Next Article in Journal
Gender-Specific Effects of Two Treatment Strategies in a Mouse Model of Niemann-Pick Disease Type C1
Next Article in Special Issue
Angiopoietin-Like Protein 3 (ANGPTL3) Modulates Lipoprotein Metabolism and Dyslipidemia
Previous Article in Journal
Identification of Predictive Biomarkers of Response to HSP90 Inhibitors in Lung Adenocarcinoma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Foam Cells as Therapeutic Targets in Atherosclerosis with a Focus on the Regulatory Roles of Non-Coding RNAs

by
Amin Javadifar
1,
Sahar Rastgoo
1,
Maciej Banach
2,3,*,
Tannaz Jamialahmadi
4,5,
Thomas P. Johnston
6 and
Amirhossein Sahebkar
7,8,9,10,*
1
Department of Allergy and Immunology, Mashhad University of Medical Sciences, Mashhad 9177948564, Iran
2
Department of Hypertension, Chair of Nephrology and Hypertension, Medical University of Lodz, 93338 Lodz, Poland
3
Polish Mother’s Memorial Hospital Research Institute (PMMHRI), 93338 Lodz, Poland
4
Department of Food Science and Technology, Quchan Branch, Islamic Azad University, Quchan 9479176135, Iran
5
Department of Nutrition, Faculty of Medicine, Mashhad University of Medical Sciences, Mashhad 9177948564, Iran
6
Division of Pharmacology and Pharmaceutical Sciences, School of Pharmacy, University of Missouri-Kansas City, Kansas City, MO 64108-2718, USA
7
Biotechnology Research Center, Pharmaceutical Technology Institute, Mashhad University of Medical Sciences, Mashhad 9177948564, Iran
8
Applied Biomedical Research Center, Mashhad University of Medical Sciences, Mashhad 9177948564, Iran
9
School of Pharmacy, Mashhad University of Medical Sciences, Mashhad 9177948954, Iran
10
Department of Medical Biotechnology and Nanotechnology, School of Medicine, Mashhad University of Medical Sciences, Mashhad 9177948564, Iran
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(5), 2529; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22052529
Submission received: 14 January 2021 / Revised: 24 February 2021 / Accepted: 24 February 2021 / Published: 3 March 2021
(This article belongs to the Special Issue Molecular Mechanisms and Pathophysiology of Atherosclerosis)

Abstract

:
Atherosclerosis is a major cause of human cardiovascular disease, which is the leading cause of mortality around the world. Various physiological and pathological processes are involved, including chronic inflammation, dysregulation of lipid metabolism, development of an environment characterized by oxidative stress and improper immune responses. Accordingly, the expansion of novel targets for the treatment of atherosclerosis is necessary. In this study, we focus on the role of foam cells in the development of atherosclerosis. The specific therapeutic goals associated with each stage in the formation of foam cells and the development of atherosclerosis will be considered. Processing and metabolism of cholesterol in the macrophage is one of the main steps in foam cell formation. Cholesterol processing involves lipid uptake, cholesterol esterification and cholesterol efflux, which ultimately leads to cholesterol equilibrium in the macrophage. Recently, many preclinical studies have appeared concerning the role of non-encoding RNAs in the formation of atherosclerotic lesions. Non-encoding RNAs, especially microRNAs, are considered regulators of lipid metabolism by affecting the expression of genes involved in the uptake (e.g., CD36 and LOX1) esterification (ACAT1) and efflux (ABCA1, ABCG1) of cholesterol. They are also able to regulate inflammatory pathways, produce cytokines and mediate foam cell apoptosis. We have reviewed important preclinical evidence of their therapeutic targeting in atherosclerosis, with a special focus on foam cell formation.

1. Introduction

Atherosclerosis is a chronic, progressive immuno-inflammatory disease that affects blood vessels and can result in the development of atherosclerotic plaques. These plaques have the potential to rupture and lead to cardiovascular disease, especially myocardial infarction (MI) or stroke, which represent two of the leading causes of death worldwide [1]. Since foam cells are a central player in the underlying pathology of atherosclerosis, reducing the formation of foam cells, or inducing their removal, might represent an effective therapeutic strategy. Recently, several groups reported that targeting foam cells could effectively ameliorate atherosclerosis, which are described in this review. Current studies have suggested that non-coding RNAs are involved in the pathophysiology and progression of atherosclerosis and are good diagnostic and prognostic biomarkers, as well as therapeutic targets [1]. In this review, we discuss non-coding RNAs as therapeutic targets in the context of either enhancing or attenuating the formation of foam cells, which ultimately affects the metabolism and overall homeostasis of cholesterol.

2. Pathophysiology of Atherosclerosis and the Key Role of Foam Cells

The pathophysiology of atherosclerosis includes a complex network of different cellular processes and is associated with risk factors such as arterial hypertension, smoking, hyperglycemia and hypercholesterolemia [1]. One of the triggers of this disease is endothelial damage, which plays an important role in the formation of atherosclerotic plaques. Endothelial damage results in arteries experiencing a decrease in nitric oxide (NO) bioavailability and an increase in the production of reactive oxygen species (ROS) [2,3]. NO is an anti-atherogenic and vasoprotective factor essential for vascular health and is obtained by conversion of arginine to citrulline by endothelial nitric oxide synthase III (ENOS) [4]. Increased ROS also produce a state of oxidative stress that assists in the modification of LDL to its oxidized form (oxLDL). Platelets can also be activated by oxLDL and induce vascular inflammation [5,6]. Eventually, oxLDL, together with chronic low-grade inflammation resulting from endothelial injury, trigger an innate immune response and increase the recruitment of immune cells, especially monocytes and neutrophils, to the subendothelial space to participate in plaque formation. Recruitment of monocytes to the subendothelial space is mediated by the upregulation of cell adhesion molecules and chemokines (such as monocyte chemoattractant protein-1 (MCP-1)) via oxLDL signaling pathways [7,8]. The recruited monocytes then differentiate into macrophages mediated by colony-stimulating cytokines such as M-CSF and GM-CSF, both of which are enhanced through oxLDL-induced signaling pathways [9]. Macrophages also differentiate into inflammatory (M1) macrophages in the presence of the Th1 cytokine, or anti-inflammatory (M2) macrophages in the presence of the Th2 cytokine, in the subendothelial environment [10,11]. No M1 and M2 macrophages have been identified as specific precursors for foam cell formation, but several studies have shown that M2 macrophages are more susceptible to foam cell formation [12].
Another aspect of the pathophysiology of atherosclerosis is the dysregulation of cholesterol metabolism in the macrophage, which is the main cell responsible for atherosclerotic plaque [13]. Oxidized-LDL signaling pathways are involved in the downregulation of cholesterol efflux transporters. By increasing both the internalization of oxLDL and the accumulation of lipid droplets in the macrophage, foam cell formation gradually occurs, which initially leads to fatty streaks and ultimately, to primary atherosclerotic lesions. Foam cells play a central role in the pathogenesis of atherosclerosis. Specifically, the formation and accumulation of foam cells in the subendothelial space of a damaged artery is one of the early key steps responsible for the development of atherosclerosis [14,15]. MCP-1 and TNF-α represent two inflammatory mediators released during foam cell formation. MCP-1 increases vascular smooth muscle cell (VSMC) proliferation and leukocyte migration and TNF-α upregulates cell adhesion molecules (CAMs), which subsequently increases the recruitment of VSMCs and immune cells [16]. The proliferation of VSMCs inside the plaque temporarily stabilizes the lesion through collagen synthesis and other extracellular matrix (ECM) proteins, which function to maintain the integrity of the fibrous cap and inhibit cap rupture. Nevertheless, an increasing number of foam cells, which have pro-atherogenic properties, ultimately leads to plaque rupture and blood vessel occlusion due to the release of matrix-degrading enzymes [17,18]. Most plaque growth, or lesion progression, is a result of macrophage activity, although other immune cells (e.g., neutrophils), have also been shown to play an important role [19]. Eventually, the plaque ruptures and results in clinical manifestations recognized as MI and stroke [20].

3. Foam Cells as Therapeutic Targets in Atherosclerosis

Foam cell formation is one of the critical processes in the development and pathogenesis of atherosclerosis. As mentioned above, fatty streaks are the first detectable “yellow “lesion in the vessel wall that indicates foam cell formation and the development of atherosclerosis. Foam cells are involved in the formation of primary atherosclerotic plaques, their continued growth and ultimately, their rupture, which finally, leads to either MI or stroke [5,12,14]. The specific therapeutic goals associated with each stage in the formation of foam cells and the development of atherosclerosis will be considered below. Processing and metabolism of cholesterol in the macrophage is one of the major steps in foam cell formation. Cholesterol processing involves lipid uptake, cholesterol esterification and cholesterol efflux, which ultimately leads to cholesterol equilibrium in the macrophage. However, dysregulation and disruption of these processes results in foam cell formation [21]. During the development of atherosclerosis, chemoattractants (especially MCP-1) recruit monocytes to the subendothelial space of the damaged endothelium and they undergo differentiation to macrophages [22]. The macrophages express scavenger receptors SRA-1, SRA-2, CD36 and LOX1. Cellular uptake of oxLDL occurs by phagocytosis and pinocytosis via scavenger receptors, which preferentially incorporate oxLDL relative to its native (unmodified/unoxidized) form [23].
Numerous studies in mice have selectively targeted either chemoattractants or adhesion molecules involved in the recruitment of monocytes, or cytokines involved in the transmigration of monocytes. This has been accomplished by inactivating the genes encoding various molecules, such as M-CSF, CCL2, CXCR1 and CCR5, which has been shown to result in less and smaller atherosclerotic lesions in ApoE−/− mice [24,25,26]. Animal studies using LDLR−/− mice have evaluated both the knockdown of VCAM-1 [27,28], or drug inhibition with a natural antioxidant AGI-1067, which showed successful and significant reductions in the development of atherosclerotic plaques in LDLR−/− mice [29]. SRA-1, 2, CD36 and LOX-1 are the primary scavenger receptors of macrophages that play a key role in lipid uptake and are controlled by several regulators such as MEKK-2, MAP kinases and STAT1 [23,30,31,32]. Numerous studies have targeted the role of these three major scavenger receptors in foam cell formation. For example, overexpression of LOX-1 in ApoE−/− mice accelerated the development of atherosclerosis, while LOX-1−/−, LDLR−/− mice had smaller atherosclerotic lesions, which suggests a proatherogenic role for LOX-1 [33]. These receptors are multifunctional in nature and their expression is not limited to macrophages, because they are also expressed in aortic endothelial cells (ECs) and VSMCs. For these reasons, as well as the fact that different lipid uptake pathways exist, such as phagocytosis and pinocytosis, it is difficult to design studies that evaluate the therapeutic targeting of these receptors [34]. Additionally, an analysis of different studies indicates that there is still a debate between whether these receptors are pro-atherogenic, or anti-atherogenic [35,36]. Thus, we would suggest that further studies are needed to better understand the lipid uptake pathways and the true functional role of each of these receptors. In cholesterol esterification, ACAT-1 and neutral cholesterol ester hydrolase (NCEH) play a key role in catalyzing the esterification of cholesterol and removing cholesterol from macrophages, respectively [37,38]. ACAT-1 is found ubiquitously in the endoplasmic reticulum of cells and is sensitive to the degree of membrane cholesterol enrichment, which is why it functions to maintain the cholesterol content of cell membranes at an optimal level by catalyzing the esterification of excess free cholesterol [39]. Studies have shown that ACAT-1 gene knockout in ApoE−/− and LDLR−/− mice resulted in no change in the development of atherosclerotic lesions and failed to inhibit foam cell formation [40]. However, pharmacological inhibition of ACAT-1 in macrophages has been shown to increase foam cell formation in mouse and rabbit models of atherosclerosis [41,42]. Moreover, animal studies evaluating the suppression of NCEH1 and hormone-sensitive lipase (LIPE), two important enzymes involved in the hydrolysis of cholesterol esters (CEs) back to free cholesterol (FC), have demonstrated a significant increase in intracellular CE accumulation when compared to control animals [38,43]. This would suggest that future studies should target all three enzymes (NCEH1, LIPE and ACAT-1) to intentionally modulate cholesterol turnover in macrophages.
Cholesterol efflux is one of the primary events in cholesterol metabolism and foam cell formation. The mechanism for cholesterol efflux is mainly attributable to ATP-binding cassette transporters ABCA1, ABCG1, as well as SR-B1, which function to maintain cholesterol and phospholipid homeostasis in the macrophage [44]. ABCA1 facilitates cholesterol and phospholipid efflux and leads to the formation of ApoA1, while ABCG1 mediates cholesterol efflux to form nascent high-density lipoprotein (HDL), which ultimately prevents the formation of foam cells. The expression of these transporters is predominantly dependent on the activation of PPAR and LXRα transcription factors [45]. Many studies have targeted these specific transporters. For example, in LDLR−/− mice treated with PPARα and PPARγ agonists, the progression of atherosclerosis decreased due to the increased expression of ABCG1 and ABCA1 [46,47].
One of the limitations of targeting these transporters involved in cholesterol efflux pathways is that other pathways, such as passive diffusion, exist to transport cholesterol throughout the cell [48]. In studies that have been conducted, including genetic ablation of both ABCA1 and SR-B1 simultaneously in ABCA1×SR-BI double knockout (dKO) mice, macrophage RCT was markedly impaired and macrophage foam cell formation was increased in both lung and Peyer’s patches of dKO mice. However, atherosclerotic lesions did not develop in these dKO mice, which was potentially due to the low levels of non-HDL-C [49]. It should be noted, however, that hepatic overexpression of ABCA1 in LDLr-KO mice leads to enhanced aortic atherosclerotic lesions [49]. Therefore, knockdown studies on these three transporters could best be described as having yielded mixed results [50].
Cholesterol efflux inhibits the accumulation of excess lipids in the foam cell. Over time, the ability of the foam cell to manage the extra lipoproteins in the foam cell decreases, which leads to endoplasmic reticulum stress and the production of ROS and, in turn, triggers the apoptotic cascade. Additionally, uncontrolled lipoprotein uptake itself causes foam cell apoptosis [51,52].
Eventually, apoptosis of foam cells leads to the release of pro-inflammatory cytokines such as IL-1, IL-6, TNF-alpha and matrix metalloproteinases (MMPs); all, of which, may aggravate atherosclerosis by promoting an inflammatory state and the infiltration of immune cells [53].
The use of either pharmacological anti-inflammatory agents such as canakinumab, adalimumab and the TNF inhibitor etanercept, or agents that attenuate oxidative stress, may hold promise as a therapeutic intervention [54,55].
Furthermore, trying to slow the death of foam cells might represent a therapeutic goal to prevent the development and worsening of atherosclerosis [56]. There are a number of strategies that could be employed to decelerate foam cell apoptosis. These include (1) targeting apoptotic pathways, such as the genetic inhibition of proapoptotic proteins such as BAX or Bcl-2 [57,58], (2) knockdown of apoptosis inhibitor of macrophages (AIM) [59], (3) targeting secondary necrosis pathways such as the clearance of apoptotic cells [60,61], (4) promoting efferocytosis of apoptotic macrophages by using LXR ligands or glucocorticoids [62,63], or (5) by activating PPARγ pathways [64]. Another strategy to either control foam cell formation or increase the clearance of foam cells is statin therapy. Statins have been extensively investigated in patients with cardiovascular disease due to their numerous pleiotropic effects such as reducing oxidative stress, enhancing plaque stability and exerting anti-inflammatory effects [65,66,67,68,69,70,71,72,73,74,75]. Following foam cell formation, the foam cells take on a profibrotic phenotype. By releasing substances such as MMPs, they increase monocyte recruitment and degradation of extracellular matrix proteins such as collagen and fibronectin. This process leads to plaque instability and ultimately, plaque rupture [53,76]. For this reason, inhibition of MMPs may represent a treatment option in atherosclerosis. Many studies have targeted MMPs, in particular MMP8, because its increase in human carotid plaque has been associated with an unstable plaque phenotype [77]. Using MMP8−/− ApoE−/− mice, it has been demonstrated that monocyte/leukocyte recruitment and macrophage lesions decrease. Moreover, either MMP-9 or MMP-12 deficiency in ApoE−/− mice has been shown to increase plaque instability [78,79,80]. Historically, monocyte-derived macrophages were considered the major source of plaque foam cells. However, other cells in the arterial wall such as ECs and VSMCs, as well as stem/progenitor cells (SPCs), show a foam cell phenotype in atherosclerotic plaques [81,82]. In fact, studies have shown that 50% of the foam cell population originates from VSMCs in human atherosclerotic lesions [83]. Therefore, this may suggest yet another strategy in the treatment of atherosclerosis; namely, targeting alternative cellular origins of foam cells [84,85,86]. Finally, one of the most important strategies for targeting foam cell formation is the use of miRNAs, which will be discussed in detail in this article.
The role and function of non-coding RNAs in foam cell formation and related processes (cholesterol efflux, cholesterol influx and cholesterol esterification) is depicted in Figure 1.

4. Non-Coding RNAs as Therapeutic Targets

The development of modern genomic and transcriptomic techniques, mass spectrometry and bioinformatics has led to a greater understanding concerning the role of non-coding RNAs (ncRNAs) in the regulation of gene expression involved in cellular and pathogenic disease processes, especially in atherosclerosis and the role of foam cells. Many studies have been conducted to identify the capacity of ncRNAs to alter or improve the progression of atherosclerosis at the foam cell stage [87,88,89,90]. ncRNAs are molecules that do not translate into proteins and the size of ncRNAs correlates with their function. There are different classes of ncRNAs in biological processes. For example, those involved in the regulation of gene expression, such as miRNAs, piRNAs and lncRNAs; those involved in the maturation of RNAs (snRNAs, snoRNAs); and those involved in the synthesis of proteins (rRNAs, tRNAs). The ncRNAs are typically classified by size. Thus, ncRNAs that are less than 200 nucleotides in length are categorized as small ncRNAs (sncRNAs), while those that are over 200 nucleotides in length are considered long ncRNAs (lncRNAs) [91,92]. The length of microRNAs is between 20–24 nucleotides, which bind to the 3’ untranslated region (3′UTR) of the target mRNA and either prevent its translation, or cause it to degrade. Most miRs reduce gene expression post-transcription, although in some instances, miRs can increase gene expression by activating the translation of the target mRNA. MiRNAs are often transcribed by RNA polymerase II (its early version was termed pri-miRNA) [93,94,95]. Biogenesis of miRNA occurs via the canonical pathway involving Drosha and Dicer, as well as through different non-canonical pathways that are independent of Drosha and even Dicer. Single mRNAs can be involved in different biological processes and can also be targeted by different miRNAs. Therefore, miRNAs play a multifunctional role due to the fact that they can participate in separate biological processes, which represents one of the limitations of using miRNAs in basic research and clinical investigations [96,97]. LncRNAs have a size between 200 bp to 100 kb, which are capable of binding to DNA, protein and RNA at various levels, exerting cellular regulation such as chromatin remodeling, mRNA splicing, or mRNA translation and multi-protein complex assembly. Additionally, lncRNAs are transcribed by RNA polymerase II and most of them undergo alternative splicing, 5′-capping and polyadenylation processes [98,99].
LncRNAs may function in both the cis and trans configuration to regulate the expression of target genes. They are also able to use a set of different molecular mechanisms to accomplish this goal. For example, they may act as a scaffold for the recruitment of chromatin modifiers or transcription factors, act as decoys for the breakdown of proteins and function as miRNA sponges to activate or deactivate target genes [100,101]. Presently, lncRNAs are classified by their genomic localization and modes of action or function, including intronic lncRNAs and intergenic lncRNAs that originate from a different region of the gene. Most mature transcribed lncRNAs are thought to have little potential for protein-coding, because some of them contain small open reading frames and encode small functional peptides. Interestingly, lncRNAs have more functional diversity due to the fact that they have less conserved sequences and are under less selective pressure [98,102,103].
In pathological conditions, therapeutic strategies involving miRNAs primarily use miRNA inhibition to reduce the expression of miRNAs whose expression has increased and use miRNA replacement to increase the expression of suppressed miRNAs.
miRNA inhibitors are chemically modified single-stranded antisense oligonucleotides (ASOs) that are complementary to the mature miRNA. To date, angtagomiRs (target-single miRNAs conjugated with cholesterol), anti-miRs (target-single miRNAs) and tiny anti-miRs (target miRNA families) have been synthesized to mediate miRNA inhibition. These ASOs can reduce pathogenic expressed miRNAs [91,92].
Nonviral delivery of miRNAs, for example, using liposomes, nanoparticles, or antibody-based methods, as well as viral delivery such as adeno-associated virus (AAV) and adenovirus have been used for the restoration of microRNA levels.
Since various studies have shown that ncRNAs have the potential to regulate different cell pathways, one promising therapeutic goal could be to manipulate their function using oligonucleotide inhibitors or miR mimics. Antisense oligonucleotides directed against specific miR sequences are effectively taken up by diverse tissues. Furthermore, miR mimics and inhibitors are relatively stable in plasma and miR mimics and inhibitors are not highly toxic and can easily reach cellular gene targets. However, the challenge to directly target a specific inflamed tissue and/or a specific cell line is still ongoing [87,88]. It should be noted that some studies using anti-miR have shown that they can target plaque macrophages and regulate gene expression in these cells [91,92].
Non-coding RNAs, which are involved in stimulating and reducing the formation of foam cells, are summarized in Table 1 and Table 2.

5. Non-Coding RNAs That Stimulate Foam Cell Formation/Function

Most of the microRNAs described in this review, either directly or indirectly, target ABCA1.
As mentioned above, ABCA1 is a membrane protein at the cell surface that regulates the transport of cholesterol and phospholipid and other metabolites [144]. In general, deficiency or downregulation of ABCA1 expression leads to a significant decrease in serum HDL levels and an increased risk of atherosclerosis. This is primarily due to the suppression of cholesterol efflux and disruption of the reverse cholesterol transport (RCT) cycle and subsequent foam cell formation. HDL has a cardioprotective role by preventing the oxidation of lipoproteins and returning cholesterol from peripheral tissues back to the liver via the RCT process [145].
Protein levels and ABCA1 activity are regulated at both the transcriptional and post-transcriptional levels, such that any downregulation of ABCA1 expression adversely affects the function of atheroprotective lipoprotein subclasses. Additionally, in APOE−/− mice, the overexpression of ABCA1 can effectively reduce the size of atherosclerotic plaques [146,147].
Direct and indirect effects of different Non-coding RNAs on the expression of ABCA1 are shown in Figure 2.
Some microRNA, such as miR-33 [104], miR-19b [106], miR-101 [106], miR-144-3p [106], miR-302a [104], miR-26 [106], miR-20a/b [106] and miR-758 [106] directly target the 3′UTR of ABCA1 in macrophages, which suppresses ABCA1 expression and disrupts cholesterol homeostasis. The disruption in cholesterol homeostasis occurs due to a decrease in the efflux of cholesterol to ApoA1 or HDL, which ultimately causes the formation of foam cells. A more detailed description is given in Table 1.

5.1. miR-33

This microRNA directly targets ABCA1 and ABCG1 in murine and human macrophages and downregulates these transporters, which leads to lower cellular cholesterol efflux and increased cholesterol accumulation and foam cell formation. In human cells, miR-33 neither downregulated the expression of ABCG1, nor interfered with cholesterol efflux to HDL in human THP-1 macrophages, which is consistent with the lack of miR-33-binding sites in the human ABCG1 3′UTR [150,151,152]. In a study using double-knockout miR-33−/−, ApoE−/− mice, cholesterol efflux increased and plaque size decreased compared to control mice. Moreover, it was demonstrated that the use of microRNA-33 antagonism increased ABCA1 expression in vitro and in vivo and increased cholesterol efflux to ApoA1 [105,152,153]. In another study, the silencing of miR-33 with antisense oligonucleotides (anti-miR-33) in LDLR−/− mice for 14 weeks did not affect lesion formation and the progression of atherosclerosis and failed to maintain elevated plasma HDL-cholesterol (HDL-C) [104]. This is probably because previous regression studies employed mice that received anti-miRs for only 4 weeks, as well as the existence of homeostatic compensatory mechanisms controlling plasma HDL. Additionally, the discrepancy between experimental designs, including the dietary composition, the gender of the mice and the possible influence of the intestinal microbiome may vary between the two models [154]. There is evidence suggesting athero-protective effects of hematopoietic loss of miR-33 including decreased number of apoptotic cells and decreased size of the necrotic cores in plaques from mice transplanted with miR-33−/− bone marrow, indicative of enhanced plaque stability. Moreover, upregulation of MERTK experssion, a kinase involved in the regulation of efferocytosis, increases ABCA1 expression, MQs cholesterol efflux capacity and RCT and decreases the accumulation of cholesterol esters (CEs) in MQs in vivo. Furthermore, loss of miR-33 remarkably decreases the amount of MQ phosphatidylethanolamines (PEs) and phosphatidylserine (PS). These changes may have important effects such as attenuating NFKB activation and innate immune response [155].
Despite athero-protective effect of anti-miR-33 therapies, global deletion of miR-33 results in disturbed responsiveness to insulin in multiple metabolic organs including the liver, white adipose tissue (WAT) and skeletal muscle through TAGs, DAGs and ceramides accumulation, PKC activation and inhibition of ERK activity. Furthermore, alterations in feeding behavior are mainly responsible for the obesity in miR-33−/− mice. Since the hypothalamic POMC and AgRP neurons, which regulate food intake, are also GABAergic neurons, they may be directly influenced by alterations in miR-33. This could elucidate some of the differences between genetic and pharmacological inhibition of miR-33, as antimiR-33 therapeutics would not be expected to pass the blood-brain barrier [156].

5.2. miR-27a/b

The miR-27 family has two isoforms; namely, miR-27a, which is an intergenic miRNA, and miR-27b, which is an intronic miRNA. miR-27a/b has been reported to regulate several genes involved in lipid metabolism, including PPARs and RXRs, which can activate the transcription of many target genes in vivo, including SR-BI, ABCG1 and ABCA1 [157]. This miRNA, by repressing CD36 expression in THP-1 macrophages, regulates cholesterol uptake. Additionally, miR-27a/b reduces the expression of ACAT1 (involved in the formation of foam cells via re-esterification of excess FC as a means to increase CEs), which, in turn, leads to a decrease in the formation of CEs in THP-1 macrophage-derived foam cells [157]. To maintain the homeostasis of cholesterol in macrophages, intracellular CE is hydrolyzed to FC as the initial step for the efflux of excess cholesterol, which occurs via the ABC transporters ABCA1 and ABCG1, as well as other non-transporter pathways such as SR-B1 [157]. Moreover, miR-27a/b inhibits LPL expression, which is involved in lipid uptake in vitro and in vivo. Importantly, this miRNA also regulates HDL-mediated cholesterol efflux in human foam cells without targeting SRBI and ABCG1, which probably indirectly affects the PPAR/RXR pathway. Additionally, miR-27a/b inhibits cholesterol efflux from macrophages via reduction in the expression of ABCA1 and subsequently causing a decrease in ApoA1 [110,158,159].
Some microRNAs, such as miR-216a [114,160], miR-486 [114,160], miR-212 [114,160], miR-497 [114,160] and LncRNA ENST00000602558.1 [118], also indirectly decrease the expression of the ABCA1 gene. The details are provided in Table 1.

5.3. miR-216a

The expression of this particular miRNA has been shown to be increased in myocardial biopsies from patients with ischemic heart failure [161].
It is known that the cystathionine gamma-lyase/hydrogen sulfide (CSE/H2S) enzymatic reaction in the trans-sulfuration pathway (a pathway that generates endogenous H2S) increases ABCA1 expression in foam cell [160,162,163,164].
Specifically, miR-216a directly targets the 3′UTR of CSE, which is one of the two key enzymes in endogenous H2S production. The CSE/H2S system has an anti-atherosclerotic role in the cardiovascular system. Downregulation of the CSE/H2S enzymatic reaction pathway results in decreased ABCA1 expression due to a reduction in the phosphorylation of PI3K and AKT, which subsequently leads to an increase in cholesterol accumulation in foam cells [114,160].

5.4. miR-382-5p

The RP5-833A20.1/miR-382-5p/NFIA pathway is essential for the regulation of cholesterol homeostasis and inflammatory responses. Interestingly, RP5-833A20.1 has been shown to decrease nuclear factor IA (NFIA) expression through the hsa-miR-382-5p pathway. NFIA is essential for adipocyte differentiation and lipid droplet formation. Specifically, NFIA inhibits atherosclerotic plaque formation in ApoE−/− mice by increasing RCT and decreasing circulating cytokine levels. Overexpression of miR-382-5p causes a reduction in the expression of NFIA, as well as decreases the expression of ABCA1 and ABCG1 and increases the expression of SRA1, CD36 and NFKB [107,165]. In vivo studies with miR-382-5p have not been performed to date.

6. Non-Coding RNAs That Attenuate Foam Cell Formation

It has been shown that some non-coding RNAs are involved in macrophage cholesterol homeostasis by acting on inflammatory pathways. Studies have shown that miR-181a [126], miR-135a [126] and miR-223 [126], target TLR4 and reduce its expression. Thisprocess regulates cholesterol homeostasis in macrophages by reducing inflammation, reducing lipid uptake and increasing cholesterol efflux. Within macrophages, MiR-23a [126], miR-16 [126] and Let-7g [126] have an inhibitory effect on NF-kB pathways, which regulate lipid metabolism by reducing inflammation. The details are summarized in Table 2.

6.1. miR-150

MiR-150 has been shown to be upregulated in a human monocyte cell line in response to oxLDL treatment. Overexpression of this microRNA was proven to decrease the endogenous expression of AdipoR2 in macrophages, which promotes cholesterol efflux by the upregulation of ABCA1 and ABCG1. This process occurs via the PPARγ- and LXRα-dependent signaling pathways and decreases CD36 and intracellular lipid accumulation.
Since miR-150 can be packaged into microvesicles (MVs) and secreted from monocytes, MVs isolated from the plasma of patients with atherosclerosis have been determined to possess greater amounts of miR-150 than those from healthy donors. This finding may be explained by the fact that during the pathogenesis of atherosclerosis, stimulation of monocytes to release MVs that contain miRNAs, such as miR-150, can prevent lipid accumulation and foam cell formation [130].

6.2. miR-155

Dual effects of miR-155 on macrophages in the context of atherosclerosis.
miR-155 is a “multi-target” molecule specifically expressed in atherosclerotic plaques and pro-inflammatory macrophages [166]. The effects of miR-155 on atherogenesis have been controversial, because it exerts dual effects on both inflammatory and apoptotic pathways in lesional macrophages. miR-155 can enhance or prevent lesion formation relevant to the “stage” of atherosclerosis. Nazari-Jahantigh et al. reported that miR-155 was anti-atherogenic in the early stage of atherosclerosis, but became pro-atherogenic as the lesions progressed to a more advanced stage [167]. OxLDL induces miR-155 expression in human macrophages, which is essential for lipid uptake and ROS production by macrophages [168]. The use of antagomiR-155 has been shown to reduce lipid levels in human macrophages and decrease the size of plaques in ApoE−/− mice. Elevated miR-155 levels increase oxLDL-induced foam cell formation by targeting HBP1. In addition, miR-155 expression is downregulated by the YY1/HDACs complex [113].
Another study showed that miR-155 mimics enhanced the expression of CEH at both the transcriptional and translational level in a dose- and time-dependent manner in human foam cells, although this effect may have occurred due to the inhibition in the expression of Tim-3, since overexpression of Tim-3 can inhibit the expression of CEH. Additionally, using a human monocyte cell line, it has been demonstrated that the overexpression of miR-155 can significantly inhibit the expression of SR-A, decrease lipid accumulation, increase the expression of ABCA1 and thereby increase cholesterol efflux [132].
A study also found that NF-κB transcription factor, which is activated by TNF-α, binds to the miR-155 promoter and stimulates transcription of miR-155. Accordingly, the expression of calcium-regulated heat-stable protein 1 (CARHSP1) is decreased. This is significant, because CARHSP1 enhances the stability of TNF-α mRNA, which is important for the efficient production of TNF-α. Furthermore, it has been demonstrated in a human monocyte cell line that miR-155 indirectly decreases TNF-α expression, decreases macrophage inflammation and lipid uptake and decreases foam cell formation [133].

7. Conclusions

Atherosclerosis is a chronic disease with a network of complex pathological processes. One of the primary mechanisms involved in this disease is the formation of foam cells, which leads to the growth and rupture of atherosclerotic plaques and finally, to its clinical manifestations (MI and stroke) [12,169]. Noncoding RNAs play an important role in various aspects of atherogenesis including foam cell formation. For example, many studies using noncoding RNAs have targeted different aspects of foam cell formation, such as cholesterol uptake, storage and efflux, indicating the well-established and effective role of non-coding RNAs, particularly miRNAs, in foam cell formation. Noncoding RNAs, especially miRNAs such, as miR-33, miR-27ab and miR-144, can have a pro-atherosclerotic role by stimulating foam cell formation, while others, such as miR-150, miR-1275 and Let-7g, exert an atheroprotective role by suppressing foam cell formation. miR-155 has shown conflicting results in different studies.
Clinical work performed using miRNAs has focused primarily on the inhibition of miRNAs through antisense oligonucleotides that complement targeting mRNA. For example, anti-miR oligonucleotides, modified antimiR oligonucleotides, anti-miR peptides, or the use of genetic vectors to replace miRs, such as with miR mimics, have been used [87,88,89,90]. Today, many preclinical studies have investigated the therapeutic anti-atherosclerotic potential of miRNAs by using miR mimics or their antagonists, which have shown promising results. These findings indicate that miRNA targeted therapy may represent a novel approach for the treatment of atherosclerosis. However, chronic treatment or genetic ablation of some of these miRNAs (miR-33), has been found to result in adverse effects including obesity and insulin resistance [156].
There are also limitations to using miRNAs, including their multifunctionality and their role in different biological processes. Moreover, there are few specific miRNAs in a tissue or organ, so most of the miRNAs affect unselected or non-targeted tissues. One of the major drawbacks to the therapeutic use of microRNAs in atherosclerosis is the fact that many molecules of this class have additional unwanted side effects and studies on the interaction between different miRNAs are relatively scarce. Secondly, miRNAs target prediction tools are not completely accurate and false positive and false negative predictions are still an issue of concern. Lastly, the full scope of elucidating their functions in vivo is limited and this fact impedes the investigation of the most interesting primate-specific non-coding RNAs in widely used atherosclerotic mouse models. Furthermore, the discrepancies in anatomy, lipid metabolism and gene expression complicate the translation of experimental results obtained in mice to humans.

Author Contributions

All authors were involved in the conception, design, drafting, revision and final approval of this review. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

Maciej Banach: speakers bureau: Abbott/Mylan, Abbott Vascular, Actavis, Akcea, Amgen, Biofarm, KRKA, MSD, Polpharma, Sanofi Aventis, Servier and Valeant; consultant to Abbott Vascular, Akcea, Amgen, Daichii Sankyo, Esperion, Freia Pharmaceuticals, Lilly, MSD, Polfarmex, Resverlogix, Sanofi-Aventis; Grants from Sanofi and Valeant. All other authors declare no conflict of interest.

References

  1. Wakabayashi, I. Associations between alcohol drinking and multiple risk factors for atherosclerosis in smokers and nonsmokers. Angiology 2010, 61, 495–503. [Google Scholar] [CrossRef] [PubMed]
  2. Sitia, S.; Tomasoni, L.; Atzeni, F.; Ambrosio, G.; Cordiano, C.; Catapano, A.; Tramontana, S.; Perticone, F.; Naccarato, P.; Camici, P. From endothelial dysfunction to atherosclerosis. Autoimmun. Rev. 2010, 9, 830–834. [Google Scholar] [CrossRef] [PubMed]
  3. Gliozzi, M.; Scicchitano, M.; Bosco, F.; Musolino, V.; Carresi, C.; Scarano, F.; Maiuolo, J.; Nucera, S.; Maretta, A.; Paone, S. Modulation of nitric oxide synthases by oxidized LDLs: Role in vascular inflammation and atherosclerosis development. Int. J. Mol. Sci. 2019, 20, 3294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Cyr, A.R.; Huckaby, L.V.; Shiva, S.S.; Zuckerbraun, B.S. Nitric Oxide and Endothelial Dysfunction. Crit. Care Clin. 2020, 36, 307–321. [Google Scholar] [CrossRef]
  5. Liu, W.; Yin, Y.; Zhou, Z.; He, M.; Dai, Y. OxLDL-induced IL-1beta secretion promoting foam cells formation was mainly via CD36 mediated ROS production leading to NLRP3 inflammasome activation. Inflamm. Res. 2014, 63, 33–43. [Google Scholar] [CrossRef]
  6. Daub, K.; Seizer, P.; Stellos, K.; Krämer, B.F.; Bigalke, B.; Schaller, M.; Fateh-Moghadam, S.; Gawaz, M.; Lindemann, S. Oxidized LDL-activated platelets induce vascular inflammation. In Seminars in Thrombosis and Hemostasis; Thieme Medical Publishers: New York, NY, USA, 2010; pp. 146–156. [Google Scholar]
  7. Chávez-Sánchez, L.; Espinosa-Luna, J.E.; Chávez-Rueda, K.; Legorreta-Haquet, M.V.; Montoya-Díaz, E.; Blanco-Favela, F. Innate immune system cells in atherosclerosis. Arch. Med. Res. 2014, 45, 1–14. [Google Scholar] [CrossRef] [PubMed]
  8. Kiyan, Y.; Tkachuk, S.; Hilfiker-Kleiner, D.; Haller, H.; Fuhrman, B.; Dumler, I. oxLDL induces inflammatory responses in vascular smooth muscle cells via urokinase receptor association with CD36 and TLR4. J. Mol. Cell. Cardiol. 2014, 66, 72–82. [Google Scholar] [CrossRef]
  9. Seo, J.-W.; Yang, E.-J.; Yoo, K.-H.; Choi, I.-H. Macrophage differentiation from monocytes is influenced by the lipid oxidation degree of low density lipoprotein. Mediat. Inflamm. 2015, 2015, 235797. [Google Scholar] [CrossRef] [Green Version]
  10. Rios, F.J.; Koga, M.M.; Pecenin, M.; Ferracini, M.; Gidlund, M.; Jancar, S. Oxidized LDL induces alternative macrophage phenotype through activation of CD36 and PAFR. Mediat. Inflamm. 2013, 2013, 198193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Wolfs, I.M.; Donners, M.M.; de Winther, M.P. Differentiation factors and cytokines in the atherosclerotic plaque micro-environment as a trigger for macrophage polarisation. Thromb. Haemost. 2011, 106, 763–771. [Google Scholar] [CrossRef]
  12. Maguire, E.M.; Pearce, S.W.; Xiao, Q. Foam cell formation: A new target for fighting atherosclerosis and cardiovascular disease. Vasc. Pharmacol. 2019, 112, 54–71. [Google Scholar] [CrossRef]
  13. Moore, K.J.; Sheedy, F.J.; Fisher, E.A. Macrophages in atherosclerosis: A dynamic balance. Nat. Rev. Immunol. 2013, 13, 709–721. [Google Scholar] [CrossRef]
  14. Vainio, S.; Ikonen, E. Macrophage cholesterol transport: A critical player in foam cell formation. Ann. Med. 2003, 35, 146–155. [Google Scholar] [CrossRef] [PubMed]
  15. Birck, M.M.; Saraste, A.; Hyttel, P.; Odermarsky, M.; Liuba, P.; Saukko, P.; Hansen, A.K.; Pesonen, E. Endothelial cell death and intimal foam cell accumulation in the coronary artery of infected hypercholesterolemic minipigs. J. Cardiovasc. Transl. Res. 2013, 6, 579–587. [Google Scholar] [CrossRef] [PubMed]
  16. Chen, C.; Khismatullin, D.B. Oxidized low-density lipoprotein contributes to atherogenesis via co-activation of macrophages and mast cells. PLoS ONE 2015, 10, e0123088. [Google Scholar] [CrossRef] [PubMed]
  17. Galis, Z.S.; Sukhova, G.K.; Lark, M.W.; Libby, P. Increased expression of matrix metalloproteinases and matrix degrading activity in vulnerable regions of human atherosclerotic plaques. J. Clin. Investig. 1994, 94, 2493–2503. [Google Scholar] [CrossRef] [Green Version]
  18. Hultgårdh-Nilsson, A.; Durbeej, M. Role of the extracellular matrix and its receptors in smooth muscle cell function: Implications in vascular development and disease. Curr. Opin. Lipidol. 2007, 18, 540–545. [Google Scholar] [CrossRef]
  19. Wezel, A.; Lagraauw, H.M.; van der Velden, D.; de Jager, S.C.; Quax, P.H.; Kuiper, J.; Bot, I. Mast cells mediate neutrophil recruitment during atherosclerotic plaque progression. Atherosclerosis 2015, 241, 289–296. [Google Scholar] [CrossRef]
  20. Van der Donckt, C.; Van Herck, J.L.; Schrijvers, D.M.; Vanhoutte, G.; Verhoye, M.; Blockx, I.; Van Der Linden, A.; Bauters, D.; Lijnen, H.R.; Sluimer, J.C. Elastin fragmentation in atherosclerotic mice leads to intraplaque neovascularization, plaque rupture, myocardial infarction, stroke, and sudden death. Eur. Heart J. 2015, 36, 1049–1058. [Google Scholar] [CrossRef] [Green Version]
  21. Choudhury, R.P.; Lee, J.M.; Greaves, D.R. Mechanisms of disease: Macrophage-derived foam cells emerging as therapeutic targets in atherosclerosis. Nat. Clin. Pract. Cardiovasc. Med. 2005, 2, 309–315. [Google Scholar] [CrossRef]
  22. Peters, W.; Charo, I.F. Involvement of chemokine receptor 2 and its ligand, monocyte chemoattractant protein-1, in the development of atherosclerosis: Lessons from knockout mice. Curr. Opin. Lipidol. 2001, 12, 175–180. [Google Scholar] [CrossRef]
  23. Greaves, D.R.; Gordon, S. Thematic review series: The immune system and atherogenesis. Recent insights into the biology of macrophage scavenger receptors. J. Lipid Res. 2005, 46, 11–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Combadière, C.; Potteaux, S.; Rodero, M.; Simon, T.; Pezard, A.; Esposito, B.; Merval, R.; Proudfoot, A.; Tedgui, A.; Mallat, Z. Combined inhibition of CCL2, CX3CR1, and CCR5 abrogates Ly6C(hi) and Ly6C(lo) monocytosis and almost abolishes atherosclerosis in hypercholesterolemic mice. Circulation 2008, 117, 1649–1657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Smith, J.D.; Trogan, E.; Ginsberg, M.; Grigaux, C.; Tian, J.; Miyata, M. Decreased atherosclerosis in mice deficient in both macrophage colony-stimulating factor (op) and apolipoprotein E. Proc. Natl. Acad. Sci. USA 1995, 92, 8264–8268. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Cipriani, S.; Francisci, D.; Mencarelli, A.; Renga, B.; Schiaroli, E.; D’Amore, C.; Baldelli, F.; Fiorucci, S. Efficacy of the CCR5 antagonist maraviroc in reducing early, ritonavir-induced atherogenesis and advanced plaque progression in mice. Circulation 2013, 127, 2114–2124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Fotis, L.; Agrogiannis, G.; Vlachos, I.S.; Pantopoulou, A.; Margoni, A.; Kostaki, M.; Verikokos, C.; Tzivras, D.; Mikhailidis, D.P.; Perrea, D. Intercellular adhesion molecule (ICAM)-1 and vascular cell adhesion molecule (VCAM)-1 at the early stages of atherosclerosis in a rat model. In Vivo 2012, 26, 243–250. [Google Scholar]
  28. Ley, K.; Huo, Y. VCAM-1 is critical in atherosclerosis. J. Clin. Investig. 2001, 107, 1209–1210. [Google Scholar] [CrossRef]
  29. Serebruany, V.; Malinin, A.; Scott, R. The in vitro effects of a novel vascular protectant, AGI-1067, on platelet aggregation and major receptor expression in subjects with multiple risk factors for vascular disease. J. Cardiovasc. Pharmacol. Ther. 2006, 11, 191–196. [Google Scholar] [CrossRef]
  30. Rahaman, S.O.; Lennon, D.J.; Febbraio, M.; Podrez, E.A.; Hazen, S.L.; Silverstein, R.L. A CD36-dependent signaling cascade is necessary for macrophage foam cell formation. Cell Metab. 2006, 4, 211–221. [Google Scholar] [CrossRef] [Green Version]
  31. Coller, S.P.; Paulnock, D.M. Signaling pathways initiated in macrophages after engagement of type A scavenger receptors. J. Leukoc. Biol. 2001, 70, 142–148. [Google Scholar]
  32. Sikorski, K.; Czerwoniec, A.; Bujnicki, J.M.; Wesoly, J.; Bluyssen, H.A. STAT1 as a novel therapeutical target in pro-atherogenic signal integration of IFNγ, TLR4 and IL-6 in vascular disease. Cytokine Growth Factor Rev. 2011, 22, 211–219. [Google Scholar] [CrossRef]
  33. De Vos, J.; Mathijs, I.; Xavier, C.; Massa, S.; Wernery, U.; Bouwens, L.; Lahoutte, T.; Muyldermans, S.; Devoogdt, N. Specific targeting of atherosclerotic plaques in ApoE−/− mice using a new camelid sdAb binding the vulnerable plaque marker LOX-1. Mol. Imaging Biol. 2014, 16, 690–698. [Google Scholar] [CrossRef] [PubMed]
  34. Xu, S.; Ogura, S.; Chen, J.; Little, P.J.; Moss, J.; Liu, P. LOX-1 in atherosclerosis: Biological functions and pharmacological modifiers. Cell. Mol. Life Sci. 2013, 70, 2859–2872. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Mäkinen, P.I.; Lappalainen, J.P.; Heinonen, S.E.; Leppänen, P.; Lähteenvuo, M.T.; Aarnio, J.V.; Heikkilä, J.; Turunen, M.P.; Ylä-Herttuala, S. Silencing of either SR-A or CD36 reduces atherosclerosis in hyperlipidaemic mice and reveals reciprocal upregulation of these receptors. Cardiovasc. Res. 2010, 88, 530–538. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Dai, X.-Y.; Cai, Y.; Mao, D.-D.; Qi, Y.-F.; Tang, C.; Xu, Q.; Zhu, Y.; Xu, M.-J.; Wang, X. Increased stability of phosphatase and tensin homolog by intermedin leading to scavenger receptor A inhibition of macrophages reduces atherosclerosis in apolipoprotein E-deficient mice. J. Mol. Cell. Cardiol. 2012, 53, 509–520. [Google Scholar] [CrossRef]
  37. Sakashita, N.; Miyazaki, A.; Takeya, M.; Horiuchi, S.; Chang, C.C.; Chang, T.-Y.; Takahashi, K. Localization of human acyl-coenzyme A: Cholesterol acyltransferase-1 (ACAT-1) in macrophages and in various tissues. Am. J. Pathol. 2000, 156, 227–236. [Google Scholar] [CrossRef] [Green Version]
  38. Sekiya, M.; Osuga, J.-I.; Igarashi, M.; Okazaki, H.; Ishibashi, S. The role of neutral cholesterol ester hydrolysis in macrophage foam cells. J. Atheroscler. Thromb. 2011, 18, 359–364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. McLaren, J.E.; Michael, D.R.; Ashlin, T.G.; Ramji, D.P. Cytokines, macrophage lipid metabolism and foam cells: Implications for cardiovascular disease therapy. Prog. Lipid Res. 2011, 50, 331–347. [Google Scholar] [CrossRef]
  40. Fazio, S.; Linton, M. Mouse models of hyperlipidemia and atherosclerosis. Front. Biosci 2001, 6, D515–D525. [Google Scholar] [CrossRef]
  41. Ikenoya, M.; Yoshinaka, Y.; Kobayashi, H.; Kawamine, K.; Shibuya, K.; Sato, F.; Sawanobori, K.; Watanabe, T.; Miyazaki, A. A selective ACAT-1 inhibitor, K-604, suppresses fatty streak lesions in fat-fed hamsters without affecting plasma cholesterol levels. Atherosclerosis 2007, 191, 290–297. [Google Scholar] [CrossRef]
  42. Perrey, S.; Legendre, C.; Matsuura, A.; Guffroy, C.; Binet, J.; Ohbayashi, S.; Tanaka, T.; Ortuno, J.C.; Matsukura, T.; Laugel, T. Preferential pharmacological inhibition of macrophage ACAT increases plaque formation in mouse and rabbit models of atherogenesis. Atherosclerosis 2001, 155, 359–370. [Google Scholar] [CrossRef]
  43. Sekiya, M.; Yamamuro, D.; Ohshiro, T.; Honda, A.; Takahashi, M.; Kumagai, M.; Sakai, K.; Nagashima, S.; Tomoda, H.; Igarashi, M. Absence of Nceh1 augments 25-hydroxycholesterol-induced ER stress and apoptosis in macrophages. J. Lipid Res. 2014, 55, 2082–2092. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Chistiakov, D.A.; Bobryshev, Y.V.; Orekhov, A.N. Macrophage-mediated cholesterol handling in atherosclerosis. J. Cell. Mol. Med. 2016, 20, 17–28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Chistiakov, D.A.; Melnichenko, A.A.; Myasoedova, V.A.; Grechko, A.V.; Orekhov, A.N. Mechanisms of foam cell formation in atherosclerosis. J. Mol. Med. 2017, 95, 1153–1165. [Google Scholar] [CrossRef]
  46. Silva, J.C.; César, F.A.; de Oliveira, E.M.; Turato, W.M.; Tripodi, G.L.; Castilho, G.; Machado-Lima, A.; de las Heras, B.; Boscá, L.; Rabello, M.M. New PPARγ partial agonist improves obesity-induced metabolic alterations and atherosclerosis in LDLr−/− mice. Pharmacol. Res. 2016, 104, 49–60. [Google Scholar] [CrossRef]
  47. Li, A.C.; Binder, C.J.; Gutierrez, A.; Brown, K.K.; Plotkin, C.R.; Pattison, J.W.; Valledor, A.F.; Davis, R.A.; Willson, T.M.; Witztum, J.L. Differential inhibition of macrophage foam-cell formation and atherosclerosis in mice by PPARα, β/δ, and γ. J. Clin. Investig. 2004, 114, 1564–1576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Phillips, M.C. Molecular mechanisms of cellular cholesterol efflux. J. Biol. Chem. 2014, 289, 24020–24029. [Google Scholar] [CrossRef] [Green Version]
  49. Joyce, C.W.; Wagner, E.M.; Basso, F.; Amar, M.J.; Freeman, L.A.; Shamburek, R.D.; Knapper, C.L.; Syed, J.; Wu, J.; Vaisman, B.L. ABCA1 overexpression in the liver of LDLr-KO mice leads to accumulation of pro-atherogenic lipoproteins and enhanced atherosclerosis. J. Biol. Chem. 2006, 281, 33053–33065. [Google Scholar] [CrossRef] [Green Version]
  50. Van Eck, M.; Bos, I.S.T.; Hildebrand, R.B.; Van Rij, B.T.; Van Berkel, T.J. Dual role for scavenger receptor class B, type I on bone marrow-derived cells in atherosclerotic lesion development. Am. J. Pathol. 2004, 165, 785–794. [Google Scholar] [CrossRef] [Green Version]
  51. Wang, H.; Yang, Y.; Chen, H.; Dan, J.; Cheng, J.; Guo, S.; Sun, X.; Wang, W.; Ai, Y.; Li, S. The predominant pathway of apoptosis in THP-1 macrophage-derived foam cells induced by 5-aminolevulinic acid-mediated sonodynamic therapy is the mitochondria-caspase pathway despite the participation of endoplasmic reticulum stress. Cell. Physiol. Biochem. 2014, 33, 1789–1801. [Google Scholar] [CrossRef]
  52. Tabas, I. Macrophage apoptosis in atherosclerosis: Consequences on plaque progression and the role of endoplasmic reticulum stress. Antioxid. Redox Signal. 2009, 11, 2333–2339. [Google Scholar] [CrossRef]
  53. Newby, A.C.; George, S.J.; Ismail, Y.; Johnson, J.L.; Sala-Newby, G.B.; Thomas, A.C. Vulnerable atherosclerotic plaque metalloproteinases and foam cell phenotypes. Thromb. Haemost. 2009, 101, 1006–1011. [Google Scholar]
  54. Tousoulis, D.; Oikonomou, E.; Economou, E.K.; Crea, F.; Kaski, J.C. Inflammatory cytokines in atherosclerosis: Current therapeutic approaches. Eur. Heart J. 2016, 37, 1723–1732. [Google Scholar] [CrossRef] [Green Version]
  55. Chistiakov, D.A.; Melnichenko, A.A.; Grechko, A.V.; Myasoedova, V.A.; Orekhov, A.N. Potential of anti-inflammatory agents for treatment of atherosclerosis. Exp. Mol. Pathol. 2018, 104, 114–124. [Google Scholar] [CrossRef]
  56. Tabas, I. Consequences and therapeutic implications of macrophage apoptosis in atherosclerosis: The importance of lesion stage and phagocytic efficiency. Arterioscler. Thromb. Vasc. Biol. 2005, 25, 2255–2264. [Google Scholar] [CrossRef]
  57. Liu, J.; Thewke, D.P.; Su, Y.R.; Linton, M.F.; Fazio, S.; Sinensky, M.S. Reduced macrophage apoptosis is associated with accelerated atherosclerosis in low-density lipoprotein receptor-null mice. Arterioscler. Thromb. Vasc. Biol. 2005, 25, 174–179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Thorp, E.; Li, Y.; Bao, L.; Yao, P.M.; Kuriakose, G.; Rong, J.; Fisher, E.A.; Tabas, I. Brief report: Increased apoptosis in advanced atherosclerotic lesions of Apoe−/− mice lacking macrophage Bcl-2. Arterioscler. Thromb. Vasc. Biol. 2009, 29, 169–172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Arai, S.; Shelton, J.M.; Chen, M.; Bradley, M.N.; Castrillo, A.; Bookout, A.L.; Mak, P.A.; Edwards, P.A.; Mangelsdorf, D.J.; Tontonoz, P. A role for the apoptosis inhibitory factor AIM/Spα/Api6 in atherosclerosis development. Cell Metab. 2005, 1, 201–213. [Google Scholar] [CrossRef] [Green Version]
  60. Grainger, D.J.; Reckless, J.; McKilligin, E. Apolipoprotein E modulates clearance of apoptotic bodies in vitro and in vivo, resulting in a systemic proinflammatory state in apolipoprotein E-deficient mice. J. Immunol. 2004, 173, 6366–6375. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Nagata, S. Apoptosis and clearance of apoptotic cells. Annu. Rev. Immunol. 2018, 36, 489–517. [Google Scholar] [CrossRef]
  62. Kratzer, A.; Buchebner, M.; Pfeifer, T.; Becker, T.M.; Uray, G.; Miyazaki, M.; Miyazaki-Anzai, S.; Ebner, B.; Chandak, P.G.; Kadam, R.S. Synthetic LXR agonist attenuates plaque formation in apoE−/− mice without inducing liver steatosis and hypertriglyceridemia. J. Lipid Res. 2009, 50, 312–326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Zahuczky, G.; Kristóf, E.; Majai, G.; Fésüs, L. Differentiation and glucocorticoid regulated apopto-phagocytic gene expression patterns in human macrophages. Role of Mertk in enhanced phagocytosis. PLoS ONE 2011, 6, e21349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Majai, G.; Sarang, Z.; Csomós, K.; Zahuczky, G.; Fésüs, L. PPARγ-dependent regulation of human macrophages in phagocytosis of apoptotic cells. Eur. J. Immunol. 2007, 37, 1343–1354. [Google Scholar] [CrossRef] [PubMed]
  65. Rosenblat, M.; Volkova, N.; Aviram, M. Pomegranate phytosterol (β-sitosterol) and polyphenolic antioxidant (punicalagin) addition to statin, significantly protected against macrophage foam cells formation. Atherosclerosis 2013, 226, 110–117. [Google Scholar] [CrossRef] [PubMed]
  66. Uitz, E.; Bahadori, B.; McCarty, M.F.; Moghadasian, M.H. Practical strategies for modulating foam cell formation and behavior. World J. Clin. Cases WJCC 2014, 2, 497. [Google Scholar] [CrossRef] [PubMed]
  67. Laufs, U.; Endres, M.; Custodis, F.; Gertz, K.; Nickenig, G.; Liao, J.K.; Böhm, M. Suppression of endothelial nitric oxide production after withdrawal of statin treatment is mediated by negative feedback regulation of rho GTPase gene transcription. Circulation 2000, 102, 3104–3110. [Google Scholar] [CrossRef] [Green Version]
  68. Afshari, A.R.; Mollazadeh, H.; Henney, N.C.; Jamialahmad, T.; Sahebkar, A. Effects of statins on brain tumors: A review. Semin. Cancer Biol. 2020. [Google Scholar] [CrossRef]
  69. Bagheri, H.; Ghasemi, F.; Barreto, G.E.; Sathyapalan, T.; Jamialahmadi, T.; Sahebkar, A. The effects of statins on microglial cells to protect against neurodegenerative disorders: A mechanistic review. BioFactors 2020, 46, 309–325. [Google Scholar] [CrossRef]
  70. Gorabi, A.M.; Kiaie, N.; Pirro, M.; Bianconi, V.; Jamialahmadi, T.; Sahebkar, A. Effects of statins on the biological features of mesenchymal stem cells and therapeutic implications. Heart Fail. Rev. 2020. [Google Scholar] [CrossRef]
  71. Kouhpeikar, H.; Delbari, Z.; Sathyapalan, T.; Simental-Mendía, L.E.; Jamialahmadi, T.; Sahebkar, A. The Effect of Statins through Mast Cells in the Pathophysiology of Atherosclerosis: A Review. Curr. Atheroscler. Rep. 2020, 22, 19. [Google Scholar] [CrossRef]
  72. Mollazadeh, H.; Tavana, E.; Fanni, G.; Bo, S.; Banach, M.; Pirro, M.; von Haehling, S.; Jamialahmadi, T.; Sahebkar, A. Effects of statins on mitochondrial pathways. J. Cachexia Sarcopenia Muscle 2021. [Google Scholar] [CrossRef] [PubMed]
  73. Sahebkar, A.; Serban, C.; Mikhailidis, D.P.; Undas, A.; Lip, G.Y.H.; Muntner, P.; Bittner, V.; Ray, K.K.; Watts, G.F.; Hovingh, G.K.; et al. Association between statin use and plasma d-dimer levels: A systematic review and meta-analysis of randomised controlled trials. Thromb. Haemost. 2015, 114, 546–557. [Google Scholar] [CrossRef]
  74. Sahebkar, A.; Serban, C.; Ursoniu, S.; Mikhailidis, D.P.; Undas, A.; Lip, G.Y.H.; Bittner, V.; Ray, K.K.; Watts, G.F.; Kees Hovingh, G.; et al. The impact of statin therapy on plasma levels of von Willebrand factor antigen: Systematic review and meta-analysis of Randomised placebo-controlled trials. Thromb. Haemost. 2016, 115, 520–532. [Google Scholar] [CrossRef] [PubMed]
  75. Serban, C.; Sahebkar, A.; Ursoniu, S.; Mikhailidis, D.P.; Rizzo, M.; Lip, G.Y.H.; Kees Hovingh, G.; Kastelein, J.J.P.; Kalinowski, L.; Rysz, J.; et al. A systematic review and meta-analysis of the effect of statins on plasma asymmetric dimethylarginine concentrations. Sci. Rep. 2015, 5, 1–8. [Google Scholar] [CrossRef] [Green Version]
  76. Thomas, A.C.; Eijgelaar, W.J.; Daemen, M.J.; Newby, A.C. Foam cell formation in vivo converts macrophages to a pro-fibrotic phenotype. PLoS ONE 2015, 10, e0128163. [Google Scholar] [CrossRef] [Green Version]
  77. Peeters, W.; Moll, F.L.; Vink, A.; van der Spek, P.J.; de Kleijn, D.P.; de Vries, J.-P.P.; Verheijen, J.H.; Newby, A.C.; Pasterkamp, G. Collagenase matrix metalloproteinase-8 expressed in atherosclerotic carotid plaques is associated with systemic cardiovascular outcome. Eur. Heart J. 2011, 32, 2314–2325. [Google Scholar] [CrossRef] [PubMed]
  78. Laxton, R.C.; Hu, Y.; Duchene, J.; Zhang, F.; Zhang, Z.; Leung, K.-Y.; Xiao, Q.; Scotland, R.S.; Hodgkinson, C.P.; Smith, K. A role of matrix metalloproteinase-8 in atherosclerosis. Circ. Res. 2009, 105, 921–929. [Google Scholar] [CrossRef]
  79. Luttun, A.; Lutgens, E.; Manderveld, A.; Maris, K.; Collen, D.; Carmeliet, P.; Moons, L. Loss of matrix metalloproteinase-9 or matrix metalloproteinase-12 protects apolipoprotein E-deficient mice against atherosclerotic media destruction but differentially affects plaque growth. Circulation 2004, 109, 1408–1414. [Google Scholar] [CrossRef] [Green Version]
  80. Johnson, J.L.; George, S.J.; Newby, A.C.; Jackson, C.L. Divergent effects of matrix metalloproteinases 3, 7, 9, and 12 on atherosclerotic plaque stability in mouse brachiocephalic arteries. Proc. Natl. Acad. Sci. USA 2005, 102, 15575–15580. [Google Scholar] [CrossRef] [Green Version]
  81. Daub, K.; Langer, H.; Seizer, P.; Stellos, K.; May, A.E.; Goyal, P.; Bigalke, B.; Schönberger, T.; Geisler, T.; Siegel-Axel, D. Platelets induce differentiation of human CD34+ progenitor cells into foam cells and endothelial cells. FASEB J. 2006, 20, 2559–2561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Feng, Y.; Schouteden, S.; Geenens, R.; Van Duppen, V.; Herijgers, P.; Holvoet, P.; Van Veldhoven, P.P.; Verfaillie, C.M. Hematopoietic stem/progenitor cell proliferation and differentiation is differentially regulated by high-density and low-density lipoproteins in mice. PLoS ONE 2012, 7, e47286. [Google Scholar] [CrossRef] [Green Version]
  83. Allahverdian, S.; Chehroudi, A.C.; McManus, B.M.; Abraham, T.; Francis, G.A. Contribution of intimal smooth muscle cells to cholesterol accumulation and macrophage-like cells in human atherosclerosis. Circulation 2014, 129, 1551–1559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Feil, S.; Fehrenbacher, B.; Lukowski, R.; Essmann, F.; Schulze-Osthoff, K.; Schaller, M.; Feil, R. Transdifferentiation of vascular smooth muscle cells to macrophage-like cells during atherogenesis. Circ. Res. 2014, 115, 662–667. [Google Scholar] [CrossRef]
  85. Yu, B.; Wong, M.M.; Potter, C.M.; Simpson, R.M.; Karamariti, E.; Zhang, Z.; Zeng, L.; Warren, D.; Hu, Y.; Wang, W. Vascular stem/progenitor cell migration induced by smooth muscle cell-derived chemokine (C-C Motif) ligand 2 and chemokine (C-X-C motif) ligand 1 contributes to neointima formation. Stem Cells 2016, 34, 2368–2380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Ivan, L.; Antohe, F. Hyperlipidemia induces endothelial-derived foam cells in culture. J. Recept. Signal. Transduct. 2010, 30, 106–114. [Google Scholar] [CrossRef]
  87. Xu, S.; Kamato, D.; Little, P.J.; Nakagawa, S.; Pelisek, J.; Jin, Z.G. Targeting epigenetics and non-coding RNAs in atherosclerosis: From mechanisms to therapeutics. Pharmacol. Ther. 2019, 196, 15–43. [Google Scholar] [CrossRef]
  88. Uszczynska-Ratajczak, B.; Lagarde, J.; Frankish, A.; Guigó, R.; Johnson, R. Towards a complete map of the human long non-coding RNA transcriptome. Nat. Rev. Genet. 2018, 19, 535–548. [Google Scholar] [CrossRef]
  89. Qureshi, I.A.; Mehler, M.F. Emerging roles of non-coding RNAs in brain evolution, development, plasticity and disease. Nat. Rev. Neurosci. 2012, 13, 528–541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Fatica, A.; Bozzoni, I. Long non-coding RNAs: New players in cell differentiation and development. Nat. Rev. Genet. 2014, 15, 7–21. [Google Scholar] [CrossRef] [PubMed]
  91. Engels, B.M.; Hutvagner, G. Principles and effects of microRNA-mediated post-transcriptional gene regulation. Oncogene 2006, 25, 6163–6169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Beermann, J.; Piccoli, M.-T.; Viereck, J.; Thum, T. Non-coding RNAs in development and disease: Background, mechanisms, and therapeutic approaches. Physiol. Rev. 2016, 96, 1297–1325. [Google Scholar] [CrossRef] [Green Version]
  93. Cech, T.R.; Steitz, J.A. The noncoding RNA revolution—Trashing old rules to forge new ones. Cell 2014, 157, 77–94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Bartel, D.P. MicroRNAs: Target recognition and regulatory functions. Cell 2009, 136, 215–233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Lee, Y.; Kim, M.; Han, J.; Yeom, K.H.; Lee, S.; Baek, S.H.; Kim, V.N. MicroRNA genes are transcribed by RNA polymerase II. EMBO J. 2004, 23, 4051–4060. [Google Scholar] [CrossRef]
  96. Cai, X.; Hagedorn, C.H.; Cullen, B.R. Human microRNAs are processed from capped, polyadenylated transcripts that can also function as mRNAs. RNA 2004, 10, 1957–1966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Ruby, J.G.; Jan, C.H.; Bartel, D.P. Intronic microRNA precursors that bypass Drosha processing. Nature 2007, 448, 83–86. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Derrien, T.; Johnson, R.; Bussotti, G.; Tanzer, A.; Djebali, S.; Tilgner, H.; Guernec, G.; Martin, D.; Merkel, A.; Knowles, D.G. The GENCODE v7 catalog of human long noncoding RNAs: Analysis of their gene structure, evolution, and expression. Genome Res. 2012, 22, 1775–1789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Ulitsky, I.; Bartel, D.P. lincRNAs: Genomics, evolution, and mechanisms. Cell 2013, 154, 26–46. [Google Scholar] [CrossRef] [Green Version]
  100. Batista, P.J.; Chang, H.Y. Long noncoding RNAs: Cellular address codes in development and disease. Cell 2013, 152, 1298–1307. [Google Scholar] [CrossRef] [Green Version]
  101. Gong, C.; Maquat, L.E. lncRNAs transactivate STAU1-mediated mRNA decay by duplexing with 3′ UTRs via Alu elements. Nature 2011, 470, 284–288. [Google Scholar] [CrossRef] [Green Version]
  102. Magny, E.G.; Pueyo, J.I.; Pearl, F.M.; Cespedes, M.A.; Niven, J.E.; Bishop, S.A.; Couso, J.P. Conserved regulation of cardiac calcium uptake by peptides encoded in small open reading frames. Science 2013, 341, 1116–1120. [Google Scholar] [CrossRef]
  103. Hu, Y.W.; Hu, Y.R.; Zhao, J.Y.; Li, S.F.; Ma, X.; Wu, S.G.; Lu, J.B.; Qiu, Y.R.; Sha, Y.H.; Wang, Y.C.; et al. An agomir of miR-144-3p accelerates plaque formation through impairing reverse cholesterol transport and promoting pro-inflammatory cytokine production. PLoS ONE 2014, 9, e94997. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Marquart, T.J.; Wu, J.; Lusis, A.J.; Baldán, Á. Anti-miR-33 therapy does not alter the progression of atherosclerosis in low-density lipoprotein receptor-deficient mice. Arterioscler. Thromb. Vasc. Biol. 2013, 33, 455–458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Rotllan, N.; Ramírez, C.M.; Aryal, B.; Esau, C.C.; Fernández-Hernando, C. Therapeutic silencing of microRNA-33 inhibits the progression of atherosclerosis in Ldlr−/− mice—Brief report. Arterioscler. Thromb. Vasc. Biol. 2013, 33, 1973–1977. [Google Scholar] [CrossRef] [Green Version]
  106. Lv, Y.-C.; Tang, Y.-Y.; Peng, J.; Zhao, G.-J.; Yang, J.; Yao, F.; Ouyang, X.-P.; He, P.-P.; Xie, W.; Tan, Y.-L. MicroRNA-19b promotes macrophage cholesterol accumulation and aortic atherosclerosis by targeting ATP-binding cassette transporter A1. Atherosclerosis 2014, 236, 215–226. [Google Scholar] [CrossRef] [PubMed]
  107. Zhang, N.; Lei, J.; Lei, H.; Ruan, X.; Liu, Q.; Chen, Y.; Huang, W. MicroRNA-101 overexpression by IL-6 and TNF-α inhibits cholesterol efflux by suppressing ATP-binding cassette transporter A1 expression. Exp. Cell Res. 2015, 336, 33–42. [Google Scholar] [CrossRef] [PubMed]
  108. Wang, D.; Yan, X.; Xia, M.; Yang, Y.; Li, D.; Li, X.; Song, F.; Ling, W. Coenzyme Q10 promotes macrophage cholesterol efflux by regulation of the activator protein-1/miR-378/ATP-binding cassette transporter G1–signaling pathway. Arterioscler. Thromb. Vasc. Biol. 2014, 34, 1860–1870. [Google Scholar] [CrossRef] [Green Version]
  109. Meiler, S.; Baumer, Y.; Toulmin, E.; Seng, K.; Boisvert, W.A. MicroRNA 302a is a novel modulator of cholesterol homeostasis and atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 2015, 35, 323–331. [Google Scholar] [CrossRef] [Green Version]
  110. Zhang, M.; Wu, J.-F.; Chen, W.-J.; Tang, S.-L.; Mo, Z.-C.; Tang, Y.-Y.; Li, Y.; Wang, J.-L.; Liu, X.-Y.; Peng, J. MicroRNA-27a/b regulates cellular cholesterol efflux, influx and esterification/hydrolysis in THP-1 macrophages. Atherosclerosis 2014, 234, 54–64. [Google Scholar] [CrossRef] [PubMed]
  111. Sun, D.; Zhang, J.; Xie, J.; Wei, W.; Chen, M.; Zhao, X. MiR-26 controls LXR-dependent cholesterol efflux by targeting ABCA1 and ARL7. FEBS Lett. 2012, 586, 1472–1479. [Google Scholar] [CrossRef] [Green Version]
  112. Lan, G.; Xie, W.; Li, L.; Zhang, M.; Liu, D.; Tan, Y.-L.; Cheng, H.-P.; Gong, D.; Huang, C.; Zheng, X.-L. MicroRNA-134 actives lipoprotein lipase-mediated lipid accumulation and inflammatory response by targeting angiopoietin-like 4 in THP-1 macrophages. Biochem. Biophys. Res. Commun. 2016, 472, 410–417. [Google Scholar] [CrossRef]
  113. Tian, F.-J.; An, L.-N.; Wang, G.-K.; Zhu, J.-Q.; Li, Q.; Zhang, Y.-Y.; Zeng, A.; Zou, J.; Zhu, R.-F.; Han, X.-S. Elevated microRNA-155 promotes foam cell formation by targeting HBP1 in atherogenesis. Cardiovasc. Res. 2014, 103, 100–110. [Google Scholar] [CrossRef]
  114. Gong, D.; Cheng, H.-P.; Xie, W.; Zhang, M.; Liu, D.; Lan, G.; Huang, C.; Zhao, Z.-W.; Chen, L.-Y.; Yao, F. Cystathionine γ-lyase (CSE)/hydrogen sulfide system is regulated by miR-216a and influences cholesterol efflux in macrophages via the PI3K/AKT/ABCA1 pathway. Biochem. Biophys. Res. Commun. 2016, 470, 107–116. [Google Scholar] [CrossRef] [PubMed]
  115. Liu, D.; Zhang, M.; Xie, W.; Lan, G.; Cheng, H.-P.; Gong, D.; Huang, C.; Lv, Y.-C.; Yao, F.; Tan, Y.-L. MiR-486 regulates cholesterol efflux by targeting HAT1. Biochem. Biophys. Res. Commun. 2016, 472, 418–424. [Google Scholar] [CrossRef] [PubMed]
  116. Miao, H.; Zeng, H.; Gong, H. microRNA-212 promotes lipid accumulation and attenuates cholesterol efflux in THP-1 human macrophages by targeting SIRT1. Gene 2018, 643, 55–60. [Google Scholar] [CrossRef]
  117. Chen, H.; Li, X.; Liu, S.; Gu, L.; Zhou, X. MircroRNA-19a promotes vascular inflammation and foam cell formation by targeting HBP-1 in atherogenesis. Sci. Rep. 2017, 7, 1–10. [Google Scholar] [CrossRef] [PubMed]
  118. Cui, J.; Ren, Z.; Zou, W.; Jiang, Y. miR-497 accelerates oxidized low-density lipoprotein-induced lipid accumulation in macrophages by repressing the expression of apelin. Cell Biol. Int. 2017, 41, 1012–1019. [Google Scholar] [CrossRef] [PubMed]
  119. Liang, B.; Wang, X.; Song, X.; Bai, R.; Yang, H.; Yang, Z.; Xiao, C.; Bian, Y. MicroRNA-20a/b regulates cholesterol efflux through post-transcriptional repression of ATP-binding cassette transporter A1. Biochim. Biophys. Acta BBA Mol. Cell Biol. Lipids 2017, 1862, 929–938. [Google Scholar] [CrossRef]
  120. Ramirez, C.; Dávalos, A.; Goedeke, L.; Salerno, A.; Warrier, N.; Cirera-Salinas, D.; Suárez, Y.; Fernández-Hernando, C. miR-758 regulates cholesterol efflux through post-transcriptional repression of ABCA1. Arter. Thromb Vasc. Biol. 2011, 31, 2707–2714. [Google Scholar] [CrossRef] [Green Version]
  121. Huangfu, N.; Xu, Z.; Zheng, W.; Wang, Y.; Cheng, J.; Chen, X. LncRNA MALAT1 regulates oxLDL-induced CD36 expression via activating β-catenin. Biochem. Biophys. Res. Commun. 2018, 495, 2111–2117. [Google Scholar] [CrossRef]
  122. Cai, C.; Zhu, H.; Ning, X.; Li, L.; Yang, B.; Chen, S.; Wang, L.; Lu, X.; Gu, D. LncRNA ENST00000602558. 1 regulates ABCG1 expression and cholesterol efflux from vascular smooth muscle cells through a p65-dependent pathway. Atherosclerosis 2019, 285, 31–39. [Google Scholar] [CrossRef] [PubMed]
  123. Hu, X.; Ma, R.; Fu, W.; Zhang, C.; Du, X. LncRNA UCA1 sponges miR-206 to exacerbate oxidative stress and apoptosis induced by ox-LDL in human macrophages. J. Cell. Physiol. 2019, 234, 14154–14160. [Google Scholar] [CrossRef] [PubMed]
  124. Li, L.; Xu, W.; Fu, X.; Huang, Y.; Wen, Y.; Xu, Q.; He, X.; Wang, K.; Huang, S.; Lv, Z. Blood miR-1275 is associated with risk of ischemic stroke and inhibits macrophage foam cell formation by targeting ApoC2 gene. Gene 2020, 731, 144364. [Google Scholar] [CrossRef] [PubMed]
  125. Li, L.; Wu, F.; Xie, Y.; Xu, W.; Xiong, G.; Xu, Y.; Huang, S.; Wu, Y.; Jiang, X. MiR-202-3p Inhibits Foam Cell Formation and is Associated with Coronary Heart Disease Risk in a Chinese Population. Int. Heart J. 2020, 61, 153–159. [Google Scholar] [CrossRef] [Green Version]
  126. Du, X.J.; Lu, J.M.; Sha, Y. MiR-181a inhibits vascular inflammation induced by ox-LDL via targeting TLR4 in human macrophages. J. Cell. Physiol. 2018, 233, 6996–7003. [Google Scholar] [CrossRef]
  127. Du, X.J.; Lu, J.M. MiR-135a represses oxidative stress and vascular inflammatory events via targeting toll-like receptor 4 in atherogenesis. J. Cell. Biochem. 2018, 119, 6154–6161. [Google Scholar] [CrossRef]
  128. Xu, J.; Hu, G.; Lu, M.; Xiong, Y.; Li, Q.; Chang, C.C.; Song, B.; Chang, T.; Li, B. MiR-9 reduces human acyl-coenzyme A: Cholesterol acyltransferase-1 to decrease THP-1 macrophage-derived foam cell formation. Acta Biochim. Biophys. Sin. 2013, 45, 953–962. [Google Scholar] [CrossRef] [Green Version]
  129. Canfrán-Duque, A.; Rotllan, N.; Zhang, X.; Fernández-Fuertes, M.; Ramírez-Hidalgo, C.; Araldi, E.; Daimiel, L.; Busto, R.; Fernández-Hernando, C.; Suárez, Y. Macrophage deficiency of miR-21 promotes apoptosis, plaque necrosis, and vascular inflammation during atherogenesis. EMBO Mol. Med. 2017, 9, 1244–1262. [Google Scholar] [CrossRef]
  130. Li, J.; Zhang, S. microRNA-150 inhibits the formation of macrophage foam cells through targeting adiponectin receptor 2. Biochem. Biophys. Res. Commun. 2016, 476, 218–224. [Google Scholar] [CrossRef]
  131. Peng, X.-P.; Huang, L.; Liu, Z.-H. miRNA-133a attenuates lipid accumulation via TR4-CD36 pathway in macrophages. Biochimie 2016, 127, 79–85. [Google Scholar] [CrossRef]
  132. Zhang, F.; Zhao, J.; Sun, D.; Wei, N. MiR-155 inhibits transformation of macrophages into foam cells via regulating CEH expression. Biomed. Pharmacother. 2018, 104, 645–651. [Google Scholar] [CrossRef]
  133. Li, X.; Kong, D.; Chen, H.; Liu, S.; Hu, H.; Wu, T.; Wang, J.; Chen, W.; Ning, Y.; Li, Y. miR-155 acts as an anti-inflammatory factor in atherosclerosis-associated foam cell formation by repressing calcium-regulated heat stable protein 1. Sci. Rep. 2016, 6, 21789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Dai, Y.; Wu, X.; Dai, D.; Li, J.; Mehta, J.L. MicroRNA-98 regulates foam cell formation and lipid accumulation through repression of LOX-1. Redox Biol. 2018, 16, 255–262. [Google Scholar] [CrossRef] [PubMed]
  135. Wang, J.; Bai, X.; Song, Q.; Fan, F.; Hu, Z.; Cheng, G.; Zhang, Y. miR-223 inhibits lipid deposition and inflammation by suppressing toll-like receptor 4 signaling in macrophages. Int. J. Mol. Sci. 2015, 16, 24965–24982. [Google Scholar] [CrossRef] [PubMed]
  136. Qiao, Y.; Wang, C.; Kou, J.; Han, D.; Huo, D.; Li, F.; Zhou, X.; Meng, D.; Xu, J.; Murtaza, G. MicroRNA-23a suppresses the apoptosis of inflammatory macrophages and foam cells in atherogenesis by targeting HSP90. Gene 2020, 729, 144319. [Google Scholar] [CrossRef] [PubMed]
  137. Lin, N.; An, Y. Blockade of 146b-5p promotes inflammation in atherosclerosis-associated foam cell formation by targeting TRAF6. Exp. Ther. Med. 2017, 14, 5087–5092. [Google Scholar] [CrossRef]
  138. Zhuang, X.; Li, R.; Maimaitijiang, A.; Liu, R.; Yan, F.; Hu, H.; Gao, X.; Shi, H. miR-221-3p inhibits oxidized low-density lipoprotein induced oxidative stress and apoptosis via targeting a disintegrin and metalloprotease-22. J. Cell. Biochem. 2019, 120, 6304–6314. [Google Scholar] [CrossRef]
  139. Liang, X.; Xu, Z.; Yuan, M.; Zhang, Y.; Zhao, B.; Wang, J.; Zhang, A.; Li, G. MicroRNA-16 suppresses the activation of inflammatory macrophages in atherosclerosis by targeting PDCD4. Int. J. Mol. Med. 2016, 37, 967–975. [Google Scholar] [CrossRef] [Green Version]
  140. Zhao, Q.; Li, S.; Li, N.; Yang, X.; Ma, S.; Yang, A.; Zhang, H.; Yang, S.; Mao, C.; Xu, L. miR-34a targets HDAC1-regulated H3K9 acetylation on lipid accumulation induced by homocysteine in foam cells. J. Cell. Biochem. 2017, 118, 4617–4627. [Google Scholar] [CrossRef]
  141. Wang, Y.-S.; Hsi, E.; Cheng, H.-Y.; Hsu, S.-H.; Liao, Y.-C.; Juo, S.-H.H. Let-7g suppresses both canonical and non-canonical NF-κB pathways in macrophages leading to anti-atherosclerosis. Oncotarget 2017, 8, 101026. [Google Scholar] [CrossRef] [Green Version]
  142. Li, H.; Han, S.; Sun, Q.; Yao, Y.; Li, S.; Yuan, C.; Zhang, B.; Jing, B.; Wu, J.; Song, Y. Long non-coding RNA CDKN2B-AS1 reduces inflammatory response and promotes cholesterol efflux in atherosclerosis by inhibiting ADAM10 expression. Aging (Albany N. Y.) 2019, 11, 1695. [Google Scholar] [CrossRef]
  143. Li, Y.; Sun, T.; Shen, S.; Wang, L.; Yan, J. LncRNA DYNLRB2-2 inhibits THP-1 macrophage foam cell formation by enhancing autophagy. Biol. Chem. 2019, 400, 1047–1057. [Google Scholar] [CrossRef] [PubMed]
  144. Oram, J.F.; Vaughan, A.M. ABCA1-mediated transport of cellular cholesterol and phospholipids to HDL apolipoproteins. Curr. Opin. Lipidol. 2000, 11, 253–260. [Google Scholar] [CrossRef] [PubMed]
  145. Attie, A.D. ABCA1: At the nexus of cholesterol, HDL and atherosclerosis. Trends Biochem. Sci. 2007, 32, 172–179. [Google Scholar] [CrossRef] [PubMed]
  146. Su, Y.R.; Dove, D.E.; Major, A.S.; Hasty, A.H.; Boone, B.; Linton, M.F.; Fazio, S. Reduced ABCA1-mediated cholesterol efflux and accelerated atherosclerosis in Apolipoprotein E–deficient mice lacking macrophage-derived ACAT1. Circulation 2005, 111, 2373–2381. [Google Scholar] [CrossRef] [Green Version]
  147. Ragozin, S.; Niemeier, A.; Laatsch, A.; Loeffler, B.; Merkel, M.; Beisiegel, U.; Heeren, J. Knockdown of hepatic ABCA1 by RNA interference decreases plasma HDL cholesterol levels and influences postprandial lipemia in mice. Arterioscler. Thromb. Vasc. Biol. 2005, 25, 1433–1438. [Google Scholar] [CrossRef] [Green Version]
  148. Liu, Y.; Tang, C. Regulation of ABCA1 functions by signaling pathways. Biochim. Biophys. Acta BBA Mol. Cell Biol. Lipids 2012, 1821, 522–529. [Google Scholar] [CrossRef] [Green Version]
  149. Yin, K.; Liao, D.-F.; Tang, C.-K. ATP-binding membrane cassette transporter A1 (ABCA1): A possible link between inflammation and reverse cholesterol transport. Mol. Med. 2010, 16, 438–449. [Google Scholar] [CrossRef] [Green Version]
  150. Yvan-Charvet, L.; Ranalletta, M.; Wang, N.; Han, S.; Terasaka, N.; Li, R.; Welch, C.; Tall, A.R. Combined deficiency of ABCA1 and ABCG1 promotes foam cell accumulation and accelerates atherosclerosis in mice. J. Clin. Investig. 2007, 117, 3900–3908. [Google Scholar] [CrossRef] [Green Version]
  151. Goedeke, L.; Vales-Lara, F.M.; Fenstermaker, M.; Cirera-Salinas, D.; Chamorro-Jorganes, A.; Ramírez, C.M.; Mattison, J.A.; de Cabo, R.; Suárez, Y.; Fernández-Hernando, C. A regulatory role for microRNA 33* in controlling lipid metabolism gene expression. Mol. Cell. Biol. 2013, 33, 2339–2352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Rayner, K.J.; Sheedy, F.J.; Esau, C.C.; Hussain, F.N.; Temel, R.E.; Parathath, S.; van Gils, J.M.; Rayner, A.J.; Chang, A.N.; Suarez, Y.; et al. Antagonism of miR-33 in mice promotes reverse cholesterol transport and regression of atherosclerosis. J. Clin. Investig. 2011, 121, 2921–2931. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Baba, O. MicroRNA-33 Deficiency Reduces the Progression of Atherosclerotic Plaque in ApoE−/− Mice. Ph.D. Thesis, Kyoto University, Kyoto, Japan, 2014. [Google Scholar]
  154. Zadelaar, S.; Kleemann, R.; Verschuren, L.; de Vries-Van der Weij, J.; van der Hoorn, J.; Princen, H.M.; Kooistra, T. Mouse models for atherosclerosis and pharmaceutical modifiers. Arter. Thromb. Vasc. Biol. 2007, 27, 1706–1721. [Google Scholar] [CrossRef] [Green Version]
  155. Price, N.L.; Rotllan, N.; Canfrán-Duque, A.; Zhang, X.; Pati, P.; Arias, N.; Moen, J.; Mayr, M.; Ford, D.A.; Baldán, Á.; et al. Genetic dissection of the impact of miR-33a and miR-33b during the progression of atherosclerosis. Cell Rep. 2017, 21, 1317–1330. [Google Scholar] [CrossRef] [Green Version]
  156. Price, N.L.; Singh, A.K.; Rotllan, N.; Goedeke, L.; Wing, A.; Canfran-Duque, A.; Diaz-Ruiz, A.; Araldi, E.; Baldan, A.; Camporez, J.P.; et al. Genetic Ablation of miR-33 Increases Food Intake, Enhances Adipose Tissue Expansion, and Promotes Obesity and Insulin Resistance. Cell Rep. 2018, 22, 2133–2145. [Google Scholar] [CrossRef] [Green Version]
  157. Chen, W.-J.; Yin, K.; Zhao, G.-J.; Fu, Y.-C.; Tang, C.-K. The magic and mystery of microRNA-27 in atherosclerosis. Atherosclerosis 2012, 222, 314–323. [Google Scholar] [CrossRef]
  158. Vickers, K.C.; Shoucri, B.M.; Levin, M.G.; Wu, H.; Pearson, D.S.; Osei-Hwedieh, D.; Collins, F.S.; Remaley, A.T.; Sethupathy, P. MicroRNA-27b is a regulatory hub in lipid metabolism and is altered in dyslipidemia. Hepatology 2013, 57, 533–542. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Shirasaki, T.; Honda, M.; Shimakami, T.; Horii, R.; Yamashita, T.; Sakai, Y.; Sakai, A.; Okada, H.; Watanabe, R.; Murakami, S. MicroRNA-27a regulates lipid metabolism and inhibits hepatitis C virus replication in human hepatoma cells. J. Virol. 2013, 87, 5270–5286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Cheung, S.H.; Kwok, W.K.; To, K.F.; Lau, J.Y.W. Anti-atherogenic effect of hydrogen sulfide by over-expression of cystathionine gamma-lyase (CSE) gene. PLoS ONE 2014, 9, e113038. [Google Scholar] [CrossRef] [Green Version]
  161. Greco, S.; Fasanaro, P.; Castelvecchio, S.; D’Alessandra, Y.; Arcelli, D.; Di Donato, M.; Malavazos, A.; Capogrossi, M.C.; Menicanti, L.; Martelli, F. MicroRNA dysregulation in diabetic ischemic heart failure patients. Diabetes 2012, 61, 1633–1641. [Google Scholar] [CrossRef] [Green Version]
  162. Mani, S.; Li, H.; Untereiner, A.; Wu, L.; Yang, G.; Austin, R.C.; Dickhout, J.G.; Lhoták, Š.; Meng, Q.H.; Wang, R. Decreased endogenous production of hydrogen sulfide accelerates atherosclerosis. Circulation 2013, 127, 2523–2534. [Google Scholar] [CrossRef] [Green Version]
  163. Peh, M.T.; Anwar, A.B.; Ng, D.S.; Atan, M.S.B.M.; Kumar, S.D.; Moore, P.K. Effect of feeding a high fat diet on hydrogen sulfide (H2S) metabolism in the mouse. Nitric Oxide 2014, 41, 138–145. [Google Scholar] [CrossRef]
  164. Guo, Z.; Li, C.S.; Wang, C.M.; Xie, Y.J.; Wang, A.L. CSE/H2S system protects mesenchymal stem cells from hypoxia and serum deprivation-induced apoptosis via mitochondrial injury, endoplasmic reticulum stress and PI3K/Akt activation pathways. Mol. Med. Rep. 2015, 12, 2128–2134. [Google Scholar] [CrossRef] [PubMed]
  165. Waki, H.; Nakamura, M.; Yamauchi, T.; Wakabayashi, K.-I.; Yu, J.; Hirose-Yotsuya, L.; Take, K.; Sun, W.; Iwabu, M.; Okada-Iwabu, M. Global mapping of cell type-specific open chromatin by FAIRE-seq reveals the regulatory role of the NFI family in adipocyte differentiation. PLoS Genet. 2011, 7, e1002311. [Google Scholar] [CrossRef] [PubMed]
  166. Chen, L.-J.; Lim, S.H.; Yeh, Y.-T.; Lien, S.-C.; Chiu, J.-J. Roles of microRNAs in atherosclerosis and restenosis. J. Biomed. Sci. 2012, 19, 79. [Google Scholar] [CrossRef] [Green Version]
  167. Nazari-Jahantigh, M.; Wei, Y.; Noels, H.; Akhtar, S.; Zhou, Z.; Koenen, R.R.; Heyll, K.; Gremse, F.; Kiessling, F.; Grommes, J. MicroRNA-155 promotes atherosclerosis by repressing Bcl6 in macrophages. J. Clin. Investig. 2012, 122, 4190–4202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Rodriguez, A.; Vigorito, E.; Clare, S.; Warren, M.V.; Couttet, P.; Soond, D.R.; Van Dongen, S.; Grocock, R.J.; Das, P.P.; Miska, E.A. Requirement of bic/microRNA-155 for normal immune function. Science 2007, 316, 608–611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Celermajer, D.S.; Chow, C.K.; Marijon, E.; Anstey, N.M.; Woo, K.S. Cardiovascular disease in the developing world: Prevalences, patterns, and the potential of early disease detection. J. Am. Coll. Cardiol. 2012, 60, 1207–1216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The mechanisms and molecules involved in cholesterol metabolism and homeostasis and the regulatory effects of Non-encoding RNAs are depicted in a foam cell. Cholesterol metabolism involves cholesterol influx, cholesterol esterification and cholesterol efflux, which ultimately leads to cholesterol homeostasis in the macrophage. Scavenger receptors SRA-1, CD36 and LOX1 are involved in oxLDL uptake. The gene expression of these receptors is downregulated in part A (see above); that is, non-encoding RNAs that have anti-atherosclerotic properties and are upregulated by part B ncRNAs, which are atherogenic and lead to lipid accumulation in macrophages. Cholesterol efflux is a pathway that transports excess free cholesterol from the cell, mainly via ABCA1, ABCG1, as well as SR-B1 transporters, leading to the formation of HDL and apoa1. The ncRNAs that reduce the formation of foam cells by increasing cholesterol efflux are specified in part C, as well as ncRNAs that lead to induction of the formation of foam cells by reducing cholesterol efflux, which are specified in part D. In cholesterol esterification, ACAT-1 and neutral cholesterol ester hydrolase (NCEH) play a key role in catalyzing the esterification of cholesterol and removing free cholesterol from foam cells, respectively. ACAT1 re-esterifies excess FC to promote the biosynthesis of CE that is stored in lipid droplets. In part E, microRNAs have been listed that reduce cholesterol esterification and increase the accumulation of free cholesterol, which ultimately promotes foam cell formation. Reducing foam cell apoptosis can stabilize atherosclerotic plaque and prevent the progression and worsening of atherosclerosis. ncRNAs that reduce foam cell apoptosis are shown in part F. Ijms 22 02529 i001 = decrease, Ijms 22 02529 i002 = increase.
Figure 1. The mechanisms and molecules involved in cholesterol metabolism and homeostasis and the regulatory effects of Non-encoding RNAs are depicted in a foam cell. Cholesterol metabolism involves cholesterol influx, cholesterol esterification and cholesterol efflux, which ultimately leads to cholesterol homeostasis in the macrophage. Scavenger receptors SRA-1, CD36 and LOX1 are involved in oxLDL uptake. The gene expression of these receptors is downregulated in part A (see above); that is, non-encoding RNAs that have anti-atherosclerotic properties and are upregulated by part B ncRNAs, which are atherogenic and lead to lipid accumulation in macrophages. Cholesterol efflux is a pathway that transports excess free cholesterol from the cell, mainly via ABCA1, ABCG1, as well as SR-B1 transporters, leading to the formation of HDL and apoa1. The ncRNAs that reduce the formation of foam cells by increasing cholesterol efflux are specified in part C, as well as ncRNAs that lead to induction of the formation of foam cells by reducing cholesterol efflux, which are specified in part D. In cholesterol esterification, ACAT-1 and neutral cholesterol ester hydrolase (NCEH) play a key role in catalyzing the esterification of cholesterol and removing free cholesterol from foam cells, respectively. ACAT1 re-esterifies excess FC to promote the biosynthesis of CE that is stored in lipid droplets. In part E, microRNAs have been listed that reduce cholesterol esterification and increase the accumulation of free cholesterol, which ultimately promotes foam cell formation. Reducing foam cell apoptosis can stabilize atherosclerotic plaque and prevent the progression and worsening of atherosclerosis. ncRNAs that reduce foam cell apoptosis are shown in part F. Ijms 22 02529 i001 = decrease, Ijms 22 02529 i002 = increase.
Ijms 22 02529 g001
Figure 2. Direct and indirect effects of different Non-coding RNAs on the expression of ABCA1, which has a large 3′UTR region and is one of the most important transporters in cholesterol efflux [144,145]. In direct mechanism: non-coding RNAs able to bind to 3′UTR of ABCA1 mRNA transcript which ABCA1 expression regulate, in indirect mechanism: non-coding RNAs were not able to bind to 3′UTR of ABCA1 mRNA transcript although these noncoding RNAs that bind to 3′UTR of other mRNA transcript in turn regulate ABCA1 expression. It has also been shown that reducing inflammation increases the expression of this transporter and has a negative role in the formation of foam cells by increasing cholesterol efflux [148,149].
Figure 2. Direct and indirect effects of different Non-coding RNAs on the expression of ABCA1, which has a large 3′UTR region and is one of the most important transporters in cholesterol efflux [144,145]. In direct mechanism: non-coding RNAs able to bind to 3′UTR of ABCA1 mRNA transcript which ABCA1 expression regulate, in indirect mechanism: non-coding RNAs were not able to bind to 3′UTR of ABCA1 mRNA transcript although these noncoding RNAs that bind to 3′UTR of other mRNA transcript in turn regulate ABCA1 expression. It has also been shown that reducing inflammation increases the expression of this transporter and has a negative role in the formation of foam cells by increasing cholesterol efflux [148,149].
Ijms 22 02529 g002
Table 1. Non-coding RNAs that stimulate foam cell formation/function.
Table 1. Non-coding RNAs that stimulate foam cell formation/function.
Non-Coding RNAsExpression TypeTarget GeneGenetic Information
(Human)
Experimental ModelEffect on Foam Cell FormationRegulation in Lipid MetabolismReference
In VivoIn Vitro
miR-144-3p Ijms 22 02529 i003ABCA1LXR and FXR control miR-144 expression, with both causing an upregulation in this miRNAApoE−/− micehuman THP-1 macrophage-derived foam cells(+)
Decrease
cholesterol efflux
to both apoAI and HDL
decreased HDL-C circulation and impaired RCT in vivo
144-3p mimics (agomir) increases the expression of inflammatory factors such as IL-1b, IL-6 and TNF-α in vivo and in vitro
[103]
miR-33(a/b) Ijms 22 02529 i003ABCA1,
NPC1
ABCG1
encoded within intron 16 of SREBF2, a gene that encodes a key transcriptional regulator of cholesterol uptake and synthesismiR-33−/− ApoE−/− miceTHP-1 macrophage-derived foam cells(+)
Decrease
cholesterol efflux
miR-33 coordinates cholesterol homeostasis[104,105]
miR-19b Ijms 22 02529 i003ABCA1located on chromosome 13q31.3
By an unknown mechanism, expression was increased compared with the control group in advanced human atherosclerotic plaques obtained from patients during endarterectomy
ApoE−/− micehuman THP-1 macrophage-derived foam cells(+)
Decrease
the efflux of cholesterol to apoAI
Decrease the levels of HDL (inhibitory role in RCT)[106]
miR-101 Ijms 22 02529 i003ABCA1located on chromosome
1p31.3
-human THP-1-derived macrophages(+)
Decrease
the efflux of cholesterol under inflammatory conditions
regulates the availability of free cholesterol for cellular efflux by inhibiting autophagy[107]
miR-378 Ijms 22 02529 i003ABCG1The level of miR-378 in the aorta is elevated during the progression of atherosclerosis in apoE−/− miceApoE−/− miceox-LDL-treated human THP-1 macrophages(+)
Decrease cholesterol efflux
activator protein-1/miRNA-378/ABCG1 is a novel cascade for CoQ10 in facilitating macrophage cholesterol efflux in vitro and in vivo[108]
miR-302a Ijms 22 02529 i003ABCA1located on chromosome 4q25Ldlr−/− miceprimary human and murine macrophages(+)
Decreases efflux of cholesterol
inhibiting miR-302a in vivo increases ABCA1 in aorta and liver of Ldlr−/− mice
and increases circulating HDL
[109]
miR-27 Ijms 22 02529 i003LPL, ACAT1, ABCA1
CD36
located on chromosome
9q22
-ox-LDL-treated THP-1 macrophages(+)
Reduced cholesterol efflux
reduces cholesteryl ester formation
blocks lipid uptake
regulates the ratio of cellular free cholesterol (FC) and cholesterol ester (CE) in THP-1 macrophages[110]
miR-26 Ijms 22 02529 i003ARL7, ABCA1located on chromosome
3p22.2
-mouse and human
LXR activated
macrophages
(+)
Downregulate LXR dependent
cholesterol efflux
blocks the expression of two important LXR target genes (ABCA1 and ARL7) required for cholesterol efflux[111]
miR-134 Ijms 22 02529 i003ANGPTL4located on chromosome
14q32.31
-THP-1 macrophages(+)
Increase LPL-mediated
lipid accumulation
By suppressing the expression of ANGPTL4, which is a regulator of lipoprotein lipase (LPL) activity, it promotes oxLDL uptake and inflammatory responses in vitro[112]
miR-155 Ijms 22 02529 i003HBP1located within a region known as the B-cell integration cluster (BIC)ApoE−/− miceox-LDL-treated human THP-1 macrophages(+)
Enhanced lipid uptake and enhanced ROS production
Silencing of miR-155 in ApoE−/− mice
by injecting antagomiR-155 decreased the lipid-laden macrophages and the formation of atherosclerotic plaques
[113]
miR-216a Ijms 22 02529 i003CSElocated on chromosome 2p16.1-THP-1 macrophages-derived foam cells(+)
Decreased ABCA1 expression and cholesterol efflux
Downregulation of CSE/H2S leads to an increase in cholesterol accumulation in foam cells[114]
miR-382-5p Ijms 22 02529 i003NFIAlocated on chromosome
14q32.31
- THP-1 macrophage-derived foam cells(+)
Reduces cholesterol efflux
Increases lipid
uptake
RP5-833A20.1/miR-382-5p/NFIA pathway regulates cholesterol homeostasis[107]
miR-486 Ijms 22 02529 i003HAT1located on chromosome
8p11.21
-THP-1 macrophage-derived foam cells(+)
Reduces cholesterol efflux
HAT1 is capable of acetylating H4K5 and H4K12 and increasing ABCA1 expression[115]
miR-212 Ijms 22 02529 i003SirT1located on chromosome
17p13.3
ApoE−/− miceTHP-1 human macrophages treated with oxLDL(+)
Suppresses ABCA1 dependent cholesterol efflux
SIRT1 has capable of inducing LXR activity to increase ABCA1 expression in human macrophages[116]
miR-19a Ijms 22 02529 i003HBP-1miR-19a is an important member of the miR-17–92 polycistronic gene cluster
MiR-19a is abundant in the blood and tissues of patients with atherosclerotic coronary artery disease
ApoE−/− miceTHP-1 derived macrophages
RAW264.7 cells
(+)
Increases lipid uptake of macrophages
HBP-1 participates in inhibiting the expression of the macrophage migration inhibitory factor (MIF)
and lipid uptake by macrophages
size of the atherosclerotic plaques in antagmiR-19a treated mice was reduced
[117]
miR-497 Ijms 22 02529 i003ApelinLocated on chromosome 17q13.1-Human THP-1 macrophages treated with oxLDL(+)
Decrease cholesterol efflux to apoA-I.
Apelin is an adipokine that is involved in the pathophysiology of
cardiovascular diseases
[118]
miR-20a/b Ijms 22 02529 i003ABCA1Mir 20a Located on chromosome 13q31.3
Mir 20b Located on chromosome Xq26.2
ApoE−/− MiceTHP-1 and RAW 264.7
Macrophage-derived foam cells.
(+)
Reducing cholesterol
Efflux,
impairs RCT in vivo
Both in in vitro studies and in ApoE−/− mice treated with miR-20a/b, the hepatic expression of ABCA1, as well as reverse cholesterol transport, are decreased[119]
miR-758 Ijms 22 02529 i003ABCA1miR-758 was widely expressed in mouse tissues and particularly abundant in the brain, heart and aorta
localized in an intergenic region within chromosome 14
Ldlr−/− mouseMouse and human
macrophage
(+)
Cholesterol efflux to apoA1
effect of miR-758 on
cholesterol efflux to ApoA1 was significantly attenuated after ABCA1 silencing
decrease in peritoneal macrophages obtained from hypercholesterolemia LDLR−/− mice
[120]
LncRNA MALAT1 Ijms 22 02529 i003CD36,
b-catenin
Located on chromosome 11q13.1-THP-1-derived macrophage(+)
Enhances lipid uptake
oxLDL induces MALAT1 transcription via the NF-kB pathway
Knockdown of MALAT1 using siRNA transfection reduces CD36 expression and affects lipid uptake in macrophages
[121]
LncRNA ENST00000602558.1 Ijms 22 02529 i003ABCG1 -vascular smooth muscle cells (VSMCs)(+)
Reduce ABCG1-mediated cholesterol efflux to HDL
ENST00000602558.1 induces p65, which is a specific inhibitor of NF-kB and mediates a decrease in the expression of ABCG1[122]
LncRNA UCA1 Ijms 22 02529 i003Mi R-206downregulation of UCA1 inhibits oxidative stress and induces apoptosis of THP-1 cells,
Located on chromosome
19p13.12
-THP-1 cells(+)
Increase oxidative stress process
In addition, CD36 levels
oxLDL greatly increases UCA1 expression,
UCA1 ‘sponges’ miR-206 to exacerbate atherosclerosis
[123]
Table 2. Non-coding RNAs that attenuate foam cell formation/function.
Table 2. Non-coding RNAs that attenuate foam cell formation/function.
Non-Coding RNAExpression TypeTarget GeneGenetic Information
(Human)
Experimental ModelEffect on Foam Cell FormationRegulation in Lipid MetabolismReference
In VivoIn Vitro
miR-1275 Ijms 22 02529 i003ApoC2located on chromosome
6p21.31
-THP-1 derived macrophages(−)
Inhibited the
cellular uptake of Ox-LDL and lipid accumulation
ApoC2 most important cofactor for LPL lipolytic activity[124]
miR-202-3p Ijms 22 02529 i003SCARB2, ABCG4, NCEH1Ilocated on chromosome
10q26.3
-Human THP-1
macrophage-derived foam cell
(−)
Attenuates lipid uptake
Enhances cholesterol efflux
Reduces the expression of scavenger receptor SCARB2
Increases the expression of NCEH1 and ABCG4
[125]
miR-181a Ijms 22 02529 i003TLR4located on chromosome
1q32.1
-THP-1(−)
Attenuates lipid uptake and apoptosis and inflammation
Inhibits protein levels of CD36 (scavenger receptor class B)[126]
miR-135a Ijms 22 02529 i003TLR4located on chromosome
3p21.2
-RAW264.7 and MOVAS cells(−)
Attenuates lipid uptake and inhibit oxidative stress and vascular inflammation
Inhibits CD36
expression
[127]
miR-9 Ijms 22 02529 i003ACAT1located on chromosome
1q22
-Human THP-1
macrophage-derived foam cell
(−)
Decrease the cholesterol ester formation
Reduces the levels of the ACAT1 protein
(Acyl-coenzyme A:cholesterol acyltransferase )
[128]
miR-21 Ijms 22 02529 i004MKK3
MERTK
located on chromosome
17q23.1
Ldlr−/− or miR21−/− micePeritoneal macrophages from adult miR21−/− mice(−)
Attenuated cholesterol efflux and promoting the lipid accumulation,
Reduce apoptotic cell
uptake
Knock down of miR-21 increases the expression of MKK3, promoting the induction of p38-CHOP and jNK signaling, which results in degradation of ABCG1,
Decreases expression of MERTK; a key receptor that mediates the clearance of apoptotic cells
[129]
miR-150 Ijms 22 02529 i003AdipoR2located on chromosome
19q13.33
-THP-1 macrophages(−)
Attenuates lipid uptake
Increases cholesterol efflux
Decreases CD36,
upregulates
ABCA1 and ABCG1 via the PPARγ- and LXRα-dependent pathways.
[130]
miR-133a Ijms 22 02529 i003TR4
(testicular orphan nuclear receptor 4)
located on chromosome
18q11.2
-RAW 264.7 macrophage cells(−)
Decreased oxLDL uptake
In addition, lipid accumulation
Decreases expression of CD36[131]
miR-155 Ijms 22 02529 i003Tim-3, CEH
(cholesterol ester hydrolase)
located on chromosome
21q21.3
-THP-1(−)
Increases cholesterol efflux,
Decreased lipid uptake
inhibits Tim-3 expression and increases the expression of CEH resulting in increased expression of ABCA1;
inhibits the expression of SR-A,
[132]
miR-155 Ijms 22 02529 i003CARHSP1located on chromosome
21q21.3
-THP-1(−)
Decrease inflammation and lipid uptake
suppresses TNF-α production by directly targeting CARHSP1, which is required for TNF-α mRNA stabilization[133]
miR-98 Ijms 22 02529 i003LOX-1
(Lectin-like ox-LDL scavenger receptor-1)
located on chromosome
Xp11.22
ApoE−/− miceMacrophages were collected from C57BL/6 mice(−)
Decreases the lipid uptake and lipid accumulation
decreases LOX-1 expression[134]
miR-223 Ijms 22 02529 i003TLR4-NF-κB signaling pathwaylocated on chromosome
Xq12
ApoE−/− mouse,
C57BL/6J wild-type mice(control)
Murine macrophage cell line RAW 264.7(−)
Increase cholesterol efflux to apoA-I,
Decrease inflammation
activates the PI3K/AKT signaling pathway[135]
miR-23a Ijms 22 02529 i003HSP90located on chromosome
19p13.12
-THP-1 macrophages(−)
Inhibit lipid accumulation
Decreased apoptosis of foam cells; decreases inflammation factors by inhibiting the activation of NF-κB pathways[136]
miR-146b-5p Ijms 22 02529 i003TRAF6located on chromosome
10q24.32
-THP-1 human monocytic and the HEK293T human embryonic kidney cell line(−)
Decrease lipid uptake
miR-146-5p inhibitor increases TRAF6-mediated activation of NF-κB (p65) and
increases inflammatory response
[137]
miR-221-3p Ijms 22 02529 i003ADAM22located on chromosome Xp11.3-RAW264.7(−)
Decrease lipid uptake
Decreases CD36; decrease oxidative stress and apoptosis[138]
miR-16 Ijms 22 02529 i003PDCD4
(Programmed cell death 4)
located on chromosome 13q14ApoE−/− micemacrophage-derived foam cell
(RAW264.7)
(−)Suppresses the activation of inflammatory macrophages and decreases release of inflammatory cytokines via the MAPK and NF-κB pathways[139]
miR-34a Ijms 22 02529 i003HDAC1
(histone deacetylase 1)
located on chromosome 1p36.22Male C57BL/6J and ApoE−/− miceHuman THP-1 macrophages(−)
Prevents the accumulation of triglyceride and total and free cholesterol
Hyperhomocysteinemia (HHcy) accelerates atherogenesis via decreased expression of miR-34a[140]
Let-7g Ijms 22 02529 i003NF-κB complex
MEKK1
located on chromosome 3p21.2ApoE−/− mouse,
IKKαf/f:MLysCre/apoE−/− mice
Human
THP-1 macrophage
(−)
increase cholesterol efflux,
Decrease intracellular lipid accumulation
up-regulates SREBF2, which is a critical regulator of cholesterol/lipid homeostasis,
up-regulates ABCA1 via suppression of miR-33a,
decreases p53-dependent apoptosis
[141]
LncRNA CDKN2B-AS1 Ijms 22 02529 i004ADAM10located on chromosome 9p21.3ApoE−/− Mice,
C57BL/6J Mice
THP-1
macrophage-derived foam cells
(−)
Increases Cholesterol Efflux,
Decrease lipid accumulation
Inhibits inflammatory responses by suppressing the transcription of ADAM10[142]
LncRNA DYNLRB2-2 Ijms 22 02529 i003miR-298located on chromosome 16q23.2-THP-1 macrophage-derived foam cells(−)
Impairing oxLDL uptake,
increases cholesterol efflux
Induces autophagy by activating the LKB1/AMPK/mTOR signaling pathway via the miR-298/SIRT3 axis;
decreases the expression of TLR2 and enhances the expression of ABCA1
[143]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Javadifar, A.; Rastgoo, S.; Banach, M.; Jamialahmadi, T.; Johnston, T.P.; Sahebkar, A. Foam Cells as Therapeutic Targets in Atherosclerosis with a Focus on the Regulatory Roles of Non-Coding RNAs. Int. J. Mol. Sci. 2021, 22, 2529. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22052529

AMA Style

Javadifar A, Rastgoo S, Banach M, Jamialahmadi T, Johnston TP, Sahebkar A. Foam Cells as Therapeutic Targets in Atherosclerosis with a Focus on the Regulatory Roles of Non-Coding RNAs. International Journal of Molecular Sciences. 2021; 22(5):2529. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22052529

Chicago/Turabian Style

Javadifar, Amin, Sahar Rastgoo, Maciej Banach, Tannaz Jamialahmadi, Thomas P. Johnston, and Amirhossein Sahebkar. 2021. "Foam Cells as Therapeutic Targets in Atherosclerosis with a Focus on the Regulatory Roles of Non-Coding RNAs" International Journal of Molecular Sciences 22, no. 5: 2529. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22052529

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

Article Metrics

Back to TopTop