Next Article in Journal
Poly(amidoamine) Dendrimers as Nanocarriers for 5-Fluorouracil: Effectiveness of Complex Formation and Cytotoxicity Studies
Previous Article in Journal
Encapsulation of Plant Biocontrol Bacteria with Alginate as a Main Polymer Material
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Discovery Profiling of PIWI-Interacting RNAs and Their Diverse Functions in Metazoans

Department of Aquatic Bioscience, Graduate School of Agricultural and Life Sciences, The University of Tokyo, Bunkyo, Tokyo 113-8657, Japan
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(20), 11166; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms222011166
Submission received: 8 September 2021 / Revised: 11 October 2021 / Accepted: 14 October 2021 / Published: 16 October 2021
(This article belongs to the Section Molecular Genetics and Genomics)

Abstract

:
PIWI-interacting RNAs (piRNAs) are a class of small non-coding RNAs (sncRNAs) that perform crucial biological functions in metazoans and defend against transposable elements (TEs) in germ lines. Recently, ubiquitously expressed piRNAs were discovered in soma and germ lines using small RNA sequencing (sRNA-seq) in humans and animals, providing new insights into the diverse functions of piRNAs. However, the role of piRNAs has not yet been fully elucidated, and sRNA-seq studies continue to reveal different piRNA activities in the genome. In this review, we summarize a set of simplified processes for piRNA analysis in order to provide a useful guide for researchers to perform piRNA research suitable for their study objectives. These processes can help expand the functional research on piRNAs from previously reported sRNA-seq results in metazoans. Ubiquitously expressed piRNAs have been discovered in the soma and germ lines in Annelida, Cnidaria, Echinodermata, Crustacea, Arthropoda, and Mollusca, but they are limited to germ lines in Chordata. The roles of piRNAs in TE silencing, gene expression regulation, epigenetic regulation, embryonic development, immune response, and associated diseases will continue to be discovered via sRNA-seq.

1. Introduction

Small non-coding RNAs (sncRNAs) engage in gene regulation at the transcriptional and post-transcriptional levels and are classified as microRNAs (miRNAs), endogenous small interfering RNAs (endo-siRNAs), and PIWI-interacting RNAs (piRNAs) based on their size and Argonaute partner in biogenesis [1]. piRNAs form the largest and most heterogeneous class of sncRNAs because they lack conserved structural motifs and sequence homology across species [2,3]. Studies on piRNAs have attracted significant attention from researchers in the last decade.
Double-stranded small RNA derived from the suppressor of the Stellate locus on the Y chromosome was first discovered in the Drosophila melanogaster germ line [4]. Repeat-associated small interfering RNAs (rasiRNAs) were first identified in Drosophila germ lines [4,5,6], and were later termed piRNA subspecies as they were found to interact with PIWI proteins [7]. Remarkable progress has been made in understanding piRNA biogenesis and function, especially in Drosophila, Caenorhabditis elegans, mice, and humans [8,9,10,11,12,13,14,15,16,17,18,19]. Next-generation sequencing (NGS) has been widely used for high-throughput characterization of sncRNAs. Increasingly, piRNAs have been discovered in the soma and germ lines of non-model organisms, including Platyhelminthes [20], Annelida [21], Cnidaria [22,23,24,25], Echinodermata [26], Mollusca [27], Crustacea [28], Arthropoda [29], Reptilia [30], and Mammals [31].
In the last decade, several studies have attempted to elucidate the biogenesis of piRNAs [32,33,34,35,36,37]. Two models of the piRNA biogenesis pathway have been demonstrated in various animals: the primary piRNA biogenesis pathway and the amplification loop or ping-pong cycle [32]. In the primary piRNA biogenesis pathway, long piRNA precursors are transcribed from piRNA clusters, cleaved and modified by complex factors in the cytoplasm, and then transported into the nucleus in complex with PIWI proteins [38]. piRNAs generated by the primary pathway may play a role in regulating gene expression [32]. Secondary piRNAs are formed in an amplification mechanism (termed the ping-pong amplification loop) to specifically enhance piRNA sequences [35,36].
The PIWI–piRNA pathway effectively suppresses transposable element (TE) activity in order to safeguard the genome from detrimental insertion mutagenesis [39]. Recent findings show that the PIWI–piRNA pathway also plays a vital role in somatic cells [40,41] and various cancer cells [42,43,44,45]. The present review aims to provide guidelines for piRNA discovery in future studies. We discuss the discovery profiling of piRNAs in model and non-model organisms using small RNA sequencing (sRNA-seq) and provide an overview of piRNA functions in animals. In addition, we rediscovered ubiquitously expressed piRNAs in the soma and germ lines of invertebrates from previously overlooked sRNA-seq data. Overall, discovering piRNAs can assist researchers in analyze their functions in non-model organisms.

2. Identification of piRNA

2.1. Discovery Workflow

Identifying piRNAs from sRNA-seq is imperative for further functional analysis (Figure 1). Samples from tissues or cells were prepared for sRNA-seq to identify the piRNA molecules. Raw data from high-throughput sequencing required trimming adapters and quality control processes, such as filtration of low-quality reads, poly(A) reads, or length, to obtain clean reads. Moreover, the clean small RNAs were aligned with genome sequences and well-known RNA databases for the filtration of infectant reads and known RNA molecules, such as ribosomal RNAs (rRNAs), miRNAs, and small interfering RNAs (siRNAs). Generally, piRNA sequences are represented by ncRNA fragments, while some piRNA databases contain a subset of sequences that correspond to piRNA-sized fragments of ncRNAs (rRNAs, transfer RNAs (tRNAs), small nuclear RNAs (snRNAs), and small nucleolar RNAs (snoRNAs)) and intermediates of miRNA biogenesis, which strongly affect the estimation of piRNA expression outside mammalian gonads [46,47]. Therefore, all known ncRNA fragments should be thoroughly filtered out when analyzing somatic piRNAs in mammals. Finally, the putative reads were processed experimentally or using bioinformatics tools to identify the piRNA molecules.
Crosslinking immunoprecipitation sequencing (CLIP-seq) and RNA immunoprecipitation sequencing (RIP-seq) are commonly used to detect piRNAs with the coprecipitation of PIWI/Argonaute. The experimental method is powerful, allowing unambiguous classification of precipitated small RNAs and elucidation of the functions of various PIWI or Argonaute proteins, but with the disadvantages of being time-consuming and expensive [48,49]. Therefore, specialized bioinformatics tools for piRNA identification and processing on a large scale are required.
Figure 1. Overview of common pipeline for PIWI-interacting RNA (piRNA) discovery and functional analysis in metazoans. The raw data of small RNA sequencing (sRNA-seq) were trimmed using adapters, quality control was performed, and the data were subsequently filtered by read lengths. Generally, microRNAs (miRNAs) and small interfering RNAs (siRNAs) are 20–25 nt in length, transfer RNAs (tRNAs) are typically ~80 nt in length, and small nuclear RNAs (snRNAs) and circular RNAs (circRNAs) are more than 100 nt in length, whereas piRNAs normally have lengths of 24–31 nt. However, species-specific piRNAs of 21 nt with a 5′ uridine (21U-RNAs) binding to p53-responsive gene 1 (PRG-1) and 22 nt piRNAs with a 5′ guanosine (22G-RNAs) loaded onto worm-specific Argonautes (WAGOs) were detected in C. elegans [50,51,52]. In the preprocessing step, the potential piRNAs and piRNA isoforms with a length of 18–35 nt were preserved for subsequent known RNA mapping and filtration. The remaining putative piRNA reads were processed for piRNA analysis using multiple tools and databases.
Figure 1. Overview of common pipeline for PIWI-interacting RNA (piRNA) discovery and functional analysis in metazoans. The raw data of small RNA sequencing (sRNA-seq) were trimmed using adapters, quality control was performed, and the data were subsequently filtered by read lengths. Generally, microRNAs (miRNAs) and small interfering RNAs (siRNAs) are 20–25 nt in length, transfer RNAs (tRNAs) are typically ~80 nt in length, and small nuclear RNAs (snRNAs) and circular RNAs (circRNAs) are more than 100 nt in length, whereas piRNAs normally have lengths of 24–31 nt. However, species-specific piRNAs of 21 nt with a 5′ uridine (21U-RNAs) binding to p53-responsive gene 1 (PRG-1) and 22 nt piRNAs with a 5′ guanosine (22G-RNAs) loaded onto worm-specific Argonautes (WAGOs) were detected in C. elegans [50,51,52]. In the preprocessing step, the potential piRNAs and piRNA isoforms with a length of 18–35 nt were preserved for subsequent known RNA mapping and filtration. The remaining putative piRNA reads were processed for piRNA analysis using multiple tools and databases.
Ijms 22 11166 g001

2.2. Processing of piRNAs

The absence of many conserved structural and sequence characteristics makes it difficult to identify piRNAs using computational methods. An exception to this is their preference for a uridine nucleotide at the first position (1U) from 5 terminal [53]. A hallmark characteristic of piRNA sequences is their appearance in genome clusters ranging from 1 to >100 kb in length, with both monodirectional and bidirectional encoding clusters [54]. Moreover, secondary piRNAs show a strong bias for adenine at position 10 (10A), complementing the 1U bias of primary piRNAs [55].
In the last decade, scientists have developed various computational methods to identify piRNAs from sRNA-seq. These methods can be classified into two categories: linear classification algorithms to predict individual piRNAs and clustering approaches to predict clustered piRNAs [3]. One aim in identifying piRNAs is to summarize the general features of known piRNAs from model organisms with complete genome sequences and use them to predict novel piRNAs [2]. Several methods have been developed to predict individual piRNAs based on their type. For example, Pinao [56], a genetic algorithm-based weighted ensemble (GA-WE) [57], and accurate piRNA prediction [58] have been used for transposon-related piRNA prediction, and two-layer integrated programs for identifying piRNAs (2L-piRNA) [59], such as 2L-piRNAPred [60], 2lpiRNApred [61], and 2L-piRNADNN [62], have been developed for mRNA-related piRNA prediction, while piRNAPredictor [2], PiRPred [3], piRNAdetect [63], IpiRId [64], piRNN [65], and piRNApred [66] have been employed for total piRNA prediction. miRanda [17], pirnaPre [67], and pirScan [18] have been used for piRNA target prediction, and three algorithms have been proposed for predicting piRNA clusters from sRNA-seq data: proTRAC [54], piClust [68], and PILFER [69]. In addition, multiple integrated platforms, such as sRNAtools [70] and Workflow for piRNAs and Beyond (WIND) [71], have been recently developed for piRNA annotation and downstream analysis from raw data to plots and statistics by sRNA-seq. The performances of most of these piRNA prediction tools have been reviewed by Liu et al. [46].
Multiple piRNA-dedicated databases have been built for piRNA annotation and downstream analysis. These can be divided into different categories according to their functions: piRNABank [72], piRBase [73], and piRNAdb for comprehensive annotation; IsopiRBank [74] for piRNA isoform identification; piRNAclusterDB [75] for piRNA cluster annotation; piRNAtarget [76] and piRTarBase [77] for piRNA-mediated target prediction; and piRDisease [78] and piRPheno [79] for piRNA-related disease analysis. However, most piRNA databases have been generated from model organisms, such as C. elegans, Drosophila, mice, and humans, which limits their use in non-model organisms.

2.3. Validation of piRNA

Northern blotting, in situ hybridization, and quantitative reverse transcription-polymerase chain reaction (qRT-PCR) are the three main approaches for the experimental validation of piRNAs. These methods have low throughput and cannot validate hundreds of piRNAs and their isoforms detected by sRNA-seq. Sequencing of PIWI-precipitated small RNAs is usually used to detect piRNAs directly. However, sRNA-seq of cell lines or tissues before and after the knockdown or knockout of piRNA biogenesis pathway genes can be used to assess the biosynthesis of the predicted piRNAs, as the biogenesis of real piRNAs can be expected to be affected. No piRNAs were detected in zebrafish PIWI (ZIWI) mutant testes in zebrafish [80] or in PIWI mutant fat bodies in Drosophila [81].
High-throughput CLIP-seq is another method employed not only for the validation of putative piRNAs but also to verify their activity [82,83]. Overlaps between CLIP-seq tags for putative piRNAs and microprocessor complex subunits or PIWI proteins provide evidence for interactions between the putative piRNAs and the microprocessor or RNA-induced silencing complexes (RISCs) [84]. To determine piRNA targets, CLIP-seq and RIP-seq can identify thousands of transcripts associated with PIWI proteins; however, it is difficult to infer the target of a specific piRNA using these methods [19]. Bioinformatics can be used to first predict the targets of a specific piRNA, but additional approaches are required to validate the predicted binding sites in vivo, such as the dual-luciferase reporter assay with co-transfected piRNA expression vector and wild-type and mutated forms of the predicted 3′ untranslated region (UTR) reporter vector [85]. The interaction between piRNA precursors and intermediate biogenesis factors has also been verified by CLIP-seq [86,87]. Crosslinking, ligation, and sequencing of hybrids (CLASH) has been used to identify small RNAs and candidate target RNA binding sites [88], thus providing direct evidence of piRNA-mediated gene regulation in RISC. CLASH was utilized to study the binding sites between piRNAs and their potential target mRNAs in C. elegans [89].
Periodate-mediated oxidation has been used to yield clean piRNA sequences during sRNA-seq processing [21,24,29,31]. The chemical structures of piRNAs were confirmed using this method, followed by β-elimination reactions [90]. It was reported that almost all the piRNAs tested were resistant to periodate treatment, indicating a modified 2′ or 3′ hydroxyl group at the 3′ terminal nucleotides of piRNA, which is methylated by the small RNA methyltransferase HUA ENHANCER1 (HEN1) [91,92].

3. Discovery of piRNAs by sRNA-seq

Most of the information on piRNAs is obtained from model organisms such as Drosophila; however, continual progress is being made with other organisms belonging to Cnidaria, Mollusca, and Chordata (Teleostean, Amphibian, Reptilia, Aves, and Mammal) (Figure 2). We acquired approximately 1424 sRNA-seq datasets for 114 animal species from public databases for piRNA identification and characterization in invertebrates and vertebrates (Table 1; Supplementary Table S1), including species with and without existing piRNA information. In the same taxa, the proportion of TEs increases with genome size [93], whereas the number of piRNA species does not increase with the size of the genome or the proportion of TEs (Figure 2). piRNAs were not detected in Protozoa but were detected in C. elegans and Halichondria panicea, belonging to Nematoda and Porifera [14,21]. In Platyhelminthes, piRNAs were detected in planarians (Schmidtea mediterranea) [20] but were absent in flukes and tapeworms [94]. Ubiquitously expressed piRNAs were discovered in the soma and germ lines of Annelida, Cnidaria, Echinodermata, Crustacea, Arthropoda, and Mollusca. piRNA expression underwent tremendous changes in the Chordata. They were mostly expressed in early embryos, mammalian testes, and ovaries of Macaca fascicularis and Oryctolagus cuniculus [31]. piRNAs were also found to exist outside the germ line, particularly in the nervous system of Aplysia species [95] and the liver of the bamboo shark (Chiloscyllium plagiosum) [96], suggesting much broader roles than previously understood. The somatic piRNA pathway plays a minor role in Drosophila, whereas in other Arthropoda somatic piRNAs are more abundant and diversified [29]. The presence of piRNAs in most lower animal species suggests that their last common ancestor had pathways active in both the soma and germ line, and several species in Chordata lost their activity in all but gonadal tissues.

4. Diverse Functions of piRNAs

The PIWI–piRNA pathway in animals is a conserved pathway that is crucial for genome defense. Its main function is to repress TEs via transcriptional or post-transcriptional silencing mechanisms, thereby maintaining germ-line genomic integrity [32,33,108]. In addition to transposon silencing, piRNAs interact with PIWI proteins to form the piRNA-induced silencing complex (piRISC), which is associated with genome rearrangement, mRNA regulation, epigenetic regulation, spermatogenesis, development, virus defense, and human diseases (Figure 3).

4.1. Silencing of Transposable Elements

The first evidence for a small RNA-based regulatory mechanism that could protect against transposon mobilization was noted in repeat-associated small interfering RNAs (rasiRNAs) [4,5,6,12,103]. Since then, abundant TE-related piRNAs have been found in the germ lines of Mollusca, Arthropoda, and Chordata, including fish, dogs, bats, horses, mice, rats, marmosets, and rhesus macaques [11,29,80,106,109,110]. The complexes of piRISC repress transposons via two mechanisms depending on the PIWI protein involved [33]. The cytoplasmic proteins Aubergine (Aub) and Argonaute3 (Ago3) in Drosophila, mouse PIWI (Miwi) and Miwi-like protein (Mili) in mice, and silkworm PIWI (Siwi) and Ago3 in silkworms participate in slicer-dependent post-transcriptional gene silencing (PTGS) via the ping-pong cycle [54,103]. In contrast, Drosophila PIWI and murine Miwi2 translocate to the nucleus when loaded with piRNAs [54,103,111]. It was found that these molecular mechanisms repress transposons through transcriptional gene silencing (TGS) [15,16,112,113,114,115]. Recent studies have identified novel components of piRNA-mediated TGS; testis expressed 15 (TEX15) and Spen paralogue and orthologue C-terminal domain containing 1 (SPOCD1) might provide a link between piRNA-guided complexes that recognize genomic targets and the molecular machinery that induces DNA methylation and transcriptional repression in mice [116,117,118] and in HP1, histone 3 lysine 9 trimethylation (H3K9me3), small ubiquitin-like modifier (SUMO), and histone deacetylase Rpd3 in Drosophila [119,120,121], which would considerably deepen our understanding of PIWI–piRNA-mediated heterochromatin formation at transposon loci. In actual analyses, piRNAs have been found to suppress transposon expression in both somatic and gonadal tissues in Hydra [122], Crassostrea gigas [27], Lymnaea stagnalis [27], and Pinctada fucata of Mollusca [123], as well as most Arthropoda [29], which indicates the main role of piRNAs in TE silencing. piRNAs tend to be antisense to transposons and display a preference for a 5 terminal uridine (1U), while piRNAs are primarily in the sense orientation and exhibit a bias for adenosine at position 10 (10A). Moreover, the 5 terminals of sense–antisense piRNA pairs overlap by precisely 10 nt, a relationship termed the ping-pong signature [33,103].

4.2. Gene Regulation and Development

In addition to having a role in transposon silencing, piRNAs are also involved in the regulation of cellular genes and pseudogenes, which do not exhibit extensive complementarity to transposons [124,125]. Pachytene piRNA-based RISC containing murine Miwi eliminates mRNA from inactivating cellular processes in preparation for sperm production in elongating spermatids [17]. Miwi–CHIL-seq, gene expression profiling, and reporter-based assays further revealed base-pairing between piRNAs and mRNA targets in mouse testes [85]. Meiotic piRNAs might partially regulate mRNA targets via the ping-pong cycle to enable successful spermatogenesis in mice [126]. RNA interference (RNAi) was used to study a single piRNA (fem piRNA) from the silkworm W chromosome, which downregulates z-linked masculinizer (Masc) mRNA in response to primary sex determination [127]. In Drosophila testes, a Y chromosome-specific piRNA induces sex- and paralog-specific gene regulation of pirate, which suggests distinct but related silencing strategies to regulate a conserved protein-coding gene [128]. piRNAs were first demonstrated to engage germ line mRNAs, while tolerating a few mismatches, through perfect pairing at the seed region via miRNA-like pairing rules to regulate gene expression in a model of C. elegans, while CLASH analyses and piRNA reporter assays were used to identify piRNA binding sites in detail [89]. The latest research also revealed a piRNA-mediated maternal mRNA decay during the maternal-to-zygotic transition in Aedes mosquito and Drosophila [98,129]. The role of PIWI–piRNA in gene regulation in development, stem cells, and germ lines has been reviewed previously [130]. Identification of non-transposon piRNA targets is difficult to study in model organisms, and few studies have reported piRNA-mediated gene regulation in non-model animals, although they also possess non-transposon piRNAs [131]. In P. fucata, the somatic piRNAs were presumed to regulate endogenous genes by using locked nucleic acid-modified oligonucleotides (LNA antagonists) to silence specific piRNAs in somatic tissues [123]. piRNA-mediated mRNA silencing will provide comprehensive insights into the post-transcriptional regulatory steps in germ-line gene expression in animals.
Recent studies have shown that piRNAs play critical roles in embryonic development in animals, which regulate transposons to maintain genome integrity from parent to offspring [24,47,132]. During the rediscovery of piRNAs from sRNA-seq, abundant piRNAs were also detected in the embryos or early larvae of diverse organisms such as Drosophila, cuttlefish, clawed frogs, chickens, and ducks. In Nematostella, piRISCs loaded with mature piRNAs cleave the transcripts derived from TEs as well as protein-coding genes in soma, demonstrating that the roles of piRNAs in transposon repression and gene regulation are likely ancestral features that evolved before the split between Cnidaria and Bilateria [24]. The changes in piRNA composition in different chicken germ line developmental stages and the potential roles of PIWI–piRNA pathways in modulating embryonic stage-dependent TE expression were also investigated [132]. In contrast to most animal species, planarian flatworms also expressed piRNAs in adult stem cells known as neoblasts, where they are required not only for germ line development during the postembryonic stage, but also for tissue renewal, regeneration, and starvation [99,133,134]. In addition, the expression of PIWI proteins and piRNAs in the nervous systems of C. elegans [135,136], Drosophila [137], Aplysia [95], and mice [138,139] may be associated with neurogenesis, learning, and memory. piRNAs also play an essential role in the assembly of telomeric chromatin in the Drosophila germ line [140,141]. With the application of sRNA-seq, the roles of piRNAs in the non-canonical functions in animals, especially in embryonic development, nervous system development, and body regeneration, will be progressively discovered.

4.3. Epigenetic Regulation

Strong evidence indicates that PIWI–piRNA pathways play a crucial role in epigenetic regulation. piRNAs guide PIWI proteins to specific target sequences in the genome by sequence complementarity to regulate epigenetic processes via histone modification or DNA methylation [142,143,144,145]. Histone modifications are the predominant means by which epigenetic regulation is transmitted from parents to offspring. DNA methylation is another epigenetic silencing marker that is functionally linked to PIWI. Analyses of mouse Mili and Miwi2 indicated that they mediate DNA methylation in the male germ line during embryogenesis [54,146,147]. It may seem that piRNAs can also direct DNA methylation on non-transposon loci, such as the Ras protein-specific guanine nucleotide-releasing factor 1 (Rasgrf1) locus in the mouse male germ line to regulate genomic imprinting [148] and the CAMP response element-binding protein 2 (CREB2) promoter in Aplysia neurons to influence long-term memory plasticity [95]. Although the molecular mechanism by which piRNAs influence DNA methyltransferases is not clear, the evolutionary conservation of this function is notable. Since piRNAs are involved in epigenetic modifications of gene expression, PIWI–piRNA pathways may play a role in maintaining genome rearrangement and transcriptional or post-transcriptional epigenetic inheritance [38].

4.4. Immune Response

Recently, sufficient evidence supporting the involvement of PIWI–piRNA pathways in protection against invading viruses has been found in mosquitoes, although little is known for other insects [149,150]. Eukaryotic genomes contain virus-derived sequences called endogenous viral elements (EVEs), the majority of which are related to retroviruses, which integrate into the host genome for replication [151]. In addition to transposon repression, recent findings support the possibility of an antiviral role for the PIWI–piRNA pathway, suggesting that piRNAs are derived from fragments of RNA viruses [152,153]. Virus-specific piRNAs have been detected in Drosophila ovarian somatic sheet (OSS) cell lines, which led to the discovery that the cells were persistently infected with several RNA viruses [154]. However, only virus-derived siRNAs were detected in in vivo studies and they mostly had no effect on viral infection in Drosophila mutated for key piRNA pathway proteins [29,155]. In contrast, virus-derived piRNAs, which have ping-pong-specific characteristics, have been reported in a plethora of viral infections, including Reoviridae, Togaviridae, Alphaviruses, and Bunyavirales [156]. However, piRNAs against flaviviruses had no ping-pong signatures, except for a slight 10A-bias [149]. An endogenous viral element from a nonretroviral RNA virus produced a set of piRNAs that provided resistance to infection with a cognate virus in the mosquito Aedes albopictus, analogous to piRNA-mediated TE silencing in the germ line [157]. Knockdown of key piRNA pathway proteins led to enhanced replication of arboviruses in mosquito cells, suggesting their potential antiviral properties in mosquitoes [158,159]. In addition, metagenomic sequencing data of small RNAs also indicated the presence of an endogenous RNA or DNA virus-derived piRNA expression in divergent animal phyla, including Cnidaria, Echinodermata, and Mollusca [21]. More evidence on endogenous viral element-derived piRNAs supports the hypothesis that they mediate antiviral immunity like clustered regularly interspaced short palindromic repeats (CRISPR) RNAs in prokaryotes [151,160].

4.5. Human Diseases (Including Cancer)

Gene expression in cancers is controlled by a variety of regulatory molecules, including small RNAs. Among the three major categories of small RNAs, miRNA profiles in cancers have been extensively characterized, but they are limited in piRNAs. The first report of PIWI expression was in seminomas, a cancer of male germ cells [161]. Since then, ectopic expression of PIWI proteins has been detected in cell lines and tissue samples of a variety of cancers, including those associated with breast, bladder, colorectal, cervical, gastric, liver, and lung cancers [42,43,44,45,162]. A loss-of-function screening for the factors responsible for malignant brain tumors has also demonstrated that PIWI and Aub contribute to tumor growth in Drosophila [163]. Furthermore, piRNAs have also been detected in these cancers [43]. Specifically, piRNAs have been found to be differentially expressed in various cancers and cardiovascular diseases [164]. An increasing number of studies have shown that aberrant PIWI and piRNA expression is a signature feature across multiple tumors, which may serve as a novel therapeutic target and biomarker for cancer detection, classification, and therapy [165]. Interestingly, not all piRNAs interact with PIWI proteins in human tumorigenesis. Depletion of piRNA-like-163 (piR-L-163) resulted in accelerated DNA synthesis and G2-M accumulation, as well as increased invasion and cell migration capabilities in human bronchial epithelial cell lines [166]. This occurred through the specific binding of piR-L-163 to phosphorylated ezrin, radixin, and moesin (ERM proteins), which indicates a novel functional role of piRNAs in tumorigenesis. Remarkably, this also reveals another dimension of the functional role of piRNAs in human cancer independent of PIWI proteins. However, the molecular mechanisms and signaling pathways involved in piRNA function in cancers and cardiovascular diseases have not been fully elucidated [43,45,94]. The study of piRNAs will provide new insights into its potential application in clinical diagnoses, prognoses, and therapeutic strategies against human diseases.

5. Conclusions

piRNAs are a complex category of small RNAs with non-conserved sequences and functions. They participate in germ-line transposon silencing, genome rearrangement, epigenetic regulation, gene regulation, embryonic development, virus defense, and associated human diseases. Existing research in this area cannot be extended to non-model organisms. The development of sRNA-seq using NGS technologies has dramatically increased the number of newly discovered piRNAs in metazoans over the last decade. In the current review, we presented a common pipeline for piRNA research especially suitable for non-model animals. piRNAs were found to be widely expressed in vertebrate and invertebrate soma and germ lines through the reanalysis of existing sRNA-seq data, suggesting that piRNA function might be broader than previously expected. Further research on piRNA processing is needed to facilitate sRNA-seq analyses in non-model animals.

Supplementary Materials

The following are available online at www.mdpi.com/article/10.3390/ijms222011166/s1.

Author Contributions

Conceptualization, S.H. and S.A.; methodology, S.H.; software, S.H. and K.Y.; validation, S.H., and K.Y.; formal analysis, S.H.; investigation, S.H.; resources, S.H.; data curation, S.H.; writing—original draft preparation, S.H.; writing—review and editing, S.A.; visualization, S.H.; supervision, S.A.; project administration, S.A.; funding acquisition, S.H. and S.A. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Japan Society for the Promotion of Science (grant number JP24248034) and the Japan Society for the Promotion of Science Postdoctoral Fellowship for Overseas Researchers (grant number P20395).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

CLIP-seq: crosslinking immunoprecipitation sequencing; circRNA: circular RNA; lncRNA: long non-coding RNA; NGS: next-generation sequencing; piRNA: PIWI-interacting RNA; PTGS: post-transcriptional gene silencing; qRT-PCR: quantitative reverse transcription-polymerase chain reaction; rasiRNA: repeat-associated small interfering RNA; RIP-seq: RNA immunoprecipitation sequencing; RISC: RNA-induced silencing complex; rRNA: ribosomal RNA; siRNA: small interfering RNA; sncRNA: small non-coding RNA; snoRNA: small nucleolar RNA; sRNA-seq: small RNA sequencing; TE: transposon element; TGS: transcriptional gene silencing; tRNA: transfer RNA.

References

  1. Carninci, P. Molecular biology: The long and short of RNAs. Nature 2009, 457, 974–975. [Google Scholar]
  2. Zhang, Y.; Wang, X.; Kang, L. A k-mer scheme to predict piRNAs and characterize locust piRNAs. Bioinformatics 2011, 27, 771–776. [Google Scholar]
  3. Brayet, J.; Zehraoui, F.; Jeanson-Leh, L.; Israeli, D.; Tahi, F. Towards a piRNA prediction using multiple kernel fusion and support vector machine. Bioinformatics 2014, 30, i364–i370. [Google Scholar]
  4. Aravin, A.A.; Naumova, N.M.; Tulin, A.V.; Vagin, V.V.; Rozovsky, Y.M.; Gvozdev, V.A. Double-stranded RNA-mediated silencing of genomic tandem repeats and transposable elements in the D. melanogaster germline. Curr. Biol. 2001, 11, 1017–1127. [Google Scholar]
  5. Aravin, A.A.; Lagos-Quintana, M.; Yalcin, A.; Zavolan, M.; Marks, D.; Snyder, B.; Gaasterland, T.; Meyer, J.; Tuschl, T. The small RNA profile during Drosophila melanogaster development. Dev. Cell 2003, 5, 337–350. [Google Scholar]
  6. Vagin, V.V.; Sigova, A.; Li, C.; Seitz, H.; Gvozdev, V.; Zamore, P.D. A distinct small RNA pathway silences selfish genetic elements in the germline. Science 2006, 313, 320–324. [Google Scholar]
  7. Klattenhoff, C.; Theurkauf, W. Biogenesis and germline functions of piRNAs. Development 2008, 135, 3–9. [Google Scholar]
  8. Kim, V.N. Small RNAs just got bigger: Piwi-interacting RNAs (piRNAs) in mammalian testes. Genes Dev. 2006, 20, 1993–1997. [Google Scholar]
  9. Aravin, A.A.; Gaidatzis, D.; Pfeffer, S.; Lagos-Quintana, M.; Landgraf, P.; Iovino, N.; Morris, P.; Brownstein, M.J.; Kuramochi-Miyagawa, S.; Nakano, T.; et al. A novel class of small RNAs bind to MILI protein in mouse testes. Nature 2006, 442, 203–207. [Google Scholar]
  10. Girard, A.; Sachidanandam, R.; Hannon, G.J.; Carmell, M.A. A germline-specific class of small RNAs binds mammalian Piwi proteins. Nature 2006, 442, 199–202. [Google Scholar]
  11. Lau, N.C.; Seto, A.G.; Kim, J.; Kuramochi-Miyagawa, S.; Nakano, T.; Bartel, D.P.; Kingston, R.E. Characterization of the piRNA complex from rat testes. Science 2006, 313, 363–367. [Google Scholar]
  12. Saito, K.; Nishida, K.M.; Mori, T.; Kawamura, Y.; Miyoshi, K.; Nagami, T.; Siomi, H.; Siomi, M.C. Specific association of Piwi with rasiRNAs derived from retrotransposon and heterochromatic regions in the Drosophila genome. Genes Dev. 2006, 20, 2214–2222. [Google Scholar]
  13. Houwing, S.; Berezikov, E.; Ketting, R.F. Zili is required for germ cell differentiation and meiosis in zebrafish. EMBO J. 2008, 27, 2702–2711. [Google Scholar]
  14. Lee, H.C.; Gu, W.; Shirayama, M.; Youngman, E.; Conte, D.; Mello, C.C. C. elegans piRNAs mediate the genome-wide surveillance of germline transcripts. Cell 2012, 150, 78–87. [Google Scholar]
  15. Sienski, G.; Donertas, D.; Brennecke, J. Transcriptional silencing of transposons by Piwi and Maelstrom and its impact on chromatin state and gene expression. Cell 2012, 151, 964–980. [Google Scholar]
  16. Weick, E.M.; Miska, E.A. piRNAs: From biogenesis to function. Development 2014, 141, 3458–3471. [Google Scholar]
  17. Gou, L.T.; Dai, P.; Yang, J.H.; Xue, Y.; Hu, Y.P.; Zhou, Y.; Kang, J.Y.; Wang, X.; Li, H.; Hua, M.M.; et al. Pachytene piRNAs instruct massive mRNA elimination during late spermiogenesis. Cell Res. 2014, 24, 680–700. [Google Scholar]
  18. Wu, W.S.; Huang, W.C.; Brown, J.S.; Zhang, D.; Song, X.; Chen, H.; Tu, S.; Weng, Z.; Lee, H.C. pirScan: A webserver to predict piRNA targeting sites and to avoid transgene silencing in C. elegans. Nucleic Acids Res. 2018, 46, W43–W48. [Google Scholar]
  19. Zhang, D.; Tu, S.; Stubna, M.; Wu, W.S.; Huang, W.C.; Weng, Z.; Lee, H.C. The piRNA targeting rules and the resistance to piRNA silencing in endogenous genes. Science 2018, 359, 587–592. [Google Scholar]
  20. Lakshmanan, V.; Sujith, T.N.; Bansal, D.; Shivaprasad, P.V.; Palakodeti, D.; Krishna, S. Comprehensive annotation and characterization of planarian tRNA and tRNA-derived fragments (tRFs). RNA 2021, 27, 477–495. [Google Scholar]
  21. Waldron, F.M.; Stone, G.N.; Obbard, D.J. Metagenomic sequencing suggests a diversity of RNA interference-like responses to viruses across multicellular eukaryotes. PLoS Genet. 2018, 14, e1007533. [Google Scholar]
  22. Lim, R.S.; Anand, A.; Nishimiya-Fujisawa, C.; Kobayashi, S.; Kai, T. Analysis of Hydra PIWI proteins and piRNAs uncover early evolutionary origins of the piRNA pathway. Dev. Biol. 2014, 386, 237–251. [Google Scholar]
  23. Juliano, C.E.; Reich, A.; Liu, N.; Götzfried, J.; Zhong, M.; Uman, S.; Reenan, R.A.; Wessel, G.M.; Steele, R.E.; Lin, H. PIWI proteins and PIWI-interacting RNAs function in Hydra somatic stem cells. Proc. Natl. Acad. Sci. USA 2014, 111, 337–342. [Google Scholar]
  24. Praher, D.; Zimmermann, B.; Genikhovich, G.; Columbus-Shenkar, Y.; Modepalli, V.; Aharoni, R.; Moran, Y.; Technau, U. Characterization of the piRNA pathway during development of the sea anemone Nematostella vectensis. RNA Biol. 2017, 14, 1727–1741. [Google Scholar]
  25. Nong, W.; Cao, J.; Li, Y.; Qu, Z.; Sun, J.; Swale, T.; Yip, H.Y.; Qian, P.Y.; Qiu, J.W.; Kwan, H.S.; et al. Jellyfish genomes reveal distinct homeobox gene clusters and conservation of small RNA processing. Nat. Commun. 2020, 11, 3051. [Google Scholar]
  26. Wei, Z.; Liu, X.; Zhang, H. Identification and characterization of piRNA-like small RNAs in the gonad of sea urchin (Strongylocentrotus nudus). Mar. Biotechnol. 2012, 14, 459–467. [Google Scholar]
  27. Jehn, J.; Gebert, D.; Pipilescu, F.; Stern, S.; Kiefer, J.; Hewel, C.; Rosenkranz, D. PIWI genes and piRNAs are ubiquitously expressed in mollusks and show patterns of lineage-specific adaptation. Commun. Biol. 2018, 1, 137. [Google Scholar]
  28. Waiho, K.; Fazhan, H.; Zhang, Y.; Li, S.; Zhang, Y.; Zheng, H.; Ikhwanuddin, M.; Ma, H. Comparative profiling of ovarian and testicular piRNAs in the mud crab Scylla paramamosain. Genomics 2020, 112, 323–331. [Google Scholar]
  29. Lewis, S.H.; Quarles, K.A.; Yang, Y.; Tanguy, M.; Frézal, L.; Smith, S.A.; Sharma, P.P.; Cordaux, R.; Gilbert, C.; Giraud, I.; et al. Pan-arthropod analysis reveals somatic piRNAs as an ancestral defence against transposable elements. Nat. Ecol. Evol. 2018, 2, 174–181. [Google Scholar]
  30. Sun, Y.H.; Zhu, J.; Xie, L.H.; Li, Z.; Meduri, R.; Zhu, X.; Song, C.; Chen, C.; Ricci, E.P.; Weng, Z.; et al. Ribosomes guide pachytene piRNA formation on long intergenic piRNA precursors. Nat. Cell. Biol. 2020, 22, 200–212. [Google Scholar]
  31. Roovers, E.F.; Rosenkranz, D.; Mahdipour, M.; Han, C.T.; He, N.; Chuva de Sousa Lopes, S.M.; van der Westerlaken, L.A.; Zischler, H.; Butter, F.; Roelen, B.A.; et al. Piwi proteins and piRNAs in mammalian oocytes and early embryos. Cell Rep. 2015, 10, 2069–2082. [Google Scholar]
  32. Iwasaki, Y.W.; Siomi, M.C.; Siomi, H. PIWI-Interacting RNA: Its biogenesis and functions. Ann. Rev. Biochem. 2015, 84, 405–433. [Google Scholar]
  33. Czech, B.; Munafò, M.; Ciabrelli, F.; Eastwood, E.L.; Fabry, M.H.; Kneuss, E.; Hannon, G.J. piRNA-guided genome defense: From biogenesis to silencing. Ann. Rev. Genet. 2018, 52, 131–157. [Google Scholar]
  34. Huang, X.; Fejes Tóth, K.; Aravin, A.A. piRNA Biogenesis in Drosophila melanogaster. Trends Genet. 2017, 33, 882–894. [Google Scholar]
  35. Ishizu, H.; Siomi, H.; Siomi, M.C. Biology of PIWI-interacting RNAs: New insights into biogenesis and function inside and outside of germlines. Genes Dev. 2012, 26, 2361–2373. [Google Scholar]
  36. Czech, B.; Hannon, G.J. One loop to rule them all: The ping-pong cycle and piRNA-guided silencing. Trends Biochem. Sci. 2016, 41, 324–337. [Google Scholar]
  37. Gainetdinov, I.; Colpan, C.; Arif, A.; Cecchini, K.; Zamore, P.D. A single mechanism of biogenesis, initiated and directed by PIWI Proteins, explains piRNA production in most animals. Mol. Cell 2018, 71, 775–790. [Google Scholar]
  38. Ross, R.J.; Weiner, M.M.; Lin, H.F. PIWI proteins and PIWI-interacting RNAs in the soma. Nature 2014, 505, 353–359. [Google Scholar]
  39. Tóth, K.F.; Pezic, D.; Stuwe, E.; Webster, A. The piRNA pathway guards the germline genome against transposable elements. Adv. Exp. Med. Biol. 2016, 886, 51–77. [Google Scholar]
  40. Teixeira, F.K.; Okuniewska, M.; Malone, C.D.; Coux, R.X.; Rio, D.C.; Lehmann, R. piRNA-mediated regulation of transposon alternative splicing in the soma and germ line. Nature 2017, 552, 268–272. [Google Scholar]
  41. Sato, K.; Siomi, M.C. The piRNA pathway in Drosophila ovarian germ and somatic cells. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 2020, 96, 32–42. [Google Scholar]
  42. Liu, Y.; Dou, M.; Song, X.; Dong, Y.; Liu, S.; Liu, H.; Tao, J.; Li, W.; Yin, X.; Xu, W. The emerging role of the piRNA/piwi complex in cancer. Mol. Cancer 2019, 18, 123. [Google Scholar]
  43. Guo, B.; Li, D.; Du, L.; Zhu, X. piRNAs: Biogenesis and their potential roles in cancer. Cancer Metastasis Rev. 2020, 39, 567–575. [Google Scholar]
  44. Xu, J.; Yang, X.; Zhou, Q.; Zhuang, J.; Han, S. Biological significance of piRNA in liver cancer, a review. Biomarkers 2020, 25, 436–440. [Google Scholar]
  45. Xin, J.; Du, M.; Jiang, X.; Wu, Y.; Ben, S.; Zheng, R.; Chu, H.; Li, S.; Zhang, Z.; Wang, M. Systematic evaluation of the effects of genetic variants on PIWI-interacting RNA expression across 33 cancer types. Nucleic Acids Res. 2021, 49, 90–97. [Google Scholar]
  46. Tosar, J.P.; Rovira, C.; Cayota, A. Non-coding RNA fragments account for the majority of annotated piRNAs expressed in somatic non-gonadal tissues. Commun. Biol. 2018, 1, 2. [Google Scholar]
  47. Barreñada, O.; Fernández-Pérez, D.; Larriba, E.; Brieño-Enriquez, M.; Del Mazo, J. Diversification of piRNAs expressed in PGCs and somatic cells during embryonic gonadal development. RNA Biol. 2020, 17, 1309–1323. [Google Scholar]
  48. Nishibu, T.; Hayashida, Y.; Tani, S.; Kurono, S.; Kojima-Kita, K.; Ukekawa, R.; Kurokawa, T.; Kuramochi-Miyagawa, S.; Nakano, T.; Inoue, K.; et al. Identification of MIWI-associated Poly(A) RNAs by immunoprecipitation with an anti-MIWI monoclonal antibody. Biosci. Trends 2012, 6, 248–261. [Google Scholar]
  49. Liu, Y.; Li, A.; Xie, G.; Liu, G.; Hei, X. Computational methods and online resources for identification of piRNA-related molecules. Interdiscip. Sci. 2021, 13, 176–191. [Google Scholar] [CrossRef]
  50. Ruby, J.G.; Jan, C.; Player, C.; Axtell, M.J.; Lee, W.; Nusbaum, C.; Ge, H.; Bartel, D.P. Large-scale sequencing reveals 21U-RNAs and additional microRNAs and endogenous siRNAs in C. Elegans. Cell 2006, 127, 1193–1207. [Google Scholar]
  51. Batista, P.J.; Ruby, J.G.; Claycomb, J.M.; Chiang, R.; Fahlgren, N.; Kasschau, K.D.; Chaves, D.A.; Gu, W.; Vasale, J.J.; Duan, S.; et al. PRG-1 and 21U-RNAs interact to form the piRNA complex required for fertility in C. Elegans. Mol. Cell 2008, 31, 67–78. [Google Scholar]
  52. Gu, W.; Shirayama, M.; Conte, D., Jr.; Vasale, J.; Batista, P.J.; Claycomb, J.M.; Moresco, J.J.; Youngman, E.M.; Keys, J.; Stoltz, M.J.; et al. Distinct argonaute-mediated 22G-RNA pathways direct genome surveillance in the C. elegans germline. Mol. Cell 2009, 36, 231–244. [Google Scholar]
  53. Stein, C.B.; Genzor, P.; Mitra, S.; Elchert, A.R.; Ipsaro, J.J.; Benner, L.; Sobti, S.; Su, Y.; Hammell, M.; Joshua-Tor, L.; et al. Decoding the 5’ nucleotide bias of PIWI-interacting RNAs. Nat. Commun. 2019, 10, 828. [Google Scholar]
  54. Rosenkranz, D.; Zischler, H. proTRAC: A software for probabilistic piRNA cluster detection, visualization and analysis. BMC Bioinform. 2012, 13, 5. [Google Scholar]
  55. Aravin, A.A.; Sachidanandam, R.; Bourc’his, D.; Schaefer, C.; Pezic, D.; Toth, K.F.; Bestor, T.; Hannon, G.J. A piRNA pathway primed by individual transposons is linked to de novo DNA methylation in mice. Mol. Cell 2008, 31, 785–799. [Google Scholar]
  56. Wang, K.; Liang, C.; Liu, J.; Xiao, H.; Huang, S.; Xu, J.; Li, F. Prediction of piRNAs using transposon interaction and a support vector machine. BMC Bioinform. 2014, 15, 419. [Google Scholar]
  57. Li, D.; Luo, L.; Zhang, W.; Liu, F.; Luo, F. A genetic algorithm- based weighted ensemble method for predicting transposon-derived piRNAs. BMC Bioinform. 2016, 17, 329. [Google Scholar]
  58. Luo, L.; Li, D.; Zhang, W.; Tu, S.K.; Zhu, X.P.; Tian, G. Accurate prediction of transposon-derived piRNAs by integrating various sequential and physicochemical features. PLoS ONE 2016, 11, e0153268. [Google Scholar]
  59. Liu, B.; Yang, F.; Chou, K.C. 2L-piRNA: A two-layer ensemble classifier for identifying Piwi-interacting RNAs and their function. Mol. Ther. Nucleic Acids 2017, 7, 267–277. [Google Scholar]
  60. Li, T.Y.; Gao, M.Y.; Song, R.Y.; Yin, Q.; Chen, Y. Support vector machine classifier for accurate identification of piRNA. Appl. Sci. 2018, 8, 2204. [Google Scholar]
  61. Zuo, Y.; Zou, Q.; Lin, J.; Jiang, M.; Jiang, M.; Liu, X. 2lpiRNApred: A two-layered integrated algorithm for identifying piRNAs and their functions based on LFE-GM feature selection. RNA Biol. 2020, 17, 892–902. [Google Scholar]
  62. Khan, S.; Khan, M.; Iqbal, N.; Hussain, T.; Khan, S.A.; Chou, K.C. A two-level computation model based on deep learning algo- rithm for identification of piRNA and their functions via Chou’s 5-steps rule. Int. J. Pept. Res. Ther. 2020, 26, 795–809. [Google Scholar]
  63. Chen, C.C.; Qian, X.; Yoon, B.J. Effective computational detection of piRNAs using n-gram models and support vector machine. BMC Bioinform. 2017, 18, 517. [Google Scholar]
  64. Boucheham, A.; Sommard, V.; Zehraoui, F.; Boualem, A.; Batouche, M.; Bendahmane, A.; Israeli, D.; Tahi, F. IpiRId: Integrative approach for piRNA prediction using genomic and epigenomic data. PLoS ONE 2017, 12, e0179787. [Google Scholar]
  65. Wang, K.; Hoeksema, J.; Liang, C. piRNN: Deep learning algorithm for piRNA prediction. PeerJ 2018, 6, e5429. [Google Scholar]
  66. Monga, I.; Banerjee, I. Computational identification of piRNAs using features based on RNA sequence, structure, thermodynamic and physicochemical properties. Curr. Genom. 2019, 20, 508–518. [Google Scholar]
  67. Yuan, J.; Zhang, P.; Cui, Y.; Wang, J.; Skogerbø, G.; Huang, D.W.; Chen, R.; He, S. Computational identification of piRNA targets on mouse mRNAs. Bioinformatics 2016, 32, 1170–1177. [Google Scholar]
  68. Jung, I.; Park, J.C.; Kim, S. piClust: A density based piRNA clustering algorithm. Comput. Biol. Chem. 2014, 50, 60–67. [Google Scholar]
  69. Ray, R.; Pandey, P. piRNA analysis framework from small RNA-Seq data by a novel cluster prediction tool-PILFER. Genomics 2018, 110, 355–365. [Google Scholar]
  70. Liu, Q.; Ding, C.; Lang, X.; Guo, G.; Chen, J.; Su, X. Small noncoding RNA discovery and profiling with sRNAtools based on high-throughput sequencing. Brief. Bioinform. 2021, 22, 463–473. [Google Scholar]
  71. Geles, K.; Palumbo, D.; Sellitto, A.; Giurato, G.; Cianflone, E.; Marino, F.; Torella, D.; Mirici Cappa, V.; Nassa, G.; Tarallo, R.; et al. WIND (Workflow for pIRNAs aNd beyonD): A strategy for in-depth analysis of small RNA-seq data. F1000Res. 2021, 10, 1. [Google Scholar]
  72. Sai, L.S.; Shipra, A. piRNABank: A web resource on classified and clustered Piwi-interacting RNAs. Nucleic Acids Res. 2008, 36, D173–D177. [Google Scholar]
  73. Wang, J.; Zhang, P.; Lu, Y.; Li, Y.; Zheng, Y.; Kan, Y.; Chen, R.; He, S. piRBase: A comprehensive database of piRNA sequences. Nucleic Acids Res. 2019, 47, D175–D180. [Google Scholar]
  74. Zhang, H.; Ali, A.; Gao, J.; Ban, R.; Jiang, X.; Zhang, Y.; Shi, Q. IsopiRBank: A research resource for tracking piRNA isoforms. Database 2018, 59, bay059. [Google Scholar]
  75. Rosenkranz, D.; Zischler, H.; Gebert, D. piRNAclusterDB 2.0: Update and expansion of the piRNA cluster database. Nucleic Acids Res. 2021, 1, 1–6. [Google Scholar]
  76. Jiang, B.R.; Wu, W.Y.; Chien, C.H.; Tsai, J.; Chan, W.L. piRNAtarget: The integrated database for mining functionality of piRNA and its targets. IEEE 2016, 7, 382–386. [Google Scholar]
  77. Wu, W.S.; Brown, J.S.; Chen, T.T.; Chu, Y.H.; Huang, W.C.; Tu, S.; Lee, H.C. piRTarBase: A database of piRNA targeting sites and their roles in gene regulation. Nucleic Acids Res. 2019, 47, D181–D187. [Google Scholar]
  78. Muhammad, A.; Waheed, R.; Khan, N.A.; Jiang, H.; Song, X. piRDisease v1.0: A manually curated database for piRNA associated diseases. Database 2019, 58, 52. [Google Scholar]
  79. Zhang, W.L.; Wu, S.; Zhang, H.Y.; Guan, W.; Zeng, B.H.; Wei, Y.J.; Chan, G.C.; Li, W.Z. piRPheno: A manually curated database to prioritize and analyze human disease related piRNAs. bioRxiv 2020. [Google Scholar] [CrossRef]
  80. Houwing, S.; Kamminga, L.M.; Berezikov, E.; Cronembold, D.; Girard, A.; van den Elst, H.; Filippov, D.V.; Blaser, H.; Raz, E.; Moens, C.B.; et al. A role for Piwi and piRNAs in germ cell maintenance and transposon silencing in zebrafish. Cell 2007, 129, 69–82. [Google Scholar]
  81. Jones, B.C.; Wood, J.G.; Chang, C.; Tam, A.D.; Franklin, M.J.; Siegel, E.R.; Helfand, S.L. A somatic piRNA pathway in the Drosophila fat body ensures metabolic homeostasis and normal lifespan. Nat. Commun. 2016, 7, 13856. [Google Scholar]
  82. Sugimoto, Y.; König, J.; Hussain, S.; Zupan, B.; Curk, T.; Frye, M.; Ule, J. Analysis of CLIP and iCLIP methods for nucleotide-resolution studies of protein-RNA interactions. Genome Biol. 2012, 13, R67. [Google Scholar]
  83. Vourekas, A.; Mourelatos, Z. HITS-CLIP (CLIP-seq) for mouse Piwi proteins. Methods Mol. Biol. 2014, 1093, 73–95. [Google Scholar]
  84. Bortolomeazzi, M.; Gaffo, E.; Bortoluzzi, S. A survey of software tools for microRNA discovery and characterization using RNA-seq. Brief. Bioinform. 2019, 20, 918–930. [Google Scholar]
  85. Zhang, P.; Kang, J.Y.; Gou, L.T.; Wang, J.; Xue, Y.; Skogerboe, G.; Dai, P.; Huang, D.W.; Chen, R.; Fu, X.D.; et al. MIWI and piRNA-mediated cleavage of messenger RNAs in mouse testes. Cell Res. 2015, 25, 193–207. [Google Scholar]
  86. Vourekas, A.; Zheng, K.; Fu, Q.; Maragkakis, M.; Alexiou, P.; Ma, J.; Pillai, R.S.; Mourelatos, Z.; Wang, P.J. The RNA helicase MOV10L1 binds piRNA precursors to initiate piRNA processing. Genes Dev. 2015, 29, 617–629. [Google Scholar]
  87. Munafò, M.; Manelli, V.; Falconio, F.A.; Sawle, A.; Kneuss, E.; Eastwood, E.L.; Seah, J.; Czech, B.; Hannon, G.J. Daedalus and Gasz recruit Armitage to mitochondria, bringing piRNA precursors to the biogenesis machinery. Genes Dev. 2019, 33, 844–856. [Google Scholar]
  88. Van Nostrand, E.L.; Pratt, G.A.; Shishkin, A.A.; Gelboin-Burkhart, C.; Fang, M.Y.; Sundararaman, B.; Blue, S.M.; Nguyen, T.B.; Surka, C.; Elkins, K.; et al. Robust transcriptome-wide discovery of RNA-binding protein binding sites with enhanced CLIP (eCLIP). Nat. Methods 2016, 13, 508–514. [Google Scholar]
  89. Shen, E.Z.; Chen, H.; Ozturk, A.R.; Tu, S.; Shirayama, M.; Tang, W.; Ding, Y.H.; Dai, S.Y.; Weng, Z.; Mello, C.C. Identification of piRNA binding sites reveals the Argonaute regulatory landscape of the C. elegans germline. Cell 2018, 172, 937–951. [Google Scholar]
  90. Ohara, T.; Sakaguchi, Y.; Suzuki, T.; Ueda, H.; Miyauchi, K.; Suzuki, T. The 3′ termini of mouse Piwi-interacting RNAs are 2′-O-methylated. Nat. Struct. Mol. Biol. 2007, 14, 349–350. [Google Scholar]
  91. Kirino, Y.; Mourelatos, Z. The mouse homolog of HEN1 is a potential methylase for Piwi-interacting RNAs. RNA 2007, 13, 1397–1401. [Google Scholar]
  92. Saito, K.; Sakaguchi, Y.; Suzuki, T.; Suzuki, T.; Siomi, H.; Siomi, M.C. Pimet, the Drosophila homolog of HEN1, mediates 2′-O-methylation of Piwi- interacting RNAs at their 3′ ends. Genes Dev. 2007, 21, 1603–1608. [Google Scholar]
  93. Canapa, A.; Barucca, M.; Biscotti, M.A.; Forconi, M.; Olmo, E. Transposons, genome size, and evolutionary insights in animals. Cytogenet. Genome Res. 2015, 147, 217–239. [Google Scholar]
  94. Skinner, D.E.; Rinaldi, G.; Koziol, U.; Brehm, K.; Brindley, P.J. How might flukes and tapeworms maintain genome integrity without a canonical piRNA pathway? Trends Parasitol. 2014, 30, 123–129. [Google Scholar]
  95. Rajasethupathy, P.; Antonov, I.; Sheridan, R.; Frey, S.; Sander, C.; Tuschl, T.; Kandel, E.R. A role for neuronal piRNAs in the epigenetic control of memory-related synaptic plasticity. Cell 2012, 149, 693–707. [Google Scholar]
  96. Yang, L.; Ge, Y.; Cheng, D.; Nie, Z.; Lv, Z. Detection of piRNAs in whitespotted bamboo shark liver. Gene 2016, 590, 51–56. [Google Scholar]
  97. Quarato, P.; Singh, M.; Cornes, E.; Li, B.; Bourdon, L.; Mueller, F.; Didier, C.; Cecere, G. Germline inherited small RNAs facilitate the clearance of untranslated maternal mRNAs in C. elegans embryos. Nat. Commun. 2021, 12, 1441. [Google Scholar]
  98. Grimson, A.; Srivastava, M.; Fahey, B.; Woodcroft, B.J.; Chiang, H.R.; King, N.; Degnan, B.M.; Rokhsar, D.S.; Bartel, D.P. Early origins and evolution of microRNAs and Piwi-interacting RNAs in animals. Nature 2008, 455, 1193–1197. [Google Scholar]
  99. Zhou, X.; Battistoni, G.; El Demerdash, O.; Gurtowski, J.; Wunderer, J.; Falciatori, I.; Ladurner, P.; Schatz, M.C.; Hannon, G.J.; Wasik, K.A. Dual functions of Macpiwi1 in transposon silencing and stem cell maintenance in the flatworm Macrostomum lignano. RNA 2015, 21, 1885–1897. [Google Scholar]
  100. Queiroz, F.R.; Portilho, L.G.; Jeremias, W.J.; Babá, É.H.; do Amaral, L.R.; Silva, L.M.; Coelho, P.; Caldeira, R.L.; Gomes, M.S. Deep sequencing of small RNAs reveals the repertoire of miRNAs and piRNAs in Biomphalaria glabrata. Memórias Inst. Oswaldo Cruz 2020, 115, e190498. [Google Scholar]
  101. Huang, S.Q.; Ichikawa, Y.; Igarashi, Y.; Yoshitake, K.; Kinoshita, S.; Omori, F.; Maeyama, K.; Nagai, K.; Watabe, S.; Asakawa, S. Piwi-interacting RNA (piRNA) expression patterns in pearl oyster (Pinctada fucata) somatic tissues. Sci. Rep. 2019, 9, 247. [Google Scholar]
  102. Toombs, J.A.; Sytnikova, Y.A.; Chirn, G.W.; Ang, I.; Lau, N.C.; Blower, M.D. Xenopus Piwi proteins interact with a broad proportion of the oocyte transcriptome. RNA 2017, 23, 504–520. [Google Scholar]
  103. Vandewege, M.W.; Platt, R.N.; Ray, D.A.; Hoffmann, F.G. Transposable element targeting by piRNAs in Laurasiatherians with distinct transposable element histories. Genome Biol. Evol. 2016, 8, 1327–1337. [Google Scholar]
  104. Li, B.; He, X.; Zhao, Y.; Bai, D.; Bou, G.; Zhang, X.; Su, S.; Dao, L.; Liu, R.; Wang, Y.; et al. Identification of piRNAs and piRNA clusters in the testes of the Mongolian horse. Sci. Rep. 2019, 9, 5022. [Google Scholar]
  105. Chirn, G.W.; Rahman, R.; Sytnikova, Y.A.; Matts, J.A.; Zeng, M.; Gerlach, D.; Yu, M.; Berger, B.; Naramura, M.; Kile, B.T.; et al. Conserved piRNA Expression from a Distinct Set of piRNA Cluster Loci in Eutherian Mammals. PLoS Genet. 2015, 11, e1005652. [Google Scholar]
  106. Hirano, T.; Iwasaki, Y.W.; Lin, Z.Y.; Imamura, M.; Seki, N.M.; Sasaki, E.; Saito, K.; Okano, H.; Siomi, M.C.; Siomi, H. Small RNA profiling and characterization of piRNA clusters in the adult testes of the common marmoset, a model primate. RNA 2014, 20, 1223–1237. [Google Scholar]
  107. Williams, Z.; Morozov, P.; Mihailovic, A.; Lin, C.; Puvvula, P.K.; Juranek, S.; Rosenwaks, Z.; Tuschl, T. Discovery and characterization of piRNAs in the human fetal ovary. Cell Rep. 2015, 13, 854–863. [Google Scholar]
  108. Siomi, M.C.; Sato, K.; Pezic, D.; Aravin, A.A. PIWI-interacting small RNAs: The vanguard of genome defence. Nat. Rev. Genet. 2011, 12, 246–258. [Google Scholar]
  109. Gan, H.; Lin, X.; Zhang, Z.; Zhang, W.; Liao, S.; Wang, L.; Han, C. piRNA profiling during specific stages of mouse spermatogenesis. RNA 2011, 17, 1191–1203. [Google Scholar]
  110. Brennecke, J.; Aravin, A.A.; Stark, A.; Dus, M.; Kellis, M.; Sachidanandam, R.; Hannon, G.J. Discrete small RNA-generating loci as master regulators of transposon activity in Drosophila. Cell 2007, 128, 1089–1103. [Google Scholar]
  111. Saito, K.; Ishizu, H.; Komai, M.; Kotani, H.; Kawamura, Y.; Nishida, K.M.; Siomi, H.; Siomi, M.C. Roles for the Yb body components Armitage and Yb in primary piRNA biogenesis in Drosophila. Genes Dev. 2010, 24, 2493–2498. [Google Scholar]
  112. Wang, S.H.; Elgin, S.C. Drosophila Piwi functions downstream of piRNA production mediating a chromatin-based transposon silencing mechanism in female germ line. Proc. Natl. Acad. Sci. USA 2011, 108, 21164–21169. [Google Scholar]
  113. Le Thomas, A.; Rogers, A.K.; Webster, A.; Marinov, G.K.; Liao, S.E.; Perkins, E.M.; Hur, J.K.; Aravin, A.A.; Tóth, K.F. Piwi induces piRNA-guided transcriptional silencing and establishment of a repressive chromatin state. Genes Dev. 2013, 27, 390–399. [Google Scholar]
  114. Rozhkov, N.V.; Hammell, M.; Hannon, G.J. Multiple roles for Piwi in silencing Drosophila transposons. Genes Dev. 2013, 27, 400–412. [Google Scholar]
  115. Onishi, R.; Yamanaka, S.; Siomi, M.C. piRNA- and siRNA-mediated transcriptional repression in Drosophila, mice, and yeast: New insights and biodiversity. EMBO Rep. 2021, 22, e53062. [Google Scholar]
  116. Yang, F.; Lan, Y.; Pandey, R.R.; Homolka, D.; Berger, S.L.; Pillai, R.S.; Bartolomei, M.S.; Wang, P.J. TEX15 associates with MILI and silences transposable elements in male germ cells. Genes Dev. 2020, 34, 745–750. [Google Scholar]
  117. Schöpp, T.; Zoch, A.; Berrens, R.V.; Auchynnikava, T.; Kabayama, Y.; Vasiliauskaitė, L.; Rappsilber, J.; Allshire, R.C.; O’Carroll, D. TEX15 is an essential executor of MIWI2-directed transposon DNA methylation and silencing. Nat. Commun. 2020, 11, 3739. [Google Scholar]
  118. Zoch, A.; Auchynnikava, T.; Berrens, R.V.; Kabayama, Y.; Schöpp, T.; Heep, M.; Vasiliauskaitė, L.; Pérez-Rico, Y.A.; Cook, A.G.; Shkumatava, A.; et al. SPOCD1 is an essential executor of piRNA-directed de novo DNA methylation. Nature 2020, 584, 635–639. [Google Scholar]
  119. Mugat, B.; Nicot, S.; Varela-Chavez, C.; Jourdan, C.; Sato, K.; Basyuk, E.; Juge, F.; Siomi, M.C.; Pélisson, A.; Chambeyron, S. The Mi-2 nucleosome remodeler and the Rpd3 histone deacetylase are involved in piRNA-guided heterochromatin formation. Nat. Commun. 2020, 11, 2818. [Google Scholar]
  120. Ninova, M.; Chen, Y.A.; Godneeva, B.; Rogers, A.K.; Luo, Y.; Fejes Tóth, K.; Aravin, A.A. Su(var)2-10 and the SUMO pathway link piRNA-guided target recognition to chromatin ailencing. Mol. Cell 2020, 77, 556–570. [Google Scholar]
  121. Wei, K.H.; Chan, C.; Bachtrog, D. Establishment of H3K9me3-dependent heterochromatin during embryogenesis in Drosophila miranda. eLife 2021, 10, e55612. [Google Scholar]
  122. Teefy, B.B.; Siebert, S.; Cazet, J.F.; Lin, H.; Juliano, C.E. PIWI-piRNA pathway-mediated transposable element repression in Hydra somatic stem cells. RNA 2020, 26, 550–563. [Google Scholar]
  123. Huang, S.Q.; Ichikawa, Y.; Yoshitake, K.; Kinoshita, S.; Asaduzzaman, M.; Omori, F.; Maeyama, K.; Nagai, K.; Watabe, S.; Asakawa, S. Conserved and widespread expression of piRNA-like molecules and PIWI-like genes reveal dual functions of transposon silencing and gene regulation in Pinctada fucata (Mollusca). Front. Mar. Sci. 2021, 8, 730556. [Google Scholar]
  124. Watanabe, T.; Lin, H.F. Posttranscriptional regulation of gene expression by Piwi proteins and piRNAs. Mol. Cell 2014, 56, 18–27. [Google Scholar]
  125. Ramat, A.; Simonelig, M. Functions of PIWI proteins in gene regulation: New arrows added to the piRNA quiver. Trends Genet. 2021, 37, 188–200. [Google Scholar]
  126. Goh, W.S.; Falciatori, I.; Tam, O.H.; Burgess, R.; Meikar, O.; Kotaja, N.; Hammell, M.; Hannon, G.J. piRNA-directed cleavage of meiotic transcripts regulates spermatogenesis. Genes Dev. 2015, 29, 1032–1044. [Google Scholar]
  127. Kiuchi, T.; Koga, H.; Kawamoto, M.; Shoji, K.; Sakai, H.; Arai, Y.; Ishihara, G.; Kawaoka, S.; Sugano, S.; Shimada, T.; et al. A single female-specific piRNA is the primary determiner of sex in the silkworm. Nature 2014, 509, 633–636. [Google Scholar]
  128. Chen, P.; Kotov, A.A.; Godneeva, B.K.; Bazylev, S.S.; Olenina, L.V.; Aravin, A.A. piRNA-mediated gene regulation and adaptation to sex-specific transposon expression in D. melanogaster male germline. Genes Dev. 2021, 35, 1–22. [Google Scholar]
  129. Halbach, R.; Miesen, P.; Joosten, J.; Taşköprü, E.; Rondeel, I.; Pennings, B.; Vogels, C.; Merkling, S.H.; Koenraadt, C.J.; Lambrechts, L.; et al. A satellite repeat-derived piRNA controls embryonic development of Aedes. Nature 2020, 580, 274–277. [Google Scholar]
  130. Rojas-Ríos, P.; Simonelig, M. piRNAs and PIWI proteins: Regulators of gene expression in development and stem cells. Development 2018, 145, dev161786. [Google Scholar]
  131. Gebert, D.; Ketting, R.F.; Zischler, H.; Rosenkranz, D. piRNAs from pig testis provide evidence for a conserved role of the Piwi pathway in posttranscriptional gene regulation in mammals. PLoS ONE 2015, 10, e0124860. [Google Scholar]
  132. Chang, K.W.; Tseng, Y.T.; Chen, Y.C.; Yu, C.Y.; Liao, H.F.; Chen, Y.C.; Tu, Y.E.; Wu, S.C.; Liu, I.H.; Pinskaya, M.; et al. Stage-dependent piRNAs in chicken implicated roles in modulating male germ cell development. BMC Genom. 2018, 19, 425. [Google Scholar]
  133. De Mulder, K.; Pfister, D.; Kuales, G.; Egger, B.; Salvenmoser, W.; Willems, M.; Steger, J.; Fauster, K.; Micura, R.; Borgonie, G.; et al. Stem cells are differentially regulated during development, regeneration and homeostasis in flatworms. Dev. Biol. 2009, 334, 198–212. [Google Scholar]
  134. Cao, Z.; Rosenkranz, D.; Wu, S.; Liu, H.; Pang, Q.; Zhang, X.; Liu, B.; Zhao, B. Different classes of small RNAs are essential for head regeneration in the planarian Dugesia japonica. BMC Genom. 2020, 21, 876. [Google Scholar]
  135. Kim, K.W.; Tang, N.H.; Andrusiak, M.G.; Wu, Z.; Chisholm, A.D.; Jin, Y. A neuronal piRNA pathway inhibits axon regeneration in C. elegans. Neuron 2018, 97, 511–519. [Google Scholar]
  136. Posner, R.; Toker, I.A.; Antonova, O.; Star, E.; Anava, S.; Azmon, E.; Hendricks, M.; Bracha, S.; Gingold, H.; Rechavi, O. Neuronal small RNAs control behavior transgenerationally. Cell 2019, 177, 1814–1826. [Google Scholar]
  137. Perrat, P.N.; DasGupta, S.; Wang, J.; Theurkauf, W.; Weng, Z.; Rosbash, M.; Waddell, S. Transposition-driven genomic heterogeneity in the Drosophila brain. Science 2013, 340, 91–95. [Google Scholar]
  138. Leighton, L.J.; Wei, W.; Marshall, P.R.; Ratnu, V.S.; Li, X.; Zajaczkowski, E.L.; Spadaro, P.A.; Khandelwal, N.; Kumar, A.; Bredy, T.W. Disrupting the hippocampal Piwi pathway enhances contextual fear memory in mice. Neurobiol. Learn. Mem. 2019, 161, 202–209. [Google Scholar]
  139. Nandi, S.; Chandramohan, D.; Fioriti, L.; Melnick, A.M.; Hebert, J.M.; Mason, C.E.; Rajasethupathy, P.; Kandel, E.R. Roles for small noncoding RNAs in silencing of retrotransposons in the mammalian brain. Proc. Natl. Acad. Sci. USA 2016, 113, 12697–12702. [Google Scholar]
  140. Radion, E.; Morgunova, V.; Ryazansky, S.; Akulenko, N.; Lavrov, S.; Abramov, Y.; Komarov, P.A.; Glukhov, S.I.; Olovnikov, I.; Kalmykova, A. Key role of piRNAs in telomeric chromatin maintenance and telomere nuclear positioning in Drosophila germline. Epigenetics Chromatin 2018, 11, 40. [Google Scholar]
  141. Kordyukova, M.; Olovnikov, I.; Kalmykova, A. Transposon control mechanisms in telomere biology. Curr. Opin. Genet. Dev. 2018, 49, 56–62. [Google Scholar]
  142. Peng, J.C.; Lin, H.F. Beyond transposons: The epigenetic and somatic functions of the Piwi-piRNA mechanism. Curr. Opin. Cell Biol. 2013, 25, 190–194. [Google Scholar]
  143. Liu, J.; Zhang, S.; Cheng, B. Epigenetic roles of PIWI-interacting RNAs (piRNAs) in cancer metastasis (Review). Oncol. Rep. 2018, 40, 2423–2434. [Google Scholar]
  144. Sadoughi, F.; Mirhashemi, S.M.; Asemi, Z. Epigenetic roles of PIWI proteins and piRNAs in colorectal cancer. Cancer Cell Int. 2021, 21, 328. [Google Scholar]
  145. Pathania, A.S.; Prathipati, P.; Pandey, M.K.; Byrareddy, S.N.; Coulter, D.W.; Gupta, S.C.; Challagundla, K.B. The emerging role of non-coding RNAs in the epigenetic regulation of pediatric cancers. Semin. Cancer Biol. 2021, in press. [Google Scholar] [CrossRef]
  146. Aravin, A.A.; Sachidanandam, R.; Girard, A.; Fejes-Toth, K.; Hannon, G.J. Developmentally regulated piRNA clusters implicate MILI in transposon control. Science 2007, 316, 744–747. [Google Scholar]
  147. Kuramochi-Miyagawa, S.; Watanabe, T.; Gotoh, K.; Totoki, Y.; Toyoda, A.; Ikawa, M.; Asada, N.; Kojima, K.; Yamaguchi, Y.; Ijiri, T.W.; et al. DNA methylation of retrotransposon genes is regulated by Piwi family members MILI and MIWI2 in murine fetal testes. Genes Dev. 2008, 22, 908–917. [Google Scholar]
  148. Watanabe, T.; Tomizawa, S.; Mitsuya, K.; Totoki, Y.; Yamamoto, Y.; Kuramochi-Miyagawa, S.; Iida, N.; Hoki, Y.; Murphy, P.J.; Toyoda, A.; et al. Role for piRNAs and noncoding RNA in de novo DNA methylation of the imprinted mouse Rasgrf1 locus. Science 2011, 332, 848–852. [Google Scholar]
  149. Leggewie, M.; Schnettler, E. RNAi-mediated antiviral immunity in insects and their possible application. Curr. Opin. Virol. 2018, 32, 108–114. [Google Scholar]
  150. Kolliopoulou, A.; Santos, D.; Taning, C.; Wynant, N.; Vanden Broeck, J.; Smagghe, G.; Swevers, L. PIWI pathway against viruses in insects. Wiley Interdiscip. Rev. RNA 2019, 10, e1555. [Google Scholar]
  151. Ophinni, Y.; Palatini, U.; Hayashi, Y.; Parrish, N.F. piRNA-guided CRISPR-like immunity in eukaryotes. Trends Immunol. 2019, 40, 998–1010. [Google Scholar]
  152. Palatini, U.; Miesen, P.; Carballar-Lejarazu, R.; Ometto, L.; Rizzo, E.; Tu, Z.; van Rij, R.P.; Bonizzoni, M. Comparative genomics shows that viral integrations are abundant and express piRNAs in the arboviral vectors Aedes aegypti and Aedes albopictus. BMC Genom. 2017, 81, 512. [Google Scholar]
  153. Palatini, U.; Masri, R.A.; Cosme, L.V.; Koren, S.; Thibaud-Nissen, F.; Biedler, J.K.; Krsticevic, F.; Johnston, J.S.; Halbach, R.; Crawford, J.E.; et al. Improved reference genome of the arboviral vector Aedes albopictus. Genome Biol. 2020, 21, 215. [Google Scholar]
  154. Wu, Q.; Luo, Y.; Lu, R.; Lau, N.; Lai, E.C.; Li, W.X.; Ding, S.W. Virus discovery by deep sequencing and assembly of virus-derived small silencing RNAs. Proc. Natl. Acad. Sci. USA 2010, 107, 1606–1611. [Google Scholar]
  155. Petit, M.; Mongelli, V.; Frangeul, L.; Blanc, H.; Jiggins, F.; Saleh, M.C. piRNA pathway is not required for antiviral defense in Drosophila melanogaster. Proc. Natl. Acad. Sci. USA 2016, 113, E4218–E4227. [Google Scholar]
  156. Miesen, P.; Joosten, J.; van Rij, R.P. PIWIs go viral: Arbovirus-derived piRNAs in vector mosquitoes. PLoS Pathog. 2016, 12, e1006017. [Google Scholar]
  157. Suzuki, Y.; Baidaliuk, A.; Miesen, P.; Frangeul, L.; Crist, A.B.; Merkling, S.H.; Fontaine, A.; Lequime, S.; Moltini-Conclois, I.; Blanc, H.; et al. Non-retroviral endogenous Viral element limits cognate virus replication in Aedes aegypti ovaries. Curr. Biol. 2020, 30, 3495–3506. [Google Scholar]
  158. Schnettler, E.; Donald, C.L.; Human, S.; Watson, M.; Siu, R.; McFarlane, M.; Fazakerley, J.K.; Kohl, A.; Fragkoudis, R. Knockdown of piRNA pathway proteins results in enhanced semliki forest virus production in mosquito cells. J. Gen. Virol. 2013, 94, 1680–1689. [Google Scholar]
  159. Miesen, P.; Girardi, E.; Van Rij, R.P. Distinct sets of PIWI proteins produce arbovirus and transposon-derived piRNAs in Aedes aegypti mosquito cells. Nucleic Acids Res. 2015, 43, 6545–6556. [Google Scholar]
  160. Crava, C.M.; Varghese, F.S.; Pischedda, E.; Halbach, R.; Palatini, U.; Marconcini, M.; Gasmi, L.; Redmond, S.; Afrane, Y.; Ayala, D.; et al. Population genomics in the arboviral vector Aedes aegypti reveals the genomic architecture and evolution of endogenous viral elements. Mol. Ecol. 2021, 30, 1594–1611. [Google Scholar]
  161. Qiao, D.; Zeeman, A.M.; Deng, W.; Looijenga, L.H.; Lin, H.F. Molecular characterization of hiwi, a human member of the piwi gene family whose over expression is correlated to seminomas. Oncogene 2002, 21, 3988–3999. [Google Scholar]
  162. Suzuki, R.; Honda, S.; Kirino, Y. PIWI expression and function in cancer. Front. Genet. 2012, 3, 204. [Google Scholar]
  163. Janic, A.; Mendizabal, L.; Llamazares, S.; Rossell, D.; Gonzalez, C. Ectopic expression of germline genes drives malignant brain tumor growth in Drosophila. Science 2010, 330, 1824–1827. [Google Scholar]
  164. Li, M.; Yang, Y.; Wang, Z.; Zong, T.; Fu, X.; Aung, L.; Wang, K.; Wang, J.X.; Yu, T. Piwi-interacting RNAs (piRNAs) as potential biomarkers and therapeutic targets for cardiovascular diseases. Angiogenesis 2021, 24, 19–34. [Google Scholar]
  165. Chalbatani, G.M.; Dana, H.; Memari, F.; Gharagozlou, E.; Ashjaei, S.; Kheirandish, P.; Marmari, V.; Mahmoudzadeh, H.; Mozayani, F.; Maleki, A.R.; et al. Biological function and molecular mechanism of piRNA in cancer. Pract. Lab. Med. 2018, 13, e00113. [Google Scholar]
  166. Mei, Y.; Wang, Y.; Kumari, P.; Shetty, A.C.; Clark, D.; Gable, T.; MacKerell, A.D.; Ma, M.Z.; Weber, D.J.; Yang, A.J.; et al. A piRNA-like small RNA interacts with and modulates p-ERM proteins in human somatic cells. Nat. Commun. 2015, 6, 7316. [Google Scholar]
Figure 2. Discovery of piRNAs in metazoans. In the same taxa, the proportion of TEs increases with the genome size, whereas the number of piRNA species does not increase with the size of the genome or the proportion of TEs.
Figure 2. Discovery of piRNAs in metazoans. In the same taxa, the proportion of TEs increases with the genome size, whereas the number of piRNA species does not increase with the size of the genome or the proportion of TEs.
Ijms 22 11166 g002
Figure 3. piRNA biogenesis and its functional roles in metazoans. In most cases, the piRNA pathway begins with transcription of piRNA clusters, which is mediated by RNA polymerase II (pol II), to generate the respective precursor piRNA (pre-piRNA) transcripts in the nucleus and drive them to cytoplasm where primary and second piRNA biogenesis takes place. The primary transcripts of piRNA clusters are shortened into piRNA intermediates and subsequently loaded onto PIWI proteins and trimmed from the 3′ end to the size of mature piRNAs and then 2′-O-methylated. The mature piRNAs interact with PIWI proteins to form piRISC, which serves various functions in the nucleus and cytoplasm. piRISC is translocated to the nucleus and targets the nascent transcripts through sequence complementarity. Upon binding, PIWI recruits the epigenetic modifier heterochromatin protein 1 (HP1a) and histone methyltransferase (HMT) to a methyl group on unmethylated histone 3 lysine 9 (H3K9) to inhibit pol II transcription, effectively silencing transcription of the gene or TE. The piRNA pathway may also start with a transcript of a protein-coding gene, viral DNA (vDNA), or an invasive viral RNA in the cytoplasm in order to silence the transcript through the ping-pong amplification loop.
Figure 3. piRNA biogenesis and its functional roles in metazoans. In most cases, the piRNA pathway begins with transcription of piRNA clusters, which is mediated by RNA polymerase II (pol II), to generate the respective precursor piRNA (pre-piRNA) transcripts in the nucleus and drive them to cytoplasm where primary and second piRNA biogenesis takes place. The primary transcripts of piRNA clusters are shortened into piRNA intermediates and subsequently loaded onto PIWI proteins and trimmed from the 3′ end to the size of mature piRNAs and then 2′-O-methylated. The mature piRNAs interact with PIWI proteins to form piRISC, which serves various functions in the nucleus and cytoplasm. piRISC is translocated to the nucleus and targets the nascent transcripts through sequence complementarity. Upon binding, PIWI recruits the epigenetic modifier heterochromatin protein 1 (HP1a) and histone methyltransferase (HMT) to a methyl group on unmethylated histone 3 lysine 9 (H3K9) to inhibit pol II transcription, effectively silencing transcription of the gene or TE. The piRNA pathway may also start with a transcript of a protein-coding gene, viral DNA (vDNA), or an invasive viral RNA in the cytoplasm in order to silence the transcript through the ping-pong amplification loop.
Ijms 22 11166 g003
Table 1. Discovery of piRNAs by sRNA-seq in metazoans.
Table 1. Discovery of piRNAs by sRNA-seq in metazoans.
PhylumCommon NameSpeciespiRNA ExpressionSources
NematodaNematodeCaenorhabditis elegansWhole organism[14,18,19,89,97]
PoriferaSpongeAmphimedon queenslandicaWhole organism[98]
SpongeHalichondria paniceaWhole organism[21]
PlatyhelminthesPlanarianSchmidtea mediterraneaWhole organism[20]
FlatwormMacrostomum lignanoWhole organism[99]
FlukeSchistosoma japonicum, S. haematobium, and S. mansoninon-piRNAsRNAome
Liver flukeClonorchis sinensisnon-piRNAsRNAome
TapewormEchinococcus canadensis and E. granulosusnon-piRNAsRNAome
AnnelidaEarthwormLumbricus and Amynthas sppsBodywall[21]
EarthwormEisenia fetidaBodywallsRNAome
CnidariaHydraHydra vulgaris and H. magnipapillataWhole organism and soma[22,23]
CoralAcropora muricata, Stylophora pistillata, Montipora capricornis, Montipora foliosa, and Pocillopora damicornisPolypssRNAome
Sea anemoneNematostella vectensisWhole organism at different stages[24]
Beadlet anenomeActinia equinaPolyps[21]
Sea anemoneAiptasia pallidaPolypssRNAome
JellyfishSanderia malayensis, Rhopilema esculentum, and Aurelia auritaAppendages, tentacles, rhopalia, oral arms, gonads[25]
EchinodermataStarfishAsterias rubensTube foot[21]
Sea cucumberApostichopus japonicusRespiratory tree, tube foot, intestine, body wallsRNAome
Sea urchinStrongylocentrotus intermedius and S. nudusTube foot, larvaesRNAome
MolluscaPacific oysterCrassostrea gigasReproductive tract, foot muscle[27]
Great pond snailLymnaea stagnalisReproductive tract, foot muscle[27]
SnailBiomphalaria glabrataAdult snail[100]
Pearl oysterPinctada fucataAdductor, gill, gonad, mantle[101]
Pacific abaloneHaliotis discus hannaiAdductor musclesRNAome
Common musselMytilus galloprovincialisHemolymphsRNAome
Manila clamRuditapes philippinarumMantlesRNAome
Blood clamScapharca broughtoniiHaemocytesRNAome
Ark shellTegillarca granosaHaemocytesRNAome
CuttlefishSepiella japonicaLarvaesRNAome
PeriwinkleLittorina littoreanon-piRNAsRNAome
Sea snailRapana venosanon-piRNAsRNAome
Garden snailHelix lucorumnon-piRNAsRNAome
Pearl musselHyriopsis cumingiinon-piRNAsRNAome
CrustaceaMud crabScylla paramamosainOvary, testis[28]
Swimming crabPortunus trituberculatusOvary, testissRNAome
Black tiger shrimpPenaeus monodonOvarysRNAome
ArthropodaFruitflyDrosophila melanogaster and D. virilisGermline, thorax, embryo[29]
FruitflyD. willistoni, D. simulans, D. sechellia, D. pseudoobscura, D. persimilis, D. mojavensis, D. grimshawi, D. erecta, D. ananassae, and D. yakubaGermline, thorax, head, embryosRNAome
HouseflyMusca domesticaGermline, thorax[29]
Pea aphidAcyrthosiphon pisumGermline, thorax[29]
MosquitoAedes aegyptiGermline, thorax[29]
Honey beeApis melliferaGermline, thorax[29]
Bumble beeBombus terrestrisGermline, thorax[29]
RootwormDiabrotica virgiferaGermline, thorax[29]
Postman butterflyHeliconius melpomeneGermline, thorax[29]
Horseshoe crabLimulus polyphemusGermline, thorax[29]
BeetleNicrophorus vespilloidesGermline, thorax[29]
Lygaeid bugOncopeltus fasciatusGermline, thorax[29]
SpiderParasteatoda tepidariorumGermline, mesosoma[29]
Diamondback mothPlutella xylostellaGermline, thorax[29]
CentipedeStrigamia maritimaFat body, nerve chord[29]
Red flour beetleTribolium castaneumGermline, thorax[29]
NoctuidTrichoplusia niGermline, thorax[29]
ScorpionCentruroides sculpturatusGermline, prosoma[29]
ChordataAmur sturgeonAcipenser schrenckiiOvary, testissRNAome
(Fish)Elephant sharkCallorhinchus miliiOvary, testissRNAome
Bamboo sharkChiloscyllium plagiosumLiver[96]
Epaulette sharkHemiscyllium ocellatumNon-piRNAsRNAome
ZebrafishDanio rerioOvary, testis[80]
MedakaOryzias melastigmaOvary, testissRNAome
PufferfishTakifugu rubripesOvary, testissRNAome
Nile tilapiaOreochromis niloticusOvary, testissRNAome
Rainbow troutOncorhynchus mykissOvary, testissRNAome
Yellow catfishTachysurus fulvidracoOvary, testissRNAome
SticklebackGasterosteus aculeatusOvary, testissRNAome
Ricefield eelMonopterus albusMix of brain, liver, and gonad sRNAome
(Amphibian)Clawed frogXenopus tropicalis and X. laevisOvary, embryo[102]
(Reptilia)AlligatorAlligator sinensisOvarysRNAome
TurtlePelodiscus sinensisOvary, testissRNAome
TortoiseMauremys reevesiiOvary, testissRNAome
Lizard Anolis carolinensisNon-piRNAsRNAome
(Aves)ChickenGallus gallusOvary, testis, embryosRNAome
BudgerigarMelopsittacus undulatusOvary, testissRNAome
DuckAnas platyrhynchosEmbryosRNAome
GooseAnser cygnoidesOvarysRNAome
PigeonColumba liviaOvarysRNAome
(Mammal)BatEptesicus fuscusTestis[103]
PlatypusOrnithorhynchus anatinusTestissRNAome
HouseEquus caballusTestis[104]
SheepOvis ariesTestissRNAome
DogCanis lupus familiarisTestissRNAome
RabbitOryctolagus cuniculusTestis, ovary (sRNAome)[105]
CowBos taurusTestissRNAome
PigSus scrofaTestissRNAome
MouseMus musculusTestis[105]
RatRattus norvegicusTestissRNAome
OpossumMonodelphis domesticaTestissRNAome
MacaqueMacaca mulattaTestissRNAome
MachinMacaca fascicularisOvary[31]
MarmosetCallithrix jacchusTestis[106]
HumanHomo sapiensTestis[31,107]
The sRNAome indicated that the piRNAs were discovered from the sRNA-seq data, which were used for the detection of miRNAs but not piRNAs. The datasets did not include all published sRNA-seq data from specific animals or all known animals. In each animal taxon, several representative species were selected for piRNA rediscovery to evaluate the type and quantity of piRNA species during the animal evolution process. The data sources for sRNA-seq are shown in Supplementary Table S1.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Huang, S.; Yoshitake, K.; Asakawa, S. A Review of Discovery Profiling of PIWI-Interacting RNAs and Their Diverse Functions in Metazoans. Int. J. Mol. Sci. 2021, 22, 11166. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms222011166

AMA Style

Huang S, Yoshitake K, Asakawa S. A Review of Discovery Profiling of PIWI-Interacting RNAs and Their Diverse Functions in Metazoans. International Journal of Molecular Sciences. 2021; 22(20):11166. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms222011166

Chicago/Turabian Style

Huang, Songqian, Kazutoshi Yoshitake, and Shuichi Asakawa. 2021. "A Review of Discovery Profiling of PIWI-Interacting RNAs and Their Diverse Functions in Metazoans" International Journal of Molecular Sciences 22, no. 20: 11166. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms222011166

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop