Elsevier

Gene

Volume 441, Issues 1–2, 15 July 2009, Pages 126-131
Gene

Tracing back EFL gene evolution in the cryptomonads–haptophytes assemblage: Separate origins of EFL genes in haptophytes, photosynthetic cryptomonads, and goniomonads

https://doi.org/10.1016/j.gene.2008.05.010 Get rights and content

Abstract

A recently identified GTPase, elongation factor-like (EFL) protein is proposed to bear the principal functions of translation elongation factor 1α (EF-1α). Pioneering studies of EF-1α/EFL evolution have revealed the phylogenetically scattered distribution of EFL amongst eukaryotes, suggesting frequent eukaryote-to-eukaryote EFL gene transfer events and subsequent replacements of EF-1α functions by EFL. We here determined/identified seven new EFL sequences of the photosynthetic cryptomonad Cryptomonas ovata, the non-photosynthetic cryptomonad (goniomonad) Goniomonas amphinema, the foraminifer Planoglabratella opecularis, the haptophyte Chrysochromulina sp., the centroheliozoan Raphidiophrys contractilis, and two red algae Chondrus crispus and Gracilaria changii. The analyses of these EFL sequences successfully brought new insights into lateral EFL gene transfer amongst eukaryotes. Of most interest is a complex EFL evolution in a monophyletic assemblage comprised of cryptomonads and haptophytes. Since our analyses rejected any phylogenetic affinity amongst the EFL sequences from Goniomonas, photosynthetic cryptomonads, and haptophytes, the EFL genes of the three lineages most likely originated from different phylogenetic sources.

Introduction

Elongation factor 1α (EF-1α) in eukaryotes and archaebacteria (homologous to EF-Tu in eubacteria) is one of the principal protein factors involved in nascent peptide synthesis (Andersen et al., 2003). The GTP-bound EF-1α can bind and deliver aminoacyl-tRNA (aa-tRNA) to the A site of the ribosome. Subsequently, EF-1α hydrolyzes GTP coupled with the release of aa-tRNA. GDP-bound EF-1α then interacts with EF-1β (or EF-Ts in eubacteria) to recharge GTP before participating in the second round of peptide elongation process. In addition to both “core” functions described above, eukaryotic EF-1α bears arbitrary functions (e.g., interaction to cytoskeleton and proteosomes, Negrutskii and El'skaya, 1998). Due to the constraints from the core and arbitrary functions in cells, the primary, secondary, and tertiary structures of EF-1α/EF-Tu are highly conserved.

It has been generally believed that the evolution of EF-1α/EF-Tu is relatively simple: EF-1α/EF-Tu genes have been vertically inherited from the last universal common ancestor, and the gene products are ubiquitous in all extant cells. However, large-scale sequence data from phylogenetically diverged organisms have exposed both duplication and lateral transfer of EF-1α/EF-Tu genes. Firstly, there is literature describing lateral EF-1α/EF-Tu gene transfers (Ke et al., 2000, Inagaki et al., 2002, Inagaki et al., 2006). In addition, EF-1α paralogues, which were produced by gene duplication followed by functional divergence, are regularly found in both prokaryotic and eukaryotic genomes (Inagaki and Doolittle, 2000, Inagaki et al., 2003). Most importantly, recent complete genome data questioned the absolute ubiquity of EF-1α. No EF-1α gene has been identified in the complete nuclear genome of the chlorophycean green alga Chlamydomonas reinhardtii (genome.jgi-psf.org/Chlre3/Chlre3.home.html), or in those of prasinophycean green algae Ostreococcus lucimarinus (genome.jgi-psf.org/Ost9901_3/Ost9901_3.home.html) and O. tauri (genome.jgi-psf.org/Ostta4/Ostta4.home.html). Instead, these genomes possess genes encoding “elongation factor-like (EFL)” proteins that bear sequence similarity to, but are clearly divergent from canonical EF-1α (Keeling and Inagaki, 2004). An in silico analysis comparing the functions of EFL and those of EF-1α suggested that EFL holds at least the primary EF-1α functions (Keeling and Inagaki, 2004). Although EFL and EF-1α are considered as functionally equivalent molecules, the evolutionary mode proposed for EFL genes is drastically different from that for EF-1α genes in light of EF-1α/EFL gene distribution in global eukaryotic phylogeny.

A current hypothetical eukaryotic tree and the EF-1α/EFL distribution including the results in this study are shown in Fig. 1. The monophylies of the supergroups Opisthokonta, Amoebozoa, Plantae, and Rhizaria have been demonstrated by multigene analyses (Bapteste et al., 2002, Burki and Pawlowski, 2006, Rodriguez-Ezpeleta et al., 2005, Seenkamp et al., 2006). On the other hand, the monophyly of the lineages included in the supergroup Excavata has been proposed mainly based on morphological and ultrastructural data (Simpson and Roger, 2004). The supergroup Chromalveolata contains four major protist groups — alveolates (dinoflagellates, ciliates, and apicomplexans), stramenopiles, haptophytes, and cryptomonads. The chromalveolate hypothesis (Cavalier-Smith, 1999) assumes that a common ancestor of “chromalveolates” acquired plastids through the endosymbiosis of a red alga, and non-photosynthetic relatives (e.g., ciliates and goniomonads) were secondarily lost the red alga-derived plastids. To our knowledge, no nucleus-encoded gene phylogeny has recovered the host monophyly of the four “chromalveolata” groups. Nevertheless, the robust sisterhood between haptophytes and cryptomonads was constantly recovered in multigene analyses (e.g., Patron et al., 2007). In addition, some eukaryotic lineages show no clear phylogenetic affinity to any of the six supergroups described above. Centrohelida is typically considered incertae sedis (Sakaguchi et al., 2005, Sakaguchi et al., 2007).

EFL genes have so far been identified in (i) the members of Chlorophyta (except the ulvophycean green alga Acetabularia), (ii) the mesostigmatophycean green alga Mesostigma viride, (iii) dinoflagellates and the perkinsid Perkinsus marinus, (iv) photosynthetic cryptomonads, (v) haptophytes, (vi) chytrid and zygomycete fungi, (vii) choanoflagellates, (viii) ichthyosporeans, and (ix) chlorarachniophytes (Keeling and Inagaki, 2004, Gile et al., 2006, Ruiz-Trillo et al., 2006, Noble et al., 2007). Keeling and Inagaki (2004) originally pointed out that “EFL-containing” organisms phylogenetically associate with “EF-1α-containing” organisms as shown in Fig. 1. In fungi, EFL genes were identified only in limited species, while the majority of fungal species utilizes EF-1α (Keeling and Inagaki, 2004). Similarly, Viridiplantae is comprised of both EF-1α-containing and EFL-containing organisms — land plants, charophycean green algae, and the ulvophycean green alga Acetabularia acetabulum utilize EF-1α, while other green algal species possess EFL (Noble et al., 2007). The large protist assemblage Alveolata is comprised of two EF-1α-containing sub-groups (apicomplexans and ciliates) and an EFL-containing sub-group (dinoflagellates plus perkinsids) (Keeling and Inagaki, 2004). Such mosaic EF-1α/EFL distribution was most likely achieved by EFL gene transfer between distantly related eukaryotes. Nevertheless, an alternative hypothesis in which invokes parallel loss events of EFL genes cannot completely be excluded. If one hesitates to invoke an extremely high rate of lateral EFL gene transfer in eukaryotic evolution (e.g., LGT amongst very closely related lineages), the “gene-loss” scenario can be an alternative explanation for complex EF-1α/EFL distribution in choanoflagellates (Keeling and Inagaki, 2004), ichthyosporeans (Ragan et al., 2003, Ruiz-Trillo et al., 2006), and ulvophycean green algae (Noble et al., 2007), which are highlighted by grey circles in Fig. 1.

In this study, we newly determined/identified seven EFL sequences, and updated the EFL distribution in global eukaryotic phylogeny. Of particular interest is the EFL evolution in a monophyletic assemblage comprised of the cryptomonads, haptophytes, and their closely related lineages (henceforth designated as the cryptomonads–haptophytes assemblage). Our EFL phylogenetic analyses failed to recover the monophyly of the sequences from the member of the cryptomonad–haptophyte assemblage, suggesting that the EFL origins of haptophytes, goniomonads, and photosynthetic cryptomonads are separate from one another.

Section snippets

Cell culture and cDNA preparation

Raphidiophrys contractilis cells were cultured as described by Sakaguchi and Suzaki (1999). Goniomonas amphinema was kindly donated by N. Yubuki (University of Tsukuba). Cryptomonas ovata (NIES-274) was given by I. Inouye (University of Tsukuba). The haptophyte Chrysochromulina sp. (NIES-1333) was purchased from the Microbial Culture Collection at the National Institute for Environmental Studies (NIES; 16-2, Onogawa, Tsukuba, Ibaraki 305-8506, Japan). Poly(A)+ RNA samples were prepared from the

Florideophycean red algal EFL

We newly identified EFL transcripts in EST data of florideophycean red algae Chondrus crispus (GenBank accession numbers; CO652990 , CO652099 , and CO652788 ) and Gracilaria changii (DV963090 , DV963156 , and DV964877 ). Importantly, no EF-1α transcript was detected in the two red algal data. The two florideophycean EFL sequences were robustly united in an EFL phylogeny [Bootstrap probability (BP) value = 73–99%], while this clade as a whole showed no apparent affinity to any sequences considered in

Discussion

In the current study, we determined/identified seven new EFL genes, and, by the analyses of the expanded EFL data, successfully elevated our understanding of EFL evolution in global eukaryotic phylogeny as discussed below.

Acknowledgements

This study was supported in part by the Japan Society for Promotion of Science grants 18770061 (M.S.), 18570214 (Y.I.), and 17370086 (T.H.). We would like to thank Drs. T. Suzaki (Kobe University) and M. Tsuchiya (JAMSTEC) for kindly providing Raphidiophrys cells and Planoglabratella cDNA, respectively. We also thank R. Kamikawa (Kyoto University) and J. D. Reimer (University of the Ryukyus) for valuable discussions and the critical reading of the manuscript, respectively.

References (42)

  • SakaguchiM. et al.

    Centrohelida is still searching for a phylogenetic home: analyses of seven Raphidiophrys contractilis genes

    Gene

    (2007)
  • SakaguchiM. et al.

    Monoxenic culture of the heliozoon Actinophrys sol

    Eur. J. Protistol.

    (1999)
  • TakishitaK. et al.

    A close relationship between Cercozoa and Foraminifera supported by phylogenetic analyses based on combined amino acid sequences of three cytoskeletal proteins (actin, a-tubulin, and b-tubulin)

    Gene

    (2005)
  • Abascal, F., Correspondence to: Zardoya, R., Posada, D., 2005. ProtTest: selection of best-fit models of protein...
  • ArchibaldJ.M. et al.

    Actin and ubiquitin protein sequences support a cercozoan/foraminiferan ancestry for the plasmodiophorid plant pathogens

    J. Eukaryot. Microbiol.

    (2004)
  • ArchibaldJ.M. et al.

    A novel polyubiquitin structure in Cercozoa and Foraminifera: evidence for a new eukaryotic supergroup

    Mol. Biol. Evol.

    (2003)
  • BaptesteE. et al.

    The analysis of 100 genes supports the grouping of three highly divergent amoebae: Dictyostelium, Entamoeba, and Mastigamoeba

    Proc. Natl. Acad. Sci. U. S. A.

    (2002)
  • BerneyC. et al.

    Revised small subunit rRNA analysis provides further evidence that Foraminifera are related to Cercozoa

    J. Mol. Evol.

    (2003)
  • BurkiF. et al.

    Monophyly of Rhizaria and multigene phylogeny of unicellular bikonts

    Mol. Biol. Evol.

    (2006)
  • BurkiF. et al.

    Phylogenomics reshuffles the eukaryotic supergroups

    PLoS ONE

    (2007)
  • Cavalier-SmithT.

    Principles of protein and lipid targeting in secondary symbiogenesis: euglenoid, dinoflagellate, and sporozoan plastid origins and the eukaryote family tree

    J. Eukaryot. Microbiol.

    (1999)
  • Cited by (14)

    • Monitoring of the toxic dinoflagellate Alexandrium catenella in Osaka Bay, Japan using a massively parallel sequencing (MPS)-based technique

      2019, Harmful Algae
      Citation Excerpt :

      Effectiveness of MPS to detect many species simultaneously is especially relevant to reveal the presence of HAB species/genera (e.g. Dzhembekova et al., 2017; Elferink et al., 2017; Moreno-Pino et al., 2018; Nagai, 2018) and gather information on their appearances as some species may be present sporadically (Nagai et al., 2017) or during specific seasons (Sildever et al., 2019). This information is prerequisite for monitoring by other molecular methods, e.g. qPCR and microarrays, which are limited by the number of species that can be detected at the same time (Medlin and Orozco, 2017) and by the need for specific primers to detect species of interest from the environmental samples (Garneau et al., 2011; Nagai et al., 2016c; Penna et al., 2013; Ruvindy et al., 2018). The microalgal genus Alexandrium Halim, 1960 (Dinophyceae, Gonyaulacales) currently contains 34 species (Balech, 1995; MacKenzie and Todd, 2002; Montressor et al., 2004; Guiry and Guiry, 2019), including 12 that are known to produce paralytic shellfish toxins (PSTs) (Anderson et al., 2012; Moestrup et al., 2019), responsible shellfish toxicity and the resulting paralytic shellfish poisoning (PSP) throughout the world (Anderson et al., 2012).

    • Massively parallel sequencing-based survey of eukaryotic community structures in Hiroshima Bay and Ishigaki Island

      2016, Gene
      Citation Excerpt :

      The relative abundance of OTUs at the supergroup levels among samples collected was compared by counting the number of sequences at each supergroup level for the 10 samples and by constructing pie charts. In this study, a supergroup was defined as Alveolata, Amoebozoa, Apusozoa, Archaeplastida, Excavata, Hacrobia, Metazoa, Opisthokonta, Rhizaria, Stramenopila, and Viridiplantae, according to Adl et al. (2005); Burki et al. (2007); Frommolt et al. (2008); Archibald (2009); Hampl et al. (2009); Okamoto et al. (2009), and Sakaguchi et al. (2009). The relative abundance of the top 20 OTUs was depicted in pie charts and compared among the samples.

    • Oxnerella micra sp. n. (Oxnerellidae fam. n.), a Tiny Naked Centrohelid, and the Diversity and Evolution of Heliozoa

      2012, Protist
      Citation Excerpt :

      It is noteworthy that we found EFL in both Microheliella and the centrohelid Marophrys marina and that EF1-α was not detected for Microheliella or Oxnerella in our EST project (among about 76,000 sequence reads). This is consistent with previous evidence that EFL but not EF1-α is found in the centrohelid Polyplacocystis (formerly Raphidiophrys) contractilis (Sakaguchi et al. 2009). Eukaryotes generally use either EF1-α or EFL for protein synthesis, but almost all have genes for only one of these apparently functionally equivalent proteins (Kamikawa et al. 2011).

    • Cercozoa comprises both EF-1α-containing and EFL-containing members

      2011, European Journal of Protistology
      Citation Excerpt :

      Moreover, there are several cases in which EF-1α-containing and EFL-containing species are closely related. For instance, we detected only EF-1α sequences (e.g., GenBank accession number ES559272) in the EST survey of Goniomonas pacifica, whereas an EFL gene was isolated from G. amphinema (Sakaguchi et al. 2009). Similarly, EF-1α-containing and EFL-containing members co-exist in choanoflagellates, ulvophycean green algae, euglenozoans, and diatoms (Keeling and Inagaki 2004; Ruiz-Trillo et al. 2006; Noble et al. 2007; Kamikawa et al. 2008; Gile et al. 2009a,b).

    • Rooting for the root of elongation factor-like protein phylogeny

      2010, Molecular Phylogenetics and Evolution
    View all citing articles on Scopus
    View full text