-Asp instead of Pro-Asp (Figure S4). We therefore propose that these allergens in mosquitoes are functional FABPs.

In contrast to relatively uniform genomic structures for mammalian FABP genes, invertebrate FABP genes were organized in a wide range of patterns with a dominance of the four-exon three-intron structure. This study indicates that invertebrate FABP genes may have tended towards loss of introns during evolution. This idea is enhanced by the fact that most of the invertebrate FABPs investigated have matched intron positions [2]. Compared to the early branching invertebrate, T. adhaerens, FABP genes from cestodes and mosquitoes were intron-poor. Our findings strongly argue against the speculation that the first and second introns of FABP genes might have evolved later [21]. Surprisingly, a number of invertebrates encoded intronless FABPs with most of the key residues that participate in lipid binding being altered. Here no evidence was obtained to suggest that these FABPs remain able to bind to lipids. However, with the exception of E. granulosus and P. humanus corporis, transcription of intronless FABPs in other species was verified by transcriptomic or/and EST data, suggesting that they are functional. Such intronless FABPs have also been reported in several mammals where they may have lost their capacity to bind lipid ligands although it has not been fully established if they are transcribed [1517].

A wealth of data has revealed that numerous introns were present generally in early multicellular organisms and alterations of intron positions occurred at a very low frequency during evolution [29]. Several mechanisms have been proposed for intron gain or loss [30,31]. In comparison with a canonical three-intron structure, the first intron (17/31) was more likely to be preferentially retained in two- or one-intron FABP genes in invertebrates. This suggests that reverse transcription followed by gene conversion may have been involved in the FABP intron loss as this mechanism tends to remove 3’ introns from genes [30]. An analysis of 684 gene introns from eight organisms has showed that loss of most ancestral introns has occurred in worms and arthropods but not in humans [32]. This result may give us some clues, but the selective forces that have driven intron loss in platyhelminths remain unclear.

Alternative splicing, a substantial mechanism for the modification of pre-mRNA, exists in nearly all eukaryotic organisms and accounts for the complexity and diversity of protein functions. In contrast to mammals, where alternative splicing of FABP genes has rarely been observed, FABP genes in some invertebrates were alternatively spliced, leading to generation of FABP variants. In particular, these various transcripts were produced by different splicing patterns. In C. elegans, only FABP genes 5, 6 and 9 were confirmed to mature by means of SL trans-splicing using spliced leader 1 (SL1). There are two distinct spliced leader sequences in C. elegans, SL1 and SL2, and the former is used to generate mainly monocistronic pre-mRNA [33]. It is estimated that approximately 70% of all genes in this free-living nematode are post-transcriptionally modified by this mechanism [34]. It is still not clear why C. elegans FABP 1, 2, 3 and 8 pre-mRNAs are not matured via SL trans-splicing. Although the SL trans-splicing mechanism is also extensively present in the Phyla Cnidaria, Platyhelminthes and Chordata [18], it was not observed in FABP transcripts in other invertebrates collected in this study. These results suggest that the SL trans-splicing modification in FABP transcripts may have been acquired during evolution of C. elegans.

Supporting Information

Figure S1.

Alignment of FABP amino acid sequences. Tadh: Trichoplax adhaerens; Ctel: Capitella teleta; Hrob: Helobdella robusta; Lgig: Lottia gigantean; Smed: Schmidtea mediterranea; Sman: Schistosoma mansoni; Sjap: Schistosoma japonicum; Egra: Echinococcus granulosus; Emul: Echinococcus multilocularis; Cele: Caenorhabditis elegans; Ppac: Pristionchus pacificus; Hbac: Heterorhabditis bacteriophora; Tspi: Trichinella spiralis; Srat: Strongyloides ratti; Bmal: Brugia malayi; Dpul: Daphnia pulex; Phum: Pediculus humanus corporis; Bmor: Bombyx mori; Tcas: Tribolium castaneum; Nvit: Nasonia vitripennis; Apis: Acyrthosiphon pisum; Amel: Apis mellifera; Dmel: Drosophila melanogaster; Agam: Anopheles gambiae; Aaeg: Aedes aegypti; Cpip: Culex pipiens quinquefasciatus; Rpro: Rhodnius prolixus; Spur: Strongylocentrotus purpuratus; Bflo: Branchiostoma floridae; Csav: Ciona savignyi; Skow: Saccoglossus kowalevskii; Has: Homo sapiens Note: to make it simpler, ‘FABP’ was omitted in every branch name. For example: Tadh1 refers to Tadh_FABP1, Tadh2 to Tadh_FABP2 and so forth.

https://doi.org/10.1371/journal.pone.0077636.s001

(TIF)

Figure S2.

Tertiary structure of Emul_FABP3 predicted using Phyre.

https://doi.org/10.1371/journal.pone.0077636.s002

(TIF)

Figure S3.

A ML tree of FABPs. Bootstrap values more than 60 were shown at nodes. Tadh: Trichoplax adhaerens; Ctel: Capitella teleta; Hrob: Helobdella robusta; Lgig: Lottia gigantean; Smed: Schmidtea mediterranea; Sman: Schistosoma mansoni; Sjap: Schistosoma japonicum; Egra: Echinococcus granulosus; Emul: Echinococcus multilocularis; Cele: Caenorhabditis elegans; Ppac: Pristionchus pacificus; Hbac: Heterorhabditis bacteriophora; Tspi: Trichinella spiralis; Srat: Strongyloides ratti; Bmal: Brugia malayi; Dpul: Daphnia pulex; Phum: Pediculus humanus corporis; Bmor: Bombyx mori; Tcas: Tribolium castaneum; Nvit: Nasonia vitripennis; Apis: Acyrthosiphon pisum; Amel: Apis mellifera; Dmel: Drosophila melanogaster; Agam: Anopheles gambiae; Aaeg: Aedes aegypti; Cpip: Culex pipiens quinquefasciatus; Rpro: Rhodnius prolixus; Spur: Strongylocentrotus purpuratus; Bflo: Branchiostoma floridae; Csav: Ciona savignyi; Skow: Saccoglossus kowalevskii; Has: Homo sapiens Note: to make it simpler, ‘FABP’ was omitted in every branch name. For example: Tadh1 refers to Tadh_FABP1, Tadh2 to Tadh_FABP2 and so forth.

https://doi.org/10.1371/journal.pone.0077636.s003

(TIF)

Figure S4.

Fatty acid binding-related residues in intronless FABP genes of invertebrates. Invertebrate intronless FABP amino acid sequences were aligned using Clustal W. The amino acids identical to the consensus are shown as dots and alignment gaps are indicated with dashes (-). Numbers above the alignment represent positions of amino acids. The key amino acids responsible for interactions with lipid ligands are directly indicated beneath the alignment. Rpro, Rhodnius prolixus; Egra, Echinococcus granulosus; Emul, E. multilocularis; Srat, Strongyloides ratti; Phum, Pediculus humanus corporis; Spur, Strongylocentrotus purpuratus; Skow, Saccoglossus kowalevskii. Note: to make it simpler, ‘FABP’ was omitted in every branch name. For example: Tadh1 refers to Tadh_FABP1, Tadh2 to Tadh_FABP2 and so forth.

https://doi.org/10.1371/journal.pone.0077636.s004

(TIF)

File S1.

Supplementary text file. Putative amino acid sequences of FABP genes.

https://doi.org/10.1371/journal.pone.0077636.s005

(TXT)

Acknowledgments

The authors are indebted to Prof. Guan Zhu from Texas A&M University, USA, for critical reading and constructive suggestions.

Author Contributions

Conceived and designed the experiments: JB DB YZ. Performed the experiments: YZ. Analyzed the data: YZ DB. Contributed reagents/materials/analysis tools: YZ. Wrote the manuscript: YZ DB JB.

References

  1. 1. Schaap FG, van der Vusse GJ, Glatz JF (2002) Evolution of the family of intracellular lipid binding proteins in vertebrates. Mol Cell Biochem 239: 69-77. doi:https://doi.org/10.1023/A:1020519011939. PubMed: 12479570.
  2. 2. Esteves A, Ehrlich R (2006) Invertebrate intracellular fatty acid binding proteins. Comp Biochem Physiol C Toxicol Pharmacol 142: 262-274. doi:https://doi.org/10.1016/j.cbpc.2005.11.006. PubMed: 16423563.
  3. 3. Storch J, Corsico B (2008) The emerging functions and mechanisms of mammalian fatty acid-binding proteins. Annu Rev Nutr 28: 73-95. doi:https://doi.org/10.1146/annurev.nutr.27.061406.093710. PubMed: 18435590.
  4. 4. Venkatachalam AB, Thisse C, Thisse B, Wright JM (2009) Differential tissue-specific distribution of transcripts for the duplicated fatty acid-binding protein 10 (fabp10) genes in embryos, larvae and adult zebrafish (Danio rerio). FEBS J 276: 6787-6797. doi:https://doi.org/10.1111/j.1742-4658.2009.07393.x. PubMed: 19843178.
  5. 5. Agulleiro MJ, André M, Morais S, Cerdà J, Babin PJ (2007) High transcript level of fatty acid-binding protein 11 but not of very low-density lipoprotein receptor is correlated to ovarian follicle atresia in a teleost fish (Solea senegalensis). Biol Reprod 77: 504-516. doi:https://doi.org/10.1095/biolreprod.107.061598. PubMed: 17554079.
  6. 6. Liu RZ, Li X, Godbout R (2008) A novel fatty acid-binding protein (FABP) gene resulting from tandem gene duplication in mammals: transcription in rat retina and testis. Genomics 92: 436-445. doi:https://doi.org/10.1016/j.ygeno.2008.08.003. PubMed: 18786628.
  7. 7. Kim SH, Bae YA, Yang HJ, Shin JH, Diaz-Camacho SP et al. (2012) Structural and Binding Properties of Two Paralogous Fatty Acid Binding Proteins of Taenia solium Metacestode. PLoS Negl Trop. Drosophila Inf Serv 6: e1868.
  8. 8. Huang L, Hu Y, Huang Y, Fang H, Li R et al. (2012) Gene/protein expression level, immunolocalization and binding characteristics of fatty acid binding protein from Clonorchis sinensis (CsFABP). Mol Cell Biochem 363: 367-376. doi:https://doi.org/10.1007/s11010-011-1189-3. PubMed: 22189506.
  9. 9. Furuhashi M, Hotamisligil GS (2008) Fatty acid-binding proteins: role in metabolic diseases and potential as drug targets. Nat Rev Drug Discov 7: 489-503. doi:https://doi.org/10.1038/nrd2589. PubMed: 18511927.
  10. 10. Jakobsson E, Alvite G, Bergfors T, Esteves A, Kleywegt GJ (2003) The crystal structure of Echinococcus granulosus fatty-acid-binding protein 1. Biochim Biophys Acta 1649: 40-50. doi:https://doi.org/10.1016/S1570-9639(03)00151-1. PubMed: 12818189.
  11. 11. Lek M, MacArthur DG, Yang N, North KN (2010) Phylogenetic analysis of gene structure and alternative splicing in alpha-actinins. Mol Biol Evol 27: 773-780. doi:https://doi.org/10.1093/molbev/msp268. PubMed: 19897525.
  12. 12. Tamura K, Dudley J, Nei M, Kumar S (2007) MEGA4: Molecular Evolutionary Genetics Analysis (MEGA) software version 4.0. Mol Biol Evol 24: 1596-1599. doi:https://doi.org/10.1093/molbev/msm092. PubMed: 17488738.
  13. 13. Bryson K, McGuffin LJ, Marsden RL, Ward JJ, Sodhi JS et al. (2005) Protein structure prediction servers at University College London. Nucleic Acids Res 33: W36-W38. doi:https://doi.org/10.1093/nar/gni035. PubMed: 15980489.
  14. 14. Milne I, Lindner D, Bayer M, Husmeier D, McGuire G et al. (2009) TOPALi v2: a rich graphical interface for evolutionary analyses of multiple alignments on HPC clusters and multi-core desktops. Bioinformatics 25: 126-127. doi:https://doi.org/10.1093/bioinformatics/btn575. PubMed: 18984599.
  15. 15. Treuner M, Kozak CA, Gallahan D, Grosse R, Müller T (1994) Cloning and characterization of the mouse gene encoding mammary-derived growth inhibitor/heart-fatty acid-binding protein. Gene 147: 237-242. doi:https://doi.org/10.1016/0378-1119(94)90073-6. PubMed: 7926807.
  16. 16. Bonné A, Gösele C, den Bieman M, Gillissen G, Kreitler T et al. (2003) Sequencing and chromosomal localization of Fabp6 and an intronless Fabp6 segment in the rat. Mol Biol Rep 30: 173-176. doi:https://doi.org/10.1023/A:1024968606162. PubMed: 12974472.
  17. 17. Prinsen CF, Weghuis DO, Kessel AG, Veerkamp JH (1997) Identification of a human heart FABP pseudogene located on chromosome 13. Gene 193: 245-251. doi:https://doi.org/10.1016/S0378-1119(97)00129-7. PubMed: 9256083.
  18. 18. Hastings KE (2005) SL trans-splicing: easy come or easy go? Trends Genet 21: 240-247. doi:https://doi.org/10.1016/j.tig.2005.02.005. PubMed: 15797620.
  19. 19. Halanych KM (2004) The new view of animal phylogeny. Annu Rev Ecol Evol Syst 35: 229-256. doi:https://doi.org/10.1146/annurev.ecolsys.35.112202.130124.
  20. 20. Dellaporta SL, Xu A, Sagasser S, Jakob W, Moreno MA et al. (2006) Mitochondrial genome of Trichoplax adhaerens supports placozoa as the basal lower metazoan phylum. Proc Natl Acad Sci U S A 103: 8751-8756. doi:https://doi.org/10.1073/pnas.0602076103. PubMed: 16731622.
  21. 21. Esteves A, Portillo V, Ehrlich R (2003) Genomic structure and expression of a gene coding for a new fatty acid binding protein from Echinococcus granulosus. Biochim Biophys Acta 1631: 26-34. doi:https://doi.org/10.1016/S1388-1981(02)00321-9. PubMed: 12573446.
  22. 22. Ganfornina MD, Gutiérrez G, Bastiani M, Sánchez D (2000) A phylogenetic analysis of the lipocalin protein family. Mol Biol Evol 17: 114-126. doi:https://doi.org/10.1093/oxfordjournals.molbev.a026224. PubMed: 10666711.
  23. 23. Nene V, Wortman JR, Lawson D, Haas B, Kodira C et al. (2007) Genome sequence of Aedes aegypti, a major arbovirus vector. Science 316: 1718-1723. doi:https://doi.org/10.1126/science.1138878. PubMed: 17510324.
  24. 24. Jeong KY, Kim WK, Lee JS, Lee J, Lee IY et al. (2005) Immunoglobulin E reactivity of recombinant allergen Tyr p 13 from Tyrophagus putrescentiae homologous to fatty acid binding protein. Clin Diagn Lab Immunol 12: 581-585. PubMed: 15879018.
  25. 25. Puerta L, Kennedy MW, Jim nez S, Caraballo L (1999) Structural and ligand binding analysis of recombinant blo t 13 allergen from Blomia tropicalis mite, a fatty acid binding protein. Int Arch Allergy Immunol 119: 181-184. doi:https://doi.org/10.1159/000024193. PubMed: 10436389.
  26. 26. Spence HJ, Moore J, Brass A, Kennedy MW (1993) A cDNA encoding repeating units of the ABA-1 allergen of Ascaris. Mol Biochem Parasitol 57: 339-343. doi:https://doi.org/10.1016/0166-6851(93)90210-O. PubMed: 8433722.
  27. 27. Kennedy MW (2000) The polyprotein lipid binding proteins of nematodes. Biochim Biophys Acta 1476: 149-164. doi:https://doi.org/10.1016/S0167-4838(99)00249-6. PubMed: 10669781.
  28. 28. Orton SM, Arasu P, Hammerberg B (2007) A novel gene from Brugia sp. that encodes a cytotoxic fatty acid binding protein allergen recognized by canine monoclonal IgE and serum IgE from infected dogs. J Parasitol 93: 1378-1387. doi:https://doi.org/10.1645/GE-1217.1. PubMed: 18314684.
  29. 29. Irimia M, Roy SW (2008) Spliceosomal introns as tools for genomic and evolutionary analysis. Nucleic Acids Res 36: 1703-1712. doi:https://doi.org/10.1093/nar/gkn012. PubMed: 18263615.
  30. 30. Belshaw R, Bensasson D (2006) The rise and falls of introns. Heredity 96: 208-213. doi:https://doi.org/10.1038/sj.hdy.6800791. PubMed: 16449982.
  31. 31. Roy SW, Gilbert W (2006) The evolution of spliceosomal introns: patterns, puzzles and progress. Nat Rev Genet 7: 211-221. doi:https://doi.org/10.1038/nrm1858. PubMed: 16485020.
  32. 32. Rogozin IB, Wolf YI, Sorokin AV, Mirkin BG, Koonin EV (2003) Remarkable interkingdom conservation of intron positions and massive, lineage-specific intron loss and gain in eukaryotic evolution. Curr Biol 13: 1512-1517. doi:https://doi.org/10.1016/S0960-9822(03)00558-X. PubMed: 12956953.
  33. 33. Pettitt J, Harrison N, Stansfield I, Connolly B, Müller B (2010) The evolution of spliced leader trans-splicing in nematodes. Biochem Soc Trans 38: 1125-1130. doi:https://doi.org/10.1042/BST0381125. PubMed: 20659016.
  34. 34. Ross LH, Freedman JH, Rubin CS (1995) Structure and expression of novel spliced leader RNA genes in Caenorhabditis elegans. J Biol Chem 270: 22066-22075. doi:https://doi.org/10.1074/jbc.270.37.22066. PubMed: 7665629.