The role of small and long non-coding RNAs in cardiac pathologies
Review Article

The role of small and long non-coding RNAs in cardiac pathologies

Vittoria Di Mauro1,2, Daniele Catalucci1,2

1Humanitas Clinical and Research Center, IRCCS, Rozzano, Milan, Italy; 2National Research Council (CNR), Institute of Genetic and Biomedical Research (IRGB), UOS Milano, Milan, Italy

Contributions: (I) Conception and design: V Di Mauro; (II) Administrative support: None; (III) Provision of study materials or patients: None; (IV) Collection and assembly of data: None; (V) Data analysis and interpretation: None; (VI) Manuscript writing: All authors; (VII) Final approval of manuscript: All authors.

Correspondence to: Vittoria Di Mauro; Daniele Catalucci. Humanitas Clinical and Research Center, IRCCS, via Manzoni 56, 20089 Rozzano (Mi), Milan, Italy; National Research Council (CNR), Institute of Genetic and Biomedical Research (IRGB), UOS Milano, via Fantoli 15/16, 20138 Milan, Italy. Email: Vittoria.Di_Mauro@humanitasresearch.it; daniele.catalucci@cnr.it.

Abstract: Cardiovascular disorders (CVDs) still remain the leading cause of death worldwide. The current lack of therapeutic approaches that efficiently improve CVD quality of care and prevent the development of life-threatening complications, prompted the scientific community to continue a deeper investigation of the molecular mechanisms driving the onset and progression of these pathologies. In this context, non-coding RNAs (ncRNAs) were demonstrated to be involved in the onset of different forms of CVDs. In this review we will discuss basic aspects of some classes of ncRNAs as well as their mechanism of action and involvement in CVDs. The potential therapeutic use of ncRNAs in the clinical practice will also be addressed.

Keywords: Cardiovascular disorders (CVDs); non-coding RNAs (ncRNAs); microRNAs (miRNAs); and long non-coding RNAs (lncRNAs); nanotechnology


Received: 20 March 2019; Accepted: 23 May 2019; Published: 05 June 2019.

doi: 10.21037/ncri.2019.05.03


Introduction

Cardiovascular disorders (CVDs) define a wide range of pathologies affecting heart muscle and its associated circulatory system (1,2). In particular, heart failure (HF) and myocardial infarction (MI) still represent the leading cause of morbidity in the industrialized world and it is also estimated to became the major source of death in the developing countries by 2020 (3,4). Additionally, the occurrence of co-morbidities such as hypertension, diabetes, and coronary artery disease threaten the efficacy of current pharmacological therapies that just alleviate the consequence of HF and fail to completely restore the functionality of injured myocardium (5). As consequence, heart-transplantation still remains the gold standard treatment for end-stage HF. However, due to the growing number of patients in combination with the lack of donors as well as the related immunological complications following organ transplantation, alternative and more efficient clinical approaches are strongly required to be discovered. In line with this, an increasing interest for deeper characterization of key regulators of molecular mechanisms underlying cardiac dysfunctions has started by the scientific community (6,7). In this context, innovative findings were indeed obtained leading to the displacement of a central dogma that has dominated the scientific landscape for half a century, the fact that only proteins represent the leading effectors of cellular functions and so responsible for the onset of diseases (8). Initiatives like encyclopaedia of DNA elements (ENCODE) (9) followed by the latest advances of deep DNA/RNA-sequencing technologies [next generation sequencing (NGS)] (10) and the large amount of public available bioinformatics tools for collecting and processing data, allowed the discover that around 98% of the entire genome is actively transcribed into RNA without coding potential, thus called non-coding RNAs (ncRNAs) (11-13). ncRNAs, initially thought to be “functionless-transcriptional noise”, were on the other hand extensively demonstrated to provide an active contribution in a plethora of biological processes. A significant correlation between their altered expression and the onset or progression of different human disorders, included CVDs, was found (11,14). Since then, the ncRNA field has increasingly received a great attention from both the research community and industry with a particular interest for their potential use as diagnostic and therapeutic tools in medicine (15).

In this review, the role of the two major groups of ncRNAs [microRNAs (miRNAs) and long ncRNAs (lncRNAs)] in cardiac pathologies will be discussed. In particular we will focus on ncRNAs participating to the pathogenesis of different CVDs and how they can be used as therapeutic tools for novel clinical strategies (16). Furthermore, there will be an overview of up-to-date challenges and possible strategies for their clinical applications.


Categories of ncRNAs and basic function

ncRNAs can be classified based on an arbitrary threshold of 200 nucleotides (nt), separating small ncRNAs (sncRNAs) and mid-size RNAs (mid-sizeRNAs) from long ncRNAs (lncRNAs) (17,18). The sncRNA family includes microRNAs (miRNAs), small interference RNAs (siRNAs), and piwi-interacting RNA (piRNA), while the mid-size category comprises transfer RNAs (tRNAs) and small nucleolar RNAs (snoRNAs). Ultimately, the long ncRNA groups mainly refers to long intergenic non-coding RNA (lincRNA) and natural antisense transcripts (NATs) (19,20). In addition, other type of ncRNAs, which fall into both sides of canonical length cut-off groups, such as enhancer-associated RNAs (eRNAs) and the more recently emerged circular RNAs (circRNAs), have been recently identified (17).

miRNAs

Among small ncRNAs, miRNAs are so far the most widely studied class (16). miRNAs are endogenous, single-stranded ncRNA molecules of 18–24 nt in length. It is estimated to be the most abundant family in the human genome, which indeed hosts more than 2000 different loci for miRNAs production (21). To date the majority of identified miRNAs are “intragenic” and processed mainly from introns rather than exons of protein coding genes, while the remaining are “intergenic”, transcribed independently of a host gene and controlled by their own promoters (22-24).

miRNAs are produced via a complex sequence of events orchestrated by a multitude of enzymes and proteins (25). In brief, miRNA gene is initially transcribed by RNA polymerase II into a primary transcript (pri-miRNA), which it is then processed into a shorter precursor (pre‐miRNAs) by the ribonuclease III termed DROSHA and exported outside the nucleus (26). In the cytosol, the RNAse III protein called Dicer, converts the pre‐miRNA into a double‐stranded mature miRNA molecule. Finally, one strand of the mature miRNA is loaded into the Argonaute (AGO) protein to form the miRNA-induced silencing complex (miRISC) complex. This complex recognizes either the 3′ or 5′ untranslated region of different mRNA genes, inducing their translational repression and/or mRNA degradation (27,28). Beyond the classical cytosolic post-transcriptional mechanism of action, numerous studies highlighted how miRNAs can regulate the expression of target gene also in different subcellular compartments such as rough endoplasmic reticulum (rER) (29,30), processing (P)-bodies (30), mitochondria (31) and the nucleus (32-34) (and Di Mauro et al, unpublished). In particular within nuclei, mature miRNAs where demonstrated to elicit novel and unconventional functions, although poorly understood, such as the induction of epigenetic changes by a direct binding to DNA (35), or even the promotion of more efficient splicing or alternative splicing profiles of nascent mRNA transcripts (32). Ultimately, when loaded into secreted vesicles, miRNAs can be released in extracellular fluids and possibly elicit their function in other cell types or tissue districts (36).

piRNAs

The piRNAs represent another large family of small ncRNAs, existing throughout the animal kingdom, particularly enriched in the germ-line tissues (37,38). piRNAs are 24–32 nt in length and are generated from the RNA transcripts of transposons, protein-coding genes, and specific intergenic loci (39). Similarly to miRNAs, piRNAs are able to form a RISC with a specific sub-group of AGO protein family, called PIWI (P-element induced wimpy testis) protein clade (40). The resulting RNA-protein complex (piRISC) is well known to negatively regulate gene transcription via endonucleolytic cleavage (slicing) of the target sequence after complementary base-pair recognition through the piRISC. Such target sequences are mainly represented by transposable elements (TEs), which are autonomous pieces of DNA that replicate and able to insert into the genome and therefore potentially prone to cause DNA damage (38,41,42). Moreover, new functions beyond transposon silencing were uncovered, such as the regulation of proteins coding gene via DNA methylation (43) or repression by mRNA deadenylation (44,45). Surprisingly, recent papers demonstrated that piRNAs are not confined in germ line tissues but they are also present in somatic cells of brain, liver, and heart (46). In the cardiovascular area, the knowledge of piRNAs is still far from being completed defined. Nevertheless there is strong evidence demonstrating that piRNAs might be involved in the control of signaling pathways activated during stress stimuli (47,48) as well as being differentially expressed in exosomes isolated from serum of HF patients (49). Altogether, these data suggest both an active role of piRNAs in the onset of pathological cardiac remodelling and their potential use as prognostic biomarkers.

siRNAs

This class of ncRNAs and their related mechanism of action, called RNA interference (RNAi), was firstly described in 1998 by a pioneering work in C.elegans (50). In this paper the authors showed in worms that the injection of double-stranded RNAs coding for specific proteins were able to silence genes carrying the same sequence (51). From a structural and functional point of view, siRNAs resemble miRNAs, however, some key differences between these two classes of ncRNAs are present:

  • siRNA molecules are shorter, with a mean length of 21nt (51);
  • siRNA molecules require a perfect complementarity with target mRNA, thus allowing the knock down of gene with less off-target exceptions (52);
  • siRNA-mediated gene silencing occurs trough an endonucleolytic cleavage of mRNA (52,53).

Initially defined as extragenomic RNA material, siRNAs are now accepted to be present endogenously in organisms such as plant, flies, and mice (54-56). However, the complete comprehension of endogenous siRNAs is still quite limited, therefore in order to further dissect the origin and biogenesis in mammals of this class of ncRNAs, an increase research in this field is essential.

Mid-size ncRNAs

tRNAs represent a type of ubiquitous RNAs, codified by tRNA genes and subsequently transcribed by RNA polymerase III into macromolecules whose function is to match an mRNA codon with the amino acid it codes for (57). As a matter of fact tRNAs transfer activated amino-acids from aminoacyl-tRNA synthetases to the ribosome, where they are used for the protein synthesis (5). However, various works pointed out additional roles for tRNAs beyond being a simple adaptor molecule as well as tRNA-associated genetic disorders (58,59). snoRNAs are a conserved class of 60–300 nt long stable RNAs, present in nucleolus of cells that play important role in ribosome biogenesis (60), as well as in the regulation of alternative splicing and posttranscriptional modification of mRNA (61).


lncRNAs

The overwhelming advance in transcriptome sequencing, epi-genomic technologies, and computational prediction techniques allowed the discovery of lncRNAs, which among all sub-categories of ncRNAs represent the most recent and thus the least characterized family (62-64). Similarly to sncRNAs, lncRNAs were found in basically all type of organisms, which include, despite a poor rate of conservation between species, virus, yeast, plants, and animals (65-68). The biogenesis of lncRNAs starts into the nucleus where RNA polymerase II generate transcripts, which bear some features resembling mRNA molecules such as the classical splice sites (GU/AG), the presence of intron and exon structures with related alternative splicing, 5’-capping and a final polyadenilation (69). However, in contrast to mRNAs, the lncRNA transcripts have little or no functional open reading frame (ORF) (62,70). So far it was estimated that 75–90% of the human genome generate a series of lncRNAs (71,72), but despite this, to date, a standard definition for this family of ncRNAs is still missing. Some empirical features can be used for a sort of classification:

  • Length: lncRNAs are transcripts whose length span from 200 nt up to 100 kilobases (kb) (72);
  • Genomic location and context: based on the genomic position relative to the nearest coding gene, lncRNA genes can be divided into:
    • Intergenic (also called lincRNAs): transcribed from introns of annotated genes (69);
    • Intronic: transcribed entirely from introns of protein-coding genes (69);
    • Bidirectional (divergent): generated from both sense and antisense direction of transcription start areas (69,73);
    • Sense or overlapped (intronic or exonic): produced from the sense strand of annotated transcription units (UTs). These transcripts can contain exons and derive from protein-coding genes, can overlap with part of protein-coding genes or cover the entire sequence of a protein-coding gene through an intron (74);
    • Antisense (intronic or NAT, natural antisense transcript) produced from the antisense strand of annotated transcription units (UTs) (75);
  • Subcellular localization: mounting lines of evidence demonstrated a link between subcellular distributions of lncRNAs and their function (76). So far the majority of annotated lncRNAs were found more enriched in the nucleus where they are involved in nuclear processes such as transcription and RNA processing (77). On the contrary, other lncRNAs were found more abundant in the cytosolic compartment, where their function can impact on mRNA stability and translation or even on localization of final protein products (76). Interestingly, in the last years different several works demonstrated the presence of lncRNAs in the extracellular space by a selective sorting into exosomes, thus revealing an additional role in mediating cell-cell communication (78,79).
  • Function: based on the evidence that many lncRNAs have a cell-type specific expression and respond to different stimuli, they are considered “signal molecules” due to their ability to integrate different hints in precise cellular and temporal contest (80). lncRNAs are also known to exert a role of “molecular decoy”, whose function consists in binding and titrating away proteins to target genes (81). Additionally, lncRNAs can act as “guides” by interacting with regulatory proteins, such us chromatin-remodelling complex, and directing them to specific DNA loci in order to induce epigenetic modifications (13,62). Other lncRNAs can play a “scaffold role” acting as platforms to bring different proteins together, in the nucleus or cytosol, at the aim to activate or repress the transcription or translation of target gene (62). Moreover, lncRNAs can counteract the action of miRNAs, by “sponging” them, therefore favoring the expression of repressed target mRNAs (5).

Despite the fast increase of data regarding the lncRNAs, this field is still far from being completely understood, especially for what concerning the function of this class of ncRNAs. Indeed, it is plausible that in the near future, more lncRNA-mediated functions will emerge and with them also many questions and challenges associated with them.


eRNAs and circRNAs

Although enhancers have been identified more than 30 years ago, only it was only in 2010 when the existence of enhancer-related RNAs transcripts was demonstrated and subsequently termed eRNAs (82,83). Despite eRNA field is still in its infancy, here we will try to briefly report the current knowledge concerning their molecular features as well as some information about their mechanism of actions.

eRNAs can be distinguishable from lncRNAs by some specific features:

  • Epigenetic signature: eRNAs are marked by histone modifications such as H3K4me1/2 and H3K27Ac unlike from promoters of lncRNAs which are specifically enriched in H3K3me3 (84);
  • Origin: eRNAs are generated from bidirectional transcription of both strands of DNA from a central non-transcribed region (82,85,86);
  • Length: eRNAs are relatively shorter than lncRNAs, with a range between 500-2000 nt (87);
  • Post-transcriptional modifications: generally after transcription, eRNAs are not subjected to full maturation processes, indeed they possess a 5′ cap but they are unspliced and rarely polyadenylated (88-90).
  • Half life: eRNAs are mainly retained into nuclei therefore they show a very short half-life with a degradation process within minutes (91).

Since eRNAs were initially considered as merely transcriptional noise, also their function was consequently underestimated. Nevertheless, in the last years different functions were demonstrated to be associated to eRNAs. As a matter of fact, eRNAs can regulate activities of target or neighbor genes in different ways, such as affecting their transcriptional elongation by releasing the negative elongation factor (NELF) from paused RNA polymerase II at specific gene promoters (92,93). In addition, eRNAs can initiate the transcription of target genes by directly opening the chromatin or trough recruitment of chromatin remodelers (94). Moreover, eRNAs can also act as bridge to facilitate a physical interaction between enhancers and promoters (E:P Loop) independently from distance and orientation (95,96).

circRNAs represent a subset of ncRNAs initially considered as byproducts of defective splicing due to their very low expression and uncommon features, therefore did not receiving much attention (97). However, the high-throughput sequencing technology coupled with development of high specific computational algorithms [e.g., circRNAs_finder (98,99), find_circ (99), CIRI (100) and CIRCexplorer (101)] overturned this vision by the discover of thousands of endogenous circRNAs (102). CircRNAs present a unique structure resulting from a 3′ to 5′ end-joining event also called backsplicing (103). CircRNAs are generated from both exons and introns of protein coding genes, although the latter are less frequent (102). Moreover, there are pieces of evidence showing that they can also took origin from intergenic regions, UTRs, and ncRNA loci (99). Once transcribed, circRNAs can be retained into nuclear compartment or exported in the cytoplasm. For what concerning their biological function, five main mechanisms have been so far proposed:

  • miRNA-sponge: by harboring binding sites for miRNAs, circRNAs can sequester these ncRNAs and therefore indirectly regulate the activity of miRNA-target genes (99,104). However, this function is controversial and nowadays considered not a general phenomenon of circRNAs (105). In fact, to feature proper miRNA-sponge activities, circRNAs must fulfill some principal criterions:
    • To harbor a large number of binding sites for the putative miRNA targets;
    • To target few number of miRNAs;
    • To show higher circRNA-miRNA affinity compared to miRNA-mRNA ones;
    • To induce degradation of miRNA targets.
  • Post-transcriptional control: circRNAs enriched into cell nuclei can interact with elongating RNA Pol II and promote transcription of related genes (106);
  • Rolling circle translation: circRNAs containing open reading frame (ORF) or internal ribosome entry site (IRES) can be translated into proteins via a rolling circle amplification mechanism. However, this mechanism is only demonstrated in vitro, and so far there are no evidence showing that natural eukaryotic endogenous circRNAs can be effectively translated into peptides (107,108);
  • Production of circRNA-derived pseudogenes: some circRNAs can be reversely transcribed to cDNA and even integrated into the genome (109);
  • Control of alternative splicing: this feature is handled by circRNAs derived from exons of protein coding genes. Indeed, biogenesis of such circRNAs can hinder the pre-mRNA splicing, resulting in lower levels of linear mRNAs or even changing the composition of processed mRNA by selective exons exclusion (110).

Despite the number of annotated circRNAs is in rapid expansion, so far only a small group of them have been clearly correlated with CVDs (111). Indeed, compared to other RNAs, the current knowledge of circRNAs still needs to face some concrete obstacles. For example, the relatively low abundance of circRNAs as well as their unique circular structure makes current techniques for detection, manipulation, and other functional studies sill poor efficient (105). As a consequence, for an effective application of circRNAs in clinical management of cardiac diseases, the scientific community must accomplish significant efforts in the discovery and characterization of this class of ncRNAs.


The role of miRNA and lncRNAs in cardiovascular disorders

Starting from the embryonic stage to fully developed adult heart, ncRNAs were detected in many cell lineages forming the cardiac tissues where they contribute to processes such as differentiation, proliferation, contractility, and electric conduction system (112). Therefore, any alterations in ncRNA expression can determine the onset or progression of cardiac pathologies. Here we will report few examples of miRNAs and lncRNAs affecting the pathophysiology of heart.

Cardiac fibrosis

Cardiac fibrosis is a pathological consequence related to heart damages, characterized by the adverse accumulation of collagens and other extracellular matrix (ECM) proteins (113). Nowadays a number of investigations demonstrated that expression of distinct miRNAs and lncRNAs, strongly correlate with the genesis, progression, and treatment of cardiac fibrosis (114). One of the first study in this context was made by the Olson’s group that found a negative correlation between the expression of miR-29 family members (miR-29a, miR-29b and miR-29c) and genes involved in ECM production and fibrosis after experimental MI (115). Another fibrotic-related miRNA was described in a work by Thum et al, where the authors demonstrated that miR-21 promotes cardiac fibrosis by targeting extracellular regulated kinase inhibitor sprouty homolog 1 (Spry1) augmenting ERK-MAP kinase signaling in cardiac fibroblasts (116). Several other miRNAs have been identified as potent anti-fibrotic molecules, such us Let-7i and miR-26a which were found to attenuate collagen deposition by targeting Col1α2 and Col1α1, respectively (117,118). Although the current knowledge regarding fibrotic-related lncRNAs remains poor, there are available scientific lines of evidence underlining their involvement in controlling cardiac fibrosis. This is the case of the conserved Wisp2 super-enhancer-associated RNA (Wisper), which was described to control the progression of this disease by up-regulating pro-fibrotic genes (119). In another work, in vivo inhibition of lncRNA Meg3 during the early phase of cardiac remodelling was able to prevent the induction of matrix metalloproteinase-2 (MMP-2), therefore to decrease cardiac fibrosis and improved diastolic function (120).

Cardiac hypertrophy

Biomechanical stress and other pathological stimuli can trigger the initiation of a phenotypic remodelling, i.e., cardiac hypertrophy, which results in an increase in myocyte size and myofibrillar volume (121). Although beneficial for a short period, prolonged hypertrophic state can lead to HF and death (122). Among the first identified hypertrophic-related miRNAs, there are miR-1 and miR-133. The inverse correlation between their expression and progression of cardiac hypertrophy in a rodent model of cardiac hypertrophy led to the identification of a relevant role in this disorder (122). Indeed, in vitro studies using neonatal rodent myocytes demonstrated that miR-1 was able to target several hypertrophic genes such as Ras GTPase-activating protein (RasGAP), cyclin-dependent kinase 9 (Cdk9), Ras homolog enriched in brain (Rheb), and fibronectin (123). In addition, the work of Carè et al, showed for the first time that an in vivo administration of a synthetic molecule (i.e., antagomir, discussed in next paragraphs) able to down-regulate the level of cardiac miR-133a, led to an hypertrophic response of the myocardial tissues (124). Other miRNAs on the contrary were demonstrated to have pro-hypertrophic effects. This is the case of miR-19a/b family, whose members downregulate the anti-hypertrophic genes atrogin 1 and muscle ring finger protein 1 (Murf1) with the consequent activation of calcineurin/nuclear factor of activated T cells (NFAT) signaling (125).

Among the first lncRNAs interrogated for their involvement in hypertrophic remodeling, there is the myosin heavy chain associated RNA transcripts (MHRT), which antagonizes the brahma-related gene 1 (Brg1) by blocking its recognition sites on cardiac stress genes, thus protecting the heart from pathological HF (126). More recently, the lncRNA cardiac-hypertrophy-associated epigenetic regulator (Chaer) was described as an important checkpoint for the progression of cardiac hypertrophy. Indeed, Chaer promotes an activation of stress-related genes, by tethering away from them the polycomb repressor complex 2 (PRC2) (127). Recently, another lncRNA called TINCR was demonstrated to attenuate cardiac hypertrophy by the epigenetic silencing of CaMKII (128).

MI

MI is a common cardiovascular event characterized by cardiomyocytes loss via programmed cell death (129). In this contest miRNA can either promote or impair cardiomyocyte survival, thus acting as angel or devil in the regulatory network of cardiac cell death (113). An example of negative regulators of cell death is represented by miR-15 family, whose chemical inhibition was demonstrated to reduce myocardial damaged area after ischaemia–reperfusion injury (130). On the contrary miR-24, when overexpressed, was reported to exert a positive effect on cardiomyocytes survival by targeting the proapoptotic Bcl-2-like protein 11 (131). Analogously to miRNAs, there are also examples of cell death-related lncRNAs. The long noncoding RNA, named autophagy promoting factor (APF), can regulate autophagic cell death by targeting miR-188-3p and ATG7 (132). More recently, the GATA1 activated lncRNA Galont, was demonstrated to interact with miR-338 and to promote ATG5-mediated autophagic cell death in murine cardiomyocytes (133) (Tables 1,2).

Table 1

miRNAs involved in cardiac disorders

miRNA/miRNA family Target Associated disease Reference
miR-29 Elastin (Eln); Fibrillin 1 (Fbn1); collagen type I, alpha 1 and 2 (Col1α1, Col1α2) and collagen type III, alpha 1 (Col3α1) Cardiac fibrosis 117
miR-21 Kinase inhibitor sprouty homolog 1 (Spry1) Cardiac fibrosis 118
Let-7i Col1α2 Cardiac fibrosis 119
Mir-26 Col1α1 Cardiac fibrosis 120
miR-133a Ras homolog gene family, member ARhoA; Cell division control protein 42 homolog (Cdc42); Wolf-Hirschhorn syndrome candidate 2Nelf-A/WHSC2 Cardiac hypertrophy 126
miR-1 Ras GTPase-activating protein (RasGAP), cyclin-dependent kinase 9 (Cdk9), Ras homolog enriched in brain (Rheb), and fibronectin (FN1) Cardiac hypertrophy 125
miR-19a/b atrogin 1 (Fbx32) and muscle ring finger protein 1 (Murf1) Cardiac hypertrophy 127
miR-15 BCL2 Apoptosis Regulator (Bcl2) MI 132
mir-34 Bcl-2-like protein 11 (Bcl2l11) MI 133

MI, myocardial infarction.

Table 2

lncRNAs involved in cardiac pathologies

lncRNA Target Associated disease Reference
Wisper Col3α2, Fn1, Tgfb2 Cardiac fibrosis 121
Meg3 matrix metalloproteinase-2 (MMP-2) Cardiac fibrosis 122
MHRT Brahma-related gene 1 (Brg1) Cardiac hypertrophy 128
Chaer Polycomb repressor complex 2 (PRC2 Cardiac hypertrophy 129
TINCR Calcium/calmodulin dependent protein kinase ii (CaMKII) Cardiac hypertrophy 130
APF miR-188-3p, Autophagy Related 7 (ATG7) MI 134
Galont miR-338 MI 135

MI, myocardial infarction.


miRNA and lncRNA therapeutics: on the road for promises and challenges of RNA-based therapies

Tools for modulation of ncRNAs

As consequence of the latest advances in next-generation sequencing technology and the huge availability of generated genomic data, the scientific community managed to identify a plethora of key regulators of various pathologies, many of which successfully translated into therapeutic compounds for the treatment of CVDs (10). So far the two main classes of Food and Drug Administration (FDA) approved drugs are divided into (I) “small molecules”, largely characterized by organic hydrophobic compounds and (II) “proteins” mainly represented by antibodies. The small molecules, thanks to their reduced size are able to cross cell membrane by rapid diffusion and directly inhibiting or activating target proteins, such as receptors or enzymes (134,135). However, the prerequisite of small molecules-druggable targets is the presence of active-site pockets, and because of only the 2–5% of proteins in the human genome possess such binding pockets, the applicability of such compounds is quite limited (135,136). In this context, antibody-based drugs can be used to overcome this limitation by directing replacing mutated or missing proteins, on the other hand a larger size does not allow the cross of cellular barrier by diffusion, thus restricting their use to extracellular targets (135). The need to overcome such limitations, combined with the discovery of altered ncRNAs in different pathologies as well as their active role in controlling disorders, fuelled the academia and industry to consider RNA‐based therapeutics suitable tools for treatment and prevention of different diseases that lead up this year to the first ever FDA-approved ncRNA based drug, Onpattro (patisiran), for the treatment of polyneuropathy caused by hereditary transthyretin-mediated amyloidosis (hATTR) (137,138).

In this paragraph we will briefly discuss the current available strategies to modulate miRNA and lncRNAs functions.

So far two approaches are commonly used to modulate miRNA activity: 1) to inhibit the miRNA function by using chemically modified anti-miR oligonucleotides; 2) to restore miRNA function trough synthetic double-stranded-based overexpression system (139). Inhibition of miRNA-targeting can be achieved by the use of antisense oligonucleotide, the so called anti-miR, and in this group are comprised the miRNA sponges, the locked nucleic acid (LNA) anti-miR constructs and the antagomir (140). The miRNA sponges approach relies on the use of overexpression vectors harbouring complementary binding sites for a miRNA (or a family) whose function is blocking it by competing for the binding with bona fide target genes (141). The last two technologies are so far the most used, and they are synthetic oligonucleotides either fully or partially complementary to a specific miRNAs, which contain chemical modifications aimed to increase their binding affinity, the biostability and pharmacokinetic properties (142). The LNA modification is a conformational restriction based on the use of a methylene bridge between the 2’ oxygen with the 4’carbon of the ribose ring, demonstrated to generate a thermodynamically strong duplex formation towards complementary single-stranded RNA molecules both in vitro and in vivo (143). The antagomir molecules are characterized by the insertion of the 2′-O-methoxyethyl modification, a partial phospho-rothioate backbone (PS) and a covalent addition of a cholesterol molecule. In particular these last two modifications were demonstrated to improve the nuclease resistance and the cellular up-take respectively (144,145).

For what concerning pathological disorders related to a low level of miRNAs, so far, the most used strategy to restore miRNA level is represented by mimic molecules. miRNA-mimic molecules are double or triple stranded synthetic miRNA oligonucleotides, which inside cells are processed into a single strand form and regulate protein coding genes via miRNA-like function (146). Similarly to the antisense oligonucleotide, also miRNA-mimics possess different chemical modifications in order to ameliorate their efficiency and decreases off target effects (147). However, the available engineered modifications for miRNA-mimic are very limited because demonstrated to interfere with the loading into RISC complex, and therefore leading to a loss of effective silencing function (5). Given the relatively recent discovery and the diversity of currently known lncRNA mechanisms (some of which still yet not characterized), the available tools for their modulation are quite limited and also have to be carefully considered in the context of how the lncRNA may function (148). For what concerning the down-regulation of specific lncRNAs the RNA interference is still widely used (149). Moreover, this technology may also be applied for the lncRNA enriched in the nuclear compartment, based on the recent discovery of key components of RNAi to be present and active in cell nuclei (34,150). An alternative to RNAi for the degradation of lncRNAs is represented by gapmer molecules, which are 15–20 nt single-stranded DNA oligomers that hybridize with target lncRNA transcripts through complementarity and induce RNaseH-mediated degradation of the target transcripts (151). However, both RNAi and gapmer are reduced in efficiency because of the formation of secondary structures typical of lncRNAs. To overcome this limitation, recently the Cas13 family of CRISPR ribonucleases technology was clearly demonstrated to efficiently lower the level of specific lncRNA molecules (152). In contrast, the up regulation of lncRNAs is more complicated and to date the available system are quite limited. A potential strategy for up regulation of lncRNAs relies on the use of recombinant adeno or lentivirus. In particular cardio-trophic adeno-associated (AAV) vector were already demonstrated to be efficient for the delivery of both protein coding genes and miRNAs, therefore may also represent a promising novel approach also for lncRNAs although so far remains to be determined (13,153-155).

Delivery approaches to improve efficiency and selectivity of ncRNA therapeutics

Despite the well-accepted possibility to be promising tools for the treatment of cardiovascular disorders, so far the number of available ncRNA-based drugs is very limited. This is due to the deficiency of reliable delivery tools aimed at increasing the drug local concentration in heart muscle reducing the side effect at systemic level (156). So far the available delivery approaches used for ncRNAs local release in cardiovascular system are: (I) assisted-device delivery system, (II) viral vectors, and (III) non-viral vectors.

For what concerning the assisted-device delivery system, Hinkel et al. provided in 2012 the first evidence that local catheter-based delivery more efficiently inhibits miR-92a expression in the heart compared with intravenous infusion (157). However, this strategy didn’t completely prevent the systemic inhibitory effect of the anti-miR since the expression of miR-92a resulted affected also in other organs such as kidney and liver (158). Another example is represented by the work of Wang et al., in which they demonstrated an efficient local silencing of miR-21, with reduced systemic side effects, using a drug eluting stent (DES) in a rat model of myointimal hyperplasia (MH) (159).

The use of viral vectors has significantly improved tissue-specific enrichment of ncRNAs. One of the first research describing the use of viral vector was represented by the work of Mauro Giacca’s lab in which AAV serotype 9 was used to selectively overexpress miR-199a and miR-590 into cardiomyocytes (160). Moreover, further studies from the same research team and concerning the use of AAV vectors for specific miRNA delivery recently resulted with the publication of a proof of concept for the use of AAV coding miRNA in large animals. Indeed in this work of Gabisonia et al., the serotype 6 (AAV6) was demonstrated to be a valid vector for transducing pig cardiomyocytes after intramyocardial injection (161). Together with a re-increased interest in the field of gene therapy, as represented also by the recent in human successes from the leading company Audentes (https://www.audentestx.com/), the clinical use of viral vector for a target delivery of ncRNAs is promising. Nevertheless the viral approach is still not free from negative aspects such as related-carcinogenesis (162), immunogenicity (163), broad tropism (164), and chronic transgene expression (165,166). To cope with such limitations, the evaluation and development of alternative vectors based on non-viral systems is on going (167). An example is provided by the latest advances in material sciences, where a plethora of biosafe and biodegradable non-viral vectors with different size, shape, surface charge and drug cargo, has been recently explored in order to achieve an improved beneficial action of delivery but with reduced side effects (168). In the work of Bellera et al., the authors demonstrated in adult pigs that a single intracoronary administration of encapsulated antagomir‐92a into microspheres prevents left ventricular remodelling with no local or distant adverse effects (169). At the clinical level, a significant effort has been done by the Nano-Athero consortium, which is currently developing efficient nanocarriers for both the imaging and the treatment of thrombus and plaque (http://www.nanoathero.eu/). A similar activity is pursued by our group, who demonstrated the potential application of nanosystems for the specific targeting and release of therapeutic compounds to the heart. In particular, we demonstrated in both mice and pigs how the inhalation of biocompatible and biodegradable calcium phosphate nanoparticles (CaPs) can represent an important tool for the release of therapeutic drugs directly into the myocardium (170,171).

Based on the concept that no one-size-fits-all solution to gene delivery, new nanoformulation are constantly being proposed and investigated, therefore opening new horizon in the clinical approach for the treatment of cardiovascular disorders (172).


Conclusions

Although ncRNAs can represent promising tools for the prevention or treatment of CVDs, the general number of approved (e.g., Mipomersen, for treatment of homozygous familial hypercholesterolemia) or still in clinical trials (e.g., Inclisiran in Phase II; IONIS ANGPTL3-LRxin Phase II) (173) therapeutic RNA-based compounds remain limited. The effort from joint ventures between scientific community and pharmaceutical industries, such as Cardior (https://www.cardior.de/) or within the H2020 Cupido project (http://www.cupidoproject.eu/), are strongly required for the establishment of new strategies to overcome obstacles related to elevated toxicity, poor cellular uptake and tissue specificity that hinder the use of RNA molecules in the modern medicine.


Acknowledgments

Funding: This work was supported in part by the H2020-NMBP-2016720834 CUPIDO (www.cupidoproject.eu) and The Italian Research Nanomaxflagship Project (MIUR funding) (PNR-CNR 2011-2013).


Footnote

Conflicts of Interest: Both authors have completed the ICMJE uniform disclosure form (available at http://dx.doi.org/10.21037/ncri.2019.05.03). The authors have no conflicts of interest to declare.

Ethical Statement: The authors are accountable for all aspects of the work in ensuring that questions related to the accuracy or integrity of any part of the work are appropriately investigated and resolved.

Open Access Statement: This is an Open Access article distributed in accordance with the Creative Commons Attribution-NonCommercial-NoDerivs 4.0 International License (CC BY-NC-ND 4.0), which permits the non-commercial replication and distribution of the article with the strict proviso that no changes or edits are made and the original work is properly cited (including links to both the formal publication through the relevant DOI and the license). See: https://creativecommons.org/licenses/by-nc-nd/4.0/.


References

  1. Gandhi S, Ruehle F, Stoll M. Evolutionary Patterns of Non-Coding RNA in Cardiovascular Biology. Noncoding RNA 2019;5:15. [Crossref] [PubMed]
  2. Das A, Samidurai A, Salloum FN. Deciphering non-coding RNAs in cardiovascular health and disease. Front Cardiovasc Med 2018;5:73. [Crossref] [PubMed]
  3. O’Connor CM, Psotka MA, Fiuzat M, et al. Improving heart failure therapeutics development in the United States: the Heart Failure Collaboratory. J Am Coll Cardiol 2018;71:443-53. [Crossref] [PubMed]
  4. De Majo F, De Windt LJ. RNA therapeutics for heart disease. Biochem Pharmacol 2018;155:468-78. [Crossref] [PubMed]
  5. Di Mauro V, Barandalla-Sobrados M, Catalucci D. The noncoding-RNA landscape in cardiovascular health and disease. Noncoding RNA Res 2018;3:12-9. [Crossref] [PubMed]
  6. Raso A, Dirkx E. Cardiac regenerative medicine: at the crossroad of microRNA function and biotechnology. Noncoding RNA Res 2017;2:27-37. [Crossref] [PubMed]
  7. Ho YT, Poinard B, Kah JCY. Nanoparticle drug delivery systems and their use in cardiac tissue therapy. Nanomedicine 2016;11:693-714. [Crossref] [PubMed]
  8. Sallam T, Sandhu J, Tontonoz P. Long noncoding RNA discovery in cardiovascular disease: decoding form to function. Circ Res 2018;122:155-66. [Crossref] [PubMed]
  9. Consortium EP. An integrated encyclopedia of DNA elements in the human genome. Nature 2012;489:57. [Crossref] [PubMed]
  10. Goodwin S, McPherson JD, McCombie WR. Coming of age: ten years of next-generation sequencing technologies. Nat Rev Genet 2016;17:333. [Crossref] [PubMed]
  11. Archer K, Broskova Z, Bayoumi A, et al. Long non-coding RNAs as master regulators in cardiovascular diseases. Int J Mol Sci 2015;16:23651-67. [Crossref] [PubMed]
  12. Ono K, Kuwabara Y, Horie T, et al. Long non-coding RNAs as key regulators of cardiovascular diseases. Circ J 2018;82:1231-6. [Crossref] [PubMed]
  13. Gomes CP, Spencer H, Ford KL, et al. The function and therapeutic potential of long non-coding RNAs in cardiovascular development and disease. Mol Ther Nucleic Acids 2017;8:494-507. [Crossref] [PubMed]
  14. Uchida S, Bolli R. Short and long noncoding RNAs regulate the epigenetic status of cells. Antioxid Redox Signal 2018;29:832-45. [Crossref] [PubMed]
  15. Poller W, Tank J, Skurk C, et al. Cardiovascular RNA interference therapy: the broadening tool and target spectrum. Circ Res 2013;113:588-602. [Crossref] [PubMed]
  16. Roy S, Trautwein C, Luedde T, et al. A General Overview on Non-coding RNA-Based Diagnostic and Therapeutic Approaches for Liver Diseases. Front Pharmacol 2018;9:805. [Crossref] [PubMed]
  17. Fu XD. Non-coding RNA: a new frontier in regulatory biology. Natl Sci Rev 2014;1:190-204. [Crossref] [PubMed]
  18. Lin CP, He L. Noncoding RNAs in cancer development. Annu Rev Cancer Biol 2017;1:163-84. [Crossref]
  19. Suter-Dick L. ’Omics in Organ Toxicity, Integrative Analysis Approaches, and Knowledge Generation. Toxicogenomics-Based Cellular Models. Elsevier, 2014:235-50.
  20. Han Li C, Chen Y. Small and long non-coding RNAs: novel targets in perspective cancer therapy. Curr Genomics 2015;16:319-26. [Crossref] [PubMed]
  21. Stȩpień E, Costa MC, Kurc S, et al. The circulating non-coding RNA landscape for biomarker research: lessons and prospects from cardiovascular diseases. Acta Pharmacol Sin 2018;39:1085-99. [Crossref] [PubMed]
  22. O'Brien J, Hayder H, Zayed Y, et al. Overview of microrna biogenesis, mechanisms of actions, and circulation. Front Endocrinol (Lausanne) 2018;9:402. [Crossref] [PubMed]
  23. De Rie D, Abugessaisa I, Alam T, et al. An integrated expression atlas of miRNAs and their promoters in human and mouse. Nat Biotechnol 2017;35:872. [Crossref] [PubMed]
  24. Kim YK, Kim VN. Processing of intronic microRNAs. EMBO J 2007;26:775-83. [Crossref] [PubMed]
  25. Nana-Sinkam S, Croce C. Clinical applications for microRNAs in cancer. Clin Pharmacol Ther 2013;93:98-104. [PubMed]
  26. Ha M, Kim VN. Regulation of microRNA biogenesis. Nat Rev Mol Cell Biol 2014;15:509. [Crossref] [PubMed]
  27. Forman JJ, Legesse-Miller A, Coller HA. A search for conserved sequences in coding regions reveals that the let-7 microRNA targets Dicer within its coding sequence. Proc Natl Acad Sci U S A 2008;105:14879-84. [Crossref] [PubMed]
  28. Zhang J, Zhou W, Liu Y, et al. Oncogenic role of microRNA-532-5p in human colorectal cancer via targeting of the 5'UTR of RUNX3. Oncol Lett 2018;15:7215-20. [PubMed]
  29. Barman B, Bhattacharyya SN. mRNA targeting to endoplasmic reticulum precedes ago protein interaction and microRNA (miRNA)-mediated translation repression in mammalian cells. J Biol Chem 2015;290:24650-6. [Crossref] [PubMed]
  30. Nishi K, Takahashi T, Suzawa M, et al. Control of the localization and function of a miRNA silencing component TNRC6A by Argonaute protein. Nucleic Acids Res 2015;43:9856-73. [PubMed]
  31. Barrey E, Saint-Auret G, Bonnamy B, et al. Pre-microRNA and mature microRNA in human mitochondria. PLoS One 2011;6:e20220 [Crossref] [PubMed]
  32. Catalanotto C, Cogoni C, Zardo G. MicroRNA in control of gene expression: an overview of nuclear functions. Int J Mol Sci 2016;17:1712. [Crossref] [PubMed]
  33. Hwang HW, Wentzel EA, Mendell JT. A hexanucleotide element directs microRNA nuclear import. Science 2007;315:97-100. [Crossref] [PubMed]
  34. Gagnon KT, Li L, Chu Y, et al. RNAi factors are present and active in human cell nuclei. Cell Rep 2014;6:211-21. [Crossref] [PubMed]
  35. Roberts TC. The microRNA biology of the mammalian nucleus. Mol Ther Nucleic Acids 2014;3:e188 [Crossref] [PubMed]
  36. Iguchi H, Kosaka N, Ochiya T. Secretory microRNAs as a versatile communication tool. Commun Integr Biol 2010;3:478-81. [Crossref] [PubMed]
  37. Moyano M, Stefani G. piRNA involvement in genome stability and human cancer. J Hematol Oncol 2015;8:38. [Crossref] [PubMed]
  38. Weick EM, Miska EA. piRNAs: from biogenesis to function. Development 2014;141:3458-71. [Crossref] [PubMed]
  39. Saito K, Nishida KM, Mori T, et al. Specific association of Piwi with rasiRNAs derived from retrotransposon and heterochromatic regions in the Drosophila genome. Genes Dev 2006;20:2214-22. [Crossref] [PubMed]
  40. Carmell MA, Xuan Z, Zhang MQ, et al. The Argonaute family: tentacles that reach into RNAi, developmental control, stem cell maintenance, and tumorigenesis. Genes Dev 2002;16:2733-42. [Crossref] [PubMed]
  41. Yin H, Lin H. An epigenetic activation role of Piwi and a Piwi-associated piRNA in Drosophila melanogaster. Nature 2007;450:304. [Crossref] [PubMed]
  42. Huang XA, Yin H, Sweeney S, et al. A major epigenetic programming mechanism guided by piRNAs. Dev Cell 2013;24:502-16. [Crossref] [PubMed]
  43. Rajasethupathy P, Antonov I, Sheridan R, et al. A role for neuronal piRNAs in the epigenetic control of memory-related synaptic plasticity. Cell 2012;149:693-707. [Crossref] [PubMed]
  44. Rouget C, Papin C, Boureux A, et al. Maternal mRNA deadenylation and decay by the piRNA pathway in the early Drosophila embryo. Nature 2010;467:1128. [Crossref] [PubMed]
  45. Gou LT, Dai P, Yang JH, et al. Pachytene piRNAs instruct massive mRNA elimination during late spermiogenesis. Cell Res 2014;24:680. [Crossref] [PubMed]
  46. Sharma AK, Nelson MC, Brandt JE, et al. Human CD34+ stem cells express the hiwigene, a human homologue of the Drosophila genepiwi. Blood 2001;97:426-34. [Crossref] [PubMed]
  47. Rajan KS, Velmurugan G, Gopal P, et al. Abundant and altered expression of PIWI-Interacting RNAs during cardiac hypertrophy. Heart Lung Circ 2016;25:1013-20. [Crossref] [PubMed]
  48. Rajan KS, Velmurugan G, Pandi G, et al. miRNA and piRNA mediated Akt pathway in heart: antisense expands to survive. Int J Biochem Cell Biol 2014;55:153-6. [Crossref] [PubMed]
  49. Yang J, Xue F, Li Y, et al. Exosomal piRNA sequencing reveals differences between heart failure and healthy patients. Eur Rev Med Pharmacol Sci 2018;22:7952-61. [PubMed]
  50. Fire A, Xu S, Montgomery MK, et al. Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature 1998;391:806. [Crossref] [PubMed]
  51. Ahmadzada T, Reid G, McKenzie DR. Fundamentals of siRNA and miRNA therapeutics and a review of targeted nanoparticle delivery systems in breast cancer. Biophys Rev 2018;10:69-86. [Crossref] [PubMed]
  52. Carthew RW, Sontheimer EJ. Origins and mechanisms of miRNAs and siRNAs. Cell 2009;136:642-55. [Crossref] [PubMed]
  53. Lam JK, Chow MY, Zhang Y, et al. siRNA versus miRNA as therapeutics for gene silencing. Mol Ther Nucleic Acids 2015;4:e252 [Crossref] [PubMed]
  54. Chung WJ, Okamura K, Martin R, et al. Endogenous RNA interference provides a somatic defense against Drosophila transposons. Curr Biol 2008;18:795-802. [Crossref] [PubMed]
  55. Tam OH, Aravin AA, Stein P, et al. Pseudogene-derived small interfering RNAs regulate gene expression in mouse oocytes. Nature 2008;453:534. [Crossref] [PubMed]
  56. Hamilton AJ, Baulcombe DC. A species of small antisense RNA in posttranscriptional gene silencing in plants. Science 1999;286:950-2. [Crossref] [PubMed]
  57. Di Giulio M. The origin of the tRNA molecule: implications for the origin of protein synthesis. J Theor Biol 2004;226:89-93. [Crossref] [PubMed]
  58. Kirchner S, Ignatova Z. Emerging roles of tRNA in adaptive translation, signalling dynamics and disease. Nat Rev Genet 2015;16:98. [Crossref] [PubMed]
  59. Schimmel P. The emerging complexity of the tRNA world: mammalian tRNAs beyond protein synthesis. Nat Rev Mol Cell Biol 2018;19:45. [Crossref] [PubMed]
  60. Kiss T. Small nucleolar RNAs: an abundant group of noncoding RNAs with diverse cellular functions. Cell 2002;109:145-8. [Crossref] [PubMed]
  61. Stepanov GA, Filippova JA, Komissarov AB, et al. Regulatory role of small nucleolar RNAs in human diseases. Biomed Res Int 2015;2015:206849 [Crossref] [PubMed]
  62. Bhat SA, Ahmad SM, Mumtaz PT, et al. Long non-coding RNAs: Mechanism of action and functional utility. Noncoding RNA Res 2016;1:43-50. [Crossref] [PubMed]
  63. Guttman M, Amit I, Garber M, et al. Chromatin signature reveals over a thousand highly conserved large non-coding RNAs in mammals. Nature 2009;458:223. [Crossref] [PubMed]
  64. Morceau F, Chateauvieux S, Gaigneaux A, et al. Long and short non-coding RNAs as regulators of hematopoietic differentiation. Int J Mol Sci 2013;14:14744-70. [Crossref] [PubMed]
  65. Reeves MB, Davies AA, McSharry BP, et al. Complex I binding by a virally encoded RNA regulates mitochondria-induced cell death. Science 2007;316:1345-8. [Crossref] [PubMed]
  66. Houseley J, Rubbi L, Grunstein M, et al. A ncRNA modulates histone modification and mRNA induction in the yeast GAL gene cluster. Mol Cell 2008;32:685-95. [Crossref] [PubMed]
  67. Swiezewski S, Liu F, Magusin A, et al. Cold-induced silencing by long antisense transcripts of an Arabidopsis Polycomb target. Nature 2009;462:799. [Crossref] [PubMed]
  68. Brown CJ, Hendrich BD, Rupert JL, et al. The human XIST gene: analysis of a 17 kb inactive X-specific RNA that contains conserved repeats and is highly localized within the nucleus. Cell 1992;71:527-42. [Crossref] [PubMed]
  69. He X, Ou C, Xiao Y, et al. LncRNAs: key players and novel insights into diabetes mellitus. Oncotarget 2017;8:71325. [PubMed]
  70. Consortium F. I RGERGP, Team I. Analysis of the mouse transcriptome based on functional annotation of 60,770 full-length cDNAs. Nature 2002;420:563. [Crossref] [PubMed]
  71. Derrien T, Johnson R, Bussotti G, et al. The GENCODE v7 catalog of human long noncoding RNAs: analysis of their gene structure, evolution, and expression. Genome Res 2012;22:1775-89. [Crossref] [PubMed]
  72. Harrow J, Frankish A, Gonzalez JM, et al. GENCODE: the reference human genome annotation for The ENCODE Project. Genome Res 2012;22:1760-74. [Crossref] [PubMed]
  73. Gutschner T, Diederichs S. The hallmarks of cancer: a long non-coding RNA point of view. RNA Biol 2012;9:703-19. [Crossref] [PubMed]
  74. Ma L, Bajic VB, Zhang Z. On the classification of long non-coding RNAs. RNA Biol 2013;10:924-33. [Crossref] [PubMed]
  75. Moran VA, Perera RJ, Khalil AM. Emerging functional and mechanistic paradigms of mammalian long non-coding RNAs. Nucleic Acids Res 2012;40:6391-400. [Crossref] [PubMed]
  76. Lennox KA, Behlke MA. Mini-review: Current strategies to knockdown long non-coding RNAs. J Rare Dis Res Treat 2016;1:66-70. [Crossref]
  77. Chen LL. Linking long noncoding RNA localization and function. Trends Biochem Sci 2016;41:761-72. [Crossref] [PubMed]
  78. Dong L, Lin W, Qi P, et al. Circulating long RNAs in serum extracellular vesicles: their characterization and potential application as biomarkers for diagnosis of colorectal cancer. Cancer Epidemiol Biomarkers Prev 2016;25:1158-66. [Crossref] [PubMed]
  79. Sun Z, Yang S, Zhou Q, et al. Emerging role of exosome-derived long non-coding RNAs in tumor microenvironment. Mol Cancer 2018;17:82. [Crossref] [PubMed]
  80. Wang KC, Chang HY. Molecular mechanisms of long noncoding RNAs. Mol Cell 2011;43:904-14. [Crossref] [PubMed]
  81. Boon RA, Jaé N, Holdt L, et al. Long noncoding RNAs: from clinical genetics to therapeutic targets? J Am Coll Cardiol 2016;67:1214-26. [Crossref] [PubMed]
  82. Kim TK, Hemberg M, Gray JM, et al. Widespread transcription at neuronal activity-regulated enhancers. Nature 2010;465:182. [Crossref] [PubMed]
  83. De Santa F, Barozzi I, Mietton F, et al. A large fraction of extragenic RNA pol II transcription sites overlap enhancers. PLoS Biol 2010;8:e1000384 [Crossref] [PubMed]
  84. Mao R, Wu Y, Ming Y, et al. Enhancer RNAs: a missing regulatory layer in gene transcription. Science China Life Sciences 2018; [Epub ahead of print]. [Crossref] [PubMed]
  85. Plank JL, Dean A. Enhancer function: mechanistic and genome-wide insights come together. Mol Cell 2014;55:5-14. [Crossref] [PubMed]
  86. Natoli G, Andrau JC. Noncoding transcription at enhancers: general principles and functional models. Annu Rev Genet 2012;46:1-19. [Crossref] [PubMed]
  87. Ding M, Liu Y, Liao X, et al. Enhancer RNAs (eRNAs): New Insights into Gene Transcription and Disease Treatment. J Cancer 2018;9:2334. [Crossref] [PubMed]
  88. Andersson R, Gebhard C, Miguel-Escalada I, et al. An atlas of active enhancers across human cell types and tissues. Nature 2014;507:455. [Crossref] [PubMed]
  89. Andersson R, Andersen PR, Valen E, et al. Nuclear stability and transcriptional directionality separate functionally distinct RNA species. Nat Commun 2014;5:5336. [Crossref] [PubMed]
  90. Djebali S, Davis CA, Merkel A, et al. Landscape of transcription in human cells. Nature 2012;489:101. [Crossref] [PubMed]
  91. Liu Y, Ding M, Gao Q, et al. Current Advances on the Important Roles of Enhancer RNAs in Gene Regulation and Cancer. Biomed Res Int 2018;2018:2405351 [PubMed]
  92. Shii L, Song L, Maurer K, et al. SERPINB2 is regulated by dynamic interactions with pause-release proteins and enhancer RNAs. Mol Immunol 2017;88:20-31. [Crossref] [PubMed]
  93. Schaukowitch K, Joo JY, Liu X, et al. Enhancer RNA facilitates NELF release from immediate early genes. Mol Cell 2014;56:29-42. [Crossref] [PubMed]
  94. Mousavi K, Zare H, Dell’Orso S, et al. eRNAs promote transcription by establishing chromatin accessibility at defined genomic loci. Mol Cell 2013;51:606-17. [Crossref] [PubMed]
  95. Hatzis P, Talianidis I. Dynamics of enhancer-promoter communication during differentiation-induced gene activation. Mol Cell 2002;10:1467-77. [Crossref] [PubMed]
  96. Wang Q, Carroll JS, Brown M. Spatial and temporal recruitment of androgen receptor and its coactivators involves chromosomal looping and polymerase tracking. Mol Cell 2005;19:631-42. [Crossref] [PubMed]
  97. Quan G, Li J. Circular RNAs: biogenesis, expression and their potential roles in reproduction. J Ovarian Res 2018;11:9. [Crossref] [PubMed]
  98. Westholm JO, Miura P, Olson S, et al. Genome-wide analysis of drosophila circular RNAs reveals their structural and sequence properties and age-dependent neural accumulation. Cell Rep 2014;9:1966-80. [Crossref] [PubMed]
  99. Memczak S, Jens M, Elefsinioti A, et al. Circular RNAs are a large class of animal RNAs with regulatory potency. Nature 2013;495:333. [Crossref] [PubMed]
  100. Gao Y, Wang J, Zhao F. CIRI: an efficient and unbiased algorithm for de novo circular RNA identification. Genome Biol 2015;16:4. [Crossref] [PubMed]
  101. Zhang XO, Wang HB, Zhang Y, et al. Complementary sequence-mediated exon circularization. Cell 2014;159:134-47. [Crossref] [PubMed]
  102. Salzman J, Gawad C, Wang PL, et al. Circular RNAs are the predominant transcript isoform from hundreds of human genes in diverse cell types. PLoS One 2012;7:e30733 [Crossref] [PubMed]
  103. Ebbesen KK, Hansen TB, Kjems J. Insights into circular RNA biology. RNA Biol 2017;14:1035-45. [Crossref] [PubMed]
  104. Hansen TB, Jensen TI, Clausen BH, et al. Natural RNA circles function as efficient microRNA sponges. Nature 2013;495:384. [Crossref] [PubMed]
  105. Li HM, Ma XL, Li HG. Intriguing circles: Conflicts and controversies in circular RNA research. Wiley Interdiscip Rev RNA 2019; [Epub ahead of print]. [Crossref] [PubMed]
  106. Zhang Y, Zhang XO, Chen T, et al. Circular intronic long noncoding RNAs. Mol Cell 2013;51:792-806. [Crossref] [PubMed]
  107. Wang Y, Wang Z. Efficient backsplicing produces translatable circular mRNAs. Rna 2015;21:172-9. [Crossref] [PubMed]
  108. Chen CY, Sarnow P. Initiation of protein synthesis by the eukaryotic translational apparatus on circular RNAs. Science 1995;268:415-7. [Crossref] [PubMed]
  109. Liu J, Liu T, Wang X, et al. Circles reshaping the RNA world: from waste to treasure. Mol Cancer 2017;16:58. [Crossref] [PubMed]
  110. Khan MA, Reckman YJ, Aufiero S, et al. RBM20 regulates circular RNA production from the titin gene. Circ Res 2016;119:996-1003. [Crossref] [PubMed]
  111. Holdt LM, Kohlmaier A, Teupser D. Molecular functions and specific roles of circRNAs in the cardiovascular system. Noncoding RNA Res 2018;3:75-98. [Crossref] [PubMed]
  112. Pokhrel S, Guotian Y. MicroRNA and Its Role in Cardiovascular Disease. World J Cardiovasc Dis 2017;7:340. [Crossref]
  113. Wang J, Liew O, Richards A, et al. Overview of microRNAs in cardiac hypertrophy, fibrosis, and apoptosis. Int J Mol Sci 2016;17:749. [Crossref] [PubMed]
  114. Zhang Y, Luo G, Zhang Y, et al. Critical effects of long non-coding RNA on fibrosis diseases. Exp Mol Med 2018;50:e428 [Crossref] [PubMed]
  115. Van Rooij E, Sutherland LB, Thatcher JE, et al. Dysregulation of microRNAs after myocardial infarction reveals a role of miR-29 in cardiac fibrosis. Proc Natl Acad Sci U S A 2008;105:13027-32. [Crossref] [PubMed]
  116. Thum T, Gross C, Fiedler J, et al. MicroRNA-21 contributes to myocardial disease by stimulating MAP kinase signalling in fibroblasts. Nature 2008;456:980. [Crossref] [PubMed]
  117. Wei C, Kim IK, Kumar S, et al. NF-κB mediated miR-26a regulation in cardiac fibrosis. J Cell Physiol 2013;228:1433-42. [Crossref] [PubMed]
  118. Wang X, Wang HX, Li YL, et al. MicroRNA Let-7i negatively regulates cardiac inflammation and fibrosis. Hypertension 2015;66:776-85. [Crossref] [PubMed]
  119. Micheletti R, Plaisance I, Abraham BJ, et al. The long noncoding RNA Wisper controls cardiac fibrosis and remodeling. Sci Transl Med 2017;9:eaai9118.
  120. Piccoli MT, Gupta SK, Viereck J, et al. Inhibition of the cardiac fibroblast–enriched lncRNA Meg3 prevents cardiac fibrosis and diastolic dysfunction. Circ Res 2017;121:575-83. [Crossref] [PubMed]
  121. Samak M, Fatullayev J, Sabashnikov A, et al. Cardiac hypertrophy: an introduction to molecular and cellular basis. Med Sci Monit Basic Res 2016;22:75. [Crossref] [PubMed]
  122. Wang J, Yang X. The function of miRNA in cardiac hypertrophy. Cell Mol Life Sci 2012;69:3561-70. [Crossref] [PubMed]
  123. Sayed D, Hong C, Chen IY, et al. MicroRNAs play an essential role in the development of cardiac hypertrophy. Circ Res 2007;100:416-24. [Crossref] [PubMed]
  124. Care A, Catalucci D, Felicetti F, et al. MicroRNA-133 controls cardiac hypertrophy. Nat Med 2007;13:613-8. [Crossref] [PubMed]
  125. Song DW, Ryu JY, Kim JO, et al. The miR-19a/b family positively regulates cardiomyocyte hypertrophy by targeting atrogin-1 and MuRF-1. Biochem J 2014;457:151-62. [Crossref] [PubMed]
  126. Hermans-Beijnsberger S, van Bilsen M, Schroen B. Long non-coding RNAs in the failing heart and vasculature. Noncoding RNA Res 2018;3:118-30. [Crossref] [PubMed]
  127. Wang Z, Zhang XJ, Ji YX, et al. The long noncoding RNA Chaer defines an epigenetic checkpoint in cardiac hypertrophy. Nat Med 2016;22:1131. [Crossref] [PubMed]
  128. Shao M, Chen G, Lv F, et al. LncRNA TINCR attenuates cardiac hypertrophy by epigenetically silencing CaMKII. Oncotarget 2017;8:47565. [Crossref] [PubMed]
  129. Boon RA, Dimmeler S. MicroRNAs in myocardial infarction. Nat Rev Cardiol 2015;12:135. [Crossref] [PubMed]
  130. Hullinger TG, Montgomery RL, Seto AG, et al. Inhibition of miR-15 protects against cardiac ischemic injury. Circ Res 2012;110:71-81. [Crossref] [PubMed]
  131. Qian L, Van Laake LW, Huang Y, et al. miR-24 inhibits apoptosis and represses Bim in mouse cardiomyocytes. J Exp Med 2011;208:549-60. [Crossref] [PubMed]
  132. Wang K, Liu CY, Zhou LY, et al. APF lncRNA regulates autophagy and myocardial infarction by targeting miR-188-3p. Nat Commun 2015;6:6779. [Crossref] [PubMed]
  133. Yin G, Yang X, Li Q, et al. GATA1 activated lncRNA (Galont) promotes anoxia/reoxygenation-induced autophagy and cell death in cardiomyocytes by sponging miR-338. J Cell Biochem 2018;119:4161-9. [Crossref] [PubMed]
  134. Lieberman J. Tapping the RNA world for therapeutics. Nat Struct Mol Biol 2018;25:357-364. [Crossref] [PubMed]
  135. Verdine GL, Walensky LD. The challenge of drugging undruggable targets in cancer: lessons learned from targeting BCL-2 family members. Clin Cancer Res 2007;13:7264-70. [Crossref] [PubMed]
  136. Kaczmarek JC, Kowalski PS, Anderson DG. Advances in the delivery of RNA therapeutics: from concept to clinical reality. Genome Med 2017;9:60. [Crossref] [PubMed]
  137. Matsui M, Corey DR. Non-coding RNAs as drug targets. Nat Rev Drug Discov 2017;16:167-79. [Crossref] [PubMed]
  138. Kristen AV, Ajroud-Driss S, Conceição I, et al. Patisiran, an RNAi therapeutic for the treatment of hereditary transthyretin-mediated amyloidosis. Neurodegener Dis Manag 2019;9:5-23. [Crossref] [PubMed]
  139. Van Rooij E, Kauppinen S. Development of microRNA therapeutics is coming of age. EMBO Mol Med 2014;6:851-64. [Crossref] [PubMed]
  140. Ling H, Fabbri M, Calin GA. MicroRNAs and other non-coding RNAs as targets for anticancer drug development. Nat Rev Drug Discov 2013;12:847. [Crossref] [PubMed]
  141. Ebert MS, Neilson JR, Sharp PA. MicroRNA sponges: competitive inhibitors of small RNAs in mammalian cells. Nat Methods 2007;4:721-6. [Crossref] [PubMed]
  142. Bernardo BC, Charchar FJ, Lin RC, et al. A microRNA guide for clinicians and basic scientists: background and experimental techniques. Heart Lung Circ 2012;21:131-42. [Crossref] [PubMed]
  143. Hagedorn PH, Persson R, Funder ED, et al. Locked nucleic acid: modality, diversity, and drug discovery. Drug Discov Today 2018;23:101-14. [Crossref] [PubMed]
  144. Tili E, Michaille JJ, Gandhi V, et al. miRNAs and their potential for use against cancer and other diseases. Future Oncol 2007;3:521-37. [Crossref] [PubMed]
  145. Stenvang J, Petri A, Lindow M, et al. Inhibition of microRNA function by antimiR oligonucleotides. Silence 2012;3:1. [Crossref] [PubMed]
  146. Wang Z. The guideline of the design and validation of MiRNA mimics. MicroRNA and Cancer. Springer, 2011:211-23.
  147. Hosseinahli N, Aghapour M, Duijf PH, et al. Treating cancer with microRNA replacement therapy: A literature review. J Cell Physiol 2018;233:5574-88. [Crossref] [PubMed]
  148. Liu SJ, Lim DA. Modulating the expression of long non-coding RNAs for functional studies. EMBO Rep 2018;19:e46955 [Crossref] [PubMed]
  149. Grote P, Wittler L, Hendrix D, et al. The tissue-specific lncRNA Fendrr is an essential regulator of heart and body wall development in the mouse. Dev Cell 2013;24:206-14. [Crossref] [PubMed]
  150. Ren S, Liu Y, Xu W, et al. Long noncoding RNA MALAT-1 is a new potential therapeutic target for castration resistant prostate cancer. J Urol 2013;190:2278-87. [Crossref] [PubMed]
  151. Wu H, Lima WF, Zhang H, et al. Determination of the role of the human RNase H1 in the pharmacology of DNA-like antisense drugs. J Biol Chem 2004;279:17181-9. [Crossref] [PubMed]
  152. Shmakov S, Abudayyeh OO, Makarova KS, et al. Discovery and functional characterization of diverse class 2 CRISPR-Cas systems. Mol Cell 2015;60:385-97. [Crossref] [PubMed]
  153. Gray SJ, Samulski RJ. Optimizing gene delivery vectors for the treatment of heart disease. Expert Opin Biol Ther 2008;8:911-22. [Crossref] [PubMed]
  154. Zsebo K, Yaroshinsky A, Rudy JJ, et al. Long-term effects of AAV1/SERCA2a gene transfer in patients with severe heart failure: analysis of recurrent cardiovascular events and mortality. Circ Res 2014;114:101-8. [Crossref] [PubMed]
  155. Kota J, Chivukula RR, O'Donnell KA, et al. Therapeutic microRNA delivery suppresses tumorigenesis in a murine liver cancer model. Cell 2009;137:1005-17. [Crossref] [PubMed]
  156. Chen C, Yang Z, Tang X. Chemical modifications of nucleic acid drugs and their delivery systems for gene-based therapy. Med Res Rev 2018;38:829-69. [Crossref] [PubMed]
  157. Hinkel R, Penzkofer D, Zühlke S, et al. Inhibition of microRNA-92a protects against ischemia/reperfusion injury in a large-animal model. Circulation 2013;128:1066-75. [Crossref] [PubMed]
  158. Lucas T, Bonauer A, Dimmeler S. RNA therapeutics in cardiovascular disease. Circ Res 2018;123:205-20. [Crossref] [PubMed]
  159. Wang D, Deuse T, Stubbendorff M, et al. Local microRNA modulation using a novel anti-miR-21-eluting stent effectively prevents experimental in-stent restenosis. Arterioscler Thromb Vasc Biol 2015;35:1945-53. [Crossref] [PubMed]
  160. Eulalio A, Mano M, Dal Ferro M, et al. Functional screening identifies miRNAs inducing cardiac regeneration. Nature 2012;492:376. [Crossref] [PubMed]
  161. Gabisonia K, Prosdocimo G, Aquaro GD, et al. MicroRNA therapy stimulates uncontrolled cardiac repair after myocardial infarction in pigs. Nature 2019;569:418-22. [Crossref] [PubMed]
  162. Baum C, Kustikova O, Modlich U, et al. Mutagenesis and oncogenesis by chromosomal insertion of gene transfer vectors. Hum Gene Ther 2006;17:253-63. [Crossref] [PubMed]
  163. Bessis N. GarciaCozar F, Boissier M. Immune responses to gene therapy vectors: influence on vector function and effector mechanisms. Gene Ther 2004;11:S10. [Crossref] [PubMed]
  164. Waehler R, Russell SJ, Curiel DT. Engineering targeted viral vectors for gene therapy. Nat Rev Genet 2007;8:573. [Crossref] [PubMed]
  165. Thomas CE, Ehrhardt A, Kay MA. Progress and problems with the use of viral vectors for gene therapy. Nat Rev Genet 2003;4:346. [Crossref] [PubMed]
  166. Yin H, Kanasty RL, Eltoukhy AA, et al. Non-viral vectors for gene-based therapy. Nat Rev Genet 2014;15:541. [Crossref] [PubMed]
  167. Ramamoorth M, Narvekar A. Non viral vectors in gene therapy-an overview. J Clin Diagn Res 2015;9:GE01-6. [PubMed]
  168. Ventola CL. Progress in nanomedicine: approved and investigational nanodrugs. P T 2017;42:742. [PubMed]
  169. Bellera N, Barba I, Rodriguez-Sinovas A, et al. Single intracoronary injection of encapsulated antagomir-92a promotes angiogenesis and prevents adverse infarct remodeling. J Am Heart Assoc 2014;3:e000946 [Crossref] [PubMed]
  170. Di Mauro V, Iafisco M, Salvarani N, et al. Bioinspired negatively charged calcium phosphate nanocarriers for cardiac delivery of MicroRNAs. Nanomedicine 2016;11:891-906. [Crossref] [PubMed]
  171. Miragoli M, Ceriotti P, Iafisco M, et al. Inhalation of peptide-loaded nanoparticles improves heart failure. Sci Transl Med 2018;10:eaan6205.
  172. Havel H, Finch G, Strode P, et al. Nanomedicines: from bench to bedside and beyond. AAPS J 2016;18:1373-8. [Crossref] [PubMed]
  173. Laina A, Gatsiou A, Georgiopoulos G, et al. RNA therapeutics in cardiovascular precision medicine. Front Physiol 2018;9:953. [Crossref] [PubMed]
doi: 10.21037/ncri.2019.05.03
Cite this article as: Di Mauro V, Catalucci D. The role of small and long non-coding RNAs in cardiac pathologies. Non-coding RNA Investig 2019;3:21.

Download Citation