Microproteins: from behind the scenes to the spotlight

来源: 发布时间:2021-08-31 16:56:35 浏览次数: 【字体:

Genome Instability & Disease volume 2pages225–239 (2021)


Abstract

Microproteins, less than 200 amino acids, have been usually regarded as non-functional for a long time. We do not recognize this “dark matter” in the proteome until developing new bioinformatic, biochemical and proteomic techniques. The fast-growing number of the identified microproteins and their emerging roles in various biological processes begin to attract more and more attention. Here, we summarize the recent progress in this field mainly from four aspects: (1) brief history and various definitions of microproteins; (2) three main strategies to identify microproteins; (3) multiple regulatory mechanisms of microproteins; (4) biological functions of microproteins, with a particular focus on genome stability maintenance and their relevance to human diseases. Finally, we discuss some open questions and therapeutic opportunities hiding in these small newcomers.

Introduction

In this review, we define “microproteins” as functional proteins from 2 amino acids (aa) and up to 200 aa, which are usually encoded by small open reading frames (sORFs/smORFs) (Basrai et al., 1997; Magnani et al., 2014; Yang et al., 2011). Nowadays, researchers in this field agree that microproteins are small in size, numerous in variety with underestimated functions thus far due to several reasons. First, microproteins are routinely discarded during the automatic gene annotation process in the past because smORFs less than 100 codons are considered untranslatable and “meaningless” in the eukaryotic genomes (Basrai et al., 1997; Chugunova et al., 2018; Frith et al., 2006). Second, eukaryotic mRNA is often mono-cistronic that encodes a single protein, while the bacterial transcript is poly-cistronic that encodes several proteins. It was well-acknowledged in eukaryotic cells that canonical proteins are translated from mRNA; thus, alternative transcription and translation are overlooked.

The pioneering systematical studies on microproteins started from yeast in 1997 (Basrai et al., 1997). Shortly after that, the Boeke group reported functional analysis of smORFs in the yeast genome and identified several microproteins that respond to different stress (Kastenmayer et al., 2006). Then, Savard et al. (2006) found that mlpt mRNA is poly-cistronic and encodes four small peptides. Three of them shared conserved motif, and depletion of mlpt impaired the embryo segmentation in Tribolium. Subsequently, two groups independently reported that the homology gene of mlpt in Drosophilatal/pri, has similar biological functions (Galindo et al., 2007; Kondo et al., 2007). These pioneer studies kick off a microprotein gold rush, especially in higher eukaryotes.

With the development of techniques including bioinformatic methods, mass spectrometry and newly developed ribosome profiling (Andrews & Rothnagel, 2014; Couso & Patraquim, 2017; Ingolia et al., 2009; Makarewich & Olson, 2017; Orr et al., 2020; Peeters & Menschaert, 2020; Plaza et al., 2017; Pueyo et al., 2016; Schlesinger & Elsasser, 2021; Sousa & Farkas, 2018; Xiao et al., 2018), thousands of translated smORFs or alternative-ORFs (alt-ORFs) in the genomes have been identified thus far. Surprisingly, microproteins are encoded by mRNAs as well as rRNA and other kinds of RNA that were previously considered noncoding (Non-coding RNA, ncRNA), including alternative transcripts and non-AUG start codon initiation transcripts, uORF (upstream ORF) located in the 5’ UTR of mRNA and circular RNA (circRNA) (Table 1).

Table 1 Names and definitions of microproteinsFull size table

As an emerging and rapidly evolving field, there are many different definitions, classifications, and nomenclatures of “smORF-coded proteins” (Table 1). For example, according to (1) their small size, from 2 to 200 aa, “microprotein” (Makarewich, 2020; Rathore et al., 2018), “small protein” (Basrai et al., 1997; Oyama et al., 2004; Storz et al., 2014), “micropeptide” (Chen et al., 2021; Makarewich & Olson, 2017; Vitorino et al., 2021), “miniproteins” (Crook et al., 2020); (2) their original sources, “SEP” (smORF-encoded polypeptide) (Pueyo et al., 2016; Saghatelian & Couso, 2015; Yin et al., 2019) or “MDPs” (mitochondria-derived peptides) (Kim et al., 2017); (3) their “microRNAs-like” functions, that they bind and inhibit targeted proteins as negative regulators at the protein level, such as “microProtein, miP” (Bhati et al., 2018; Staudt & Wenkel, 2011), “inhibitor of DNA binding, Id proteins” (Roschger & Cabrele, 2017); (4) the unique structure or function, “cyclotide” was named after its head-to-tail cyclized backbone (Camarero, 2017; Craik & Du, 2017); (5) personalized naming like “CASIMO1”, “CYREN”, “DYNLL1” (Arnoult et al., 2017; He et al., 2018; Polycarpou-Schwarz et al., 2018).

Identification of microproteins

Bioinformatic prediction

Bioinformatic prediction is often based on characteristics of the coding regions distinct from randomly generated short fragments in the genome. The coding regions are more likely conserved than random parts, which can be evaluated by three parameters. First, the cross-species comparison is a useful strategy for gene prediction. Metrics like the Ka/Ks ratio test were developed to evaluate the conservation of the coding sequence (Hurst, 2002). The Ka/Ks ratio test is used based on the fact that synonymous substitutions (Ks) occur much more frequently than nonsynonymous ones (Ka) in most of the coding regions. If the protein sequence is conserved between species, the Ka/Ks ratio is usually less than 1. Second, proteins often contain conserved domain/motif(s). Third, to analyze the sequence content and characteristics features, such as Kozak sequence, codon usage bias, nucleotide composition and so on (Gelfand, 1995). Thus, stringent selection can be implemented according to these sequence signatures. Based on the protein characters mentioned above, several smORF prediction tools have been successfully developed to improve the sensitivity and accuracy of smORF prediction, such as sORF Finder (Hanada et al., 2010), FSPP (Li et al., 2018), PhyloCSF (Lin et al., 2011) and CRITICA (Frith et al., 2006). However, whether such in silico predicted smORFs could be translated into bioactive microproteins in vivo needs validation by wet-lab experiments.

Mass spectrometry (MS)

In principle, MS is a widely used and practical proteomic approach. However, the traditional MS has two major disadvantages in microprotein identification. First, MS relies on mapping the MS spectra of digested peptides to the reference database generated by the theoretical spectra of annotated proteins (Zhang et al., 2013). Most of the translated microproteins are undetected just because they are unannotated. Second, the properties of microproteins, e.g., low abundance, low molecular weight (Andrews & Rothnagel, 2014) and instability, make them hard to be detected by MS (Slavoff et al., 2013). Third, sample preparation also impacts the MS results. Insufficient separation of analytes and inefficient enrichment of target proteins always comprise the detection of microproteins.

To overcome these challenges, researchers have adopted multiple strategies to improve the traditional MS. One advancement is the application of electrostatic repulsion hydrophilic interaction chromatography (ERLIC); it helps to separate the peptide mixture into a single component better and enrich the target peptides (Chugunova et al., 2018; Hao et al., 2011). The Slavoff group utilized the MWCO (Molecular weight cut-off) + ERLIC + LC–MS/MS workflow, that combined with a custom proteomic database based on the six-frame translation information generated from the RNA-seq data of K562 cells, and have successfully identified 195 novel microproteins in K562 cells (Ma et al., 2014). To find the most effective workflow, they also optimize the isolation and enrichment procedures before MS detection. By comparing four different extraction procedures and three enrichment steps, Slavoff et al. recommend combining acid precipitation and reverse-phase (C8) cartridges to enrich microproteins (Ma et al., 2016). These improvements lead to better identification of microproteins in human cells.

Yang et al. (2011) developed a three-step strategy that combines computational, transcriptional, and translational methods to identify protein-coding smORF in the Populus genome. They first utilized expressed sequence tags (EST) to construct a custom database, then filtered all candidates by three computational programs based on the coding potential, cross-species conservation, and protein sequence clustering. After a validation step by mass spectrometry, they finally designated 611 high confidence protein-coding smORFs, 56 of which are new to the Populus genome.

Ribosome profiling (Ribo-seq)

Ribo-seq is a powerful and emerging biochemical technology for quantifying global translation conditions in cells. It is developed by Ingolia et al. (2009), based on the fact that ribosomes often bind actively translated mRNAs. Ribo-seq utilizes next-generation sequencing (NGS) to analyze the 22–35 nt mRNA protected from nuclease digestion by the ribosome. Then, the ribosome protected fragment (RPF) is mapped back to the genomic loci with density information reflecting the translation efficiency (Aspden et al., 2014).

Ribo-seq has been used to identify a large number of smORFs in various species like yeast (Ingolia et al., 2009), fruit fly (Aspden et al., 2014), zebrafish (Bazzini et al., 2014), mouse (Ingolia et al., 2011) and human (Gonzalez et al., 2014; Raj et al., 2016). Although ribosome profiling has several advantages over other methods in smORF detection, it still has some limitations. For example, RNA binding proteins and the stalled or inactive ribosomes can also bind mRNA that may result in false positives. Therefore, poly-Ribo-seq was developed to detect smORFs specifically. The actively translating canonical CDS will be covered by multiple ribosomes called polysomes, while smORF is very short in length, so it usually accommodates no more than six ribosomes. Ultracentrifugation was used to separate and enrich the small polysomes (2–6) from large polysomes (Aspden et al., 2014). Aspden et al. applied this method in Drosophila S2 cells and identified 228 translated smORFs, compared to only 60 small peptides identified by mass spectrometry. Other derivative approaches such as TARP (Translating ribosome affinity purification)-seq (Heiman et al., 2014), RiboLace (Clamer et al., 2018) have been developed for cell-type-specific translational analysis. Through optimizing sample preparation and combining multiple strategies, we have witnessed a fast-growing number of identified microproteins in the last decade.

Biological function of microproteins

Although the functional study of microproteins is still in its infancy, microproteins have been implicated in a wide range of biological processes including embryonic development (Kondo et al., 20072010), muscle formation and regeneration (Anderson et al., 2015; Magny et al., 2013; Matsumoto et al., 2017; Nelson et al., 2016), cellular stress response (Chu et al., 2019; Khitun et al., 2019; Starck et al., 2016), RNA regulation (D'Lima et al., 2017) and DNA repair (Arnoult et al., 2017; Slavoff et al., 2014). This section will elaborate on the regulation of protein function by microproteins at multiple levels, including transcription, translation and post-translational modifications, with a particular focus on their contribution to genome stability maintenance processes including DNA replication and DNA double-strand break (DSB) repair.

Microproteins function at the transcriptional level

Microproteins can regulate gene expression at the transcriptional level by forming inactive complexes or competing for DNA/RNA binding (Staudt & Wenkel, 2011).

Id proteins

One of the most representative examples is Id proteins that inhibit the DNA binding activity of the basic Helix-Loop-Helix (bHLH) transcription factors (TFs). The bHLH TFs contain a basic region that binds to DNA and an HLH region that forms a homodimer or heterodimer with another bHLH TF. The monomeric bHLH TF has no transcription activity until forming homodimers or heterodimers. Id proteins contain only an HLH domain but lack the basic region. Therefore, Id proteins can form a non-functional dimer with a bHLH TF and result in a dominant-negative effect (Ruzinova & Benezra, 2003) (Fig. 1a). For example, Id2 interacts with the bHLH TFs CLOCK and BMAL to prevent these proteins from forming a functional dimer that binds to the E-box element responsible for downstream transcription of PER1 and CRY in the core mammalian circadian clock (Duffield et al., 2009).

Fig. 1figure1

The biological functions of microproteins. a Id proteins form a non-functional dimer with a bHLH TF and result in transcription repression (left). Id2 interacts with pRb and releases E2F for transcription (right). b PINT-87aa directly binds the PAF1c and disrupts the RNA Pol II-PAF1c complex formation. c Microproteins cause ribosome stalling or dissociation. d FBXW7-185aa binds USP28 to antagonize its de-ubiquitination of c-Myc and free FBWX7α to targeting c-Myc for ubiquitination degradation. e Hug1, Sml1 and Dif1 regulate the dNTP level upon replicative stress. Under low or intermediate replication stress, both Dun1 and Mck1 can phosphorylate the transcriptional repressor Crt1 and thereby de-repress the RNR and HUG1 gene transcription. Dun1 also triggers a phosphorylation-dependent degradation of RNR inhibitors Sml1 and Dif1. In the presence of more severe stress, Mck1 inhibits the HUG1 transcription to boost dNTP levels further. f SHPRH-146aa protects full-length SHPRH from degradation mediated by E3 ligase DTL and promotes SHPRH-mediated ubiquitination and degradation of PCNA. g The recruitment of Ku70/80 to the DSB sites is diminished by CYREN association, which explains why cNHEJ is inhibited in S and G2 phases. h DYNLL1 interacts with Mre11 to impair the DNA end resection activities in BRCA1-deficient cancer cells

Full size image

Besides bHLH proteins, Id proteins also bind to and inhibit other transcription factors. Several studies have reported that Id2 interacts with the retinoblastoma protein (pRb), a transcriptional repressor controlling the genes required for G1/S transition. These indicate that Id2 can promote cell cycle progression and thereby antagonize the tumor suppressor function of pRb (Iavarone et al., 1994). This finding exemplifies that some microproteins can be exploited by cancer cells to avoiding growth suppressors (Fig. 1a).

PINT-87aa

Following the pioneering work in Id proteins, other microproteins like PINT-87aa have been found to regulate transcription through distinct mechanisms. The long intergenic non-protein-coding RNA p53-induced transcript (LINC-PINT) was found to have a circular form that encodes an 87-aa microprotein, PINT-87aa. PINT-87aa can directly bind the polymerase-associated factor complex (PAF1c) and prevent the RNA PolII-PAF1c complex formation (Fig. 1b), which is required for the transcription elongation of many oncogenes including Cyclin D1, c-Myc and Sox2. Consistently, PINT-87aa depletion increases tumor burden in vivo, suggesting that the microprotein PINT-87aa might be a candidate tumor suppressor (Zhang et al., 2018b).

Microproteins function at the translational level

In addition to the transcriptional control, accumulating evidence supports a prominent contribution of translational regulation in gene expression especially in eukaryotes. Microproteins encoded by uORFs have been found to pose a critical role in the mRNA translation process (Zhang et al., 2021). In the human genome, about 35% of genes contain uORFs. Recent findings suggest that microproteins translated from the uORFs often interact and interfere with ribosomes, leading to ribosome dissociation (start codon skipping) or stalling (Orr et al., 2020) (Fig. 1c). For example, the 5’ UTR of EPHX1 (Epoxide hydrolase 1) transcripts E1-b’ encode two microproteins of 26 and 17 aa in length, respectively. These two microproteins, in cis and in trans configurations, inhibit the EPHX1 translation (Nguyen et al., 2013). There is also a report that 5’ UTR-encoded microproteins (e.g. uORF of the McKusick-Kaufman syndromegene, uMKKS protein) may be directly involved in cellular processes, rather than translational regulation, albeit with unknown mechanisms (Akimoto et al., 2013).

Microproteins function at the post-translational level

The degradation and half-life of c-Myc proto-oncoprotein are strictly controlled by ubiquitination ligase FBXW7 (F-box and WD repeat domain containing 7) and de-ubiquitin ligase USP28. CircRNA FBXW7 encodes a microprotein called FBXW7-185aa, which has a higher affinity for USP28 and inhibits the interaction between USP28 and FBXW7α, and indirectly increases FBXW7-mediated ubiquitination of c-Myc (Fig. 1d). By this means, FBXW7-185aa limits the c-Myc protein level and expression of its downstream cell proliferation genes (Yang et al., 2018). Together with the abovementioned PINT-87aa, FBXW7-185aa exemplifies that microproteins can inhibit oncogene overexpression and proliferative signals required by cancer cells.

Microproteins function in DNA replication

Hug1, Sml1 and Dif1

Several microproteins have been found to fine-tune the cellular levels of deoxynucleotide triphosphates (dNTPs), crucial for the high fidelity of DNA replication and genomic stability maintenance. The rate-limiting step in the de novo dNTP synthesis is catalyzing the reduction of nucleotide diphosphates (NDPs) to deoxynucleotide diphosphates (dNDPs) by the ribonucleotide reductase (RNR). Yeast RNR is a hetero-tetrameric complex comprising two large subunits (Rnr1 or Rnr3) and two small subunits (Rnr2 and Rnr4). Rnr3 is usually repressed by Crt1 and only induced by replication stress or DNA damage.

Several microproteins regulate RNR at the protein level (Fig. 1e). Hug1, a 68 aa microprotein, binds the small subunit Rnr2 and inhibits the RNR activity (Meurisse et al., 2014). Recently, our group showed that the basal Hug1 protein level is very low under normal conditions. When cells encounter moderated replication stress, the Mec1-Rad53 checkpoint pathway phosphorylates Dun1 and the GSK-3 family kinase, Mck1. Both Mck1 and Dun1 phosphorylate Crt1 to release its suppression of RNR2-4 and HUG1 transcription. The RNR inhibitory role of Hug1 may function as a rheostat to prevent excessive dNTP production, which increases genome instability as well. When cells suffer the more severe or persistent stress, Mck1 can inhibit the HUG1 transcription in a Crt1-independent manner to further augment the RNR activity and dNTP levels (Li et al., 2019).

Besides Hug1, two paralogous microproteins, Sml1 and Dif1, regulate RNR as well. Sml1 hijacks Rnr1 in the cytoplasm and prevents the formation of RNR holoenzyme (Chabes et al., 1999; Zhao et al., 2001). Dif1 bears a HUG motif similar to Hug1. It can bind the Rnr2/Rnr4 heterodimer and mediate its nucleus import, preventing its association with large subunits to form a functional RNR complex in the cytoplasm (Lee et al., 2008; Wu & Huang, 2008).

In summary, these microproteins regulate the RNR activity and cellular dNTP levels in a tunable and scalable manner and thus make a significant contribution to maintaining genome integrity under both unperturbed and perturbed conditions.

SHPRH-146aa

Besides the building blocks for DNA replication, some key replisome components are regulated by microproteins as well. PCNA (Proliferating cell nuclear antigen) plays indispensable roles as the clamp of DNA polymerases and a hub for nearly all DNA-related activities. Its protein level is regulated by SHPRH (SNF2 histone linker PHD RING helicase) ubiquitin ligase E3. Recently, Zhang et al. identified a circular RNA (circ-SHPRH) transcript in the SHPRH gene locus encoding a 146 aa microprotein, SHPRH-146aa. SHPRH-146aa interacts with the E3 ligase denticleless (DTL) and inhibits DTL-mediated ubiquitin–proteasome degradation of SHPRH (Fig. 1f). By this mechanism, SHPRH-146aa negatively regulates the PCNA protein level. SHPRH-146aa is downregulated in glioblastoma. Overexpression of SHPRH-146aa in glioblastoma cells inhibits its viability and tumorigenicity (Zhang et al., 2018a). These findings indicate that microproteins like SHPRH-146aa and PINT-87aa can function as tumor suppressors and therefore represent potential therapeutic implications.

Microproteins function in the DNA double-strand break (DSB) repair

DSBs, the most severe type of DNA damage, are mainly repaired by the canonical non-homologous end joining (cNHEJ) and homologous recombination (HR) repair pathways. To date, several microproteins have been reported as critical regulators of the DSB repair pathway choices, such as CYREN1/2/3, DYNLL1 and AltMRVI1.

CYREN-1/2/3

Three isoforms of CYREN (cell cycle regulator of NHEJ)/MRI (modulator of retroviral infection) are encoded by C7orf49 in mammals. CYREN/MRI-2 was first shown to stimulate cNHEJ through interaction with the key cNHEJ determinant factor Ku70/80 (Slavoff et al., 2014). Karlseder and colleagues later found that CYREN-1 and CYREN-2 bind Ku70/80 only in S and G2 phases. Instead, the association of CYREN diminishes the recruitment of Ku70/80 to the DSB sites, which creates an environment favorable for error-free homologous recombination between sister chromatids while inhibiting cNHEJ (Fig. 1g). The deletion of CYREN enables cNHEJ to occur, promoting chromosome fusion at telomeres, specifically during S and G2 phases (Arnoult et al., 2017).

DYNLL1

In contrast to the HR-promoting role of CYREN, DYNLL1 (Dynein light chain 1 protein) directly inhibits the HR pathway during DSB repair. DYNLL1 is an 8 kDa light chain protein of the cytoplasmic dynein protein complex. The most well-acknowledged function of DYNLL1 is to act as a cargo adaptor of the motor complexes like dynein and myosin (Rapali et al., 2011).

In a genome-scale screen using CRISPR Cas9-based knockout library, Chowdhury and his colleague uncovered that loss of DYNLL1 shows strong resistance to PARPi and platinum drugs in BRCA1-deficient cancer cells. As BRCA1 is vital in the HR repair pathway, this result implicates that DYNLL1 plays a negative role in regulating the HR pathway. In molecular detail, they demonstrate that DYNLL1 directly interacts with MRE11 to inhibit its endonuclease activities, which is required for DNA end resection and subsequent HR (He et al., 2018). DYNLL1 functions as a potent inhibitor of DNA end resection when HR is impaired in BRCA1-deficient cancer cells. These findings suggest that DYNLL1 is responsible for resistance to PARPi or platinum chemotherapy and might be a possible candidate target for synthetic lethality-based cancer therapy (Fig. 1h).

Another microprotein, AltMRVI1 encoded by the 3’ UTR of the MRVI1 mRNA, was reported to interact with BRCA1 (Vanderperre et al., 2013). However, the exact role of AltMRVI1 awaits to be investigated.

The studies mentioned above establish the critical contribution of microproteins to avoiding genome instability and uncontrolled cell proliferation, which are hallmarks of many human diseases like cancer. Indeed, the abnormal expression of microproteins has been found to be associated with calcium homeostasis in muscles, obesity, diabetes, tumors and neurodegenerative diseases (Table 2). Interested readers can refer to the recent excellent reviews on this topic (Chen et al., 2021; Merino-Valverde et al., 2020; Wang et al., 2019; Wu et al., 2020a; Zhou et al., 2021).

Table 2 Microproteins in human diseasesFull size table

Conclusions and future perspectives

As the “dark matter” in the proteome, microproteins are not recognized till the recent development of cutting-edge techniques (Fig. 2a). In this minireview, we have briefly summarized the recent technique improvements and research progresses in identifying and characterizing microproteins, with a particular focus on their function in maintaining genome integrity (Fig. 2b).

Fig. 2figure2

A brief depiction of the microprotein research (a) and biological functions discussed in this review (b)

Full size image

As an emergent field, many aspects of microproteins wait to be addressed in the future.

  1. 1.

    Standard nomenclature and guidelines are necessary for future research on microproteins with different origins in different organisms.

  2. 2.

    Compare to an increasing number of microproteins that have been identified, our knowledge of their functions and molecular mechanisms is still very limited. For example, the Hieter group identified 302 unannotated ORFs (NORFs) in Saccharomyces cerevisiae in 1999 (Basrai et al., 1999). Among them, only the function of NORF5/HUG1 has been studied. The true challenge in this field is a focus shift from identification to functional studies.

  3. 3.

    Microproteins usually lack a conserved domain and more likely contain flexible structures, making their functions more challenging to address. Moreover, epitope tagging, as one of the most common methods for protein functional analysis, will more likely interfere with the physiological properties of microproteins than that of regular proteins.

  4. 4.

    Besides their physiological function, the pathological significance of microproteins will be of great interest to explore. For instance, the understanding of disease-specific microproteome, dysregulation and mutation landscape of microproteins will hold promise for novel markers, drugs and therapeutic targets.

  5. 5.

    Microproteins are very small in size, and many of them are cell membrane permeable. This unique character will endow them with unparalleled advantages in drug discovery or delivery.

Abbreviations

  • aa:

  • Amino acids

  • alt-ORFs:

  • Alternative-ORFs

  • bHLH:

  • Basic helix-loop-helix

  • CASIMO1:

  • Cancer-associated small integral membrane open reading frame 1

  • circRNA:

  • Circular RNA

  • cNHEJ:

  • Canonical non-homologous end joining

  • CYREN:

  • Cell cycle regulator of NHEJ

  • dNDPs:

  • Deoxynucleotide diphosphates

  • dNTPs:

  • Deoxynucleotide triphosphates

  • DTL:

  • Denticleless

  • DSB:

  • DNA double-strand break

  • DYNLL1:

  • Dynein light chain LC8-type 1

  • EPHX1:

  • Epoxide hydrolase 1

  • ER:

  • Endoplasmic reticulum

  • ERLIC:

  • Electrostatic repulsion hydrophilic interaction chromatography

  • EST:

  • Expressed sequence tags

  • FBXW7:

  • F-box and WD repeat domain containing 7

  • HR:

  • Homologous recombination

  • Id:

  • Inhibitor of DNA binding and/or differentiation

  • LINC-PINT:

  • The long intergenic non-protein-coding RNA p53-induced transcript

  • MDPs:

  • Mitochondria-derived peptides

  • miPs:

  • MicroProteins

  • MRI:

  • Modulator of retroviral infection

  • MS:

  • Mass spectrometry

  • MWCO:

  • Molecular weight cut-off

  • ncRNA:

  • Non-coding RNA

  • NDPs:

  • Nucleotide diphosphates

  • NGS:

  • Next-generation sequencing

  • NORFs:

  • Nonannotated ORFs

  • NOPs:

  • 3' Neo open reading frame peptides

  • PAF1c:

  • Polymerase-associated factor complex

  • PCNA:

  • Proliferating cell nuclear antigen

  • pRb:

  • The retinoblastoma protein

  • RNR:

  • Ribonucleotide reductase

  • RPF:

  • Ribosome protected fragment

  • SEP:

  • smORF-encoded polypeptide

  • SHPRH:

  • SNF2 histone linker PHD RING helicase

  • sORFs/smORFs:

  • Small open reading frames

  • SR:

  • Sarcoplasmic reticulum

  • TARP:

  • Translating ribosome affinity purification

  • TFs:

  • Transcription factors

  • uORF:

  • Upstream ORF

  • uMKKS:

  • uORF of the McKusick-Kaufman syndrome gene

References

  1. Akimoto, C., Sakashita, E., Kasashima, K., Kuroiwa, K., Tominaga, K., Hamamoto, T., & Endo, H. (2013). Translational repression of the McKusick-Kaufman syndrome transcript by unique upstream open reading frames encoding mitochondrial proteins with alternative polyadenylation sites. Biochimica Et Biophysica Acta, 1830, 2728–2738. https://doi.org/10.1016/j.bbagen.2012.12.010

    CAS Article PubMed Google Scholar 

  2. Anderson, D. M., Anderson, K. M., Chang, C. L., Makarewich, C. A., Nelson, B. R., McAnally, J. R., Kasaragod, P., Shelton, J. M., Liou, J., Bassel-Duby, R., & Olson, E. N. (2015). A micropeptide encoded by a putative long noncoding RNA regulates muscle performance. Cell, 160, 595–606. https://doi.org/10.1016/j.cell.2015.01.009

    CAS Article PubMed PubMed Central Google Scholar 

  3. Anderson, D. M., Makarewich, C. A., Anderson, K. M., Shelton, J. M., Bezprozvannaya, S., Bassel-Duby, R., & Olson, E. N. (2016). Widespread control of calcium signaling by a family of SERCA-inhibiting micropeptides. Science Signaling, 9, ra119. https://doi.org/10.1126/scisignal.aaj1460

    CAS Article PubMed PubMed Central Google Scholar 

  4. Andrews, S. J., & Rothnagel, J. A. (2014). Emerging evidence for functional peptides encoded by short open reading frames. Nature Reviews Genetics, 15, 193–204. https://doi.org/10.1038/nrg3520

    CAS Article PubMed Google Scholar 

  5. Arnoult, N., Correia, A., Ma, J., Merlo, A., Garcia-Gomez, S., Maric, M., Tognetti, M., Benner, C. W., Boulton, S. J., Saghatelian, A., & Karlseder, J. (2017). Regulation of DNA repair pathway choice in S and G2 phases by the NHEJ inhibitor CYREN. Nature, 549, 548–552. https://doi.org/10.1038/nature24023

    CAS Article PubMed PubMed Central Google Scholar 

  6. Aspden, J. L., Eyre-Walker, Y. C., Phillips, R. J., Amin, U., Mumtaz, M. A., Brocard, M., & Couso, J. P. (2014). Extensive translation of small open reading frames revealed by poly-ribo-seq. eLife, 3, e03528. https://doi.org/10.7554/eLife.03528

    Article PubMed PubMed Central Google Scholar 

  7. Basrai, M. A., Hieter, P., & Boeke, J. D. (1997). Small open reading frames: beautiful needles in the haystack. Genome Research, 7, 768–771. https://doi.org/10.1101/gr.7.8.768

    CAS Article PubMed Google Scholar 

  8. Basrai, M. A., Velculescu, V. E., Kinzler, K. W., & Hieter, P. (1999). NORF5/HUG1 is a component of the MEC1-mediated checkpoint response to DNA damage and replication arrest in Saccharomyces cerevisiaeMolecular and Cellular Biology, 19, 7041–7049. https://doi.org/10.1128/mcb.19.10.7041

    CAS Article PubMed PubMed Central Google Scholar 

  9. Bazzini, A. A., Johnstone, T. G., Christiano, R., Mackowiak, S. D., Obermayer, B., Fleming, E. S., Vejnar, C. E., Lee, M. T., Rajewsky, N., Walther, T. C., & Giraldez, A. J. (2014). Identification of small ORFs in vertebrates using ribosome footprinting and evolutionary conservation. The EMBO Journal, 33, 981–993. https://doi.org/10.1002/embj.201488411

    CAS Article PubMed PubMed Central Google Scholar 

  10. Bhati, K. K., Blaakmeer, A., Paredes, E. B., Dolde, U., Eguen, T., Hong, S. Y., Rodrigues, V., Straub, D., Sun, B., & Wenkel, S. (2018). Approaches to identify and characterize microProteins and their potential uses in biotechnology. Cellular and Molecular Life Sciences, 75, 2529–2536. https://doi.org/10.1007/s00018-018-2818-8

    CAS Article PubMed PubMed Central Google Scholar 

  11. Bi, P., Ramirez-Martinez, A., Li, H., Cannavino, J., McAnally, J. R., Shelton, J. M., Sanchez-Ortiz, E., Bassel-Duby, R., & Olson, E. N. (2017). Control of muscle formation by the fusogenic micropeptide myomixer. Science, 356, 323–327. https://doi.org/10.1126/science.aam9361

    CAS Article PubMed PubMed Central Google Scholar 

  12. Brunet, M. A., Leblanc, S., & Roucou, X. (2020). Reconsidering proteomic diversity with functional investigation of small ORFs and alternative ORFs. Experimental Cell Research, 393, 112057. https://doi.org/10.1016/j.yexcr.2020.112057

    CAS Article PubMed Google Scholar 

  13. Cabrera-Quio, L. E., Herberg, S., & Pauli, A. (2016). Decoding sORF translation - from small proteins to gene regulation. RNA Biology, 13, 1051–1059. https://doi.org/10.1080/15476286.2016.1218589

    Article PubMed PubMed Central Google Scholar 

  14. Camarero, J. A. (2017). Cyclotides, a versatile ultrastable micro-protein scaffold for biotechnological applications. Bioorganic and Medicinal Chemistry Letters, 27, 5089–5099. https://doi.org/10.1016/j.bmcl.2017.10.051

    CAS Article PubMed Google Scholar 

  15. Cao, X., Khitun, A., Luo, Y., Na, Z., Phoodokmai, T., Sappakhaw, K., Olatunji, E., Uttamapinant, C., & Slavoff, S. A. (2021). Alt-RPL36 downregulates the PI3K-AKT-mTOR signaling pathway by interacting with TMEM24. Nature Communications, 12, 508. https://doi.org/10.1038/s41467-020-20841-6

    CAS Article PubMed PubMed Central Google Scholar 

  16. Cardon, T., Fournier, I., & Salzet, M. (2021). Shedding light on the ghost proteome. Trends in Biochemical Sciences, 46, 239–250. https://doi.org/10.1016/j.tibs.2020.10.003

    CAS Article PubMed Google Scholar 

  17. Chabes, A., Domkin, V., & Thelander, L. (1999). Yeast Sml1, a protein inhibitor of ribonucleotide reductase. The Journal of Biological Chemistry, 274, 36679–36683. https://doi.org/10.1074/jbc.274.51.36679

    CAS Article PubMed Google Scholar 

  18. Chen, Y., Ho, L., & Tergaonkar, V. (2021). sORF-Encoded MicroPeptides: New players in inflammation, metabolism, and precision medicine. Cancer Letters, 500, 263–270. https://doi.org/10.1016/j.canlet.2020.10.038

    CAS Article PubMed Google Scholar 

  19. Chu, Q., Martinez, T. F., Novak, S. W., Donaldson, C. J., Tan, D., Vaughan, J. M., Chang, T., Diedrich, J. K., Andrade, L., Kim, A., Zhang, T., Manor, U., & Saghatelian, A. (2019). Regulation of the ER stress response by a mitochondrial microprotein. Nature Communications, 10, 4883. https://doi.org/10.1038/s41467-019-12816-z

    CAS Article PubMed PubMed Central Google Scholar 

  20. Chu, Q., & Saghatelian, A. (2019). A hidden ORF reveals an immune protector. Biochemistry, 58, 1022–1023. https://doi.org/10.1021/acs.biochem.8b01317

    CAS Article PubMed Google Scholar 

  21. Chugunova, A., Navalayeu, T., Dontsova, O., & Sergiev, P. (2018). Mining for small translated ORFs. Journal of Proteome Research, 17, 1–11. https://doi.org/10.1021/acs.jproteome.7b00707

    CAS Article PubMed Google Scholar 

  22. Clamer, M., Tebaldi, T., Lauria, F., Bernabo, P., Gomez-Biagi, R. F., Marchioretto, M., Kandala, D. T., Minati, L., Perenthaler, E., Gubert, D., Pasquardini, L., Guella, G., Groen, E. J. N., Gillingwater, T. H., Quattrone, A., & Viero, G. (2018). Active ribosome profiling with RiboLace. Cell Reports, 25(1097–1108), e1095. https://doi.org/10.1016/j.celrep.2018.09.084

    CAS Article Google Scholar 

  23. Cobb, L. J., Lee, C. H., Xiao, J. L., Yen, K., Wong, R. G., Nakamura, H. K., Mehta, H. H., Gao, Q. L., Ashur, C., Huffman, D. M., Wan, J. X., Muzumdar, R., Barzilai, N., & Cohen, P. (2016). Naturally occurring mitochondrial-derived peptides are age-dependent regulators of apoptosis, insulin sensitivity, and inflammatory markers. Aging, 8, 796–809. https://doi.org/10.18632/aging.100943

    CAS Article PubMed PubMed Central Google Scholar 

  24. Couso, J.-P., & Patraquim, P. (2017). Classification and function of small open reading frames. Nature Reviews Molecular Cell Biology, 18, 575–589. https://doi.org/10.1038/nrm.2017.58

    CAS Article PubMed Google Scholar 

  25. Craik, D. J., & Du, J. (2017). Cyclotides as drug design scaffolds. Current Opinion in Chemical Biology, 38, 8–16. https://doi.org/10.1016/j.cbpa.2017.01.018

    CAS Article PubMed Google Scholar 

  26. Crook, Z. R., Nairn, N. W., & Olson, J. M. (2020). Miniproteins as a powerful modality in drug development. Trends in Biochemical Sciences, 45, 332–346. https://doi.org/10.1016/j.tibs.2019.12.008

    CAS Article PubMed PubMed Central Google Scholar 

  27. D’Lima, N. G., Ma, J., Winkler, L., Chu, Q., Loh, K. H., Corpuz, E. O., Budnik, B. A., Lykke-Andersen, J., Saghatelian, A., & Slavoff, S. A. (2017). A human microprotein that interacts with the mRNA decapping complex. Nature Chemical Biology, 13, 174–180. https://doi.org/10.1038/nchembio.2249

    CAS Article PubMed Google Scholar 

  28. Duffield, G. E., Watson, N. P., Mantani, A., Peirson, S. N., Robles-Murguia, M., Loros, J. J., Israel, M. A., & Dunlap, J. C. (2009). A role for Id2 in regulating photic entrainment of the mammalian circadian system. Current Biology, 19, 297–304. https://doi.org/10.1016/j.cub.2008.12.052

    CAS Article PubMed PubMed Central Google Scholar 

  29. Eguen, T., Straub, D., Graeff, M., & Wenkel, S. (2015). MicroProteins: small size-big impact. Trends in Plant Science, 20, 477–482. https://doi.org/10.1016/j.tplants.2015.05.011

    CAS Article PubMed Google Scholar 

  30. Fang, J., Morsalin, S., Rao, V. N., & Reddy, E. S. (2017). Decoding of non-coding DNA and non-coding RNA: pri-micro RNA-encoded novel peptides regulate migration of cancer cells. Journal of Pharmaceutical Sciences and Pharmacology, 3, 23. https://doi.org/10.1166/jpsp.2017.1070

    Article Google Scholar 

  31. Friesen, M., Warren, C. R., Yu, H. J., Toyohara, T., Ding, Q. R., Florido, M. H. C., Sayre, C., Pope, B. D., Goff, L. A., Rinn, J. L., & Cowan, C. A. (2020). Mitoregulin controls beta-oxidation in human and mouse adipocytes. Stem Cell Reports, 14, 590–602. https://doi.org/10.1016/j.stemcr.2020.03.002

    CAS Article PubMed PubMed Central Google Scholar 

  32. Frith, M. C., Forrest, A. R., Nourbakhsh, E., Pang, K. C., Kai, C., Kawai, J., Carninci, P., Hayashizaki, Y., Bailey, T. L., & Grimmond, S. M. (2006). The abundance of short proteins in the mammalian proteome. PLoS Genetics, 2, e52. https://doi.org/10.1371/journal.pgen.0020052

    CAS Article PubMed PubMed Central Google Scholar 

  33. Galindo, M. I., Pueyo, J. I., Fouix, S., Bishop, S. A., & Couso, J. P. (2007). Peptides encoded by short ORFs control development and define a new eukaryotic gene family. PLoS Biology, 5, e106. https://doi.org/10.1371/journal.pbio.0050106

    CAS Article PubMed PubMed Central Google Scholar 

  34. Gelfand, M. S. (1995). Prediction of function in DNA sequence analysis. Journal of Computational Biology, 2, 87–115. https://doi.org/10.1089/cmb.1995.2.87

    CAS Article PubMed Google Scholar 

  35. Gonzalez, C., Sims, J. S., Hornstein, N., Mela, A., Garcia, F., Lei, L., Gass, D. A., Amendolara, B., Bruce, J. N., Canoll, P., & Sims, P. A. (2014). Ribosome profiling reveals a cell-type-specific translational landscape in brain tumors. The Journal of Neuroscience, 34, 10924–10936. https://doi.org/10.1523/JNEUROSCI.0084-14.2014

    CAS Article PubMed PubMed Central Google Scholar 

  36. Guo, B., Wu, S., Zhu, X., Zhang, L., Deng, J., Li, F., Wang, Y., Zhang, S., Wu, R., Lu, J., & Zhou, Y. (2020). Micropeptide CIP2A-BP encoded by LINC00665 inhibits triple-negative breast cancer progression. The EMBO Journal, 39, e102190. https://doi.org/10.15252/embj.2019102190

    CAS Article PubMed Google Scholar 

  37. Guo, Z. W., Meng, Y., Zhai, X. M., Xie, C., Zhao, N., Li, M., Zhou, C. L., Li, K., Liu, T. C., Yang, X. X., & Wu, Y. S. (2019). Translated long non-coding ribonucleic acid ZFAS1 promotes cancer cell migration by elevating reactive oxygen species production in hepatocellular carcinoma. Frontiers in Genetics, 10, 12. https://doi.org/10.3389/fgene.2019.01111

    CAS Article Google Scholar 

  38. Hanada, K., Akiyama, K., Sakurai, T., Toyoda, T., Shinozaki, K., & Shiu, S. H. (2010). sORF finder: a program package to identify small open reading frames with high coding potential. Bioinformatics, 26, 399–400. https://doi.org/10.1093/bioinformatics/btp688

    CAS Article PubMed Google Scholar 

  39. Hao, P., Zhang, H., Sze, & S. K. (2011). Application of electrostatic repulsion hydrophilic interaction chromatography to the characterization of proteome, glycoproteome, and phosphoproteome using nano LC–MS/MS. In: S. Toms, R. Weil (Eds.), Nanoproteomics. Methods in molecular biology (Methods and Protocols), (1th ed., pp. 305–318). Humana Press. https://doi.org/10.1007/978-1-61779-319-6_23

  40. Hashimoto, Y., Ito, Y., Niikura, T., Shao, Z., Hata, M., Oyama, F., & Nishimoto, I. (2001). Mechanisms of neuroprotection by a novel rescue factor humanin from Swedish mutant amyloid precursor protein. Biochemical and Biophysical Research Communications, 283, 460–468. https://doi.org/10.1006/bbrc.2001.4765

    CAS Article PubMed Google Scholar 

  41. He, Y. J., Meghani, K., Caron, M. C., Yang, C., Ronato, D. A., Bian, J., Sharma, A., Moore, J., Niraj, J., Detappe, A., Doench, J. G., Legube, G., Root, D. E., D’Andrea, A. D., Drane, P., De, S., Konstantinopoulos, P. A., Masson, J. Y., & Chowdhury, D. (2018). DYNLL1 binds to MRE11 to limit DNA end resection in BRCA1-deficient cells. Nature, 563, 522–526. https://doi.org/10.1038/s41586-018-0670-5

    CAS Article PubMed PubMed Central Google Scholar 

  42. Heiman, M., Kulicke, R., Fenster, R. J., Greengard, P., & Heintz, N. (2014). Cell type-specific mRNA purification by translating ribosome affinity purification (TRAP). Nature Protocols, 9, 1282–1291. https://doi.org/10.1038/nprot.2014.085

    CAS Article PubMed PubMed Central Google Scholar 

  43. Ho, L., Tan, S. Y. X., Wee, S., Wu, Y. X., Tan, S. J. C., Ramakrishna, N. B., Chng, S. C., Nama, S., Szczerbinska, I., Chan, Y. S., Avery, S., Tsuneyoshi, N., Ng, H. H., Gunaratne, J., Dunn, N. R., & Reversade, B. (2015). ELABELA is an endogenous growth factor that sustains hESC self-renewal via the PI3K/AKT pathway. Cell Stem Cell, 17, 435–447. https://doi.org/10.1016/j.stem.2015.08.010

    CAS Article PubMed Google Scholar 

  44. Huang, J. Z., Chen, M., Chen, D. E., Gao, X. C., Zhu, S., Huang, H., Hu, M., Zhu, H., & Yan, G. R. (2017). A peptide encoded by a putative lncRNA HOXB-AS3 suppresses colon cancer growth. Molecular Cell, 68, 171-184 e176. https://doi.org/10.1016/j.molcel.2017.09.015

    CAS Article PubMed Google Scholar 

  45. Hurst, L. D. (2002). The Ka/Ks ratio: diagnosing the form of sequence evolution. Trends in Genetics, 18, 486–487. https://doi.org/10.1016/S01689525(02)027221

    Article PubMed Google Scholar 

  46. Iavarone, A., Garg, P., Lasorella, A., Hsu, J., & Israel, M. A. (1994). The helix-loop-helix protein Id-2 enhances cell-proliferation and binds to the retinoblastoma protein. Genes and Development, 8, 1270–1284. https://doi.org/10.1101/gad.8.11.1270

    CAS Article PubMed Google Scholar 

  47. Ingolia, N. T., Ghaemmaghami, S., Newman, J. R., & Weissman, J. S. (2009). Genome-wide analysis in vivo of translation with nucleotide resolution using ribosome profiling. Science, 324, 218–223. https://doi.org/10.1126/science.1168978

    CAS Article PubMed PubMed Central Google Scholar 

  48. Ingolia, N. T., Lareau, L. F., & Weissman, J. S. (2011). Ribosome profiling of mouse embryonic stem cells reveals the complexity and dynamics of mammalian proteomes. Cell, 147, 789–802. https://doi.org/10.1016/j.cell.2011.10.002

    CAS Article PubMed PubMed Central Google Scholar 

  49. Jackson, R., Kroehling, L., Khitun, A., Bailis, W., Jarret, A., York, A. G., Khan, O. M., Brewer, J. R., Skadow, M. H., Duizer, C., Harman, C. C. D., Chang, L., Bielecki, P., Solis, A. G., Steach, H. R., Slavoff, S., & Flavell, R. A. (2018). The translation of non-canonical open reading frames controls mucosal immunity. Nature, 564, 434–438. https://doi.org/10.1038/s41586-018-0794-7

    CAS Article PubMed PubMed Central Google Scholar 

  50. Jorgensen, R. A., & Dorantes-Acosta, A. E. (2012). Conserved peptide upstream open reading frames are associated with regulatory genes in angiosperms. Frontiers in Plant Science, 3, 191. https://doi.org/10.3389/fpls.2012.00191

    CAS Article PubMed PubMed Central Google Scholar 

  51. Kang, M., Tang, B., Li, J., Zhou, Z., Liu, K., Wang, R., Jiang, Z., Bi, F., Patrick, D., Kim, D., Mitra, A. K., & Yang-Hartwich, Y. (2020). Identification of miPEP133 as a novel tumor-suppressor microprotein encoded by miR-34a pri-miRNA. Molecular Cancer, 19, 143. https://doi.org/10.1186/s12943-020-01248-9

    CAS Article PubMed PubMed Central Google Scholar 

  52. Kastenmayer, J. P., Ni, L., Chu, A., Kitchen, L. E., Au, W. C., Yang, H., Carter, C. D., Wheeler, D., Davis, R. W., Boeke, J. D., Snyder, M. A., & Basrai, M. A. (2006). Functional genomics of genes with small open reading frames (sORFs) in S-cerevisiaeGenome Research, 16, 365–373. https://doi.org/10.1101/gr.4355406

    CAS Article PubMed PubMed Central Google Scholar 

  53. Khitun, A., Ness, T. J., & Slavoff, S. A. (2019). Small open reading frames and cellular stress responses. Molecular Omics, 15, 108–116. https://doi.org/10.1039/c8mo00283e

    CAS Article PubMed PubMed Central Google Scholar 

  54. Khitun, A., & Slavoff, S. A. (2019). Proteomic detection and validation of translated small open reading frames. Current Protocols in Chemical Biology, 11, e77. https://doi.org/10.1002/cpch.77

    CAS Article PubMed PubMed Central Google Scholar 

  55. Kim, S. J., Xiao, J. L., Wan, J. X., Cohen, P., & Yen, K. (2017). Mitochondrially derived peptides as novel regulators of metabolism. Journal of Physiology, 595, 6613–6621. https://doi.org/10.1113/Jp274472

    CAS Article Google Scholar 

  56. Kondo, T., Hashimoto, Y., Kato, K., Inagaki, S., Hayashi, S., & Kageyama, Y. (2007). Small peptide regulators of actin-based cell morphogenesis encoded by a polycistronic mRNA. Nature Cell Biology, 9, 660–665. https://doi.org/10.1038/ncb1595

    CAS Article PubMed Google Scholar 

  57. Kondo, T., Plaza, S., Zanet, J., Benrabah, E., Valenti, P., Hashimoto, Y., Kobayashi, S., Payre, F., & Kageyama, Y. (2010). Small peptides switch the transcriptional activity of shavenbaby during drosophila embryogenesis. Science, 329, 336–339. https://doi.org/10.1126/science.1188158

    CAS Article PubMed Google Scholar 

  58. Koster, J., & Plasterk, R. H. A. (2019). A library of Neo Open Reading Frame peptides (NOPs) as a sustainable resource of common neoantigens in up to 50% of cancer patients. Scientific Reports, 9, 6577. https://doi.org/10.1038/s41598-019-42729-2

    CAS Article PubMed PubMed Central Google Scholar 

  59. Lee, C., Zeng, J., Drew, B. G., Sallam, T., Martin-Montalvo, A., Wan, J., Kim, S. J., Mehta, H., Hevener, A. L., de Cabo, R., & Cohen, P. (2015). The mitochondrial-derived peptide MOTS-c promotes metabolic homeostasis and reduces obesity and insulin resistance. Cell Metabolism, 21, 443–454. https://doi.org/10.1016/j.cmet.2015.02.009

    CAS Article PubMed PubMed Central Google Scholar 

  60. Lee, Y. D., Wang, J., Stubbe, J., & Elledge, S. J. (2008). Dif1 is a DNA-damage-regulated facilitator of nuclear import for ribonucleotide reductase. Molecular Cell, 32, 70–80. https://doi.org/10.1016/j.molcel.2008.08.018

    CAS Article PubMed PubMed Central Google Scholar 

  61. Li, H., Xiao, L., Zhang, L. L., Wu, J. R., Wei, B., Sun, N. H., & Zhao, Y. (2018). FSPP: a tool for genome-wide prediction of smORF-encoded peptides and their functions. Frontiers in Genetics, 9, 8. https://doi.org/10.3389/fgene.2018.00096

    CAS Article Google Scholar 

  62. Li, M. W., Li, X., Zhang, Y. N., Wu, H. M., Zhou, H. Z., Ding, X., Zhang, X. M., Jin, X. R., Wang, Y., Yin, X. Q., Li, C. C., Yang, P. W., & Xu, H. M. (2020a). Micropeptide MIAC inhibits HNSCC progression by interacting with aquaporin 2. Journal of the American Chemical Society, 142, 6708–6716. https://doi.org/10.1021/jacs.0c00706

    CAS Article PubMed Google Scholar 

  63. Li, X., Jin, X., Sharma, S., Liu, X., Zhang, J., Niu, Y., Li, J., Li, Z., Zhang, J., Cao, Q., Hou, W., Du, L. L., Liu, B., & Lou, H. (2019). Mck1 defines a key S-phase checkpoint effector in response to various degrees of replication threats. PLoS Genetics, 15, e1008136. https://doi.org/10.1371/journal.pgen.1008136

    CAS Article PubMed PubMed Central Google Scholar 

  64. Li, X. L., Pongor, L., Tang, W., Das, S., Muys, B. R., Jones, M. F., Lazar, S. B., Dangelmaier, E. A., Hartford, C. C., Grammatikakis, I., Hao, Q., Sun, Q., Schetter, A., Martindale, J. L., Tang, B., Jenkins, L. M., Robles, A. I., Walker, R. L., Ambs, S., et al. (2020b). A small protein encoded by a putative lncRNA regulates apoptosis and tumorigenicity in human colorectal cancer cells. eLife, 9, e53734. https://doi.org/10.7554/eLife.53734

    CAS Article PubMed PubMed Central Google Scholar 

  65. Lin, M. F., Jungreis, I., & Kellis, M. (2011). PhyloCSF: a comparative genomics method to distinguish protein coding and non-coding regions. Bioinformatics, 27, i275-282. https://doi.org/10.1093/bioinformatics/btr209

    CAS Article PubMed PubMed Central Google Scholar 

  66. Lin, Y. F., Xiao, M. H., Chen, H. X., Meng, Y., Zhao, N., Yang, L., Tang, H., Wang, J. L., Liu, X., Zhu, Y., & Zhuang, S. M. (2019). A novel mitochondrial micropeptide MPM enhances mitochondrial respiratory activity and promotes myogenic differentiation. Cell Death & Disease, 10, 528. https://doi.org/10.1038/s41419-019-1767-y

    CAS Article PubMed PubMed Central Google Scholar 

  67. Lu, S. H., Zhang, J., Lian, X. L., Sun, L., Meng, K., Chen, Y., Sun, Z. H., Yin, X. F., Li, Y. X., Zhao, J., Wang, T., Zhang, G., & He, Q. Y. (2019). A hidden human proteome encoded by “non-coding” genes. Nucleic Acids Research, 47, 8111–8125. https://doi.org/10.1093/nar/gkz646

    CAS Article PubMed PubMed Central Google Scholar 

  68. Luo, W. S., Grupp, I. L., Harrer, J., Ponniah, S., Grupp, G., Duffy, J. J., Doetschman, T., & Kranias, E. G. (1994). Targeted ablation of the phospholamban gene is associated with markedly enhanced myocardial-contractility and loss of beta-agonist stimulation. Circulation Research, 75, 401–409. https://doi.org/10.1161/01.RES.75.3.401

    CAS Article PubMed Google Scholar 

  69. Ma, J., Diedrich, J. K., Jungreis, I., Donaldson, C., Vaughan, J., Kellis, M., Yates, J. R., 3rd., & Saghatelian, A. (2016). Improved identification and analysis of small open reading frame encoded polypeptides. Analytical Chemistry, 88, 3967–3975. https://doi.org/10.1021/acs.analchem.6b00191

    CAS Article PubMed PubMed Central Google Scholar 

  70. Ma, J., Ward, C. C., Jungreis, I., Slavoff, S. A., Schwaid, A. G., Neveu, J., Budnik, B. A., Kellis, M., & Saghatelian, A. (2014). Discovery of human sORF-encoded polypeptides (SEPs) in cell lines and tissue. Journal of Proteome Research, 13, 1757–1765. https://doi.org/10.1021/pr401280w

    CAS Article PubMed PubMed Central Google Scholar 

  71. Magnani, E., de Klein, N., Nam, H. I., Kim, J. G., Pham, K., Fiume, E., Mudgett, M. B., & Rhee, S. Y. (2014). A comprehensive analysis of microProteins reveals their potentially widespread mechanism of transcriptional regulation. Plant Physiology, 165, 149–159. https://doi.org/10.1104/pp.114.235903

    CAS Article PubMed PubMed Central Google Scholar 

  72. Magny, E. G., Pueyo, J. I., Pearl, F. M. G., Cespedes, M. A., Niven, J. E., Bishop, S. A., & Couso, J. P. (2013). Conserved regulation of cardiac calcium uptake by peptides encoded in small open reading frames. Science, 341, 1116–1120. https://doi.org/10.1126/science.1238802

    CAS Article PubMed Google Scholar 

  73. Makarewich, C. A. (2020). The hidden world of membrane microproteins. Experimental Cell Research, 388, 111853. https://doi.org/10.1016/j.yexcr.2020.111853

    CAS Article PubMed PubMed Central Google Scholar 

  74. Makarewich, C. A., Baskin, K. K., Munir, A. Z., Bezprozvannaya, S., Sharma, G., Khemtong, C., Shah, A. M., McAnally, J. R., Malloy, C. R., Szweda, L. I., Bassel-Duby, R., & Olson, E. N. (2018). MOXI is a mitochondrial micropeptide that enhances fatty acid beta-oxidation. Cell Reports, 23, 3701–3709. https://doi.org/10.1016/j.celrep.2018.05.058

    CAS Article PubMed Google Scholar 

  75. Makarewich, C. A., & Olson, E. N. (2017). Mining for micropeptides. Trends in Cell Biology, 27, 685–696. https://doi.org/10.1016/j.tcb.2017.04.006

    CAS Article PubMed PubMed Central Google Scholar 

  76. Matsumoto, A., Pasut, A., Matsumoto, M., Yamashita, R., Fung, J., Monteleone, E., Saghatelian, A., Nakayama, K. I., Clohessy, J. G., & Pandolfi, P. P. (2017). mTORC1 and muscle regeneration are regulated by the LINC00961-encoded SPAR polypeptide. Nature, 541, 228–232. https://doi.org/10.1038/nature21034

    CAS Article PubMed Google Scholar 

  77. Meng, N., Chen, M., Chen, D., Chen, X. H., Wang, J. Z., Zhu, S., He, Y. T., Zhang, X. L., Lu, R. X., & Yan, G. R. (2020). Small protein hidden in lncRNA LOC90024 promotes “cancerous” RNA splicing and tumorigenesis. Advanced Science, 7, 1903233. https://doi.org/10.1002/advs.201903233

    CAS Article PubMed PubMed Central Google Scholar 

  78. Merino-Valverde, I., Greco, E., & Abad, M. (2020). The microproteome of cancer: From invisibility to relevance. Experimental Cell Research, 392, 111997. https://doi.org/10.1016/j.yexcr.2020.111997

    CAS Article PubMed Google Scholar 

  79. Meurisse, J., Bacquin, A., Richet, N., Charbonnier, J. B., Ochsenbein, F., & Peyroche, A. (2014). Hug1 is an intrinsically disordered protein that inhibits ribonucleotide reductase activity by directly binding Rnr2 subunit. Nucleic Acids Research, 42, 13174–13185. https://doi.org/10.1093/nar/gku1095

    CAS Article PubMed PubMed Central Google Scholar 

  80. Min, K. W., Davila, S., Zealy, R. W., Lloyd, L. T., Lee, I. Y., Lee, R., Roh, K. H., Jung, A., Jemielity, J., Choi, E. J., Chang, J. H., & Yoon, J. H. (2017). eIF4E phosphorylation by MST1 reduces translation of a subset of mRNAs, but increases lncRNA translation. Biochimica Et Biophysica Acta Gene Regulatory Mechanisms, 1860, 761–772. https://doi.org/10.1016/j.bbagrm.2017.05.002

    CAS Article PubMed Google Scholar 

  81. Nelson, B. R., Makarewich, C. A., Anderson, D. M., Winders, B. R., Troupes, C. D., Wu, F., Reese, A. L., McAnally, J. R., Chen, X., Kavalali, E. T., Cannon, S. C., Houser, S. R., Bassel-Duby, R., & Olson, E. N. (2016). A peptide encoded by a transcript annotated as long noncoding RNA enhances SERCA activity in muscle. Science, 351, 271–275. https://doi.org/10.1126/science.aad4076

    CAS Article PubMed PubMed Central Google Scholar 

  82. Nguyen, H. L., Yang, X., & Omiecinski, C. J. (2013). Expression of a novel mRNA transcript for human microsomal epoxide hydrolase (EPHX1) is regulated by short open reading frames within its 5’-untranslated region. RNA, 19, 752–766. https://doi.org/10.1261/rna.037036.112

    CAS Article PubMed PubMed Central Google Scholar 

  83. Odermatt, A., Taschner, P. E., Scherer, S. W., Beatty, B., Khanna, V. K., Cornblath, D. R., Chaudhry, V., Yee, W. C., Schrank, B., Karpati, G., Breuning, M. H., Knoers, N., & MacLennan, D. H. (1997). Characterization of the gene encoding human sarcolipin (SLN), a proteolipid associated with SERCA1: absence of structural mutations in five patients with Brody disease. Genomics, 45, 541–553. https://doi.org/10.1006/geno.1997.4967

    CAS Article PubMed Google Scholar 

  84. Orr, M. W., Mao, Y., Storz, G., & Qian, S. B. (2020). Alternative ORFs and small ORFs: shedding light on the dark proteome. Nucleic Acids Research, 48, 1029–1042. https://doi.org/10.1093/nar/gkz734

    CAS Article PubMed Google Scholar 

  85. Oyama, M., Itagaki, C., Hata, H., Suzuki, Y., Izumi, T., Natsume, T., Isobe, T., & Sugano, S. (2004). Analysis of small human proteins reveals the translation of upstream open reading frames of mRNAs. Genome Research, 14, 2048–2052. https://doi.org/10.1101/gr.2384604

    CAS Article PubMed PubMed Central Google Scholar 

  86. Pang, Y. N., Liu, Z. Y., Han, H., Wang, B. L., Li, W., Mao, C. B., & Liu, S. R. (2020). Peptide SMIM30 promotes HCC development by inducing SRC/YES1 membrane anchoring and MAPK pathway activation. Journal of Hepatology, 73, 1155–1169. https://doi.org/10.1016/j.jhep.2020.05.028

    CAS Article PubMed Google Scholar 

  87. Peeters, M. K. R., & Menschaert, G. (2020). The hunt for sORFs: A multidisciplinary strategy. Experimental Cell Research, 391, 111923. https://doi.org/10.1016/j.yexcr.2020.111923

    CAS Article PubMed Google Scholar 

  88. Plaza, S., Menschaert, G., & Payre, F. (2017). In search of lost small peptides. Annual Review of Cell and Developmental Biology, 33, 391–416. https://doi.org/10.1146/annurev-cellbio-100616-060516

    CAS Article PubMed Google Scholar 

  89. Polycarpou-Schwarz, M., Gross, M., Mestdagh, P., Schott, J., Grund, S. E., Hildenbrand, C., Rom, J., Aulmann, S., Sinn, H. P., Vandesompele, J., & Diederichs, S. (2018). The cancer-associated microprotein CASIMO1 controls cell proliferation and interacts with squalene epoxidase modulating lipid droplet formation. Oncogene, 37, 4750–4768. https://doi.org/10.1038/s41388-018-0281-5

    CAS Article PubMed Google Scholar 

  90. Pueyo, J. I., Magny, E. G., & Couso, J. P. (2016). New peptides under the s(ORF)ace of the genome. Trends in Biochemical Sciences, 41, 665–678. https://doi.org/10.1016/j.tibs.2016.05.003

    CAS Article PubMed Google Scholar 

  91. Raj, A., Wang, S. H., Shim, H., Harpak, A., Li, Y. I., Engelmann, B., Stephens, M., Gilad, Y., & Pritchard, J. K. (2016). Thousands of novel translated open reading frames in humans inferred by ribosome footprint profiling. eLife, 5, e13328. https://doi.org/10.7554/eLife.13328

    CAS Article PubMed PubMed Central Google Scholar 

  92. Rapali, P., Szenes, A., Radnai, L., Bakos, A., Pal, G., & Nyitray, L. (2011). DYNLL/LC8: a light chain subunit of the dynein motor complex and beyond. The FEBS Journal, 278, 2980–2996. https://doi.org/10.1111/j.1742-4658.2011.08254.x

    CAS Article PubMed Google Scholar 

  93. Rathore, A., Martinez, T. F., Chu, Q., & Saghatelian, A. (2018). Small, but mighty? Searching for human microproteins and their potential for understanding health and disease. Expert Review of Proteomics, 15, 963–965. https://doi.org/10.1080/14789450.2018.1547194

    CAS Article PubMed PubMed Central Google Scholar 

  94. Renz, P. F., Valdivia-Francia, F., & Sendoel, A. (2020). Some like it translated: small ORFs in the 5’ UTR. Experimental Cell Research, 396, 112229. https://doi.org/10.1016/j.yexcr.2020.112229

    CAS Article PubMed Google Scholar 

  95. Roschger, C., & Cabrele, C. (2017). The Id-protein family in developmental and cancer-associated pathways. Cell Communication and Signaling, 15, 7. https://doi.org/10.1186/s12964-016-0161-y

    CAS Article PubMed PubMed Central Google Scholar 

  96. Ruzinova, M. B., & Benezra, R. (2003). Id proteins in development, cell cycle and cancer. Trends in Cell Biology, 13, 410–418. https://doi.org/10.1016/s0962-8924(03)00147-8

    CAS Article PubMed Google Scholar 

  97. Saghatelian, A., & Couso, J. P. (2015). Discovery and characterization of smORF-encoded bioactive polypeptides. Nature Chemical Biology, 11, 909–916. https://doi.org/10.1038/nchembio.1964

    CAS Article PubMed PubMed Central Google Scholar 

  98. Savard, J., Marques-Souza, H., Aranda, M., & Tautz, D. (2006). A segmentation gene in Tribolium produces a polycistronic mRNA that codes for multiple conserved peptides. Cell, 126, 559–569. https://doi.org/10.1016/j.cell.2006.05.053

    CAS Article PubMed Google Scholar 

  99. Schlesinger, D., & Elsasser, S. J. (2021). Revisiting sORFs: overcoming challenges to identify and characterize functional microproteins. The FEBS Journalhttps://doi.org/10.1111/febs.15769

    Article PubMed Google Scholar 

  100. Slavoff, S. A., Heo, J., Budnik, B. A., Hanakahi, L. A., & Saghatelian, A. (2014). A human short open reading frame (sORF)-encoded polypeptide that stimulates DNA end joining. The Journal of Biological Chemistry, 289, 10950–10957. https://doi.org/10.1074/jbc.C113.533968

    CAS Article PubMed PubMed Central Google Scholar 

  101. Slavoff, S. A., Mitchell, A. J., Schwaid, A. G., Cabili, M. N., Ma, J., Levin, J. Z., Karger, A. D., Budnik, B. A., Rinn, J. L., & Saghatelian, A. (2013). Peptidomic discovery of short open reading frame-encoded peptides in human cells. Nature Chemical Biology, 9, 59–64. https://doi.org/10.1038/nchembio.1120

    CAS Article PubMed Google Scholar 

  102. Sousa, M. E., & Farkas, M. H. (2018). Micropeptide. PLoS Genetics, 14, e1007764. https://doi.org/10.1371/journal.pgen.1007764

    CAS Article PubMed PubMed Central Google Scholar 

  103. Starck, S. R., Tsai, J. C., Chen, K., Shodiya, M., Wang, L., Yahiro, K., Martins-Green, M., Shastri, N., & Walter, P. (2016). Translation from the 5’ untranslated region shapes the integrated stress response. Science, 351, aad3867. https://doi.org/10.1126/science.aad3867

    CAS Article PubMed PubMed Central Google Scholar 

  104. Staudt, A. C., & Wenkel, S. (2011). Regulation of protein function by “microProteins.” EMBO Reports, 12, 35–42. https://doi.org/10.1038/embor.2010.196

    CAS Article PubMed Google Scholar 

  105. Stein, C. S., Jadiya, P., Zhang, X. M., McLendon, J. M., Abouassaly, G. M., Witmer, N. H., Anderson, E. J., Elrod, J. W., & Boudreau, R. L. (2018). Mitoregulin: a lncRNA-encoded microprotein that supports mitochondrial supercomplexes and respiratory efficiency. Cell Reports, 23, 3710–3720

    CAS Article Google Scholar 

  106. Storz, G., Wolf, Y. I., & Ramamurthi, K. S. (2014). Small proteins can no longer be ignored. Annual Review of Biochemistry, 83, 753–777. https://doi.org/10.1146/annurev-biochem-070611-102400

    CAS Article PubMed PubMed Central Google Scholar 

  107. Vanderperre, B., Lucier, J. F., Bissonnette, C., Motard, J., Tremblay, G., Vanderperre, S., Wisztorski, M., Salzet, M., Boisvert, F. M., & Roucou, X. (2013). Direct detection of alternative open reading frames translation products in human significantly expands the proteome. PLoS ONE, 8, e70698. https://doi.org/10.1371/journal.pone.0070698

    CAS Article PubMed PubMed Central Google Scholar 

  108. Vitorino, R., Guedes, S., Amado, F., Santos, M., & Akimitsu, N. (2021). The role of micropeptides in biology. Cellular and Molecular Life Scienceshttps://doi.org/10.1007/s00018-020-03740-3

    Article PubMed Google Scholar 

  109. Wang, J., Zhu, S., Meng, N., He, Y., Lu, R., & Yan, G. R. (2019). ncRNA-encoded peptides or proteins and cancer. Molecular Therapy, 27, 1718–1725. https://doi.org/10.1016/j.ymthe.2019.09.001

    CAS Article PubMed PubMed Central Google Scholar 

  110. Wu, P., Mo, Y., Peng, M., Tang, T., Zhong, Y., Deng, X., Xiong, F., Guo, C., Wu, X., Li, Y., Li, X., Li, G., Zeng, Z., & Xiong, W. (2020a). Emerging role of tumor-related functional peptides encoded by lncRNA and circRNA. Molecular Cancer, 19, 22. https://doi.org/10.1186/s12943-020-1147-3

    CAS Article PubMed PubMed Central Google Scholar 

  111. Wu, S., Zhang, L., Deng, J., Guo, B., Li, F., Wang, Y., Wu, R., Zhang, S., Lu, J., & Zhou, Y. (2020b). A Novel micropeptide encoded by Y-Linked LINC00278 links cigarette smoking and AR signaling in male esophageal squamous cell carcinoma. Cancer Research, 80, 2790–2803. https://doi.org/10.1158/0008-5472.CAN-19-3440

    CAS Article PubMed Google Scholar 

  112. Wu, X., & Huang, M. (2008). Dif1 controls subcellular localization of ribonucleotide reductase by mediating nuclear import of the R2 subunit. Molecular and Cell Biology28, 7156–7167 https://doi.org/10.1128/MCB.01388-08

    CAS Article Google Scholar 

  113. Xia, X., Li, X. X., Li, F. Y., Wu, X. J., Zhang, M. L., Zhou, H. K., Huang, N. N., Yang, X. S., Xiao, F. Z., Liu, D. W., Yang, L. X., & Zhang, N. (2019). A novel tumor suppressor protein encoded by circular AKT3 RNA inhibits glioblastoma tumorigenicity by competing with active phosphoinositide-dependent Kinase-1. Molecular Cancer, 18, 131. https://doi.org/10.1186/s12943-019-1083-2

    Article PubMed PubMed Central Google Scholar 

  114. Xiao, Z. T., Huang, R. Y., Xing, X. D., Chen, Y. L., Deng, H. T., & Yang, X. R. (2018). De novo annotation and characterization of the translatome with ribosome profiling data. Nucleic Acids Research, 46, e61. https://doi.org/10.1093/nar/gky179

    CAS Article PubMed PubMed Central Google Scholar 

  115. Xu, W. L., Deng, B., Lin, P. H., Liu, C., Li, B., Huang, Q. J., Zhou, H., Yang, J. H., & Qu, L. H. (2020). Ribosome profiling analysis identified a KRAS-interacting microprotein that represses oncogenic signaling in hepatocellular carcinoma cells. Science China-Life Sciences, 63, 529–542. https://doi.org/10.1007/s11427-019-9580-5

    CAS Article PubMed Google Scholar 

  116. Yang, X., Tschaplinski, T. J., Hurst, G. B., Jawdy, S. S., Abraham, P. E., Lankford, P. K., Adams, R. M., Shah, M. B., Hettich, R. L., Lindquist, E., Kalluri, U.C., Gunter, L.E., Pennacchio, C., & Tuskan GA. (2011). Discovery and annotation of small proteins using genomics, proteomics, and computational approaches. Genome Research, 21, 634–641. https://doi.org/10.1101/gr.109280.110

    CAS Article PubMed PubMed Central Google Scholar 

  117. Yang, Y., Gao, X., Zhang, M., Yan, S., Sun, C., Xiao, F., Huang, N., Yang, X., Zhao, K., Zhou, H., Huang, S., Xie, B., & Zhang, N. (2018). Novel role of FBXW7 circular RNA in repressing glioma tumorigenesis. Journal of the National Cancer Institutehttps://doi.org/10.1093/jnci/djx166

    Article PubMed PubMed Central Google Scholar 

  118. Yin, X. Q., Jing, Y. Y., & Xu, H. M. (2019). Mining for missed sORF-encoded peptides. Expert Review of Proteomics, 16, 257–266. https://doi.org/10.1080/14789450.2019.1571919

    CAS Article PubMed Google Scholar 

  119. Zanet, J., Chanut-Delalande, H., Plaza, S., & Payre, F. (2016). Small peptides as newcomers in the control of drosophila development. Current Topics in Developmental Biology, 117, 199–219. https://doi.org/10.1016/bs.ctdb.2015.11.004

    CAS Article PubMed Google Scholar 

  120. Zhang, H., Wang, Y. R., Wu, X. K., Tang, X. L., Wu, C. C., & Lu, J. (2021). Determinants of genome-wide distribution and evolution of uORFs in eukaryotes. Nature Communications, 12, 1076. https://doi.org/10.1038/s41467-021-21394-y

    CAS Article PubMed PubMed Central Google Scholar 

  121. Zhang, M., Huang, N., Yang, X., Luo, J., Yan, S., Xiao, F., Chen, W., Gao, X., Zhao, K., Zhou, H., Li, Z., Ming, L., Xie, B., & Zhang, N. (2018a). A novel protein encoded by the circular form of the SHPRH gene suppresses glioma tumorigenesis. Oncogene, 37, 1805–1814. https://doi.org/10.1038/s41388-017-0019-9

    CAS Article PubMed Google Scholar 

  122. Zhang, M., Zhao, K., Xu, X., Yang, Y., Yan, S., Wei, P., Liu, H., Xu, J., Xiao, F., Zhou, H., Yang, X., Huang, N., Liu, J., He, K., Xie, K., Zhang, G., Huang, S., & Zhang, N. (2018b). A peptide encoded by circular form of LINC-PINT suppresses oncogenic transcriptional elongation in glioblastoma. Nature Communications, 9, 4475. https://doi.org/10.1038/s41467-018-06862-2

    CAS Article PubMed PubMed Central Google Scholar 

  123. Zhang, Q., Vashisht, A. A., O’Rourke, J., Corbel, S. Y., Moran, R., Romero, A., Miraglia, L., Zhang, J., Durrant, E., Schmedt, C., Sampath, S. C., & Sampath, S. C. (2017). The microprotein Minion controls cell fusion and muscle formation. Nature Communications, 8, 15664. https://doi.org/10.1038/ncomms15664

    CAS Article PubMed PubMed Central Google Scholar 

  124. Zhang, S., Reljic, B., Liang, C., Kerouanton, B., Francisco, J. C., Peh, J. H., Mary, C., Jagannathan, N. S., Olexiouk, V., Tang, C., Fidelito, G., Nama, S., Cheng, R. K., Wee, C. L., Wang, L. C., Roggli, P. D., Sampath, P., Lane, L., Petretto, E., et al. (2020). Mitochondrial peptide BRAWNIN is essential for vertebrate respiratory complex III assembly. Nature Communications, 11, 1312. https://doi.org/10.1038/s41467-020-14999-2

    CAS Article PubMed PubMed Central Google Scholar 

  125. Zhang, Y., Fonslow, B. R., Shan, B., Baek, M. C., & Yates, J. R., 3rd. (2013). Protein analysis by shotgun/bottom-up proteomics. Chemical Reviews, 113, 2343–2394. https://doi.org/10.1021/cr3003533

    CAS Article PubMed PubMed Central Google Scholar 

  126. Zhao, J. W., Lee, E. E., Kim, J., Yang, R., Chamseddin, B., Ni, C. Y., Gusho, E., Xie, Y., Chiang, C. M., Buszczak, M., Zhan, X. W., Laimins, L., & Wang, R. C. (2019). Transforming activity of an oncoprotein-encoding circular RNA from human papillomavirus. Nature Communications, 10, 2300. https://doi.org/10.1038/s41467-019-10246-5

    CAS Article PubMed PubMed Central Google Scholar 

  127. Zhao, X., Chabes, A., Domkin, V., Thelander, L., & Rothstein, R. (2001). The ribonucleotide reductase inhibitor Sml1 is a new target of the Mec1-Rad53 kinase cascade during growth and in response to DNA damage. The EMBO Journal, 20, 3544–3553. https://doi.org/10.1093/emboj/20.13.3544

    CAS Article PubMed PubMed Central Google Scholar 

  128. Zheng, X., Chen, L. J., Zhou, Y., Wang, Q., Zheng, Z. J., Xu, B., Wu, C., Zhou, Q., Hu, W. W., Wu, C. P., & Jiang, J. T. (2019). A novel protein encoded by a circular RNA circPPP1R12A promotes tumor pathogenesis and metastasis of colon cancer via Hippo-YAP signaling. Molecular Cancer, 18, 47. https://doi.org/10.1186/s12943-019-1010-6

    CAS Article PubMed PubMed Central Google Scholar 

  129. Zhi, X., Zhang, J., Cheng, Z., Bian, L., & Qin, J. (2019). circLgr4 drives colorectal tumorigenesis and invasion through Lgr4-targeting peptide. International Journal of Cancerhttps://doi.org/10.1002/ijc.32549

    Article PubMed Google Scholar 

  130. Zhou, B., Yang, H., Yang, C., Bao, Y. L., Yang, S. M., Liu, J., & Xiao, Y. F. (2021). Translation of noncoding RNAs and cancer. Cancer Letters, 497, 89–99. https://doi.org/10.1016/j.canlet.2020.10.002

    CAS Article PubMed Google Scholar 

  131. Zhu, S., Wang, J. Z., Chen, D., He, Y. T., Meng, N., Chen, M., Lu, R. X., Chen, X. H., Zhang, X. L., & Yan, G. R. (2020). An oncopeptide regulates m(6)A recognition by the m(6)A reader IGF2BP1 and tumorigenesis. Nature Communications, 11, 1685. https://doi.org/10.1038/s41467-020-15403-9

    CAS Article PubMed PubMed Central Google Scholar 

Download references

Acknowledgements

We thank Dr. Lili Li for help in figure preparation. This work was funded by the Beijing Municipal Natural Science Foundation (Grant no. 5204036); National Key R&D Program of China (Grant no. 2019YFA0903900); National Natural Science Foundation of China (Grant no. 31630005, 31770084).

Author information

Affiliations

  1. Guangdong Key Laboratory for Genome Stability & Disease Prevention, Shenzhen University School of Medicine, Shenzhen, 518060, China

    Meiqian Jiang, Huiqiang Lou & Wenya Hou

Corresponding author

Correspondence to Wenya Hou.

Rights and permissions

Reprints and Permissions

About this article

Cite this article

Jiang, M., Lou, H. & Hou, W. Microproteins: from behind the scenes to the spotlight. GENOME INSTAB. DIS. 2, 225–239 (2021). https://doi.org/10.1007/s42764-021-00040-3

Download citation

Share this article

Anyone you share the following link with will be able to read this content:

Get shareable link

Provided by the Springer Nature SharedIt content-sharing initiative

Keywords

  • Microproteins

  • Small open reading frames

  • Genome stability

  • DNA repair

  • DNA replication

  • Cancer



About this journal

Electronic ISSN
2524-7662
Abstracted and indexed in
  1. CNKI
  2. Dimensions
  3. EBSCO Discovery Service
  4. Google Scholar
  5. Institute of Scientific and Technical Information of China
  6. Meta
  7. Naver
  8. OCLC WorldCat Discovery Service
  9. ProQuest-ExLibris Primo
  10. ProQuest-ExLibris Summon
  11. TD Net Discovery Service
Copyright information

Rights and permissions

Springer policies

© Shenzhen University School of Medicine; Fondazione Istituto FIRC di Oncologia Molecolare

×

用户登录