Elsevier

Gene

Volume 678, 15 December 2018, Pages 318-323
Gene

Research paper
Transcriptional activation of a chimeric retrogene PIPSL in a hominoid ancestor

https://doi.org/10.1016/j.gene.2018.08.033Get rights and content

Highlights

  • We verified the expression of the PIPSL retrogene in white-handed gibbon.

  • The promoter is located upstream of the PIPSL TSS, similar to human PIPSL.

  • Promoter development occurred early in hominoid evolution before gibbon divergence.

Abstract

Retrogenes are a class of functional genes derived from the mRNA of various intron-containing genes. PIPSL was created through a unique mechanism, whereby distinct genes were assembled at the RNA level, and the resulting chimera was then reverse transcribed and integrated into the genome by the L1 retrotransposon. Expression of PIPSL RNA via its transcription start sites (TSSs) has been confirmed in the testes of humans and chimpanzee. Here, we demonstrated that PIPSL RNA is expressed in the testis of the white-handed gibbon. The 5′-end positions of gibbon RNAs were confined to a narrow range upstream of the PIPSL start codon and overlapped with those of orangutan and human, suggesting that PIPSL TSSs are similar among hominoid species. Reporter assays using a luciferase gene and the flanking sequences of human PIPSL showed that an upstream sequence exhibits weak promoter activity in human cells. Our findings suggest that PIPSL might have acquired a promoter at an early stage of hominoid evolution before the divergence of gibbons and ultimately retained similar TSSs in all of the lineages. Moreover, the upstream sequence derived from the phosphatidylinositol-4-phosphate 5-kinase, type I, alpha 5′ untranslated region and/or neighboring repetitive sequences in the genome possibly exhibits promoter activity. Furthermore, we observed that a TATA-box-like sequence has emerged by nucleotide substitution in a lineage leading to humans, with this possibly responsible for a broader distribution of the human PIPSL TSSs.

Introduction

Retrogenes are a class of functional genes derived from the mRNA of various intron-containing genes (Brosius, 1999; Wang, 2004; Casola and Betrán, 2017; Kubiak and Makałowska, 2017). They are postulated to have emerged through a mechanism similar to processed pseudogene (PP) formation; however, the detailed mechanism and their biological functions have yet to be elucidated. PPs are reverse-transcribed, intron-less cDNA copies of mRNA that have been reinserted into the genome (Vanin, 1985; Weiner et al., 1986) and are especially abundant in mammalian genomes (Ohshima et al., 2003; Zhang et al., 2003; Abyzov et al., 2013; Ewing et al., 2013; Schrider et al., 2013; Zhang, 2013; Kabza et al., 2014; Navarro and Galante, 2015; Wang, 2017). PPs are not usually transcribed, because they lack an external promoter; therefore, they have long been viewed as evolutionary dead ends with little biological relevance. In addition to PPs, there are two types of pseudogenes: DNA-duplicated (or non-processed) pseudogenes, which arose by segmental duplication (Torrents et al., 2003); and unitary pseudogenes, which are non-processed pseudogenes with no functional counterparts in the genome. Both pseudogene variants are generated by disruptive mutations occurring in functional genes (Matsui et al., 2010; Zhang et al., 2010).

Because pseudogenes generally accumulate many disruptive mutations during the course of evolution, many are considered to have lost transcriptional activity and coding potential (Mighell et al., 2000; Douglas et al., 2016). However, an increasing number of studies offer evidence that PPs can be transcribed by acquiring regulatory elements (Harrison et al., 2005; Sakai et al., 2007; Sorourian et al., 2014), with some having retained coding potential leading to new proteins (Shashidharan et al., 1994; Betrán et al., 2002; Betrán and Long, 2003; Rosso et al., 2008; Parker et al., 2009; Young et al., 2010; Ciomborowska et al., 2013; Abdelsamad and Pecinka, 2014). Others have acquired the regulatory function associated with their parental gene expression as an RNA molecule (Poliseno et al., 2010; Cheetham et al., 2013; Ha et al., 2014; Hirano et al., 2014). A recent large-scale survey of retrogene promoters revealed several patterns associated with promoter acquisition (Carelli et al., 2016).

Previously, we and our collaborators (Babushok et al., 2007) characterized a novel gene, PIPSL, created through a unique mechanism, whereby new combinations of functional domains were assembled at the RNA level from distinct genes [phosphatidylinositol-4-phosphate 5-kinase, type I, alpha (PIP5K1A) and proteasome 26S subunit, non-ATPase 4 (PSMD4)], with the resulting chimera (Akiva et al., 2006) then reverse transcribed and integrated into the genome by the L1 retrotransposon (Babushok et al., 2007; Doucet-O'Hare et al., 2015). The PIPSL gene encodes a chimeric protein with distinct cellular localization and minimal lipid kinase activity while exhibiting significant affinity for cellular ubiquitinated proteins (Babushok et al., 2007). The PIPSL sequence emerged in the genome of the common ancestor of extant hominoid species (Hylobatidae and Hominidae) (Ohshima and Igarashi, 2010), and the PIPSL locus is conserved within human populations and retains a novel haplotype with prominent amino acid changes (Ohshima and Igarashi, 2010). During hominoid diversification, ubiquitin-interacting motif (UIM)1 in the PSMD4-derived domain underwent critical amino acid replacements at an early stage, which were conserved under subsequent high levels of nonsynonymous substitutions to UIM2 and other domains, suggesting that adaptive evolution diversified these functional compartments (Ohshima and Igarashi, 2010).

Relatively high expression of PIPSL RNA along with its transcription start sites (TSSs) was confirmed in the testes of humans and chimpanzees (Babushok et al., 2007; Zhang et al., 2009); however, a full-length PIPSL protein has not been detected (Babushok et al., 2007), although partial amino acid sequences of PIPSL are available in a proteomics database (Ohshima and Igarashi, 2010). Because PIPSL has been inserted in an intergenic region without contact with neighboring genes, it is likely that PIPSL acquired its own transcription-control mechanism at an early stage of its evolution. In this study, to elucidate such processes, we investigated PIPSL expression in species other than human and chimpanzee and searched for possible promoter sequences.

Section snippets

Primate tissue samples and RNA isolation

Testes from a white-handed gibbon (Hylobates lar; Taki, GAIN#0164) and a chimpanzee (Pan troglodytes; Umetaro, GAIN#0599) were obtained from the Great Ape Information Network (http://www.shigen.nig.ac.jp/gain/) and the Primate Research Institute, Kyoto University (Aichi, Japan). Total RNA was isolated using an RNeasy mini kit (Qiagen, Tokyo, Japan) according to manufacturer instructions.

Reverse transcription polymerase chain reaction (RT-PCR)

The following primers were used for RT-PCR, specific to PIPSL RNA, and applicable to both gibbon and

PIPSL RNA is expressed in the testis of white-handed gibbon

To elucidate the molecular machinery driving initial activation of PIPSL transcription, we investigated PIPSL expression in the most distantly related species to human and chimpanzee in hominoids (i.e., the gibbon). Fig. 1 shows RT-PCR results using an RNA sample from the testis of a white-handed gibbon. We observed a discrete signal using primers specific for PIPSL on samples of gibbon testis RNA (lanes 1 and 2). We also performed RT-PCR on samples of chimpanzee testis RNA as a control (lanes

Early acquisition of promoter activity in the new retrogene

Here, we demonstrated PIPSL RNA expression in the testis of a gibbon, which is the most distantly related species to human and chimpanzee in hominoids. Moreover, the RNA sequence of gibbon PIPSL strongly suggested that the PIPSL TSSs are close in proximity between gibbons and humans. Based on these results, we propose an early acquisition hypothesis, which explains the molecular machinery associated with initial activation of PIPSL transcription (Fig. 4C). After divergence from old-world

Acknowledgments

We thank the following facilities: Fukuoka City Zoological Garden for providing a gibbon sample and Nasu World Monkey Park for providing a chimpanzee. These samples were obtained through the Great Ape Information Network and cooperative research with the Primate Research Institute, Kyoto University. We thank Drs. Masanori Imamura and Takashi Hayakawa of the Primate Research Institute for preparing samples. We thank Prof. Tamio Mizukami of the Nagahama Institute of Bio-Science and Technology for

Funding

This work was supported by an institutional grant from the Nagahama Institute of Bio-Science and Technology to K.O.

Declarations of interest

None.

References (57)

  • P. Akiva et al.

    Transcription-mediated gene fusion in the human genome

    Genome Res.

    (2006)
  • D.V. Babushok et al.

    A novel testis ubiquitin-binding protein gene arose by exon shuffling in hominoids

    Genome Res.

    (2007)
  • E. Betrán et al.

    Dntf-2r, a young Drosophila retroposed gene with specific male expression under positive Darwinian selection

    Genetics

    (2003)
  • E. Betrán et al.

    Evolution of the phosphoglycerate mutase processed gene in human and chimpanzee revealing the origin of a new primate gene

    Mol. Biol. Evol.

    (2002)
  • A. Buzdin et al.

    At least 50% of human-specific HERV-K (HML-2) long terminal repeats serve in vivo as active promoters for host nonrepetitive DNA transcription

    J. Virol.

    (2006)
  • L. Carbone et al.

    Gibbon genome and the fast karyotype evolution of small apes

    Nature

    (2014)
  • F.N. Carelli et al.

    The life history of retrocopies illuminates the evolution of new mammalian genes

    Genome Res.

    (2016)
  • C. Casola et al.

    The genomic impact of gene retrocopies: what have we learned from comparative genomics, population genomics, and transcriptomic analyses?

    Genome Biol. Evol.

    (2017)
  • S.W. Cheetham et al.

    Long noncoding RNAs and the genetics of cancer

    Br. J. Cancer

    (2013)
  • J. Ciomborowska et al.

    “Orphan” retrogenes in the human genome

    Mol. Biol. Evol.

    (2013)
  • T.T. Doucet-O'Hare et al.

    LINE-1 expression and retrotransposition in Barrett's esophagus and esophageal carcinoma

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

    (2015)
  • G.M. Douglas et al.

    Decreased transcription factor binding levels nearby primate pseudogenes suggest regulatory degeneration

    Mol. Biol. Evol.

    (2016)
  • A.D. Ewing et al.

    Retrotransposition of gene transcripts leads to structural variation in mammalian genomes

    Genome Biol.

    (2013)
  • N.V. Fuchs et al.

    Expression of the human endogenous retrovirus (HERV) group HML-2/HERV-K does not depend on canonical promoter elements but is regulated by transcription factors Sp1 and Sp3

    J. Virol.

    (2011)
  • P. Gagniuc et al.

    Eukaryotic genomes may exhibit up to 10 generic classes of gene promoters

    BMC Genomics

    (2012)
  • H. Ha et al.

    A comprehensive analysis of piRNAs from adult human testis and their relationship with genes and mobile elements

    BMC Genomics

    (2014)
  • A. Hachem et al.

    PLCζ is the physiological trigger of the Ca2+ oscillations that induce embryogenesis in mammals but conception can occur in its absence

    Development

    (2017)
  • P.M. Harrison et al.

    Transcribed processed pseudogenes in the human genome: an intermediate form of expressed retrosequence lacking protein-coding ability

    Nucleic Acids Res.

    (2005)
  • View full text