Skip to main content

The translational potential of miR-26 in atherosclerosis and development of agents for its target genes ACC1/2, COL1A1, CPT1A, FBP1, DGAT2, and SMAD7

Abstract

Atherosclerosis is one of the leading causes of death worldwide. miR-26 is a potential biomarker of atherosclerosis. Standardized diagnostic tests for miR-26 (MIR26-DX) have been developed, but the fastest progress has been in predicting the efficacy of IFN-α therapy for hepatocellular carcinoma (HCC, phase 3). MiR-26 slows atherosclerosis development by suppressing ACC1/2, ACLY, ACSL3/4, ALDH3A2, ALPL, BMP2, CD36, COL1A1, CPT1A, CTGF, DGAT2, EHHADH, FAS, FBP1, GATA4, GSK3β, G6PC, Gys2, HMGA1, HMGB1, LDLR, LIPC, IL-1β, IL-6, JAG2, KCNJ2, MALT1, β-MHC, NF-κB, PCK1, PLCβ1, PYGL, RUNX2, SCD1, SMAD1/4/5/7, SREBF1, TAB3, TAK1, TCF7L2, and TNF-α expression. Many agents targeting these genes, such as the ACC1/2 inhibitors GS-0976, PF-05221304, and MK-4074; the DGAT2 inhibitors IONIS-DGAT2Rx, PF-06427878, PF-0685571, and PF-07202954; the COL1A1 inhibitor HT-100; the stimulants 68Ga-CBP8 and RCT-01; the CPT1A inhibitors etomoxir, perhexiline, and teglicar; the FBP1 inhibitors CS-917 and MB07803; and the SMAD7 inhibitor mongersen, have been investigated in clinical trials. Interestingly, miR-26 better reduced intima-media thickness (IMT) than PCSK9 or CT-1 knockout. Many PCSK9 inhibitors, including alirocumab, evolocumab, inclisiran, AZD8233, Civi-007, MK-0616, and LIB003, have been investigated in clinical trials. Recombinant CT-1 was also investigated in clinical trials. Therefore, miR-26 is a promising target for agent development. miR-26 promotes foam cell formation by reducing ABCA1 and ARL4C expression. Multiple materials can be used to deliver miR-26, but it is unclear which material is most suitable for mass production and clinical applications. This review focuses on the potential use of miR-26 in treating atherosclerosis to support the development of agents targeting it.

Introduction

Cardiovascular disease (CVD) is one of the leading causes of death worldwide [1]. The formation of atherosclerotic plaques obstructs blood flow to organs, the clinical manifestations of which include myocardial infarction (MI), ischemic heart disease (IHD), and stroke, which are the main causes of CVD [2,3,4,5]. Atherosclerosis is a chronic inflammatory disease caused by the dysregulation of lipid metabolism. When the blood low-density lipoprotein (LDL)-c concentration is higher than the physiological level, LDL-c passively diffuses from the vascular lumen to the intima and is oxidized to form ox-LDL in response to reactive oxygen species (ROS). Macrophages engulf a large amount of ox-LDL, and their efflux is blocked, resulting in the accumulation of cholesterol in macrophages and the formation of foam cells [6, 7]. Foam cells recruit vascular smooth muscle cells (VSMCs) from the middle membrane. Migrating VSMCs form fibrous caps on plaques by secreting matrix metalloproteinase 2 (MMP2), MMP9, collagen, and elastin, leading to persistent inflammation and endothelial dysfunction. Over time, the plaque gradually becomes large enough to damage the arterial cavity, blocking blood flow to the tissue and causing tissue ischemia or plaque rupture to form blood clots. Therefore, timely inhibition of lipid metabolism disorders, the inflammatory response, and VSMC migration is the focus of atherosclerosis prevention and treatment of [8,9,10,11,12].

MiR-26, a vertebrate-specific miRNA, includes miR-26a (also named miR-26a-5p) and miR-26b (also named miR-26b-5p). MiR-26a/b is expressed in different tissues, including the spleen, brain, kidney, liver, lung, heart, testis, adipose tissue, and macrophages [13]. The mature sequences of miR-26a and miR-26b are identical except for two base differences. MiR-26a includes miR-26a-1 and miR-26a-2. miR-26a-1 is located in the intron of the C-terminal domain of RNA polymerase II polypeptide A small phosphatase-like (CTDSPL; also known as C3orf8, HYA22, PSR1, RBSP3, and SCP3). MiR-26a-2 is located in the intron of the C-terminal domain of RNA polymerase II polypeptide A small phosphatase 2 (CTDSP2; also known as OS4, PSR2, and SCP2). MiR-26b is located in the intron of CTDSP1 (also known as NIF3, NLI-IF, NLIIF, and SCP1). The mature sequences of miR-26a-1 and miR-26a-2 are identical [14]. MiR-26 decreases atherosclerosis development by regulating many genes, including ATP-binding cassette transporter A1 (ABCA1), acetyl-CoA-carboxylase 1 (ACC1), ACC2, ATP-citrate lyase (ACLY), Acyl-CoA synthetase long-chain 3 (ACSL3), ACSL4, aldehyde dehydrogenase 3 family member A2 (ALDH3A2, also named fatty aldehyde dehydrogenase (FALDH)), alkaline phosphatase (ALPL), ARF-like 7 (ARL7, also named ADP-ribosylation factor-like 4C (ARL4C)), β-myosin heavy chain (β-MHC, also named MYH7), bone morphogenetic protein 2 (BMP2), cluster of differentiation 36 (CD36), collagen I (COL1A1), carnitine palmitoyl-transferase 1A (CPT1A), connective tissue growth factor (CTGF), diacylglycerol acyltransferase 2 (DGAT2), enoyl-CoA hydratase/3-hydroxyacyl-CoA dehydrogenase (EHHADH), fatty acid synthase (FASN, also named FAS), fructose-1,6-bisphosphatase (FBP1, also named FBPase), GATA binding protein 4 (GATA4), glucose-6-phosphatase (G6PC, also named G6PC1, G6Pase, and G6Pase-α), glycogen synthase kinase-3β (GSK-3β), glycogen synthase 2 (Gys2), high mobility group A1 (HMGA1), HMGB1, interleukin (IL)-1 beta (IL-1β), IL-6, jagged canonical Notch ligand 2 (JAG2), potassium inwardly rectifying channel subfamily J member 2 (KCNJ2), low-density lipoprotein receptor (LDLR), lipase C hepatic type (LIPC), mucosa-associated lymphoid tissue lymphoma translocation protein 1 (MALT1), nuclear factor κB (NF-κB), carboxykinase (PCK1), phospholipase C beta (PLCβ1), glycogen phosphorylase L (PYGL), runt-related protein 2 (RUNX2), stearoyl-coenzyme A desaturase 1 (SCD1), mothers against decapentaplegic homolog 1 (SMAD1), SMAD4, SMAD5, SMAD7, sterol regulatory element-binding transcription factor 1 (SREBF1), TGFβ-activated kinase binding protein 3 (TAB3), transforming growth factor β (TGFβ)-activated kinase-1 (TAK1, also named MAP3K7), transcription factor 7 like 2 (TCF7L2, also named TCF-4), and tumor necrosis factor alpha (TNF-α) [15,16,17,18,19,20,21,22,23,24,25]. Many drugs have been brought to market or developed in clinical trials by targeting these genes, suggesting that targeting these genes can greatly improve the success rate of drug development [22, 26,27,28,29]. The value of miR-26 has also been investigated in clinical trials. This review focuses on the role and mechanism of action of miR-26 in atherosclerosis, including the detection value of miR-26 and the potential development of drugs targeting its downstream genes, such as ACC1/2, COL1A1, CPT1A, FBP1, DGAT2, and SMAD7, to provide insight for drug development.

The clinical value of miR-26 quantification

Atherosclerosis

The diagnostic value of miR-26 in atherosclerosis has been investigated in clinical trials and preclinical models [30,31,32,33,34,35]. The serum miR-26 level was negatively associated with serum total cholesterol (TC), triglycerides (TG), and LDL-C and positively associated with HDL-C in patients with carotid atherosclerosis (CAS) and in apoE−/− mice [36]. The miR-26 level in exosomes produced by adipose-derived stem cells (ADSC-exos) was negatively associated with TNF-α, IL-6, and IL-1β, suggesting that the ADSC-exo miR-26 level is a potential biomarker for the diagnosis of lipids and inflammatory factors (Table 1) [36]. Clinical studies on the diagnostic value of miR-26 in atherosclerosis are limited. Biomarkers should be sensitive and specific for diagnosing the disease state. More studies are needed to confirm the clinical translatability of miR-26.

Table 1 The diagnostic value of miR-26 in diseases in clinical trials

LIPC, a miR-26 target gene, encodes hepatic lipase (HL) and promotes the hydrolysis of TG. LIPC not only remodels LDL and HDL but also enhances lipid and lipoprotein uptake [37]. Interestingly, LIPC gene mutation is the second most common cause of familial hypocholesterolemia, after only angiopoietin-like 3 (ANGPTL3) [38], suggesting that LIPC is a promising target for the diagnosis of familial hypocholesterolemia.

Prediction of the efficacy of IFN-α

The effectiveness of interferon-α (IFN-α) in patients with low miR-26 expression in tumors after curative resection of hepatocellular carcinoma (HCC) is under phase 3 clinical investigation. IFN-α therapy is expected to improve disease-free survival in patients with low miR-26 expression by inhibiting tumor recurrence [39]. Compared with patients with HCC with high expression of miR-26, patients with low-miR-26-expressing tumors had shorter overall survival but a better response to IFN-α therapy (n = 489) [40,41,42], suggesting that IFN-α therapy is more effective in HCC patients with low miR-26 expression. The use of miR-26 as a predictive marker of therapeutic responsiveness has also received patent protection (WO2009152300A1). Standardized diagnostic tests for miR-26 (MIR26-DX) have also been developed [42]. IFN-α therapy has been used in more than 40 countries to treat viral infections, such as hepatitis B and C; condyloma; shingles; hematological diseases such as leukemia, multiple myeloma, T-cell lymphoma and essential thrombocytosis; and cancers such as renal cell cancer (RCC), melanoma, HCC, hairy cell leukemia, myeloma, bladder cancer, ovarian cancer, and cervical cancer (Fig. 1) [43,44,45]. The downside is that IFN-α therapy can promote hyperlipidemia and atherosclerosis [46]. The cardiovascular side effects of IFN-α therapy should be noted if it is to be applied clinically [47]. Therefore, miR-26 may be the first miRNA whose expression is considered in the decision to use IFN-α treatment for HCC.

Fig. 1
figure 1

The clinical detection value of miR-26 in HCC. IFN-α therapy was more effective in HCC patients with low miR-26 expression. HCC, hepatocellular carcinoma; IFN-α, interferon-α; MIR26-DX, standardized diagnostic tests for miR-26; RCC, renal cell cancer

In fact, IFN-α suppressed HCC proliferation and invasion by inducing miR-26a and suppressing the expression of its target enhancer of zeste homolog 2 (EZH2) [48]. miR-26a also increased IFN-α expression by targeting ubiquitin-specific protease 3 (USP3), which is a negative regulator of IFN-α [49], suggesting that IFN-α and miR-26a inhibit the development of HCC by forming positive feedback loops. However, IFN-α also promoted miR-26a degradation by increasing polyribonucleotide nucleotidyltransferase 1 (PNPT1) and its thus the DNA demethylation that it mediated. More studies are needed to confirm the role of IFN-α in regulating miR-26a expression.

T2DM, T1DM, MINOCA, leukoencephalopathy, EHOA, PsA, prostate cancer, and atopy

The plasma-derived exosomal miR-26a level was reduced in patients with type 2 diabetes mellitus (T2DM). miR-26a also promoted insulin sensitivity. Therefore, exosomal miR-26a is not only a potential biomarker for T2DM but also a potential therapeutic target. However, serum miR-26a was increased in children with T1DM [50]. The diagnostic value of miR-26 in diseases, including myocardial infarction with nonobstructive coronary arteries (MINOCA) [51], leukoencephalopathy [52], erosive hand osteoarthritis (EHOA) and psoriatic arthritis (PsA) [53], prostate cancer [54], and atopy in children over time (including atopy, atopic dermatitis eczema, wheezing and food allergy in infants), is under clinical research (Table 1) [55], the results of which have not been disclosed.

The role and mechanism of action of miR-26 in atherosclerosis

MiR-26 has become a novel target for the treatment of atherosclerosis. Overexpression of miR-26 in ADSC-Exos decreased the size of atherosclerotic plaques by correcting the changes in the serum lipid and proinflammatory factor levels in the apoE−/− mice. Mechanistically, miR-26 decreases IL-1β, IL-6, and TNF-α levels by suppressing NF-κB activity through the targeting of COL1A1, CTGF, HMGA1, HMGB1, MALT1, TAB3, and TAK1 [15,16,17, 26, 27]. MiR-26 suppresses endothelial cell (EC) growth, angiogenesis, and VSMC differentiation by targeting SMAD1 and SMAD4 [18, 28, 56, 57]. MiR-26 can also suppress atrial fibrillation (AF) and cardiomyocyte hypertrophy by targeting β-MHC, GSK3β, GATA4, KCNJ2, and PLCβ1 [58]. MiR-26a suppresses vascular and aortic valve calcification by suppressing the expression of pro-calcification genes, including ALPL, BMP2, CTGF, SMAD1, SMAD5, and RUNX2, and increasing the expression of anti-calcification genes, including SMAD7 and JAG2 [19,20,21]. MiR-26a has decreased hepatic glucose production by suppressing the expression of gluconeogenesis- and glycogen metabolism-related genes, such as FBP1, G6PC, Gys2, PCK1, PYGL, SCD1, and TCF7L2 [22] and suppressed fatty acid (FA) synthesis and oxidation by suppressing the expression of ACC1, ACC2, ACLY, ACSL3, ACSL4, ALDH3A2, CPT1A, DGAT2, EHHADH, FAS, LIPC, SCD1, and SREBF1. MiR-26a can suppress cholesterol uptake by suppressing CD36 and LDLR expression [22]. Therefore, miR-26 decreases atherosclerotic plaque size by regulating lipid metabolism, the inflammatory response, angiogenesis, cell differentiation, calcification, glucose metabolism, and insulin signaling-related genes (Fig. 2). However, miR-26 also suppresses cholesterol efflux to promote foam cell formation by binding and suppressing ABCA1 and ARL4C in vitro, as observed in RAW264.7 cells, THP-1 cells, and HepG2 cells [23, 59,60,61], suggesting that the formation of foam cells induced by miR-26 may weaken its ability to reduce atherosclerosis. Notably, the role of miR-26 in foam cell formation has not been investigated in vivo.

Fig. 2
figure 2

The role and mechanism of miR-26 in atherosclerosis. MiR-26 regulates the development of atherosclerosis through multiple mechanisms, including suppressing vascular calcification, EC growth, angiogenesis, and VSMC differentiation, cholesterol uptake, inflammatory response, glucose production, and FA synthesis and oxidation. ABCA1 ATP-binding cassette transporter A1, ACC1 acetyl-CoA-carboxylase 1, ACLY ATP-citrate lyase, ACSL3 Acyl-CoA synthetase long-chain 3, ALDH3A2 aldehyde dehydrogenase 3 family member A2, ALPL alkaline phosphatase, ARL4C ARF-like 7, β-MHC β-myosin heavy chain, BMP2 bone morphogenetic protein 2, CD36 cluster of differentiation 36, CPT1A carnitine palmitoyl-transferase 1A, CTGF connective tissue growth factor, DGAT2 diacylglycerol acyltransferase 2, EHHADH enoyl-CoA hydratase/3-hydroxyacyl-CoA dehydrogenase, FAS fatty acid synthase, FBP1 fructose-1,6-bisphosphatase, GATA4 GATA binding protein 4, G6PC glucose-6-phosphatase, GSK3β glycogen synthase kinase-3β, Gys2 glycogen synthase 2, HMGA1 high mobility group A1, IL-1β interleukin (IL)-1 beta, JAG2 jagged canonical Notch ligand 2, KCNJ2 potassium inwardly rectifying channel subfamily J member 2, LDLR low-density lipoprotein receptor, LIPC lipase C hepatic type, MALT1 mucosa-associated lymphoid tissue lymphoma translocation protein 1, NF-κB nuclear factor κB, PCK1 carboxykinase, PLCβ1 phospholipase C beta, PYGL glycogen phosphorylase L, RUNX2 runt-related protein 2, SCD1 stearoyl-coenzyme A desaturase 1, SMAD1 mothers against decapentaplegic homolog 1, SREBF1 sterol regulatory element-binding transcription factor 1, TAB3 TGFβ-activated kinase binding protein 3, TCF7L2 transcription factor 7 like 2, TNF-α tumor necrosis factor alpha

The clinical application of miR-26 target genes, with a focus on ACC1/2, COL1A1, CPT1A, DGAT2, FBP1, and SMAD7

ACC1, ACC2, and DGAT2

ACC1/2 catalyzes the conversion of acetyl-CoA to malonyl-CoA, which is the raw material for TG biosynthesis [62, 63]. Many inhibitors, such as GS-0976 (also named firsocostat, NDI-010976) [64, 65], PF-05221304 (also named clesacostat) [66, 67], MK-4074 [68], ND-654 [69], and ND-646 [70, 71], have been investigated in preclinical and clinical trials (Table 2), suggesting that ACC1/2 is a promising target for drug development. However, hypertriglyceridemia is a common adverse effect of ACC1/2 inhibitors that limits their clinical development [72].

Table 2 The drugs were approved or investigated in clinical trials by targeting ACC1/2, COL1A1, CPT1A, CT-1, FBP1, PCSK9, and SMAD7

DGAT2 is an important enzyme for TG biosynthesis (up to 90%) [73, 74]. Many DGAT2 inhibitors, such as PF-06865571 (also named ervogastat) [75,76,77], IONIS-DGAT2Rx (also named ISIS 484137) [78], PF-07202954 [79], PF-06427878 [80, 81], PF-06424439 [82], and benzimidazolone derivatives [83], have been investigated in preclinical and clinical trials. Importantly, DGAT2 inhibitors can reduce the side effects of ACC1/2 inhibitors. PF-06865571 combined with PF-05221304 showed good efficacy and safety in a phase II trial for the treatment of nonalcoholic steatohepatitis (NASH) [66, 75], suggesting that the development of multitarget inhibitors is the future direction for drugs targeting DGAT2 and ACC1/2 (Fig. 3). MiR-26 suppressed ACC1/2 expression and DGAT2 expression. Therefore, miR-26 is a potential candidate for development as a multitarget inhibitor of ACC1, ACC2, and DGAT2 expression.

Fig. 3
figure 3

MiR-26 delivery systems and targeting ACC1/2, CT-1, DGAT2, and PCSK9 with drugs that were approved or investigated in clinical trials. ADSCs overexpressing miR-26 reduced IMT more than those with PCSK9 or CT-1 knockout. DGAT2 inhibitors can reduce the side effects of ACC1/2 inhibitors. Multiple materials have been used to deliver miR-26. ACC1 acetyl-CoA-carboxylase 1, CT-1 cardiotrophin-1, PCSK9 Proprotein convertase subtilisin/kexin type 9

COL1A1

COL1A1, the main fibrous collagen in the extracellular matrix (ECM), is located mainly in tendons, bones, teeth, skin, lungs, heart, and vasculature. COL1A1 regulates MMP-2, MMP-8 and MMP-9 through its bioactive peptide/fragments proline-glycine-proline (PGP), α1 C1158/59 (C-1158/59), and C-propeptide [84,85,86]. COL1A1 promotes tumor cell proliferation, epithelial–mesenchymal transition, and metastasis by binding to its receptor. COL1A1 also modulates the efficacy of tumor treatments, such as chemotherapy, radiation, and immunotherapy [87], suggesting that COL1A1 plays a role in tumor development. In fact, COL1A1 is key to many diseases, such as atherosclerosis, myocardial fibrosis, heart failure, and osteogenesis imperfecta [88,89,90]. Many COL1A1-specific agents, such as HT-100 (also named Halofuginone Hydrobromide, PCS-100, Stenorol, and Tempostatin), 68Ga-Collagen Binding Probe #8 (CBP8) (68Ga-CBP8), and RCT-01, have been investigated in clinical trials, and the results suggested that COL1A1 is a promising target for drug development (Fig. 4).

Fig. 4
figure 4

The COL1A1-specific agents entering clinical trials and their targets. COL1A1 collagen I, Egr-1 early growth response-1, IFNγ Interferon-γ, IGFBP-1 insulin-like growth factor binding protein 1, IL-1β Interleukin (IL)-1 beta, MMP2 matrix metalloproteinase 2, NF-κB Nuclear factor κB, p-eIF2α phosphorylated eIF2α, PRL-1 phosphatase of regenerating liver-1, SMAD3 mothers against decapentaplegic homolog 1, TNF-α tumor necrosis factor alpha, WT1 Wilms’ tumor gene 1, DMD duchenne muscular dystrophy

68Ga-CBP8

68Ga-CBP8 is a gallium-68-labeled collagen-binding PET imaging probe that can selectively bind to COL1A1. 68Ga-CBP8 has a high affinity for COL1A1 in a mouse lung injury model of radiation lung injury and has good pharmacological and pharmacokinetic characteristics, with high target uptake and low retention in background tissues and organs. 8Ga-CBP8 has high specificity and a high target background ratio for pulmonary fibrosis in diseased animals, and it can be used to effectively detect pulmonary fibrosis and its response to therapy [91,92,93]. 68Ga-CBP8 can be used for noninvasive COL1A1 imaging in a range of human fibrotic diseases. Many clinical trials have investigated 68Ga-CBP8 (Table 2) [94,95,96,97,98,99]. In phase 1 clinical trials, the average injection activity of 68Ga-CBP8 in healthy volunteers (5 males and 4 females) was 220 Mbp. No adverse reactions were associated with the probe injection. 68Ga-CBP8 exhibited an extracellular distribution and was rapidly cleared, primarily by the kidneys. The initial distribution half-life and elimination half-life of 68Ga-CBP8 were 4.9 min and 72 min, respectively [100]. These results supported the further development of 68Ga-CBP8.

HT-100

HT-100, a COL1A1 inhibitor, was developed for the treatment of Duchenne muscular dystrophy (DMD) by the Akashi Therapeutics pipeline and Collgard Biopharmaceuticals [101,102,103]. The interim phase 1b/2a clinical data of the HT-100 trial showed improved muscle strength at low doses (N = 10). The average increase in total muscle strength was 11.7%, which was 22.3% higher than that in the steroid treatment group. No serious adverse events were associated with the drugs [104]. However, at higher doses, one patient experienced serious adverse events leading to death [105]. The further development of HT-100 for the treatment of DMD was discontinued [105,106,107,108]. HT-100 for the treatment of HIV-related Kaposi’s sarcoma (phase 2) and solid tumors (phase 1) was administered on 5 June 2013 and 24 July 2012, respectively [109, 110], but no results were disclosed. The further development of HT-100 in many diseases, including bladder cancer, cancer, coronary artery restenosis, hepatic fibrosis, kidney disorders, renal fibrosis, and scleroderma, has been discontinued [103]. Notably, further development of HT-100 in DMD patients may still be in phase 2 trials as of September 2022 [103]. Despite our best efforts, we did not find any information on the outcomes. In addition, HT-100 reduced the expression of multiple genes, including cyclins D1 and B1, IL-1β, IL-6, IL-8, MMP2, NF-κB, p38 MAPK, SMAD3, TGFβ, TNF-α, and Wilms’ tumor gene 1 (WT1), and increased the expression of multiple genes, including cyclin A, early growth response-1 (Egr-1)/phosphatase of regenerating liver-1 (PRL-1), IFNγ, IL-2, and phosphorylated eIF2α (p-eIF2α) [111,112,113,114,115,116,117]. Therefore, we should be wary of the low specificity and off-target effects of HT-100, which may result in high toxicity.

RCT-01

RCT-01, an autologous cell therapy, consists nonbulbar dermal sheath (NBDS) cells that expressed COL1A1 and collagen II (COL2A1). NBDS cells are fibroblasts isolated from a small tissue sample at the back of a patient's scalp (hair follicles). NBDS cells express the highest levels of COL1A1 and COL2A1 among fibroblasts anywhere in the body. These cells can repair and regenerate tissues. RCT-01 was developed for the treatment of Achilles tendinosis by RepliCel and YOFOTO in Greater China and by RepliCel elsewhere in the world. A phase 1 clinical study of RCT-01 demonstrated its safety. RCT-01 regenerated tendon tissue, restored blood flow, reduced pain, and improved function in patients with chronic tendinopathy [118, 119]. However, RCT-01 did not have significantly better efficacy than standard therapy in phase 1 studies [119]. The study was also discontinued due to slow enrollment on 28 September 2017 [120]. Notably, in 2021, RepliCel announced that it would begin a second clinical study of RCT-01 in Japan [121]. RCT-01 was reviewed by the Pharmaceutical and Medical Devices Agency (PMDA) of Japan on 28 July 2022 [122] and is expected to receive clinical registration soon. This clinical protocol for the treatment of chronic tendinopathies (including Achilles tendinopathy) was nearing completion as of 28 November 2022 [123].

CPT1A

CPT1A is the rate-limiting enzyme of fatty acid β-oxidation and long-chain fatty acid (LC-FAO) oxidation. CPT1A-mediated LC-FAO is essential for the development of CD8+ T-cell memory and protective immunity, which play key roles in the adaptive immune response against infection and cancer [124, 125]. Many CPT1A inhibitors, such as etomoxir (ETO), perhexiline (also named Pexsig), and teglicar (also named ST1326), have been investigated in clinical trials, suggesting that CPT1A is a promising target for drug development (Fig. 5).

Fig. 5
figure 5

The CPT1A inhibitors entering clinical trials and their targets. Acaa1a acetyl-CoA acyltransferase, ACOX1 acyl-CoA oxidase 1, ADH1 alcohol dehydrogenase 1, AP1S1 adaptor related protein complex 1 subunit sigma 1, ASAH1 N-acylsphingosine amidohydrolase 1, ASNS asparagine synthetase, ATF4 activating transcription factor 4, CLU clusterin, COX4I1 matrix cytochrome C oxidase subunit 4I1, CRAT carnitine acetyltransferase, DEPP1 DEPP autophagy regulator 1, EHHADH enoyl-CoA hydratase and 3-hydroxyacyl CoA dehydrogenase, EIF4EBP1 eukaryotic translation initiation factor 4E binding protein 1, FABP1 fatty acid binding protein 1, FNDC4 fibronectin type III domain containing 4, GDPD3 glycerophosphodiester phosphodiesterase domain containing 3, G6PD glucose-6-phosphate dehydrogenase, HPN hepsin, HPX hemopexin, INHBE inhibin subunit beta E, KCNH2 potassium voltage-gated channel subfamily H member 2, LSS lanosterol synthase, MT1A metallothionein 1A, NROB2 nuclear receptor subfamily 0 group B member 2, NUPR1 nuclear protein 1, transcriptional regulator, PRKCB protein kinase C beta, RPS9 ribosomal protein S9, SERCA2 sarco-endoplasmic reticulum calcium ATPase 2, SERPINA3 serpin family A member 3, SLC10A1 solute carrier family 10 member 1, TAGLN transgelin, TMPRSS2 transmembrane serine protease 2, TP53 tumor protein p53, UCP2 uncoupling protein 2, WIPI1 WD repeat domain, phosphoinositide interacting 1, T2DM type 2 diabetes mellitus

ETO

ETO inhibits CPT1A expression by binding to its active site. ETO also inhibits the expression of CPT1B, suggesting that ETO has weak specificity for CPT1A [125]. The use of ETO for the treatment of heart failure has been investigated in phase 2 clinical trials, but further development of ETO in heart failure and T2DM patients was discontinued due to hepatotoxicity [125, 126]. ETO can increase the expression of multiple genes, including acetyl-CoA acyltransferase (Acaa1a), Acaa1b, acyl-CoA oxidase 1 (ACOX1), alcohol dehydrogenase 1 (ADH1), clusterin (CLU), matrix cytochrome C oxidase subunit 4I1 (COX4I1), carnitine acetyltransferase (CRAT), enoyl-CoA hydratase and 3-hydroxyacyl CoA dehydrogenase (EHHADH), fatty acid binding protein 1 (FABP1), glucose-6-phosphate dehydrogenase (G6PD), metallothionein 1A (MT1A), protein kinase C beta (PRKCB), ribosomal protein S9 (RPS9), sarco-endoplasmic reticulum calcium ATPase 2 (SERCA2, encoding ATP2A2), tumor protein p53 (TP53), uncoupling protein 2 (UCP2), and UCP3, and decrease hemopexin (HPX) expression [127,128,129,130]. Therefore, the low specificity of perhexiline and its off-target effects may have led to its high toxicity and cessation of development.

Perhexiline

Perhexiline is a CPT1A and CPT2 inhibitor but has less of an effect on CPT2. Perhexiline was approved for the treatment of angina pectoris. However, long-term use of perhexiline can cause neurotoxicity and hepatotoxicity [87]. The use of perhexiline for the treatment of hypertrophic cardiomyopathy was also investigated in clinical trials. Its further study in a phase 2 clinical trial was terminated due to a lack of efficacy on 31 August 2017 [131]. In fact, perhexiline decreased multiple genes expression, including ABCB4, ABCB11, ABCC3, ABCG5, adaptor related protein complex 1 subunit sigma 1 (AP1S1), solute carrier family 10 member 1 (SLC10A1), solute carrier organic anion transporter family, member 1a1 (SLCO1A1), and transgelin (TAGLN), and increased multiple genes expression, including ABCB2, N-acylsphingosine amidohydrolase 1 (ASAH1), asparagine synthetase (ASNS), activating transcription factor 4 (ATF4), ATF6, DEPP autophagy regulator 1 (DEPP1), eukaryotic translation initiation factor 4E (EIF4E)/EIF4E binding protein 1 (EIF4EBP1), FABP1, fibronectin type III domain containing 4 (FNDC4), glycerophosphodiester phosphodiesterase domain containing 3 (GDPD3), hepsin (HPN), inhibin subunit beta E (INHBE), potassium voltage-gated channel subfamily H member 2 (KCNH2), lanosterol synthase (LSS), nuclear receptor subfamily 0 group B member 2 (NROB2), nuclear protein 1, transcriptional regulator (NUPR1), serpin family A member 3 (SERPINA3), transmembrane serine protease 2 (TMPRSS2), WD repeat domain, phosphoinositide interacting 1 (WIPI1) [132,133,134,135,136,137]. The low specificity of perhexiline and its off-target effects may have led to its high toxicity.

Teglicar

Teglicar, an aminocarnitine analog, is a formylcarnitine derivative. Teglicar was more selective for CPT1A than for ETO. Teglicar significantly improved hyperglycemia and regulated glucose homeostasis in obesity and type 2 diabetes mouse models [138]. Teglicar also inhibited the growth of canine breast cancer cells by inducing caspase-3/8/9-mediated apoptosis [87]. Teglicar for the treatment of T2DM (40 males and females) was well tolerated and safe in phase 1 clinical trials. The half-life of teglicar was 25 h. Teglicar (450 mg/day) significantly improved insulin resistance (from 4.1 on day 1 to 3.0 on day 16) and blood sugar levels (reducing 16 mg/dl on day 16 vs. 4 mg/dl placebo group) [139]. These results supported the further development of Teglicar, but its further development for the treatment of T2DM in phase 2 clinical trials was discontinued [140].

FBP1

FBP1 is a rate-controlling enzyme of gluconeogenesis [141]. Many FBP1 inhibitors, such as CS-917 (also named MB06322) and MB07803 (also named VK0612), were investigated in clinical trials, suggesting that FBP1 is a promising target for drug development (Fig. 6).

Fig. 6
figure 6

The FBP1 inhibitors and SMAD7 inhibitors entering clinical trials. FBP1 fructose-1,6-bisphosphatase, SMAD7 mothers against decapentaplegic homolog 7, T2DM type 2 diabetes mellitus, CD Crohn’s disease

CS-917

CS-917 is a prodrug of MB05032 (also named R-125338). MB05032 inhibits FBP1 expression by binding to its active AMP site [142]. MB05032 has low cell penetration and oral absorption due to its phosphate group. Thanks to its dialanyl amide group, CS-917 administration leads to much higher bioavailability of MB05032 than direct administration of MB05032. CS-917 is catalyzed to an intermediate form, R-134450, by the esterase enzyme and then to an active form, MB05032, by the phosphoramidase enzyme [143]. CS-917 has inhibited gluconeogenesis and improved fasting and postprandial hyperglycemia in preclinical trials [142, 144, 145]. Six phase 1/2 clinical trials of MB07803 have been completed [146]. The main theoretical side effect of FBP1 inhibitors is an increased risk of hypoglycemia. However, CS-917 exhibited good tolerability and safety in overnight-fasted healthy volunteers (in phase 1 clinical trials) and patients with T2DM (in phase 2 clinical trials). CS-917 at doses of 50, 200, and 400 mg/d reduced glucose levels in patients with T2DM. However, at a dose of 100 mg/d, CS-917 did not affect glucose, suggesting that the efficacy of CS-917 was not dose dependent. CS-917 at doses of 50 and 100 mg/d also did not affect glucose levels in phase 2a clinical trials. The lactic acidosis caused by CS-917 in combination with metformin was significantly resolved after withdrawal in phase 1 clinical trials, suggesting a possible interaction between metformin and CS-917. Further development of CS-917 has been halted due to concerns about its efficacy and interactions [141]. CS-917 also decreased the FBP2 level [147], suggesting that the low specificity of CS-917 and its off-target effects may have led to the suspension of its development.

MB07803

MB07803, a derivative of CS-917, was developed by Viking Therapeutics. MB07803 has better pharmacokinetic properties than CS-917, including higher oral bioavailability, longer active metabolite half-life, and lower inactive N-acetylated metabolite production. Six phase 1 clinical trials and one phase 2a clinical trial on MB07803 have been done [148]. MB07803 exhibited good safety and tolerability in healthy volunteers (in phase 1 clinical trials) and patients with T2DM (in phase 2a clinical trials). MB07803 reduced fasting blood sugar (FBG) at the highest dose (200 mg/day). P.O. 28 days). After optimal dosing, MB07803 also exhibited good tolerability and safety. The maximum tolerated dose of MB07803 in phase 1b was 200 mg twice daily. MB07803 reduced FBG in a dose-dependent manner (50, 200, and 400 mg BID) [149]. These results supported the further development of MB07803. Even so, no FBP1 inhibitors have been approved. More studies are needed to confirm the efficacy of MB07803 and the risks associated with its combination with other drugs, such as metformin.

SMAD7

SMAD7, a member of the I-Smad family, enhances proinflammatory cytokine expression by inhibiting the NF-κB and TGF-β1 signaling pathways and controlling DNA promoter activity [150, 151]. SMAD7 is involved in many diseases, such as cardiovascular diseases, autoimmune diseases, inflammatory diseases, cancers, and kidney diseases [152, 153]. Mongersen (also named GED-0301), an anti-SMAD7 oligonucleotide, decreased the production and activity of SMAD7 by promoting RNase H-mediated degradation of SMAD7 mRNA (Fig. 6). Mongersen was investigated for the treatment of Crohn’s disease (CD) and ulcerative colitis (UC) in clinical trials [154]. Mongersen exhibited good efficacy and safety for the treatment of CD in one phase 1 and three phase 2 clinical trials [155,156,157,158], but its development was suspended due to a lack of efficacy in phase 3 clinical trials. Notably, the chemical and pharmacological properties of mongersen are extremely unstable. Phase 1 clinical trials and phase 2 clinical trials used the same batch of mongersen, while phase 3 clinical trials used batches that were mostly different from phases 1 and 2. Most mongersen from phase 3 clinical trials did not inhibit SMAD7 expression in vitro [152, 159, 160]. In particular, the results of one phase 2 clinical trial were released after the phase 3 clinical trial. Mongersen showed good efficacy and safety in this trial. Mongersen decreased the percentage of CCR9−expressing CD45+ cells [157]. CCR9, a chemokine receptor induced by TGFβ1, induces white blood cells in lymphoid tissue to home to the intestinal mucosa, promoting the development of inflammation. CCR9-positive cells are markers of CD [161]. Mongersen inhibited SMAD7 expression because the batch used was also different from that used in the phase 3 clinical trial. This solved the stability problem of mongersen. Its development may be able to continue.

The potential value of miR-26 in the development of drugs to treat atherosclerosis

The administration of ADSCs overexpressing miR-26 (IV, 1 × 1010/mouse) for two weeks in a CAS model (apoE−/− mice fed a rich-fat diet (21% fat and 0.2% cholesterol) for 12 weeks) reduced carotid intima–media thickness (IMT), TC, TG, and LDL-C by 34.3%, 47.7%, 35.7%, and 22.9%, respectively [36]. Proprotein convertase subtilisin/kexin type 9 (PCSK9) knockout in apoE−/− mice fed a regular diet (6% fat and 0% cholesterol) for 6 months did not change IMT or TC levels. PCSK9 knockout in apoE−/− mice fed a Western diet (21% fat and 0.2% cholesterol) for 6 months reduced the IMT of valves, the root, the ascending aorta, and the brachiocephalic artery (BCA) by 25%, 18%, 9%, and 0%, respectively, but did not change TC levels [162]. Knockout of cardiotrophin-1 (CT-1), an IL-6 family member, reduced the media thickness of the carotid and aortic arteries by 33.1% and 20.5%, respectively, in mice [163]. These results suggest that the efficacy of ADSCs overexpressing miR-26 in reducing IMT is higher than that of those with PCSK9 or CT-1 knockout. Interestingly, many drugs targeting PCSK9, including alirocumab, evolocumab, inclisiran (also named ALN-60212), AZD8233, Civi-007, MK-0616, and LIB003 (also named lerodalcibep), have been approved or investigated in clinical trials [164,165,166,167,168,169,170]. The targeting of CT-1 with recombinant human CT-1 was investigated in clinical trials [171], and the results suggested that, like CT-1 and PCSK9, miR-26 has potential for drug development (Fig. 3). MiR-26 has not been studied in depth in atherosclerosis in vivo, so more studies are needed.

Materials for delivering miR-26

Multiple materials have been used to deliver miR-26 (Table 3), including ADSCs, hpMSC-EVs, adeno-associated virus (AAV) [172, 173], aptamer chimeras [174], EC-derived exos (EC-exos) [175], engineered exos [176], enhancer delivery systems/hydrogels [177], exos derived from 293T cells [178], HA-SS-PGEA [179], Lamp2b-exos [180], light polyethylene glycol (PEG) chimera nanoparticles [181], a mesoporous silicon nanoparticle (MSN)-polyethylenimine (PEI)-KALA peptide system [182], anti-Glypican 3 (GPC3) single-stranded variable fragment (scFv)-modified exos [183], a poly(l-lactic acid) (PLLA) nanofibrous (NF) scaffold attached to poly(lactic-co-glycolic acid) (PLGA) microspheres [184], PLGA nanocomplexes [185], polyetherimide-conjugated PEGylated gold nanocage ternary nanocomplexes (PPHAuNCs-TNCs) [186], a PEI/Quantum dot (QD) nanoparticle delivery vehicle [187], single-wall carbon nanotubes (SWCNTs) [188], PEG hydrogels [189], self-complementary AAV serotype 8 (scAAV8) [190], skeletal muscle satellite cell-derived exos (SMSCs-exos) engineered with RVG peptide [191], RALA peptide [192], and tumor cell-derived exos (TEXs) [193]. More studies are needed to confirm which materials are more suitable for mass production and clinical application.

Table 3 MiR-26 delivery systems and concentrations

Patents related to miR-26

Targeting miR-26 (a mimic) for the treatment of bone injury was investigated in preclinical trials by Air Force Medical University in China (patent number: CN104694542A). The delivery of miR-26 via human placental mesenchymal stem cell (hpMSC)-derived extracellular vesicles (EVs) for the treatment of coronavirus infection and autoimmune disease was investigated in preclinical trials by CHA University and Ts Cell Bio Co., Ltd. (patent number: WO2021225214). Targeting miR-26 (an inhibitor) for the treatment of ocular disease was investigated in preclinical trials by the University of Massachusetts (patent number: WO2021178668). MiR-26 has many target genes, and avoiding the occurrence of off-target effects requires further study. In addition, the diagnostic value of CTDSPL (host gene of miR-26a-1) in patients with diabetes (patent number: WO2015084862A1) or the chemotherapeutic resistance of triple-negative breast cancer (TNBC) (patent number: WO2023099889A1) was investigated at Wayne State University and Queens University of Belfast. Targeted inhibition of CTDSP1 (host gene of miR-26b) for the treatment of glioblastoma [194], brain tumors (patent number: CN113004356A), and neurodegenerative diseases (patent number: CN113004356A) was investigated at Purdue University, the University of California San Diego, the University of Texas at Austin, and Shanghai Jiao Tong University. However, the effect of miR-26 on host genes was not investigated.

Conclusions and future directions

MiR-26 decreases atherosclerosis development in vivo by suppressing the expression of multiple genes, including ACC1, ACC2, ACLY, ACSL3, ACSL4, ALDH3A2, ALPL, BMP2, CD36, COL1A1, CPT1A, CTGF, DGAT2, EHHADH, FAS, FBP1, GATA4, GSK3β, G6PC, Gys2, HMGA1, HMGB1, LDLR, LIPC, IL-1β, IL-6, JAG2, KCNJ2, MALT1, β-MHC, NF-κB, PCK1, PLCβ1, PYGL, RUNX2, SCD1, SMAD1, SMAD4, SMAD5, SMAD7, SREBF1, TAB3, TAK1, TCF7L2, and TNF-α. Several interesting and critical tasks remain to be explored: (1) miR-26 also promotes foam cell formation by reducing ABCA1 and ARL4C expression in vitro. More studies are needed to confirm the role of miR-26 in foam cell formation in vivo. (2) The miR-26 level in ADSC-exos is a potential biomarker for the diagnosis of lipid metabolism disorders and inflammatory states. However, biomarkers should be chosen considering disease status, the predisease state, or prognosis and should also be sensitive, specific, and superior to existing markers. (3) IFN-α therapy is effective in patients with HCC with low expression of miR-26. Therefore, miR-26 is expected to be a marker of IFN-α therapy for the treatment of HCC. In the future, we should determine the range of miR-26 expression levels in HCC to confirm when IFN-α therapy should be used. (4) IFN-α therapy has been approved for the treatment of multiple diseases, such as viral infections, hematological diseases, and various cancers. However, the role of miR-26 in IFN-α therapy for the treatment of these diseases is unclear. (5) The delivery of miR-26 to specific tissue- or cell-specific niches has been investigated by using different materials. It is still unclear which materials are most suitable for mass production and clinical applications. (6) MiR-26 target genes, such as ACC1, ACC2, COL1A1, CPT1A, DGAT2, FBP1, and SMAD7, are promising targets for drug development. Hypertriglyceridemia is a common adverse effect of ACC1/2 inhibitors that limits their clinical development. DGAT2 inhibitors can reduce the side effects of ACC1/2 inhibitors. Therefore, miR-26 is a potential candidate for development as a multitarget inhibitor of ACC1, ACC2, and DGAT2 expression. Like PCSK9 and CT-1, miR-26 is a promising candidate for drug development, but evidence on it is still limited. (7) COL1A1-specific agents, such as HT-100, RCT-01, and 68Ga-CBP8, are still being tested in clinical trials. No COL1A1-specific agents have been approved for use. (8) Perhexiline, a CPT1A and CPT2 inhibitor, has been approved, but the development of CPT1A inhibitors has not been successful. For example, the CPT1A inhibitor ETO was discontinued due to safety concerns, and teglicar was discontinued for unknown reasons. Investigating the underlying mechanism is conducive to the further development of teglicar. (9) The FBP1 inhibitor MB07803 exhibited good efficacy in phase 1b clinical trials, supporting its further development. Further development of its derivative CS-917 was discontinued due to efficacy and drug interactions. It is not known whether MB07803 causes drug interactions. (10) The SMAD7 inhibitor mongersen was discontinued due to poor efficacy. The quality of mongersen drug batches is not uniform, and some even have no inhibitory effect on SMAD7. Improving the quality of mongersen is the key to its development. (11) No miR-26 agents have been investigated in clinical trials. In addition, miR-26 has many target genes. We should consider which of these are the main target genes, where they are expressed, whether low doses of miR-26 cause changes in their expression, and how to deliver miR-26.

In summary, the detection of miR-26 in IFN-α therapy is worthy of further development. Targeting ACC1, ACC2, COL1A1, CPT1A, DGAT2, FBP1, and SMAD7 could improve the success rate of drug development. As research continues and technology advances, we believe that new drugs will be developed to combat atherosclerotic CVD. We have done our best not to overlook important contributions and to present the cited results as accurately as possible. If have, we apologize for the omission or error.

Availability of data and materials

Not applicable.

Abbreviations

CVD:

Cardiovascular disease

MI:

Myocardial infarction

IHD:

Ischemic heart disease

LDL:

Low-density lipoprotein

ROS:

Reactive oxygen species

VSMCs:

Vascular smooth muscle cells

MMP2:

Metalloproteinase 2

CTDSPL:

C-terminal domain RNA polymerase II polypeptide A small phosphatase-like

CTDSP2:

C-terminal domain RNA polymerase II polypeptide A small phosphatase 2

ABCA1:

ATP-binding cassette transporter A1

ACC1:

Acetyl-CoA-carboxylase 1

ACLY:

ATP-citrate lyase

ACSL3:

Acyl-CoA synthetase long-chain 3

ALDH3A2:

Aldehyde dehydrogenase 3 family member A2

FALDH:

Fatty aldehyde dehydrogenase

ALPL:

Alkaline phosphatase

ARL7:

ARF-like 7

ARL4C:

ADP-ribosylation factor-like 4C

β-MHC:

β-Myosin heavy chain

BMP2:

Bone morphogenetic protein 2

CD36:

Cluster of differentiation 36

COL1A1:

Collagen I

CPT1A:

Carnitine palmitoyl-transferase 1A

CTGF:

Connective tissue growth factor

DGAT2:

Diacylglycerol acyltransferase 2

EHHADH:

Enoyl-CoA hydratase/3-hydroxyacyl-CoA dehydrogenase

FASN:

Fatty acid synthase

FBP1:

Fructose-1,6-bisphosphatase

GATA4:

GATA binding protein 4

G6PC:

Glucose-6-phosphatase

GSK-3β:

Glycogen synthase kinase-3β

Gys2:

Glycogen synthase 2

HMGA1:

High mobility group A1

IL-1β:

Interleukin (IL)-1 beta

JAG2:

Jagged canonical Notch ligand 2

KCNJ2 :

Potassium inwardly rectifying channel subfamily J member 2

LDLR:

Low-density lipoprotein receptor

LIPC:

Lipase C hepatic type

MALT1:

Mucosa-associated lymphoid tissue lymphoma translocation protein 1

NF-κB:

Nuclear factor κB

PCK1:

Carboxykinase

PLCβ1:

Phospholipase C beta

PYGL:

Glycogen phosphorylase L

RUNX2:

Runt-related protein 2

SCD1:

Stearoyl-coenzyme A desaturase 1

SMAD1:

Mothers against decapentaplegic homolog 1

SREBF1:

Sterol regulatory element-binding transcription factor 1

TAB3:

TGFβ-activated kinase binding protein 3

TAK1:

TGFβ-activated kinase-1

TCF7L2:

Transcription factor 7 like 2

TNF-α:

Tumor necrosis factor alpha

TC:

Total cholesterol

CAS:

Carotid atherosclerosis

ADSCs-exos:

Adipose-derived stem cells

HL:

Hepatic lipase

ANGPTL3:

Angiopoietin-like 3

IFN-α:

Interferon-α

HCC:

Hepatocellular carcinoma

RCC:

Renal cell cancer

EZH2:

Zeste homolog 2

USP3:

Ubiquitin-specific protease 3

PNPT1:

Polyribonucleotide nucleotidyltransferase 1

MINOCA:

Nonobstructive coronary arteries

EHOA:

Erosive hand osteoarthritis

PsA:

Psoriatic arthritis

EC:

Endothelial cell

AF:

Atrial fibrillation

FA:

Fatty acid

IMT:

Carotid intima-media thickness

PCSK9:

Proprotein convertase subtilisin/kexin type 9

BCA:

Brachiocephalic artery

CT-1:

Cardiotrophin-1

AAV:

Adeno-associated virus

EC-exos:

EC-derived exos

PEG:

Light_polyethylene glycol

MSN:

Mesoporous silicon nanoparticle

PEI:

Polyethylenimine

GPC3:

Glypican 3

scFv:

Single-stranded variable fragment

PLLA:

Poly(l-lactic acid)

NF:

Nanofibrous

PLGA:

Poly(lactic-co-glycolic acid)

PPHAuNCs-TNCs:

PEGylated gold nanocage ternary nanocomplexes

QD:

Quantum dot

SWCNTs:

Single-wall carbon nanotubes

scAAV8:

Self-complementary AAV serotype 8

SMSCs-exos:

Skeletal muscle satellite cell-derived exos

TEXs:

Tumor cell-derived exos

hpMSC:

Human placental mesenchymal stem cell

EVs:

Extracellular vesicles

TNBC:

Triple-negative breast cancer

NASH:

Nonalcoholic steatohepatitis

ECM:

Extracellular matrix

PGP:

Peptide/fragments proline–glycine–proline

68Ga-CBP8:

68Ga-Collagen Binding Probe #8

DMD:

Duchenne muscular dystrophy

WT1:

Wilms’ tumor gene 1

Egr-1:

Early growth response-1

PRL-1:

Phosphatase of regenerating liver-1

IGFBP-1:

Insulin-like growth factor binding protein 1

p-eIF2α:

Phosphorylated eIF2α

NBDS:

Nonbulbar dermal sheath

COL2A1:

Collagen II

PMDA:

Pharmaceutical and Medical Devices Agency

LC-FAO:

Long-chain fatty acid

ETO:

Etomoxir

T2DM:

Type 2 diabetes mellitus

Acaa1a:

Acetyl-CoA acyltransferase

ACOX1:

Acyl-CoA oxidase 1

ADH1:

Alcohol dehydrogenase 1

CLU:

Clusterin

COX4I1:

Matrix cytochrome C oxidase subunit 4I1

CRAT:

Carnitine acetyltransferase

EHHADH:

Enoyl-CoA hydratase and 3-hydroxyacyl CoA dehydrogenase

FABP1:

Fatty acid binding protein 1

G6PD:

Glucose-6-phosphate dehydrogenase

MT1A:

Metallothionein 1A

PRKCB:

Protein kinase C beta

RPS9:

Ribosomal protein S9

SERCA2:

Sarco-endoplasmic reticulum calcium ATPase 2

TP53:

Tumor protein p53

UCP2:

Uncoupling protein 2

HPX:

Hemopexin

AP1S1:

Adaptor related protein complex 1 subunit sigma 1

SLC10A1:

Solute carrier family 10 member 1

SLCO1A1:

Solute carrier organic anion transporter family, member 1a1

TAGLN:

Transgelin

ASAH1:

N-Acylsphingosine amidohydrolase 1

ASNS:

Asparagine synthetase

ATF4:

Activating transcription factor 4

DEPP1:

DEPP autophagy regulator 1

EIF4E:

Eukaryotic translation initiation factor 4E

EIF4EBP1:

EIF4E binding protein 1

FNDC4:

Fibronectin type III domain containing 4

GDPD3:

Glycerophosphodiester phosphodiesterase domain containing 3

HPN:

Hepsin

INHBE:

Inhibin subunit beta E

KCNH2:

Potassium voltage-gated channel subfamily H member 2

LSS:

Lanosterol synthase

NROB2:

Nuclear receptor subfamily 0 group B member 2

NUPR1:

Nuclear protein 1, transcriptional regulator

SERPINA3:

Serpin family A member 3

TMPRSS2:

Transmembrane serine protease 2

WIPI1:

WD repeat domain, phosphoinositide interacting 1

FBG:

Fasting blood sugar

CD:

Crohn’s disease

UC:

Ulcerative colitis

References

  1. Schiano C, Balbi C, de Nigris F, Napoli C. Basic pathogenic mechanisms and epigenetic players promoted by extracellular vesicles in vascular damage. Int J Mol Sci. 2023;24(8):7509.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Chen W, Zhong Y, Feng N, Guo Z, Wang S, Xing D. New horizons in the roles and associations of COX-2 and novel natural inhibitors in cardiovascular diseases. Mol Med. 2021;27(1):123.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Chen W, Wang Y, Ren C, Yu S, Wang C, Xing J, Xu J, Yan S, Zhang T, Li Q, et al. The role of TNC in atherosclerosis and drug development opportunities. Int J Biol Sci. 2024; 20(1):127–136.

  4. Shinge SAU, Zhang D, Din AU, Yu F, Nie Y. Emerging Piezo1 signaling in inflammation and atherosclerosis; a potential therapeutic target. Int J Biol Sci. 2022;18(3):923–41.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Schiano C, Balbi C, Burrello J, Ruocco A, Infante T, Fiorito C, Panella S, Barile L, Mauro C, Vassalli G, et al. DNA methylation induced by circulating extracellular vesicles from acute coronary syndrome patients. Atherosclerosis. 2022;354:41–52.

    Article  CAS  PubMed  Google Scholar 

  6. Schiano C, D’Armiento M, Franzese M, Castaldo R, Saccone G, de Nigris F, Grimaldi V, Soricelli A, D’Armiento FP, Zullo F, et al. DNA methylation profile of the SREBF2 gene in human fetal aortas. J Vasc Res. 2022;59(1):61–8.

    Article  CAS  PubMed  Google Scholar 

  7. Denimal D, Monier S, Simoneau I, Duvillard L, Verges B, Bouillet B. HDL functionality in type 1 diabetes: enhancement of cholesterol efflux capacity in relationship with decreased HDL carbamylation after improvement of glycemic control. Cardiovasc Diabetol. 2022;21(1):154.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Chen W, Xing J, Liu X, Wang S, Xing D. The role and transformative potential of IL-19 in atherosclerosis. Cytokine Growth Factor Rev. 2021;62:70–82.

    Article  CAS  PubMed  Google Scholar 

  9. Chen W, Zhong Y, Yuan Y, Zhu M, Hu W, Liu N, Xing D. New insights into the suppression of inflammation and lipid accumulation by JAZF1. Genes Dis. 2023;10(6):2457–69.

    Article  CAS  PubMed  Google Scholar 

  10. Li X, Yang Y, Wang Z, Jiang S, Meng Y, Song X, Zhao L, Zou L, Li M, Yu T. Targeting non-coding RNAs in unstable atherosclerotic plaques: mechanism, regulation, possibilities, and limitations. Int J Biol Sci. 2021;17(13):3413–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Yang K, Xiao Q, Niu M, Pan X, Zhu X. Exosomes in atherosclerosis: convergence on macrophages. Int J Biol Sci. 2022;18(8):3266–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Rouland A, Masson D, Lagrost L, Verges B, Gautier T, Bouillet B. Role of apolipoprotein C1 in lipoprotein metabolism, atherosclerosis and diabetes: a systematic review. Cardiovasc Diabetol. 2022;21(1):272.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Li X, Pan X, Fu X, Yang Y, Chen J, Lin W. MicroRNA-26a: an emerging regulator of renal biology and disease. Kidney Blood Press Res. 2019;44(3):287–97.

    Article  CAS  PubMed  Google Scholar 

  14. Zhu Y, Lu Y, Zhang Q, Liu JJ, Li TJ, Yang JR, Zeng C, Zhuang SM. MicroRNA-26a/b and their host genes cooperate to inhibit the G1/S transition by activating the pRb protein. Nucleic Acids Res. 2012;40(10):4615–25.

    Article  CAS  PubMed  Google Scholar 

  15. Li H, Wang Y, Song Y. MicroRNA-26b inhibits the immune response to Mycobacterium tuberculosis (M.tb) infection in THP-1 cells via targeting TGFbeta-activated kinase-1 (TAK1), a promoter of the NF-kappaB pathway. Int J Clin Exp Pathol. 2018;11(3):1218–27.

    PubMed  PubMed Central  Google Scholar 

  16. Zhao N, Wang R, Zhou L, Zhu Y, Gong J, Zhuang SM. MicroRNA-26b suppresses the NF-kappaB signaling and enhances the chemosensitivity of hepatocellular carcinoma cells by targeting TAK1 and TAB3. Mol Cancer. 2014;13:35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Wei C, Kim IK, Kumar S, Jayasinghe S, Hong N, Castoldi G, Catalucci D, Jones WK, Gupta S. NF-kappaB mediated miR-26a regulation in cardiac fibrosis. J Cell Physiol. 2013;228(7):1433–42.

    Article  CAS  PubMed  Google Scholar 

  18. Cai ZG, Wu HB, Xu XP, Li W. Down-regulation of miR-26 plays essential roles in TGFbeta-induced EMT. Cell Biol Int. 2018. https://doi.org/10.1002/cbin.11029.

    Article  PubMed  Google Scholar 

  19. Wu W, Shang YQ, Dai SL, Yi F, Wang XC. MiR-26a regulates vascular smooth muscle cell calcification in vitro through targeting CTGF. Bratisl Lek Listy. 2017;118(8):499–503.

    CAS  PubMed  Google Scholar 

  20. Nigam V, Sievers HH, Jensen BC, Sier HA, Simpson PC, Srivastava D, Mohamed SA. Altered microRNAs in bicuspid aortic valve: a comparison between stenotic and insufficient valves. J Heart Valve Dis. 2010;19(4):459–65.

    PubMed  PubMed Central  Google Scholar 

  21. Wu W, Cheng L, Wang J, Yang C, Shang Y. miRNA-26a reduces vascular smooth muscle cell calcification by regulating connective tissue growth factor. Nan Fang Yi Ke Da Xue Xue Bao. 2022;42(9):1303–8.

    CAS  PubMed  Google Scholar 

  22. Fu X, Dong B, Tian Y, Lefebvre P, Meng Z, Wang X, Pattou F, Han W, Wang X, Lou F, et al. MicroRNA-26a regulates insulin sensitivity and metabolism of glucose and lipids. J Clin Invest. 2015;125(6):2497–509.

    Article  PubMed  PubMed Central  Google Scholar 

  23. 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(10):1472–9.

    Article  CAS  PubMed  Google Scholar 

  24. He Y, Liu H, Jiang L, Rui B, Mei J, Xiao H. miR-26 induces apoptosis and inhibits autophagy in non-small cell lung cancer cells by suppressing TGF-beta1-JNK signaling pathway. Front Pharmacol. 2018;9:1509.

    Article  CAS  PubMed  Google Scholar 

  25. Yin J, Zhao X, Chen X, Shen G. Emodin suppresses hepatocellular carcinoma growth by regulating macrophage polarization via microRNA-26a/transforming growth factor beta 1/protein kinase B. Bioengineered. 2022;13(4):9548–63.

    Article  PubMed  Google Scholar 

  26. Chen CY, Chang JT, Ho YF, Shyu AB. MiR-26 down-regulates TNF-alpha/NF-kappaB signalling and IL-6 expression by silencing HMGA1 and MALT1. Nucleic Acids Res. 2016;44(8):3772–87.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Chen W, Wu Y, Wang J, Yu W, Shen X, Zhao K, Liang B, Hu X, Wang S, Jiang H, et al. Clinical advances in TNC delivery vectors and their conjugate agents. Pharmacol Ther. 2023;253:108577.

  28. Liu C, Lou CH, Shah V, Ritter R, Talley J, Soibam B, Benham A, Zhu H, Perez E, Shieh YE, et al. Identification of microRNAs and microRNA targets in Xenopus gastrulae: The role of miR-26 in the regulation of Smad1. Dev Biol. 2016;409(1):26–38.

    Article  CAS  PubMed  Google Scholar 

  29. Kwon Y, Kim Y, Eom S, Kim M, Park D, Kim H, Noh K, Lee H, Lee YS, Choe J, et al. MicroRNA-26a/-26b-COX-2-MIP-2 loop regulates allergic inflammation and allergic inflammation-promoted enhanced tumorigenic and metastatic potential of cancer cells. J Biol Chem. 2015;290(22):14245–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Banik SK, Baishya S, Das Talukdar A, Choudhury MD. Network analysis of atherosclerotic genes elucidates druggable targets. BMC Med Genomics. 2022;15(1):42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Dlouha D, Blaha M, Huckova P, Lanska V, Hubacek JA, Blaha V. Long-term LDL-apheresis treatment and dynamics of circulating miRNAs in patients with severe familial hypercholesterolemia. Genes (Basel). 2023;14(8):1571.

    Article  CAS  PubMed  Google Scholar 

  32. Feng M, Xu D, Wang L. miR-26a inhibits atherosclerosis progression by targeting TRPC3. Cell Biosci. 2018;8:4.

    Article  PubMed  PubMed Central  Google Scholar 

  33. Marketou M, Kontaraki J, Kalogerakos P, Plevritaki A, Chlouverakis G, Kassotakis S, Maragkoudakis S, Danelatos C, Zervakis S, Savva E, et al. Differences in MicroRNA expression in pericoronary adipose tissue in coronary artery disease compared to severe valve dysfunction. Angiology. 2023;74(1):22–30.

    Article  CAS  PubMed  Google Scholar 

  34. Volny O, Kasickova L, Coufalova D, Cimflova P, Novak J. microRNAs in cerebrovascular disease. Adv Exp Med Biol. 2015;888:155–95.

    Article  PubMed  Google Scholar 

  35. Zhang Y, Qin W, Zhang L, Wu X, Du N, Hu Y, Li X, Shen N, Xiao D, Zhang H, et al. MicroRNA-26a prevents endothelial cell apoptosis by directly targeting TRPC6 in the setting of atherosclerosis. Sci Rep. 2015;5:9401.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Han G, Li H, Guo H, Yi C, Yu B, Lin Y, Zheng B, He D. The roles and mechanisms of miR-26 derived from exosomes of adipose-derived stem cells in the formation of carotid atherosclerotic plaque. Ann Transl Med. 2022;10(20):1134.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Santamarina-Fojo S, Gonzalez-Navarro H, Freeman L, Wagner E, Nong Z. Hepatic lipase, lipoprotein metabolism, and atherogenesis. Arterioscler Thromb Vasc Biol. 2004;24(10):1750–4.

    Article  CAS  PubMed  Google Scholar 

  38. Dijk W, Di Filippo M, Kooijman S, van Eenige R, Rimbert A, Caillaud A, Thedrez A, Arnaud L, Pronk A, Garcon D, et al. Identification of a gain-of-function LIPC variant as a novel cause of familial combined hypocholesterolemia. Circulation. 2022;146(10):724–39.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Adjuvant IFN-α for patients underwent curative surgery for HCC with a low miR-26 expression. https://classic.clinicaltrials.gov/ct2/show/NCT01681446?term=NCT01681446&draw=2&rank=1.

  40. Ji J, Shi J, Budhu A, Yu Z, Forgues M, Roessler S, Ambs S, Chen Y, Meltzer PS, Croce CM, et al. MicroRNA expression, survival, and response to interferon in liver cancer. N Engl J Med. 2009;361(15):1437–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Li C, Li Y, Lu Y, Niu Z, Zhao H, Peng Y, Li M. miR-26 family and its target genes in tumorigenesis and development. Crit Rev Oncol Hematol. 2021;157: 103124.

    Article  PubMed  Google Scholar 

  42. Ji J, Yu L, Yu Z, Forgues M, Uenishi T, Kubo S, Wakasa K, Zhou J, Fan J, Tang ZY, et al. Development of a miR-26 companion diagnostic test for adjuvant interferon-alpha therapy in hepatocellular carcinoma. Int J Biol Sci. 2013;9(3):303–12.

    Article  PubMed  PubMed Central  Google Scholar 

  43. Vidal P. Interferon alpha in cancer immunoediting: From elimination to escape. Scand J Immunol. 2020;91(5): e12863.

    Article  PubMed  Google Scholar 

  44. Shi W, Yao X, Fu Y, Wang Y. Interferon-alpha and its effects on cancer cell apoptosis. Oncol Lett. 2022;24(1):235.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Lewczuk N, Zdebik A, Boguslawska J. Interferon alpha 2a and 2b in ophthalmology: a review. J Interferon Cytokine Res. 2019;39(5):259–72.

    Article  CAS  PubMed  Google Scholar 

  46. Levy Z, Rachmani R, Trestman S, Dvir A, Shaish A, Ravid M, Harats D. Low-dose interferon-alpha accelerates atherosclerosis in an LDL receptor-deficient mouse model. Eur J Intern Med. 2003;14(8):479–83.

    Article  CAS  PubMed  Google Scholar 

  47. Kirchler C, Husar-Memmer E, Rappersberger K, Thaler K, Fritsch-Stork R. Type I Interferon as cardiovascular risk factor in systemic and cutaneous lupus erythematosus: a systematic review. Autoimmun Rev. 2021;20(5): 102794.

    Article  CAS  PubMed  Google Scholar 

  48. Wang G, Sun Y, He Y, Ji C, Hu B, Sun Y. miR-26a promoted by interferon-alpha inhibits hepatocellular carcinoma proliferation and migration by blocking EZH2. Genet Test Mol Biomarkers. 2015;19(1):30–6.

    Article  PubMed  Google Scholar 

  49. Gao S, Li J, Song L, Wu J, Huang W. Influenza A virus-induced downregulation of miR-26a contributes to reduced IFNalpha/beta production. Virol Sin. 2017;32(4):261–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Gauthier BR, Cobo-Vuilleumier N, Lopez-Noriega L. Roles of extracellular vesicles associated non-coding RNAs in Diabetes Mellitus. Front Endocrinol (Lausanne). 2022;13:1057407.

    Article  PubMed  Google Scholar 

  51. Montone RA, Cosentino N, Graziani F, Gorla R, Del Buono MG, La Vecchia G, Rinaldi R, Marenzi G, Bartorelli AL, De Marco F, et al. Precision medicine versus standard of care for patients with myocardial infarction with non-obstructive coronary arteries (MINOCA): rationale and design of the multicentre, randomised PROMISE trial. EuroIntervention. 2022;18(11):e933–9.

    Article  PubMed  PubMed Central  Google Scholar 

  52. Durand T, Jacob S, Lebouil L, Douzane H, Lestaevel P, Rahimian A, Psimaras D, Feuvret L, Leclercq D, Brochet B, et al. EpiBrainRad: an epidemiologic study of the neurotoxicity induced by radiotherapy in high grade glioma patients. BMC Neurol. 2015;15:261.

    Article  PubMed  PubMed Central  Google Scholar 

  53. microRNA in erosive hand osteoarthritis and psoriatic arthritis. https://clinicaltrials.gov/study/NCT05683769?term=NCT05683769&rank=1.

  54. PSMA Guided Approach for bIoCHEmical relapse after prostatectomy-PSICHE (PSICHE). https://classic.clinicaltrials.gov/ct2/show/NCT05022914?term=NCT05022914&draw=2&rank=1.

  55. Breast milk: influence of the micro-transcriptome profile on atopy in children over time (IMPACT). https://classic.clinicaltrials.gov/ct2/show/NCT04017520?term=NCT04017520&draw=2&rank=1.

  56. Li X, Wang J, Wu C, Lu X, Huang J. MicroRNAs involved in the TGF-beta signaling pathway in atherosclerosis. Biomed Pharmacother. 2022;146: 112499.

    Article  CAS  PubMed  Google Scholar 

  57. Leeper NJ, Raiesdana A, Kojima Y, Chun HJ, Azuma J, Maegdefessel L, Kundu RK, Quertermous T, Tsao PS, Spin JM. MicroRNA-26a is a novel regulator of vascular smooth muscle cell function. J Cell Physiol. 2011;226(4):1035–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Icli B, Dorbala P, Feinberg MW. An emerging role for the miR-26 family in cardiovascular disease. Trends Cardiovasc Med. 2014;24(6):241–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Vickers KC, Rader DJ. Nuclear receptors and microRNA-144 coordinately regulate cholesterol efflux. Circ Res. 2013;112(12):1529–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Davalos A, Fernandez-Hernando C. From evolution to revolution: miRNAs as pharmacological targets for modulating cholesterol efflux and reverse cholesterol transport. Pharmacol Res. 2013;75:60–72.

    Article  CAS  PubMed  Google Scholar 

  61. Novak J, Olejnickova V, Tkacova N, Santulli G. Mechanistic role of MicroRNAs in coupling lipid metabolism and atherosclerosis. Adv Exp Med Biol. 2015;887:79–100.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Hu M, Han T, Pan Q, Ni D, Gao F, Wang L, Ren H, Zhang X, Jiao H, Wang Y, et al. The GR-gp78 Pathway is involved in Hepatic Lipid Accumulation Induced by Overexpression of 11beta-HSD1. Int J Biol Sci. 2022;18(8):3107–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Liu F, Zhu X, Jiang X, Li S, Lv Y. Transcriptional control by HNF-1: emerging evidence showing its role in lipid metabolism and lipid metabolism disorders. Genes Dis. 2022;9(5):1248–57.

    Article  CAS  PubMed  Google Scholar 

  64. Alkhouri N, Herring R, Kabler H, Kayali Z, Hassanein T, Kohli A, Huss RS, Zhu Y, Billin AN, Damgaard LH, et al. Safety and efficacy of combination therapy with semaglutide, cilofexor and firsocostat in patients with non-alcoholic steatohepatitis: a randomised, open-label phase II trial. J Hepatol. 2022;77(3):607–18.

    Article  CAS  PubMed  Google Scholar 

  65. Study to evaluate the pharmacodynamic effects of a single oral dose of GS-0976 (NDI-010976) in healthy adult subjects. https://classic.clinicaltrials.gov/ct2/show/NCT02876796?term=NCT02876796&draw=2&rank=1.

  66. Calle RA, Amin NB, Carvajal-Gonzalez S, Ross TT, Bergman A, Aggarwal S, Crowley C, Rinaldi A, Mancuso J, Aggarwal N, et al. ACC inhibitor alone or co-administered with a DGAT2 inhibitor in patients with non-alcoholic fatty liver disease: two parallel, placebo-controlled, randomized phase 2a trials. Nat Med. 2021;27(10):1836–48.

    Article  CAS  PubMed  Google Scholar 

  67. Study of PF-05221304 in subjects with varying degrees of hepatic impairment. https://classic.clinicaltrials.gov/ct2/show/NCT03309202?term=NCT03309202&draw=1&rank=1.

  68. Li ZY, Wu G, Qiu C, Zhou ZJ, Wang YP, Song GH, Xiao C, Zhang X, Deng GL, Wang RT, et al. Mechanism and therapeutic strategy of hepatic TM6SF2-deficient non-alcoholic fatty liver diseases via in vivo and in vitro experiments. World J Gastroenterol. 2022;28(25):2937–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Lally JSV, Ghoshal S, DePeralta DK, Moaven O, Wei L, Masia R, Erstad DJ, Fujiwara N, Leong V, Houde VP, et al. Inhibition of acetyl-CoA carboxylase by phosphorylation or the inhibitor ND-654 suppresses lipogenesis and hepatocellular carcinoma. Cell Metab. 2019;29(1):174-182 e175.

    Article  CAS  PubMed  Google Scholar 

  70. Chu Q, An J, Liu P, Song Y, Zhai X, Yang R, Niu J, Yang C, Li B. Repurposing a tricyclic antidepressant in tumor and metabolism disease treatment through fatty acid uptake inhibition. J Exp Med. 2023;220(3): e20221316.

    Article  CAS  PubMed  Google Scholar 

  71. Svensson RU, Parker SJ, Eichner LJ, Kolar MJ, Wallace M, Brun SN, Lombardo PS, Van Nostrand JL, Hutchins A, Vera L, et al. Inhibition of acetyl-CoA carboxylase suppresses fatty acid synthesis and tumor growth of non-small-cell lung cancer in preclinical models. Nat Med. 2016;22(10):1108–19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Zhang XJ, Cai J, Li H. Targeting ACC for NASH resolution. Trends Mol Med. 2022;28(1):5–7.

    Article  PubMed  Google Scholar 

  73. Chitraju C, Walther TC, Farese RV Jr. The triglyceride synthesis enzymes DGAT1 and DGAT2 have distinct and overlapping functions in adipocytes. J Lipid Res. 2019;60(6):1112–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Luo W, Wang H, Ren L, Lu Z, Zheng Q, Ding L, Xie H, Wang R, Yu C, Lin Y, et al. Adding fuel to the fire: the lipid droplet and its associated proteins in cancer progression. Int J Biol Sci. 2022;18(16):6020–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Amin NB, Darekar A, Anstee QM, Wong VW, Tacke F, Vourvahis M, Lee DS, Charlton M, Alkhouri N, Nakajima A, et al. Efficacy and safety of an orally administered DGAT2 inhibitor alone or coadministered with a liver-targeted ACC inhibitor in adults with non-alcoholic steatohepatitis (NASH): rationale and design of the phase II, dose-ranging, dose-finding, randomised, placebo-controlled MIRNA (Metabolic Interventions to Resolve NASH with fibrosis) study. BMJ Open. 2022;12(3): e056159.

    Article  PubMed  PubMed Central  Google Scholar 

  76. Effect of hepatic impairment on the pharmacokinetics, safety and tolerability of PF-06865571 in subjects with hepatic impairment and in healthy subjects. https://classic.clinicaltrials.gov/ct2/show/NCT04091061?term=NCT04091061&draw=2&rank=1.

  77. Drug-drug interaction study to evaluate the effect of PF-06865571 on metformin pharmacokinetics. https://classic.clinicaltrials.gov/ct2/show/NCT03593707?term=NCT03593707&draw=1&rank=1.

  78. Loomba R, Morgan E, Watts L, Xia S, Hannan LA, Geary RS, Baker BF, Bhanot S. Novel antisense inhibition of diacylglycerol O-acyltransferase 2 for treatment of non-alcoholic fatty liver disease: a multicentre, double-blind, randomised, placebo-controlled phase 2 trial. Lancet Gastroenterol Hepatol. 2020;5(9):829–38.

    Article  PubMed  Google Scholar 

  79. Filipski KJ, Edmonds DJ, Garnsey MR, Smaltz DJ, Coffman K, Futatsugi K, Lee J, O’Neil SV, Wright A, Nason D, et al. Design of next-generation DGAT2 inhibitor PF-07202954 with longer predicted half-life. ACS Med Chem Lett. 2023;14(10):1427–33.

    Article  CAS  PubMed  Google Scholar 

  80. Futatsugi K, Cabral S, Kung DW, Huard K, Lee E, Boehm M, Bauman J, Clark RW, Coffey SB, Crowley C, et al. Discovery of ervogastat (PF-06865571): a potent and selective inhibitor of diacylglycerol acyltransferase 2 for the treatment of non-alcoholic steatohepatitis. J Med Chem. 2022;65(22):15000–13.

    Article  CAS  PubMed  Google Scholar 

  81. Amin NB, Carvajal-Gonzalez S, Purkal J, Zhu T, Crowley C, Perez S, Chidsey K, Kim AM, Goodwin B. Targeting diacylglycerol acyltransferase 2 for the treatment of nonalcoholic steatohepatitis. Sci Transl Med. 2019;11(520):eaav9701.

    Article  CAS  PubMed  Google Scholar 

  82. Almanza A, Mnich K, Blomme A, Robinson CM, Rodriguez-Blanco G, Kierszniowska S, McGrath EP, Le Gallo M, Pilalis E, Swinnen JV, et al. Regulated IRE1alpha-dependent decay (RIDD)-mediated reprograming of lipid metabolism in cancer. Nat Commun. 2022;13(1):2493.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Sabnis RW. Benzimidazolone derivatives as DGAT2 inhibitors for treating diseases. ACS Med Chem Lett. 2022;13(7):1004–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Cao T, Ni R, Ding W, Ji X, Li L, Liao G, Lu Y, Fan GC, Zhang Z, Peng T. MLKL-mediated necroptosis is a target for cardiac protection in mouse models of type-1 diabetes. Cardiovasc Diabetol. 2022;21(1):165.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Kisling A, Lust RM, Katwa LC. What is the role of peptide fragments of collagen I and IV in health and disease? Life Sci. 2019;228:30–4.

    Article  CAS  PubMed  Google Scholar 

  86. You M, Liu Y, Wang B, Li L, Zhang H, He H, Zhou Q, Cao T, Wang L, Zhao Z, et al. Asprosin induces vascular endothelial-to-mesenchymal transition in diabetic lower extremity peripheral artery disease. Cardiovasc Diabetol. 2022;21(1):25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Shi R, Zhang Z, Zhu A, Xiong X, Zhang J, Xu J, Sy MS, Li C. Targeting type I collagen for cancer treatment. Int J Cancer. 2022;151(5):665–83.

    Article  CAS  PubMed  Google Scholar 

  88. Ponticos M, Smith BD. Extracellular matrix synthesis in vascular disease: hypertension, and atherosclerosis. J Biomed Res. 2014;28(1):25–39.

    Article  CAS  PubMed  Google Scholar 

  89. Nikolov A, Popovski N. Extracellular matrix in heart disease: focus on circulating collagen type I and III derived peptides as biomarkers of myocardial fibrosis and their potential in the prognosis of heart failure: a concise review. Metabolites. 2022;12(4):297.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Marini JC, Forlino A, Bachinger HP, Bishop NJ, Byers PH, Paepe A, Fassier F, Fratzl-Zelman N, Kozloff KM, Krakow D, et al. Osteogenesis imperfecta. Nat Rev Dis Primers. 2017;3:17052.

    Article  PubMed  Google Scholar 

  91. Montesi SB, Desogere P, Fuchs BC, Caravan P. Molecular imaging of fibrosis: recent advances and future directions. J Clin Invest. 2019;129(1):24–33.

    Article  PubMed  PubMed Central  Google Scholar 

  92. Desogere P, Tapias LF, Hariri LP, Rotile NJ, Rietz TA, Probst CK, Blasi F, Day H, Mino-Kenudson M, Weinreb P, et al. Type I collagen-targeted PET probe for pulmonary fibrosis detection and staging in preclinical models. Sci Transl Med. 2017;9(384):eaaf4696.

    Article  PubMed  PubMed Central  Google Scholar 

  93. Bellaye PS, Beltramo G, Burgy O, Collin B, Cochet A, Bonniaud P. Measurement of hypoxia in the lung in idiopathic pulmonary fibrosis: a matter of control. Eur Respir J. 2022;59(3):2102711.

    Article  PubMed  PubMed Central  Google Scholar 

  94. Evaluation of PET probe [68Ga]CBP8 in the detection of radiation induced tissue injury. https://classic.clinicaltrials.gov/ct2/show/NCT04485286?term=NCT04485286&draw=2&rank=1.

  95. Preliminary evaluation of [68Ga]CBP8 in healthy individuals, lung cancer, and idiopathic pulmonary fibrosis patients. https://classic.clinicaltrials.gov/ct2/show/NCT03535545?term=NCT03535545&draw=2&rank=1.

  96. Molecular imaging of myocardial fibrosis in cardiac amyloidosis. https://classic.clinicaltrials.gov/ct2/show/NCT05619302?term=NCT05619302&draw=2&rank=1.

  97. Collagen-targeted PET imaging for early interstitial lung disease. https://classic.clinicaltrials.gov/ct2/show/NCT05417776?term=NCT05417776&draw=2&rank=1.

  98. Imaging evaluation of PLN-74809 in participants with IPF (PLN-74809) (PLN-74809). https://classic.clinicaltrials.gov/ct2/show/NCT05621252?term=NCT05621252&draw=2&rank=1.

  99. Evaluation of PET probe [64]Cu-macrin in cardiovascular disease, cancer and sarcoidosis. https://classic.clinicaltrials.gov/ct2/show/NCT04843891?term=NCT04843891&draw=2&rank=1.

  100. Izquierdo-Garcia D, Desogere P, Fur ML, Shuvaev S, Zhou IY, Ramsay I, Lanuti M, Catalano OA, Catana C, Caravan P, et al. Biodistribution, Dosimetry, and pharmacokinetics of (68)Ga-CBP8: A type I collagen-targeted PET probe. J Nucl Med. 2023;64(5):775–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Granot I, Halevy O, Hurwitz S, Pines M. Halofuginone: an inhibitor of collagen type I synthesis. Biochim Biophys Acta. 1993;1156(2):107–12.

    Article  CAS  PubMed  Google Scholar 

  102. Pines M, Knopov V, Genina O, Lavelin I, Nagler A. Halofuginone, a specific inhibitor of collagen type I synthesis, prevents dimethylnitrosamine-induced liver cirrhosis. J Hepatol. 1997;27(2):391–8.

    Article  CAS  PubMed  Google Scholar 

  103. Halofuginone—processa pharmaceuticals. https://0-adisinsight-springer-com.brum.beds.ac.uk/drugs/800013520.

  104. Akashi Reports Positive Clinical Data on HT-100. https://www.akashirx.com/akashi-reports-positive-clinical-data-on-ht-100/.

  105. Dosing and enrollment in HT-100 trial suspended. https://www.akashirx.com/dosing-and-enrollment-in-ht-100-trial-suspended/.

  106. HT-100 Long-term Study in DMD patients who completed HALO-DMD-02. https://classic.clinicaltrials.gov/ct2/show/NCT02525302?term=HT-100&draw=2&rank=1.

  107. Safety, tolerability, and pharmacokinetics of single and multiple doses of HT-100 in Duchenne muscular dystrophy. https://classic.clinicaltrials.gov/ct2/show/NCT01847573?term=HT-100&draw=2&rank=3.

  108. Open label extension study of HT-100 in patients with DMD. https://classic.clinicaltrials.gov/ct2/show/NCT01978366?term=HT-100&draw=2&rank=2.

  109. Halofuginone hydrobromide in treating patients with progressive advanced solid tumors. https://classic.clinicaltrials.gov/ct2/show/NCT00027677?term=halofuginone+hydrobromide&draw=2&rank=2.

  110. Halofuginone hydrobromide in treating patients with HIV-related Kaposi's sarcoma. https://classic.clinicaltrials.gov/ct2/show/NCT00064142?term=halofuginone+hydrobromide&draw=2&rank=1.

  111. Luo L, Gao Y, Yang C, Shao Z, Wu X, Li S, Xiong L, Chen C. Halofuginone attenuates intervertebral discs degeneration by suppressing collagen I production and inactivating TGFbeta and NF-small ka, CyrillicB pathway. Biomed Pharmacother. 2018;101:745–53.

    Article  CAS  PubMed  Google Scholar 

  112. Pinthus JH, Sheffer Y, Nagler A, Fridman E, Mor Y, Genina O, Pines M. Inhibition of Wilms tumor xenograft progression by halofuginone is accompanied by activation of WT-1 gene expression. J Urol. 2005;174(4 Pt 2):1527–31.

    Article  CAS  PubMed  Google Scholar 

  113. Nagler A, Ohana M, Shibolet O, Shapira MY, Alper R, Vlodavsky I, Pines M, Ilan Y. Suppression of hepatocellular carcinoma growth in mice by the alkaloid coccidiostat halofuginone. Eur J Cancer. 2004;40(9):1397–403.

    Article  CAS  PubMed  Google Scholar 

  114. McGaha TL, Phelps RG, Spiera H, Bona C. Halofuginone, an inhibitor of type-I collagen synthesis and skin sclerosis, blocks transforming-growth-factor-beta-mediated Smad3 activation in fibroblasts. J Invest Dermatol. 2002;118(3):461–70.

    Article  CAS  PubMed  Google Scholar 

  115. Leiba M, Cahalon L, Shimoni A, Lider O, Zanin-Zhorov A, Hecht I, Sela U, Vlodavsky I, Nagler A. Halofuginone inhibits NF-kappaB and p38 MAPK in activated T cells. J Leukoc Biol. 2006;80(2):399–406.

    Article  CAS  PubMed  Google Scholar 

  116. Herman JD, Pepper LR, Cortese JF, Estiu G, Galinsky K, Zuzarte-Luis V, Derbyshire ER, Ribacke U, Lukens AK, Santos SA, et al. The cytoplasmic prolyl-tRNA synthetase of the malaria parasite is a dual-stage target of febrifugine and its analogs. Sci Transl Med. 2015;7(288):288ra277.

    Article  Google Scholar 

  117. Gnainsky Y, Spira G, Paizi M, Bruck R, Nagler A, Genina O, Taub R, Halevy O, Pines M. Involvement of the tyrosine phosphatase early gene of liver regeneration (PRL-1) in cell cycle and in liver regeneration and fibrosis effect of halofuginone. Cell Tissue Res. 2006;324(3):385–94.

    Article  CAS  PubMed  Google Scholar 

  118. RCT-01 FOR repairing tendons. https://replicel.com/products/rct-01.

  119. RepliCel's successful RCT-01 tendon repair clinical trial shows signs of healing chronic tendon problems. https://www.prnewswire.com/news-releases/replicels-successful-rct-01-tendon-repair-clinical-trial-shows-signs-of-healing-chronic-tendon-problems-617272013.html.

  120. Safety and Efficacy of RCT-01 in men and women with unilateral, chronic achilles tendinosis (ReaCT). https://classic.clinicaltrials.gov/ct2/show/NCT02330146?term=NCT02330146&draw=2&rank=1.

  121. RepliCel launches preparations for second skin rejuvenation and tendon regeneration clinical studies. https://replicel.com/news/replicel-launches-preparations-for-second-skin-rejuvenation-and-tendon-regeneration-clinical-studies.

  122. RepliCel’s manufacturing inspected by Japan’s PMDA. https://replicel.com/news/replicel-s-manufacturing-inspected-by-japan-s-pmda.

  123. RepliCel provides corporate update. https://replicel.com/news/replicel-provides-corporate-update.

  124. Raud B, McGuire PJ, Jones RG, Sparwasser T, Berod L. Fatty acid metabolism in CD8(+) T cell memory: challenging current concepts. Immunol Rev. 2018;283(1):213–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Liang K. Mitochondrial CPT1A: insights into structure, function, and basis for drug development. Front Pharmacol. 2023;14:1160440.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Etomoxir. https://0-adisinsight-springer-com.brum.beds.ac.uk/drugs/800000032.

  127. Vickers AE. Characterization of hepatic mitochondrial injury induced by fatty acid oxidation inhibitors. Toxicol Pathol. 2009;37(1):78–88.

    Article  CAS  PubMed  Google Scholar 

  128. Rupp H, Zarain-Herzberg A, Maisch B. The use of partial fatty acid oxidation inhibitors for metabolic therapy of angina pectoris and heart failure. Herz. 2002;27(7):621–36.

    Article  PubMed  Google Scholar 

  129. Murray AJ, Panagia M, Hauton D, Gibbons GF, Clarke K. Plasma free fatty acids and peroxisome proliferator-activated receptor alpha in the control of myocardial uncoupling protein levels. Diabetes. 2005;54(12):3496–502.

    Article  CAS  PubMed  Google Scholar 

  130. Gerondaes P, Alberti KG, Agius L. Interactions of inhibitors of carnitine palmitoyltransferase I and fibrates in cultured hepatocytes. Biochem J. 1988;253(1):169–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Open-label study of perhexiline in patients with hypertrophic cardiomyopathy and moderate to severe heart failure. https://classic.clinicaltrials.gov/ct2/show/study/NCT02862600?term=NCT02862600&draw=2&rank=1.

  132. Sawada H, Takami K, Asahi S. A toxicogenomic approach to drug-induced phospholipidosis: analysis of its induction mechanism and establishment of a novel in vitro screening system. Toxicol Sci. 2005;83(2):282–92.

    Article  CAS  PubMed  Google Scholar 

  133. Saheb Sharif-Askari N, Saheb Sharif-Askari F, Mdkhana B, Al Heialy S, Ratemi E, Alghamdi M, Abusnana S, Kashour T, Hamid Q, Halwani R. Effect of common medications on the expression of SARS-CoV-2 entry receptors in liver tissue. Arch Toxicol. 2020;94(12):4037–41.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Ren Z, Chen S, Qin X, Li F, Guo L. Study of the roles of cytochrome P450 (CYPs) in the metabolism and cytotoxicity of perhexiline. Arch Toxicol. 2022;96(12):3219–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Perrin MJ, Kuchel PW, Campbell TJ, Vandenberg JI. Drug binding to the inactivated state is necessary but not sufficient for high-affinity binding to human ether-a-go-go-related gene channels. Mol Pharmacol. 2008;74(5):1443–52.

    Article  CAS  PubMed  Google Scholar 

  136. Balgi AD, Fonseca BD, Donohue E, Tsang TC, Lajoie P, Proud CG, Nabi IR, Roberge M. Screen for chemical modulators of autophagy reveals novel therapeutic inhibitors of mTORC1 signaling. PLoS ONE. 2009;4(9): e7124.

    Article  PubMed  PubMed Central  Google Scholar 

  137. Atienzar F, Gerets H, Dufrane S, Tilmant K, Cornet M, Dhalluin S, Ruty B, Rose G, Canning M. Determination of phospholipidosis potential based on gene expression analysis in HepG2 cells. Toxicol Sci. 2007;96(1):101–14.

    Article  CAS  PubMed  Google Scholar 

  138. Cacciola NA, Sepe F, Fioriniello S, Petillo O, Margarucci S, Scivicco M, Peluso G, Balestrieri A, Bifulco G, Restucci B, et al. The carnitine palmitoyltransferase 1A inhibitor teglicar shows promising antitumour activity against canine mammary cancer cells by inducing apoptosis. Pharmaceuticals (Basel). 2023;16(7):987.

    Article  CAS  PubMed  Google Scholar 

  139. Abstracts of the 44th Annual Meeting of the European Association for the Study of Diabetes, 8–11 September 2008, Rome, Italy. Diabetologia 2008, 51(Suppl 1):S5–564.

  140. Teglicar. https://0-adisinsight-springer-com.brum.beds.ac.uk/drugs/800017026.

  141. van Poelje PD, Potter SC, Erion MD. Fructose-1, 6-bisphosphatase inhibitors for reducing excessive endogenous glucose production in type 2 diabetes. Handb Exp Pharmacol. 2011;203:279–301.

    Article  Google Scholar 

  142. Erion MD, van Poelje PD, Dang Q, Kasibhatla SR, Potter SC, Reddy MR, Reddy KR, Jiang T, Lipscomb WN. MB06322 (CS-917): a potent and selective inhibitor of fructose 1,6-bisphosphatase for controlling gluconeogenesis in type 2 diabetes. Proc Natl Acad Sci U S A. 2005;102(22):7970–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Kubota K, Inaba S, Nakano R, Watanabe M, Sakurai H, Fukushima Y, Ichikawa K, Takahashi T, Izumi T, Shinagawa A. Identification of activating enzymes of a novel FBPase inhibitor prodrug, CS-917. Pharmacol Res Perspect. 2015;3(3): e00138.

    Article  PubMed  PubMed Central  Google Scholar 

  144. Yoshida T, Okuno A, Takahashi K, Ogawa J, Hagisawa Y, Kanda S, Fujiwara T. Contributions of hepatic gluconeogenesis suppression and compensative glycogenolysis on the glucose-lowering effect of CS-917, a fructose 1,6-bisphosphatase inhibitor, in non-obese type 2 diabetes Goto-Kakizaki rats. J Pharmacol Sci. 2011;115(3):329–35.

    Article  CAS  PubMed  Google Scholar 

  145. Yoshida T, Okuno A, Izumi M, Takahashi K, Hagisawa Y, Ohsumi J, Fujiwara T. CS-917, a fructose 1,6-bisphosphatase inhibitor, improves postprandial hyperglycemia after meal loading in non-obese type 2 diabetic Goto-Kakizaki rats. Eur J Pharmacol. 2008;601(1–3):192–7.

    Article  CAS  PubMed  Google Scholar 

  146. Kastrissios H, Walker JR, Carrothers TJ, Kshirsagar S, Khariton T, Habtemariam B, Mager DE, Rohatagi S. Population pharmacokinetic model for a novel oral hypoglycemic formed in vivo: comparing the use of active metabolite data alone versus using data of upstream and downstream metabolites. J Clin Pharmacol. 2012;52(3):404–15.

    Article  CAS  PubMed  Google Scholar 

  147. Dang Q, Liu Y, Cashion DK, Kasibhatla SR, Jiang T, Taplin F, Jacintho JD, Li H, Sun Z, Fan Y, et al. Discovery of a series of phosphonic acid-containing thiazoles and orally bioavailable diamide prodrugs that lower glucose in diabetic animals through inhibition of fructose-1,6-bisphosphatase. J Med Chem. 2011;54(1):153–65.

    Article  CAS  PubMed  Google Scholar 

  148. Johansson KS, Sonne DP, Knop FK, Christensen MB. What is on the horizon for type 2 diabetes pharmacotherapy?—An overview of the antidiabetic drug development pipeline. Expert Opin Drug Discov. 2020;15(11):1253–65.

    Article  CAS  PubMed  Google Scholar 

  149. VK0612 Fructose-1,6-bisphosphatase (FBPase) inhibitor. https://vikingtherapeutics.com/pipeline/other-pipeline-programs/vk0612/.

  150. Gao Y, Liu Y, Zheng D, Ho C, Wen D, Sun J, Huang L, Liu Y, Li Q, Zhang Y. HDAC5-mediated Smad7 silencing through MEF2A is critical for fibroblast activation and hypertrophic scar formation. Int J Biol Sci. 2022;18(15):5724–39.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Ren L, Jiang M, Xue D, Wang H, Lu Z, Ding L, Xie H, Wang R, Luo W, Xu L, et al. Nitroxoline suppresses metastasis in bladder cancer via EGR1/circNDRG1/miR-520h/smad7/EMT signaling pathway. Int J Biol Sci. 2022;18(13):5207–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Monteleone G, Laudisi F, Stolfi C. Smad7 as a positive regulator of intestinal inflammatory diseases. Curr Res Immunol. 2023;4: 100055.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Jankauskas SS, Mone P, Avvisato R, Varzideh F, De Gennaro S, Salemme L, Macina G, Kansakar U, Cioppa A, Frullone S, et al. miR-181c targets Parkin and SMAD7 in human cardiac fibroblasts: Validation of differential microRNA expression in patients with diabetes and heart failure with preserved ejection fraction. Mech Ageing Dev. 2023;212: 111818.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. An efficacy and safety study of mongersen (GED-0301) in subjects with active ulcerative colitis. https://classic.clinicaltrials.gov/ct2/show/results/NCT02601300?term=NCT02601300&draw=2&rank=1.

  155. Monteleone G, Neurath MF, Ardizzone S, Di Sabatino A, Fantini MC, Castiglione F, Scribano ML, Armuzzi A, Caprioli F, Sturniolo GC, et al. Mongersen, an oral SMAD7 antisense oligonucleotide, and Crohn’s disease. N Engl J Med. 2015;372(12):1104–13.

    Article  CAS  PubMed  Google Scholar 

  156. Monteleone G, Fantini MC, Onali S, Zorzi F, Sancesario G, Bernardini S, Calabrese E, Viti F, Monteleone I, Biancone L, et al. Phase I clinical trial of Smad7 knockdown using antisense oligonucleotide in patients with active Crohn’s disease. Mol Ther. 2012;20(4):870–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Marafini I, Stolfi C, Troncone E, Lolli E, Onali S, Paoluzi OA, Fantini MC, Biancone L, Calabrese E, Di Grazia A, et al. A Pharmacological batch of mongersen that downregulates Smad7 is Effective as induction therapy in active Crohn’s disease: a phase II. Open-Label Study BioDrugs. 2021;35(3):325–36.

    Article  CAS  PubMed  Google Scholar 

  158. Feagan BG, Sands BE, Rossiter G, Li X, Usiskin K, Zhan X, Colombel JF. Effects of Mongersen (GED-0301) on endoscopic and clinical outcomes in patients with active Crohn’s disease. Gastroenterology. 2018;154(1):61-64 e66.

    Article  PubMed  Google Scholar 

  159. Monteleone G, Stolfi C, Marafini I, Atreya R, Neurath MF. Smad7 antisense oligonucleotide-based therapy in Crohn’s disease: is it time to re-evaluate? Mol Diagn Ther. 2022;26(5):477–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Arrico L, Stolfi C, Marafini I, Monteleone G, Demartis S, Bellinvia S, Viti F, McNulty M, Cabani I, Falezza A, et al. Inhomogeneous diastereomeric composition of mongersen antisense phosphorothioate oligonucleotide preparations and related pharmacological activity impairment. Nucleic Acid Ther. 2022;32(4):312–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Papadakis KA, Prehn J, Moreno ST, Cheng L, Kouroumalis EA, Deem R, Breaverman T, Ponath PD, Andrew DP, Green PH, et al. CCR9-positive lymphocytes and thymus-expressed chemokine distinguish small bowel from colonic Crohn’s disease. Gastroenterology. 2001;121(2):246–54.

    Article  CAS  PubMed  Google Scholar 

  162. Denis M, Marcinkiewicz J, Zaid A, Gauthier D, Poirier S, Lazure C, Seidah NG, Prat A. Gene inactivation of proprotein convertase subtilisin/kexin type 9 reduces atherosclerosis in mice. Circulation. 2012;125(7):894–901.

    Article  CAS  PubMed  Google Scholar 

  163. Lopez-Andres N, Calvier L, Labat C, Fay R, Diez J, Benetos A, Zannad F, Lacolley P, Rossignol P. Absence of cardiotrophin 1 is associated with decreased age-dependent arterial stiffness and increased longevity in mice. Hypertension. 2013;61(1):120–9.

    Article  CAS  PubMed  Google Scholar 

  164. Gill PK, Hegele RA. New biological therapies for low-density lipoprotein cholesterol. Can J Cardiol. 2023;39:1913–30.

    Article  PubMed  Google Scholar 

  165. Banach M, Surma S, Reiner Z, Katsiki N, Penson PE, Fras Z, Sahebkar A, Paneni F, Rizzo M, Kastelein J. Personalized management of dyslipidemias in patients with diabetes-it is time for a new approach (2022). Cardiovasc Diabetol. 2022;21(1):263.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Samuel E, Watford M, Egolum UO, Ombengi DN, Ling H, Cates DW. Inclisiran: a first-in-class siRNA therapy for lowering low-density lipoprotein cholesterol. Ann Pharmacother. 2023;57(3):317–24.

    Article  CAS  PubMed  Google Scholar 

  167. PCSK9 inhibitors: a new era for cholesterol management. https://www.delveinsight.com/blog/pcsk9-inhibitors-for-cholesterol-management.

  168. Sposito AC, Breder I, Barreto J, Breder J, Bonilha I, Lima M, Oliveira A, Wolf V, Luchiari B, do Carmo HR, et al. Evolocumab on top of empagliflozin improves endothelial function of individuals with diabetes: randomized active-controlled trial. Cardiovasc Diabetol. 2022;21(1):147.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  169. Raal F, Fourie N, Scott R, Blom D, De Vries BM, Kayikcioglu M, Caldwell K, Kallend D, Stein E. Investigators LI-H: Long-term efficacy and safety of lerodalcibep in heterozygous familial hypercholesterolaemia: the LIBerate-HeFH trial. Eur Heart J. 2023;44:4272–80.

    Article  PubMed  PubMed Central  Google Scholar 

  170. Wang X, Wen D, Chen Y, Ma L, You C. PCSK9 inhibitors for secondary prevention in patients with cardiovascular diseases: a bayesian network meta-analysis. Cardiovasc Diabetol. 2022;21(1):107.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Cardiotrophin-1—digna biotech/biotecnol. https://0-adisinsight-springer-com.brum.beds.ac.uk/drugs/800030150.

  172. Kota J, Chivukula RR, O’Donnell KA, Wentzel EA, Montgomery CL, Hwang HW, Chang TC, Vivekanandan P, Torbenson M, Clark KR, et al. Therapeutic microRNA delivery suppresses tumorigenesis in a murine liver cancer model. Cell. 2009;137(6):1005–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Zhang X, Zhang X, Wang T, Wang L, Tan Z, Wei W, Yan B, Zhao J, Wu K, Yang A, et al. MicroRNA-26a is a key regulon that inhibits progression and metastasis of c-Myc/EZH2 double high advanced hepatocellular carcinoma. Cancer Lett. 2018;426:98–108.

    Article  CAS  PubMed  Google Scholar 

  174. Tanno T, Zhang P, Lazarski CA, Liu Y, Zheng P. An aptamer-based targeted delivery of miR-26a protects mice against chemotherapy toxicity while suppressing tumor growth. Blood Adv. 2017;1(15):1107–19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Mi B, Chen L, Xiong Y, Yang Y, Panayi AC, Xue H, Hu Y, Yan C, Hu L, Xie X, et al. Osteoblast/osteoclast and immune cocktail therapy of an exosome/drug delivery multifunctional hydrogel accelerates fracture repair. ACS Nano. 2022;16(1):771–82.

    Article  CAS  PubMed  Google Scholar 

  176. Liang G, Kan S, Zhu Y, Feng S, Feng W, Gao S. Engineered exosome-mediated delivery of functionally active miR-26a and its enhanced suppression effect in HepG2 cells. Int J Nanomed. 2018;13:585–99.

    Article  CAS  Google Scholar 

  177. Li Y, Fan L, Liu S, Liu W, Zhang H, Zhou T, Wu D, Yang P, Shen L, Chen J, et al. The promotion of bone regeneration through positive regulation of angiogenic-osteogenic coupling using microRNA-26a. Biomaterials. 2013;34(21):5048–58.

    Article  CAS  PubMed  Google Scholar 

  178. Zheng H, Ji J, Zhao T, Wang E, Zhang A. Exosome-encapsulated miR-26a attenuates aldosterone-induced tubulointerstitial fibrosis by inhibiting the CTGF/SMAD3 signaling pathway. Int J Mol Med. 2023;51(2):1–14.

    Google Scholar 

  179. Li R, Wang H, John JV, Song H, Teusink MJ, Xie J. 3D hybrid nanofiber aerogels combining with nanoparticles made of a biocleavable and targeting polycation and MiR-26a for bone repair. Adv Funct Mater. 2020;30(49):2005531.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  180. Wang B, Zhang A, Wang H, Klein JD, Tan L, Wang ZM, Du J, Naqvi N, Liu BC, Wang XH. miR-26a limits muscle wasting and cardiac fibrosis through exosome-mediated microRNA transfer in chronic kidney disease. Theranostics. 2019;9(7):1864–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  181. Tanno T, Zhang P, Bailey C, Wang Y, Ittiprasert W, Devenport M, Zheng P, Liu Y. A novel aptamer-based small RNA delivery platform and its application to cancer therapy. Genes Dis. 2023;10(3):1075–89.

    Article  CAS  PubMed  Google Scholar 

  182. Yan J, Lu X, Zhu X, Hu X, Wang L, Qian J, Zhang F, Liu M. Effects of miR-26a on osteogenic differentiation of bone marrow mesenchymal stem cells by a mesoporous silica nanoparticle—PEI—peptide system. Int J Nanomedicine. 2020;15:497–511.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. Mahati S, Fu X, Ma X, Zhang H, Xiao L. Delivery of miR-26a using an exosomes-based nanosystem inhibited proliferation of hepatocellular carcinoma. Front Mol Biosci. 2021;8: 738219.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. Zhang X, Li Y, Chen YE, Chen J, Ma PX. Cell-free 3D scaffold with two-stage delivery of miRNA-26a to regenerate critical-sized bone defects. Nat Commun. 2016;7:10376.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Liang GF, Zhu YL, Sun B, Hu FH, Tian T, Li SC, Xiao ZD. PLGA-based gene delivering nanoparticle enhance suppression effect of miRNA in HePG2 cells. Nanoscale Res Lett. 2011;6(1):447.

    Article  PubMed  PubMed Central  Google Scholar 

  186. Bao S, Huang S, Liu Y, Hu Y, Wang W, Ji M, Li H, Zhang NX, Song C, Duan S. Gold nanocages with dual modality for image-guided therapeutics. Nanoscale. 2017;9(21):7284–96.

    Article  CAS  PubMed  Google Scholar 

  187. Liang G, Li Y, Feng W, Wang X, Jing A, Li J, Ma K. Polyethyleneimine-coated quantum dots for miRNA delivery and its enhanced suppression in HepG2 cells. Int J Nanomedicine. 2016;11:6079–88.

    Article  PubMed  PubMed Central  Google Scholar 

  188. Hu Y, Liu H, Fang C, Li C, Xhyliu F, Dysert H, Bodo J, Habermehl G, Russell BE, Li W, et al. Targeting of CD38 by the tumor suppressor miR-26a serves as a novel potential therapeutic agent in multiple myeloma. Cancer Res. 2020;80(10):2031–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Gan M, Zhou Q, Ge J, Zhao J, Wang Y, Yan Q, Wu C, Yu H, Xiao Q, Wang W, et al. Precise in-situ release of microRNA from an injectable hydrogel induces bone regeneration. Acta Biomater. 2021;135:289–303.

    Article  CAS  PubMed  Google Scholar 

  190. Greene CM, Varley RB, Lawless MW. MicroRNAs and liver cancer associated with iron overload: therapeutic targets unravelled. World J Gastroenterol. 2013;19(32):5212–26.

    Article  PubMed  PubMed Central  Google Scholar 

  191. Ji JL, Shi HM, Li ZL, Jin R, Qu GT, Zheng H, Wang E, Qiao YY, Li XY, Ding L, et al. Satellite cell-derived exosome-mediated delivery of microRNA-23a/27a/26a cluster ameliorates the renal tubulointerstitial fibrosis in mouse diabetic nephropathy. Acta Pharmacol Sin. 2023;44:2455–68.

    Article  CAS  PubMed  Google Scholar 

  192. Chambers P, Ziminska M, Elkashif A, Wilson J, Redmond J, Tzagiollari A, Ferreira C, Balouch A, Bogle J, Donahue SW, et al. The osteogenic and angiogenic potential of microRNA-26a delivered via a non-viral delivery peptide for bone repair. J Control Release. 2023;362:489–501.

    Article  CAS  PubMed  Google Scholar 

  193. Hu J, Liu WF, Zhang XY, Shi GM, Yang XR, Zhou KQ, Hu B, Chen FY, Zhou C, Lau WY, et al. Synthetic miR-26a mimics delivered by tumor exosomes repress hepatocellular carcinoma through downregulating lymphoid enhancer factor 1. Hepatol Int. 2023;17:1265–78.

    Article  PubMed  Google Scholar 

  194. CTDSP1. https://pharma.bcpmdata.com/database/1/detail/441/kBKziIoBZV4GBptZv4r5.

Download references

Acknowledgements

We thank Dr. Dongming Xing’s laboratory for technical support and stimulating discussions.

Funding

The authors are grateful for the financial support from the Qingdao Major Scientific and Technological Project for Distinguished Scholars (20170103), the Laoshan Major Scientific and Technological Project for Distinguished Scholars (20181030), the Natural Science Foundation of Hunan Province (2021JJ40485), the Youth Innovation Team Development Program of Shandong Province (2023KJ227), and the Natural Science Foundation of Shandong Province (ZR2020MH369).

Author information

Authors and Affiliations

Authors

Contributions

DX, JW, and YW conceived the idea and planned the study. WC, XW, and CW wrote the manuscript; JH, XL, ZG, JF, XC, and MH contributed substantially to the discussion of its content; and SZ, WH, YW, MZ and MZ reviewed and edited the manuscript before submission. All authors have read and approved the final version of the manuscript for submission.

Corresponding authors

Correspondence to Chao Wang, Yudong Wu, Jie Wang or Dongming Xing.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

No individual personal data were included in this review.

Competing interests

The authors declare no conflicts of interest associated with this paper.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chen, W., Wu, X., Hu, J. et al. The translational potential of miR-26 in atherosclerosis and development of agents for its target genes ACC1/2, COL1A1, CPT1A, FBP1, DGAT2, and SMAD7. Cardiovasc Diabetol 23, 21 (2024). https://0-doi-org.brum.beds.ac.uk/10.1186/s12933-024-02119-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/s12933-024-02119-z

Keywords