Next Article in Journal
Pathophysiology and Treatment Options for Hepatic Fibrosis: Can It Be Completely Cured?
Next Article in Special Issue
Multifaceted Role of AMPK in Viral Infections
Previous Article in Journal
Mechanisms and Functions of Pexophagy in Mammalian Cells
Previous Article in Special Issue
Comparative Analysis of CTRP-Mediated Effects on Cardiomyocyte Glucose Metabolism: Cross Talk between AMPK and Akt Signaling Pathway
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

AICAr, a Widely Used AMPK Activator with Important AMPK-Independent Effects: A Systematic Review

1
Laboratory of Cell Biology, Croatian Institute for Brain Research, University of Zagreb School of Medicine, 10000 Zagreb, Croatia
2
Department of Physiology, University of Zagreb School of Medicine, 10000 Zagreb, Croatia
*
Author to whom correspondence should be addressed.
Cells 2021, 10(5), 1095; https://doi.org/10.3390/cells10051095
Submission received: 15 March 2021 / Revised: 21 April 2021 / Accepted: 1 May 2021 / Published: 4 May 2021
(This article belongs to the Special Issue Advances in AMPK Research: Basic and Translational Aspects)

Abstract

:
5-Aminoimidazole-4-carboxamide ribonucleoside (AICAr) has been one of the most commonly used pharmacological modulators of AMPK activity. The majority of early studies on the role of AMPK, both in the physiological regulation of metabolism and in cancer pathogenesis, were based solely on the use of AICAr as an AMPK-activator. Even with more complex models of AMPK downregulation and knockout being introduced, AICAr remained a regular starting point for many studies focusing on AMPK biology. However, there is an increasing number of studies showing that numerous AICAr effects, previously attributed to AMPK activation, are in fact AMPK-independent. This review aims to give an overview of the present knowledge on AMPK-dependent and AMPK-independent effects of AICAr on metabolism, hypoxia, exercise, nucleotide synthesis, and cancer, calling for caution in the interpretation of AICAr-based studies in the context of understanding AMPK signaling pathway.

1. Introduction

In 1995, 5-amino-4-imidazolecarboxamide (AICA) ribonucleoside or riboside was first proposed to be used as the activator of AMP-kinase (AMPK) in intact cells or, in other words, to play the same role that phorbol esters had in dissecting signaling pathways regulated by protein kinase C [1]. The study by Corton et al. was based on previous work showing that administration of the AICA riboside to intact cells causes AICA ribonucleotide or ribotide to accumulate inside the cell [2], and a study by Sullivan et al. [3] demonstrated that AICA ribotide mimicked the effect of AMP on allosteric activation of rat liver AMPK. Since then, the compound that has been widely used as an AMPK-agonist was an exogenous dephosphorylated AICA riboside that should be properly abbreviated AICAr. However, in more than 1700 articles that can be retrieved from PubMed on AMPK and AICA riboside, AICAr is often abbreviated as AICAR, although the AICAR acronym should be reserved for AICA ribotide or the phosphorylated form that is a physiological, endogenous precursor in de novo purine synthesis [4]. The nomenclature is additionally complicated because the other name used for the endogenous substance or AICAR is ZMP [5]. Moreover, the search of the literature reveals the common use of acadesine instead of AICAr [4,6,7,8,9,10,11,12,13,14,15,16,17,18,19].
As shown in Figure 1, AICAR or ZMP is an AMP analog. As a cell-permeable nucleotide, AICAr enters the cells through adenosine transporters [20] and becomes phosphorylated by adenosine kinase into AICAR [21]. AICAR or ZMP activates AMPK but it is 40- to 50- fold less potent than AMP in AMPK activation and accumulates in high concentrations in the cytoplasm [1], so that it was always likely that AICAr may have several AMPK-independent effects. Similar to AMP, AICAR binds to the γ subunit of AMPK, allosterically activates the enzyme, stimulates phosphorylation at Thr172 by liver kinase B1 (LKB1), and protects against pThr172 dephosphorylation [22,23]. Therefore, the most common method to test for AICAr-mediated activation of AMPK in particular tissues or cells is to detect the level of pThr172 AMPK by Western blot in lysates upon AICAr treatment.
In this article, we will give a brief overview of the present knowledge on AMPK-dependent and AMPK-independent effects of AICAr.

2. AICAr, Metabolism and Diabetes

Since 1995, AICAr has been a valuable tool to identify metabolic pathways associated with AMPK activation in various tissues and cells in vivo and in vitro. In liver cells, AICAr was found to inhibit fatty acid and cholesterol synthesis [24], to increase fatty acid oxidation [25], and to inhibit gluconeogenesis by decreasing transcription of PEPCK [26]. In skeletal muscle cells, AICAr was found to activate glycogen phosphorylase and increase glycogenolysis [27], stimulate fatty acid oxidation and glucose uptake [28], and inhibit protein synthesis [29]. All of these studies supported the role of AMPK as a central metabolic hub of the cell that is activated whenever the ratio of AMP/ATP is high and which acts to increase catabolic and inhibit anabolic processes (Figure 2).
However, investigators soon realized that ZMP in millimolar concentrations could exert many other effects, including modulation of other AMP-sensitive enzymes, like glycogen phosphorylase [30], and fructose-1-6-biphosphatase, a rate-controlling enzyme in gluconeogenesis, which is likely to explain why the effects of AICAr on gluconeogenesis in the liver are AMPK-independent [31]. The best way to assess the role of AMPK in the effects of AICAr in vivo could be provided by AMPK knockout mice. As shown in Table 1, data obtained by transgenic mice models revealed that AICAr-mediated effects on glucose uptake in skeletal muscle cells [32,33,34,35], lipogenesis and fatty acid oxidation in the liver [36], and decreased fat synthesis in adipose cells were AMPK-dependent [37]. However, a decrease in gluconeogenesis and an inhibition of oxidative phosphorylation (OXPHOS) can be observed in response to AICAr in mice lacking both AMPKα1 and α2 isoforms [38,39,40].
No matter whether being AMPK-dependent or independent, metabolic effects of AICAr may be of relevance for the potential treatment of type 2 diabetes [41]. AICAr induces hypoglycemia in vivo [42,43] and the effect is abolished in mice lacking AMPK [32,33,35], suggesting that the effect can be more ascribed to AMPK-dependent entry of glucose than to AMPK-independent effects of AICAr on the inhibition of gluconeogenesis. In addition, AICAr may help to reduce peripheral resistance to insulin action because AICAr acts to reduce the storage of fatty acids in adipose tissue [37]. In various animal models of insulin resistance, AICAr administration has been shown to improve metabolic disturbances and to enhance insulin sensitivity in peripheral tissues [44,45,46,47]. Systemic AICAr administration in humans exerted beneficial effects by reducing hepatic glucose output and increasing glucose uptake in skeletal muscle [43,48]. However, as shown in Table 2, AICAr has shown a very poor oral bioavailability in clinical trials [49]. The preferred route of administration is through continuous intravenous injection and that renders it quite unsuitable for chronic treatment of metabolic disorders like diabetes.
Table 1. AMPK-dependent and AMPK-independent effects of AICAr in transgenic mice models.
Table 1. AMPK-dependent and AMPK-independent effects of AICAr in transgenic mice models.
TissueAICAr EffectAMPK-DependencyModel
Skeletal muscleIncreased glucose uptake AMPK-dependentSkeletal and cardiac muscle-specific expression of ampkα2-KD [32]
Total ampkα2-KO [33]
Total ampky3-KO [34]
Muscle-specific expression of ampkα2-DN [35]
Increased FAO AMPK-independentMuscle-specific ampkα2-KD [50]
Increased mrna HKII and (PGC)-1αAMPK-dependentTotal ampkα2-KO [51]
LiverDecreased glucose phosphorylationAMPK-independentLiver-specific ampkα1α2-KO [39]
Decreased gluconeogenesisAMPK-independentLiver-specific ampkα1α2-KO [38]
Inhibited OXPHOSAMPK-independentLiver-specific ampkα1α2-KO [40]
Decreased lipogenesisAMPK-dependentLiver-specific ampkα1α2-KO [36]
Increased FAOAMPK-dependentLiver-specific ampkα1α2-KO [36]
Adipose tissueDecreased FA synthesisAMPK-dependentAdipose tissue-specific ampkα1/α2-KO [37]
Decreased lipolysis AMPK-dependentAMPK α1-KO [52]
Whole bodyAcute hypoglycemiaAMPK-dependentSkeletal and cardiac muscle-specific ampkα2-KD [32]
Total ampkα2-KO [33]
Muscle-specific ampkα2-DN [35]
DN, dominant negative; FA, fatty acid; FAO, fatty acid oxidation; HKII, hexokinase II; KD, kinase dead; KO, knockout; OXPHOS, oxidative phosphorylation; (PGC)-1α, peroxisome-proliferator-activated receptor γ coactivator.

3. AICAr, Adenosine, and Ischemic Heart

The first study of the safety and tolerance of AICAr was done in 1991, much before the recognition of AICAr as an AMPK agonist to establish pharmacokinetics of a drug that raised interest as a novel adenosine-regulating agent [49]. Adenosine is a potent vasodilator that plays a key role in reducing ischemia/reperfusion injury, but the applications for systemic adenosine are limited owing to peripheral hemodynamic actions [13]. As shown in Figure 1, AICAr shares structural similarities with adenosine, and therefore, can increase the extracellular concentrations of adenosine by competing for the nucleoside transporter [20]. In addition, AICAR increases intracellular concentrations by inhibiting adenosine deaminase and increasing the production of adenosine rather than inosine from ATP catabolism. Several animal studies performed in the 1980s demonstrated that AICAr or acadesine infusion improved postischemic recovery in the heart [53,54], and prompted the first international randomized studies in human participants undergoing coronary artery bypass graft surgery (CAGS). As shown in Table 2, several clinical studies testing the effects of AICAr were performed [16,18,19], and outcomes of these studies were examined in a 1997 meta-analysis that revealed that AICAr can reduce early cardiac death, myocardial infarction, and combined adverse cardiovascular outcomes [14]. However, these promising meta-analysis results were not confirmed by later clinical trials. In 2012, the RED-CABG trial was stopped early after interim data failed to indicate a reduction in morbidity or mortality among intermediate- to high-risk patients receiving AICAr versus placebo [15].
Before the mode of action via AMPK was appreciated, AICAr-mediated protection of myocardium was ascribed only to the effects of adenosine on vasodilation and inhibition of platelet aggregation and neutrophil activation [13,54]. Later studies provided the link between the activation of AMPK and AICAr-mediated effects on glucose and glycogen metabolism in heart muscle [30,55]. Although several issues have complicated the interpretation of pharmacological manipulation of AMPK by AICAR in cardiomyocytes [56], AICAr has been described to exert several beneficial effects, including improved cardiac function in a model of endotoxin-induced myocardial inflammation [57], prevention of cardiac fibrosis [58], and metabolic changes in cardiomyocytes exposed to free fatty acids that could protect the hearts of patients with type 2 diabetes against ischemia-reperfusion injury [59].
In human aortic endothelial cells, AICAr stimulated AMPK activity and nitric oxide (NO) production, and the effects were proved to be AMPK-dependent since the effects were inhibited by the expression of a dominant-negative (DN) AMPK mutant [60]. Similar AMPK-dependent effects on NO production were observed in response to hypoxia [61], and studies performed in the knockout of the upstream kinase LKB1 confirmed the important role of AMPK in angiogenesis [62].
Table 2. AICAr in clinical trials.
Table 2. AICAr in clinical trials.
YearConditionTrial typeDosesToxicityOutcome
1991Healthy volunteersPhase IPO and IV: 10, 25, 50, and 100 mg/kgWell tolerated, only mild and transient side effectsPoor oral bioavailability, the rapid decline of post-infusion plasma concentrations [49]
1994Lesch-Nyhan SyndromeCase ReportPO: 30 mg/kg/day for 4 days followed by 100 mg/kg/day for 4 daysNo adverse eventsNo changes in plasma levels of AICAr, confirm the estimate of <5% oral bioavailability of AICAr in humans [63]
1994Coronary artery bypass grafting Multicenter RCT, Phase IIContinuous IV for 7 h: 0.19 and 0.38 mg/kg/minWell tolerated, mild hyperuricemiaLimits the severity of post-bypass myocardial ischemia at higher dose [16]
1994Exercise-induced myocardial ischemia in patients with chronic stable angina pectorisSingle center RCT, Phase IIIV: 6–48 mg/kgWell tolerated, mild asymptomatic hyperuricemia, mild asymptomatic hypoglycemiaNo significant difference in comparison to placebo [17]
1995Coronary artery bypass grafting Multicenter RCT, Phase IIContinuous IV for 7h: 0.1 mg/kg/min; in cardioplegic solution 5 mg/mLNo adverse eventsNo significant difference in MI in the overall study group; significantly reduced the incidence of Q-wave MI in high-risk patients [18]
1995Coronary artery bypass grafting Multicenter RCT, Phase IIContinuous IV for 7 h: 0.05 and 0.1 mg/kg/minWell tolerated, mild hyperuricemiaNo significant difference in comparison to the placebo may reduce the incidence of larger Q-wave MI [19]
1997Coronary artery diseaseSingle center RCT, Phase IIContinuous IV: 5, 10, 20, 50 mg/kg Well tolerated, hyperlactacidemiaAt higher doses, minor beneficial effects on ejection fraction and myocardial lactate metabolism were observed [8]
2006Coronary artery bypass grafting Multicenter RCT, Phase IIIContinuous IV for 7 h: 0.1 mg/kg/minNo adverse eventsReduces the severity of acute post-reperfusion MI, substantially reducing the risk of dying over the 2 years after infarction [9]
2007Healthy volunteersPhase IContinuous IV for 3 h: 10 mg/kg/hNo adverse eventsAcutely stimulates muscle 2-DG uptake with a minor effect on whole-body glucose disposal [43]
2009Healthy volunteersPhase IContinuous intra-arterial infusion for 110 min: 1, 2, 4, or 8 mg/min/dL forearm tissueNo adverse eventsPotent vasodilation in the skeletal muscle vascular bed; does not increase skeletal muscle glucose uptake [64]
2012Coronary artery bypass grafting Multicenter RCT, Phase IIIContinuous IV for 7 h: 0.1 mg/kg/minNo adverse eventsNo significant reduction in the composite of all-cause mortality, nonfatal stroke, or severe left ventricular dysfunction through 28 days [15]
Hematologic malignancies
2013Relapsed/refractory chronic lymphocytic leukemia (CLL)Multicenter open-label clinical study, Phase I/IIContinuous IV: single doses of 50–315 mg/kg; two doses at 210 mg/kg; five doses at 210 mg/kgGrade ≥2 hyperuricemia (not clinically significant), transient anemia and/or thrombocytopenia (not clinically significant), renal impairment, and transient infusion-related hypotension (clinically significant)210 mg/kg was the MTD and OBD. Multiple-dose administrations at the OBD have an acceptable safety profile [6]
2019Azacytidine refractory MDS/AML patientsPhase I/IIContinuous IV: 140 mg/kg or 210 mg/kgTrial stopped after 2 to 3 cycles due to serious renal toxicitiesSide effects of high doses preclude its use in patients with strong comorbidities; one patient exhibited a very strong reduction (50%) of his blast count after only 2 cycles of AICAr and more than 70% after 6 cycles [10]
2-DG, 2-Deoxyglucose; IV, intravenous; MI, myocardial infarction; MTD, maximum tolerated dose; OBD, optimal biological dose; PO, per os.

4. AICAr as an “Exercise in a Pill”

In 2008, Narkar et al. reported that, even in sedentary mice, 4 weeks of AICAr treatment alone enhanced running endurance by 44% and induced genes linked to oxidative metabolism in muscle cells. AICAr induced fatigue-resistant type I (slow-twitch) fiber specification, and AMPK activation by AICAr was sufficient to increase running endurance without additional exercise signals [65]. As would be expected, AICAr was immediately set into the limelight of not only the scientific but sports and wider community as a new exercise mimetic or an “exercise in a pill.” In less than a year, the French anti-doping agency raised concerns about a new metabolic substance being used on Tour de France, and AICAr soon appeared on the list of substances banned by the World Anti-Doping Agency [66]. In 2012, a sports doctor and nine others of the Spanish cycling team were arrested in connection with an international network supplying the synthetic AMPK activator AICAR as a “next generation superdrug” performance-enhancing drug [67].
As shown in Table 1, the majority of the effects of AICAr on skeletal muscles are AMPK-dependent. AICAr-induced glucose uptake in skeletal muscle was abolished in the knockout of the α 2 [32,33,35] and α 3 isoforms of AMPK [34]. Both AICAr and treadmill exercise increased insulin sensitivity to stimulate glucose uptake, and these effects were not observed in mice with reduced or ablated AMPK activity in skeletal muscle [68,69]. However, the mechanisms of exercise- and AICAr-mediated glucose transport diverge at some point downstream of AMPK since AICAr-induced effects were absent in muscle-specific knockout of atypical PKC, and atypical PKC was not required in treadmill exercise [70]. Both AICAr and exercise induce AMPK activation and metabolic stress, but the mechanical stress is only caused by exercise, so that the combination of two may be useful in some conditions. In chronic inflammatory myopathy model mice, the combination of AICAr and exercise reverse apoptosis of fibro-adipogenic progenitors and improves muscle function and regeneration [70]. To add another layer of intersection between the exercise and AICAr, a recent study of daytime variance in exercise capacity revealed that exercise itself may induce an increase in the level of endogenous ZMP (AICA ribotide or AICAR). Moreover, endogenous ZMP was induced by exercise in a time-dependent manner and had the same effects as exogenous AICAr on AMPK activation, glycolysis, and fatty acid oxidation [71].
There is an increasing number of papers focusing on the effect of aging on mitochondrial biogenesis and the beneficial effects of exercise and exercise mimetic in delaying aging-mediated decline in AMPK activity. Reznick et al. examined AMPK activity in young and old rats and found that acute stimulation of AMPK α 2 activity by AICAr and exercise was blunted in skeletal muscle of old rats, suggesting that reductions in AMPK activity may be an important contributing factor in aging-associated reductions in mitochondrial biogenesis and dysregulated lipid metabolism [72]. However, 500 mg/kg of AICAr has beneficial effects on cognition and motor coordination in both young and old mice, and these effects were absent in the muscle-specific mutated AMPK α 2-subunit, providing support for a muscle-mediated mechanism [73]. In another study, the same AICAr concentration had no beneficial impact on skeletal muscle of wild-type mice and could not be considered as a mimetic of exercise in aged mice, but improved aerobic running capacity and mitochondrial function in aged myostatin KO mice [74]. In a mouse model for Duchene muscular dystrophy, AICAr treatment enhanced the effects of exercise [75] and improved muscle function, probably by stimulating autophagy [76].

5. AICAr, AMPK, Proliferation and Cell Cycle

As a central metabolic regulator that reacts to an increase in AMP/ATP ratio, AMPK restricts growth and proliferation in response to energetic or nutritional stress. AICAr was first reported to suppress protein synthesis in rat skeletal muscle through down-regulation of the mechanistic target of rapamycin (mTOR) signaling, including p70 S6 Kinase 1 (S6K1) and eukaryotic translation initiation factor 4E binding protein 1 (4E-BP1), which are involved in the regulation of protein translation [29]. As shown in Figure 2, mTOR is a catalytic subunit of two functionally distinct protein complexes, mTORC1 (mTOR complex 1) and mTORC2 (mTOR complex 2), and both S6K1 and 4E-BP1 lie downstream of mTORC1. AICAr inhibits mTORC1 but activates mTORC2, and both effects are AMPK-dependent. AICAr-mediated inhibition of mTORC1 requires the phosphorylation of protein raptor [77]. AICAr-mediated activation of mTORC2 did not result from AMPK-mediated suppression of mTORC1, and thus, reduced negative feedback on phosphatidylinositol 3-kinase (PI3K) flux, but rather on direct phosphorylation of mTOR in complex with rictor and phosphorylated Akt as a downstream target [78].
While mTORC1 inhibition ensures restriction of cell growth, stabilization of p53, another major target for activated AMPK, causes cell cycle arrest [79]. In many cell types, AICAr-mediated inhibition of proliferation occurs concomitantly with AMPK phosphorylation and is associated with inhibition of mTOR activity and/or p53 accumulation [80,81,82,83,84,85,86]. However, to prove that AICAr effects are dependent on AMPK, small interference RNAs targeted at AMPK or the expression of constitutively active or DN AMPK mutants should be used to test for the role of AMPK in the effects of AICAr [87]. In addition, if AICAr effects are truly AMPK-dependent, they should be mimicked by other AMPK agonists, and particularly with AMPK agonists which are more specific AMPK activators, like A-769662 [86], MK-8722 [88], and PF-739 [89]. Another more selective agonist is C13, a prodrug that is converted by cellular metabolism into C2, an AMP analog whose binding site overlaps with that of AMP and AICAR, but is orders of magnitude more potent in AMPK activation [90].
Using these approaches, AICAr-induced proliferative arrest of primary mouse embryonic fibroblasts (MEF) [81], various cancer cell lines [82], and aneuploid MEF cells [84] were shown to be AMPK-dependent as the effects were more or less inhibited in cells infected with a retroviral vector expressing DN AMPK. However, in MEF cells lacking both catalytic isoforms of AMPK, AICAr effects on proliferation and cell cycle arrest were AMPK-independent, and the effects on apoptosis were even more pronounced, while A-769662 exhibited specific AMPK-dependent effects on cell growth and metabolism [86]. In glioma cells, AICAr inhibited proliferation in vitro and in vivo independent of AMPK, while A-769662 had no effects on proliferation [91], and similar AMPK-independent effects of AICAr on proliferation have been observed in various leukemia cells [92,93].
The cell cycle analyses of AICAr-arrested cells in some studies revealed an increase in the proportion of cells in the G0/G1 phase, as would be expected from the mechanism of cell cycle arrest in response to AMPK activation and mTORC1 inhibition [23]. However, in embryonic stem cells, AICAr increased the cell population at both G1 and non-cycling S phases [85]. Furthermore, an arrest in the S phase has been observed in MEFs [86], cancer cell lines [94], and leukemia cells [95]. In glioma cell lines, AICAr-induced an AMPK-independent arrest in the G2/M phase [91]. To understand the mechanism responsible for proliferation arrest in response to AICAr, we have to go back to the role of endogenous AICAR in de novo purine synthesis that has been well-known to affect cell growth much before the discovery of the role of AICAr in AMPK activation.

6. AICAr, ZMP, and Purine Synthesis

As shown in Figure 3, AICA ribotide (AICAR) or ZMP is a normal cellular intermediate in de novo purine synthesis. AICAR or ZMP is increased in Lesch-Nyhan syndrome, one of the most common disorders of purine and pyrimidine metabolism. The Lesch-Nyhan results from a deficiency of hypoxanthine-guanine phosphoribosyltransferase (HGPRT), so that the activity of the salvage pathway is diminished and the de novo pathway of purine nucleotide synthesis accelerated, leading to an accumulation of ZMP or AICAR [96]. Recent studies reported an increase in the levels of AICAr, which normally arises from dephosphorylation of ZMP, in urine and cerebrospinal fluid of patients with deficiency of HGPRT [97], suggesting that the accumulation of ZMP may be the cause of neurological abnormalities, similar to the toxicity of ZMP accumulation in yeast [98]. Yeast is a good experimental system to study the effects of AICAr that are AMPK-independent as the yeast AMPK orthologue SNF1 is activated by ADP rather than AMP, and genes strongly regulated by Snf1p are not identical to AICAr-regulated transcription. In the yeast model, disruption of nucleotide homeostasis was identified as a crucial feature of AICAr toxicity [99], suggesting the similar role of nucleotide metabolism in AMPK-independent growth arrest induced by an exogenous AICAr in human cell lines.
When two acute lymphoblastic leukemia (ALL) cell lines were exposed to AICAr and A-769662, metabolomics analysis revealed that AICAr-treated cells excreted significant amounts of metabolites involved in purine and pyrimidine metabolism, and these metabolites were not detected in the samples treated with A-769662. AICAr-mediated increase in the level of hypoxanthine, adenine, guanine, and inosine can be ascribed to an increase of the purine metabolism generated by AICAr entering these pathways. Moreover, pyrimidine metabolism was also significantly affected by AICAr, and an increase in the levels of orotate, dihydroorotate, and carbamoylaspartate suggested a possible inhibition at the level of UMP synthase [100]. A possible role for inhibition of UMP synthase and pyrimidine starvation in AICAr-mediated apoptosis was further corroborated with metabolomics screen in multiple myeloma cells. In these cells, AICAr induced an accumulation in the S phase, an increase in purine, a decrease in pyrimidine metabolites, and the most striking increase in the level of orotate [101]. The same effects on the levels of purine and pyrimidine metabolites were observed in AICAr-treated ALL cell lines, in which AICAr-mediated arrest in the S phase was AMPK-independent, but attenuated with the addition of uridine [102]. The inhibition of pyrimidine synthesis at the level of UMP synthase was ascribed to the lack of phosphoribosyl pyrophosphate (PRPP), and a decrease in PRPP in response to AICAr has been described in other model systems [103,104].
However, most of the drugs that increase the level of endogenous ZMP act to activate AMPK so that it is difficult to completely rule out the possible involvement of AMPK in antiproliferative effects. An inhibitor of AICAR transformylase (AICART), an enzyme that catalyzes the last two steps of purine de novo synthesis and metabolizes AICAR, induces an increase in the level of AICAR or ZMP, and endogenous ZMP was capable of activating AMPK and its downstream signaling pathways [105]. The antifolate pemetrexed inhibits the folate-dependent enzyme in de novo purine biosynthesis, increases ZMP, and activates AMPK [106]. Methotrexate, a well-known cytostatic drug, inhibits purine de novo synthesis and potentiates the ability of exogenous AICAr to increase the level of ZMP by inhibiting AICART (Figure 3). Consequently, methotrexate enhances the ability of AICAr to activate AMPK and to inhibit the growth of human cancer cell lines [107], and promote glucose uptake and lipid oxidation in skeletal muscle [108].

7. AICAr, AMPK, Cancer, and Leukemia

In 2003, Campas et al. reported that AICAr activates AMPK and induces apoptosis in primary samples of B-cell chronic lymphocytic leukemia (CLL) in vitro [11]. Two years later, an epidemiological study revealed that metformin, another AMPK activator had a protective role in the development of cancer, and thus, invigorated interest in the possible use of AMPK agonists in the treatment of cancer [109]. However, although it would be expected that AMPK, acting as a downstream target of LKB1, has tumor suppressor activity, the results from many studies point that the role of AMPK in cancer is much more complex; AMPK may suppress tumor growth before tumorigenesis, but once cancer has arisen, AMPK may instead support the survival of cancer cells [23]. In the meantime, many other studies described the beneficial effects of AICAr, especially in hematological malignancies, and most of these effects turned out to be AMPK-independent.
AICAr-induced apoptosis in CLL cells was independent of AMPK since AMPK activation with phenformin or A-769662 failed to induce apoptosis, and AICAr had effects in B lymphocytes from AMPKα1(−/−) mice [92]. The effects of AICAr on CLL cells in vitro were demonstrated at doses that are tolerated well when achieved in plasma after intravenous injection [11] so that the multicenter study testing the effects of AICAr for patients with relapsed/refractory CLL had been initiated. The results of phase I/II revealed that AICAr had an acceptable safety profile and antileukemic activity in patients with poor prognosis [6].
Cytotoxic effects have been reported in other hematological malignancies. AICAr-induced apoptosis and concurrent activation of AMPK were described in childhood acute lymphoblastic leukemia (ALL) cell lines [110], as well as in B cells isolated from patients with mantle cell lymphoma and splenic marginal zone lymphoma [7]. However, the role of AMPK in cytotoxic effects of AICAr was not investigated. In chronic myelogenous leukemia (CML) cell lines [12] and primary samples [111], AICAr had antiproliferative effects, but AMPK knock-down by shRNA failed to prevent the effect of AICAr, indicating an AMPK-independent mechanism [12]. In azacytidine (Aza)-resistant myelodysplastic syndrome and acute myeloid leukemia (MS/AML) cell lines and primary samples, 2 mM AICAr blocked proliferation, and these initial findings led to a phase I/II clinical trial using AICAr in 12 patients with Aza-refractory MDS/AML patients. Despite a very good response in one out of four patients, the trial was stopped because the highest dose of AICAr caused serious renal side effects in patients with severe comorbidities [10].
Our recent studies demonstrated that lower concentrations of AICAr alone are capable of inducing differentiation of various monocytic AML cell lines, as well as a subset of primary blasts isolated from the bone marrow of AML patients [93,95,112]. These effects were not mimicked by metformin [93] or A-769662 [113]. Differentiation therapy of acute promyelocytic leukemia, a particular subtype of AML, with all-trans retinoic acid provides the most successful pharmacological therapy of AML so that the approach of treating cancer cells by differentiation instead of killing is well appreciated, especially after the recent success of AML differentiation with inhibitors of mutated isocitrate dehydrogenase [114]. Of note, differentiation in response to AICAr has been described in various other cell systems, including erythroid differentiation of embryonic stem cells [85], mineralization of osteoblastic MC3T3-E1 cells [115], and astroglial differentiation of neural stem cells [116], but the mechanisms have not been completely elucidated. In our model of monocytic leukemia cells, the effects of AICAr were not AMPK-dependent, since both growth arrest and differentiation were preserved in AML cells with siRNA-downregulated AMPK [93]. In CML cell lines, an AMPK-independent cell death involved AICAr-induced autophagy [12], but our study showed that, although AICAr induced autophagy in parallel with differentiation, AICAr-mediated differentiation did not depend on the presence of key proteins of canonical autophagy pathway [113]. However, AICAr-induced differentiation was inhibited by downregulation of checkpoint kinase 1 (Chk1), and activation of Chk1 was induced by defects in pyrimidine synthesis [95]. As shown in Figure 3, AICAr inhibits pyrimidine synthesis at the level of UMP-synthase, and thus, shares the same mechanism of leukemia differentiation as the one that has been recently described in response to brequinar, a well-known inhibitor of dihydroorotate dehydrogenase [117]. These results further corroborate the hypothesis that antiproliferative effects of AICAr are more related to the effects of AICAr on nucleotide synthesis and cell cycle arrest than on the activation of AMPK that is usually measured to be increased concomitantly.

8. Conclusions

Over the last 25 years, AICAr has been used in hundreds of studies as an activator of AMPK. The results of these initial studies pointed to the important roles of AMPK, and many of them have been later confirmed by studies in transgenic mice or by using models of cells with overexpression or down-regulation of AMPK. However, AICAr accumulates in cells in millimolar concentrations and exerts many AMPK-independent or “off-target“ effects so that allowances must be made for the possible use of AICAr. Although AICAr is no longer recommended as a specific AMPK agonist [118], mostly because there are many more specific activators of AMPK available nowadays, AICAr can be still useful in an initial screen to test for AMPK activation, especially when combined with other AMPK agonists and proper methods for AMPK downregulation. In addition, AICAr is still a highly promising pharmacological agent having many beneficial effects in metabolism, hypoxia, exercise, and cancer.

Funding

This work has been funded by Croatian Science Foundation under the projects IP-2016-06-4581, DOK-2018-01-9599, and DOK-2020-01-2873 by the European Union through the ESF Operational Programme Efficient Human Resources 2014-2020 (to D.V.), and co-financed by the Scientific Centre of Excellence for Basic, Clinical and Translational Neuroscience (project “Experimental and clinical research of hypoxic-ischemic damage in perinatal and adult brain”; GA KK01.1.1.01.0007) funded by the European Union through the European Regional Development Fund).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Corton, J.M.; Gillespie, J.G.; Hawley, S.A.; Hardie, D.G. 5-Aminoimidazole-4-Carboxamide Ribonucleoside. A Specific Method for Activating AMP-Activated Protein Kinase in Intact Cells? JBIC J. Biol. Inorg. Chem. 1995, 229, 558–565. [Google Scholar] [CrossRef]
  2. Sabina, R.L.; Patterson, D.; Holmes, E.W. 5-Amino-4-imidazolecarboxamide riboside (Z-riboside) metabolism in eukaryotic cells. J. Biol. Chem. 1985, 260, 6107–6114. [Google Scholar] [CrossRef]
  3. Sullivan, J.E.; Brocklehurst, K.J.; Marley, A.E.; Carey, F.; Carling, D.; Beri, R.K. Inhibition of lipolysis and lipogenesis in isolated rat adipocytes with AICAR, a cell-permeable activator of AMP-activated protein kinase. FEBS Lett. 1994, 353, 33–36. [Google Scholar] [CrossRef] [Green Version]
  4. Neste, E.V.D.; Berghe, G.V.D.; Bontemps, F. AICA-riboside (acadesine), an activator of AMP-activated protein kinase with potential for application in hematologic malignancies. Expert Opin. Investig. Drugs 2010, 19, 571–578. [Google Scholar] [CrossRef] [PubMed]
  5. Dembitz, V.; Lalic, H. The Role of AMPK/mTOR Modulators in the Therapy of Acute Myeloid Leukemia. Curr. Med. Chem. 2019, 26, 2208–2229. [Google Scholar] [CrossRef]
  6. Neste, E.V.D.; Cazin, B.; Janssens, A.; González-Barca, E.; Terol, M.J.; Levy, V.; De Oteyza, J.P.; Zachee, P.; Saunders, A.; De Frias, M.; et al. Acadesine for patients with relapsed/refractory chronic lymphocytic leukemia (CLL): A multicenter phase I/II study. Cancer Chemother. Pharmacol. 2013, 71, 581–591. [Google Scholar] [CrossRef] [Green Version]
  7. Campàs, C.; Santidrián, A.F.; Domingo, A.; Gil, J. Acadesine induces apoptosis in B cells from mantle cell lymphoma and splenic marginal zone lymphoma. Leukemia 2004, 19, 292–294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. De Jonge, R.; MacLeod, D.C.; Suryapranata, H.; A Van Es, G.; Friedman, J.; Serruys, P.W.; De Jong, J.W. Effect of acadesine on myocardial ischaemia in patients with coronary artery disease. Eur. J. Pharmacol. 1997, 337, 41–44. [Google Scholar] [CrossRef]
  9. Mangano, D.T.; Miao, Y.; Tudor, I.C.; Dietzel, C. Post-Reperfusion Myocardial Infarction: Long-Term Survival Improvement Using Adenosine Regulation With Acadesine. J. Am. Coll. Cardiol. 2006, 48, 206–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Cluzeau, T.; Furstoss, N.; Savy, C.; El Manaa, W.; Zerhouni, M.; Blot, L.; Calleja, A.; Dufies, M.; Dubois, A.; Ginet, C.; et al. Acadesine Circumvents Azacitidine Resistance in Myelodysplastic Syndrome and Acute Myeloid Leukemia. Int. J. Mol. Sci. 2019, 21, 164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Campàs, C.; Lopez, J.M.; Santidrián, A.F.; Barragán, M.; Bellosillo, B.; Colomer, D.; Gil, J. Acadesine activates AMPK and induces apoptosis in B-cell chronic lymphocytic leukemia cells but not in T lymphocytes. Blood 2003, 101, 3674–3680. [Google Scholar] [CrossRef] [PubMed]
  12. Robert, G.; Ben Sahra, I.; Puissant, A.; Colosetti, P.; Belhacene, N.; Gounon, P.; Hofman, P.; Bost, F.; Cassuto, J.-P.; Auberger, P. Acadesine Kills Chronic Myelogenous Leukemia (CML) Cells through PKC-Dependent Induction of Autophagic Cell Death. PLoS ONE 2009, 4, e7889. [Google Scholar] [CrossRef]
  13. Drew, B.G.; A Kingwell, B. Acadesine, an adenosine-regulating agent with the potential for widespread indications. Expert Opin. Pharmacother. 2008, 9, 2137–2144. [Google Scholar] [CrossRef] [PubMed]
  14. Mangano, D.T. Effects of Acadesine on Myocardial Infarction, Stroke, and Death Following SurgeryA Meta-analysis of the 5 International Randomized Trials. JAMA 1997, 277, 325–332. [Google Scholar] [CrossRef] [PubMed]
  15. Newman, M.F.; Ferguson, T.B.; White, J.A.; Ambrosio, G.; Koglin, J.; Nussmeier, N.A.; Pearl, R.G.; Pitt, B.; Wechsler, A.S.; Weisel, R.D.; et al. Effect of Adenosine-Regulating Agent Acadesine on Morbidity and Mortality Associated With Coronary Artery Bypass Grafting: The RED-CABG Randomized Controlled Trial. JAMA 2012, 308, 157–164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Leung, J.M.; Stanley, T.; Mathew, J.; Curling, P.; Barash, P.; Salmenpera, M.; Reves, J.G.; Hollenberg, M.; Mangano, D.T. An Initial Multicenter, Randomized Controlled Trial on the Safety and Efficacy of Acadesine in Patients Undergoing Coronary Artery Bypass Graft Surgery. Anesthesia Analg. 1994, 78, 420–434. [Google Scholar] [CrossRef]
  17. Holdright, D.R.; Sparrow, J.L.; Wright, C.L.; Steiner, J.; Fox, K.M. Effect of acadesine, a new metabolic agent, on exercise-induced myocardial ischemia in chronic stable angina. Cardiovasc. Drugs Ther. 1994, 8, 193–197. [Google Scholar] [CrossRef]
  18. Menasché, P.; Jamieson, W.E.; Flameng, W.; Davies, M.K. Acadesine: A new drug that may improve myocardial protection in coronary artery bypass grafting. J. Thorac. Cardiovasc. Surg. 1995, 110, 1096–1106. [Google Scholar] [CrossRef] [Green Version]
  19. The Multicenter Study of Perioperative Ischemia (McSPI) Research Group. Effects of Acadesine on the Incidence of Myocardial Infarction and Adverse Cardiac Outcomes after Coronary Artery Bypass Graft Surgery. Anesthesiology 1995, 83, 658–673. [Google Scholar] [CrossRef]
  20. Gadalla, A.E.; Pearson, T.; Currie, A.J.; Dale, N.; Hawley, S.A.; Sheehan, M.; Hirst, W.; Michel, A.D.; Randall, A.; Hardie, D.G.; et al. AICA riboside both activates AMP-activated protein kinase and competes with adenosine for the nucleoside transporter in the CA1 region of the rat hippocampus. J. Neurochem. 2004, 88, 1272–1282. [Google Scholar] [CrossRef] [Green Version]
  21. Hawley, S.A.; Ross, F.A.; Russell, F.M.; Atrih, A.; Lamont, D.J.; Hardie, D.G. Mechanism of Activation of AMPK by Cordycepin. Cell Chem. Biol. 2020, 27, 214–222.e4. [Google Scholar] [CrossRef] [Green Version]
  22. Hawley, S.A.; Ross, F.A.; Chevtzoff, C.; Green, K.A.; Evans, A.; Fogarty, S.; Towler, M.C.; Brown, L.J.; Ogunbayo, O.A.; Evans, A.M.; et al. Use of Cells Expressing γ Subunit Variants to Identify Diverse Mechanisms of AMPK Activation. Cell Metab. 2010, 11, 554–565. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Russell, F.M.; Hardie, D.G. AMP-Activated Protein Kinase: Do We Need Activators or Inhibitors to Treat or Prevent Cancer? Int. J. Mol. Sci. 2021, 22, 186. [Google Scholar] [CrossRef]
  24. Henin, N.; Vincent, M.-F.; Gruber, H.E.; Van Den Berghe, G. Inhibition of fatty acid and cholesterol synthesis by stimulation of AMP-activated protein kinase. FASEB J. 1995, 9, 541–546. [Google Scholar] [CrossRef] [PubMed]
  25. Velasco, G.; Geelen, M.J.H.; Guzmán, M. Control of Hepatic Fatty Acid Oxidation by 5′-AMP-Activated Protein Kinase Involves a Malonyl-CoA-Dependent and a Malonyl-CoA-Independent Mechanism. Arch. Biochem. Biophys. 1997, 337, 169–175. [Google Scholar] [CrossRef] [PubMed]
  26. Lochhead, P.A.; Salt, I.P.; Walker, K.S.; Hardie, D.G.; Sutherland, C. 5-aminoimidazole-4-carboxamide riboside mimics the effects of insulin on the expression of the 2 key gluconeogenic genes PEPCK and glucose-6-phosphatase. Diabetes 2000, 49, 896–903. [Google Scholar] [CrossRef] [Green Version]
  27. Young, M.E.; Radda, G.K.; Leighton, B. Activation of glycogen phosphorylase and glycogenolysis in rat skeletal muscle by AICAR—An activator of AMP-activated protein kinase. FEBS Lett. 1996, 382, 43–47. [Google Scholar] [CrossRef] [Green Version]
  28. Merrill, G.F.; Kurth, E.J.; Hardie, D.G.; Winder, W.W. AICA riboside increases AMP-activated protein kinase, fatty acid oxidation, and glucose uptake in rat muscle. Am. J. Physiol. Metab. 1997, 273, E1107–E1112. [Google Scholar] [CrossRef] [PubMed]
  29. Bolster, D.R.; Crozier, S.J.; Kimball, S.R.; Jefferson, L.S. AMP-activated Protein Kinase Suppresses Protein Synthesis in Rat Skeletal Muscle through Down-regulated Mammalian Target of Rapamycin (mTOR) Signaling. J. Biol. Chem. 2002, 277, 23977–23980. [Google Scholar] [CrossRef] [Green Version]
  30. Longnus, S.L.; Wambolt, R.B.; Parsons, H.L.; Brownsey, R.W.; Allard, M.F. 5-Aminoimidazole-4-carboxamide 1-β-d-ribofuranoside (AICAR) stimulates myocardial glycogenolysis by allosteric mechanisms. Am. J. Physiol. Integr. Comp. Physiol. 2003, 284, R936–R944. [Google Scholar] [CrossRef]
  31. Hunter, R.W.; Hughey, C.C.; Lantier, L.; Sundelin, E.I.; Peggie, M.; Zeqiraj, E.; Sicheri, F.; Jessen, N.; Wasserman, D.H.; Sakamoto, K. Metformin reduces liver glucose production by inhibition of fructose-1-6-bisphosphatase. Nat. Med. 2018, 24, 1395–1406. [Google Scholar] [CrossRef] [Green Version]
  32. Mu, J.; Brozinick, J.T.; Valladares, O.; Bucan, M.; Birnbaum, M.J. A Role for AMP-Activated Protein Kinase in Contraction- and Hypoxia-Regulated Glucose Transport in Skeletal Muscle. Mol. Cell 2001, 7, 1085–1094. [Google Scholar] [CrossRef]
  33. Jørgensen, S.B.; Viollet, B.; Andreelli, F.; Frøsig, C.; Birk, J.B.; Schjerling, P.; Vaulont, S.; Richter, E.A.; Wojtaszewski, J.F.P. Knockout of the α2 but Not α1 5′-AMP-activated Protein Kinase Isoform Abolishes 5-Aminoimidazole-4-carboxamide-1-β-4-ribofuranosidebut Not Contraction-induced Glucose Uptake in Skeletal Muscle. J. Biol. Chem. 2004, 279, 1070–1079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Barnes, B.R.; Marklund, S.; Steiler, T.L.; Walter, M.; Göran, H.; Amarger, V.; Mahlapuu, M.; Leng, Y.; Johansson, C.; Galuska, D.; et al. The 5′-AMP-activated Protein Kinase γ3 Isoform Has a Key Role in Carbohydrate and Lipid Metabolism in Glycolytic Skeletal Muscle. J. Biol. Chem. 2004, 279, 38441–38447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Fujii, N.; Hirshman, M.F.; Kane, E.M.; Ho, R.C.; Peter, L.E.; Seifert, M.M.; Goodyear, L.J. AMP-activated Protein Kinase α2 Activity Is Not Essential for Contraction- and Hyperosmolarity-induced Glucose Transport in Skeletal Muscle. J. Biol. Chem. 2005, 280, 39033–39041. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Boudaba, N.; Marion, A.; Huet, C.; Pierre, R.; Viollet, B.; Foretz, M. AMPK Re-Activation Suppresses Hepatic Steatosis but its Downregulation Does Not Promote Fatty Liver Development. EBioMedicine 2018, 28, 194–209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Choi, R.H.; McConahay, A.; Johnson, M.B.; Jeong, H.-W.; Koh, H.-J. Adipose tissue-specific knockout of AMPKα1/α2 results in normal AICAR tolerance and glucose metabolism. Biochem. Biophys. Res. Commun. 2019, 519, 633–638. [Google Scholar] [CrossRef]
  38. Foretz, M.; Hébrard, S.; Leclerc, J.; Zarrinpashneh, E.; Soty, M.; Mithieux, G.; Sakamoto, K.; Andreelli, F.; Viollet, B. Metformin inhibits hepatic gluconeogenesis in mice independently of the LKB1/AMPK pathway via a decrease in hepatic energy state. J. Clin. Investig. 2010, 120, 2355–2369. [Google Scholar] [CrossRef] [Green Version]
  39. Guigas, B.; Bertrand, L.; Taleux, N.; Foretz, M.; Wiernsperger, N.; Vertommen, D.; Andreelli, F.; Viollet, B.; Hue, L. 5-Aminoimidazole-4-Carboxamide-1-β-Ribofuranoside and Metformin Inhibit Hepatic Glucose Phosphorylation by an AMP-Activated Protein Kinase–Independent Effect on Glucokinase Translocation. Diabetes 2006, 55, 865–874. [Google Scholar] [CrossRef] [Green Version]
  40. Guigas, B.; Taleux, N.; Foretz, M.; Detaille, D.; Andreelli, F.; Viollet, B.; Hue, L. AMP-activated protein kinase-independent inhibition of hepatic mitochondrial oxidative phosphorylation by AICA riboside. Biochem. J. 2007, 404, 499–507. [Google Scholar] [CrossRef]
  41. Viollet, B.; Lantier, L.; Devin-Leclerc, J.; Hebrard, S.; Amouyal, C.; Mounier, R.; Foretz, M.; Andreelli, F. Targeting the AMPK pathway for the treatment of Type 2 diabetes. Front. Biosci. (Landmark Ed.) 2009, 14, 3380–3400. [Google Scholar] [CrossRef] [Green Version]
  42. Bergeron, R.; Previs, S.F.; Cline, G.W.; Perret, P.; Russell, R.R., III; Young, L.H.; Shulman, G.I. Effect of 5-Aminoimidazole-4-Carboxamide-1-β-Ribofuranoside Infusion on In Vivo Glucose and Lipid Metabolism in Lean and Obese Zucker Rats. Diabetes 2001, 50, 1076–1082. [Google Scholar] [CrossRef] [Green Version]
  43. Cuthbertson, D.J.; Babraj, J.A.; Mustard, K.J.W.; Towler, M.C.; Green, K.A.; Wackerhage, H.; Leese, G.P.; Baar, K.; Thomason-Hughes, M.; Sutherland, C.; et al. 5-Aminoimidazole-4-Carboxamide 1-β-Ribofuranoside Acutely Stimulates Skeletal Muscle 2-Deoxyglucose Uptake in Healthy Men. Diabetes 2007, 56, 2078–2084. [Google Scholar] [CrossRef] [Green Version]
  44. Song, X.M.; Fiedler, M.; Galuska, D.; Ryder, J.W.; Fernström, M.; Chibalin, A.V.; Wallberg-Henriksson, H.; Zierath, J.R. 5-Aminoimidazole-4-carboxamide ribonucleoside treatment improves glucose homeostasis in insulin-resistant diabetic (ob/ob) mice. Diabetologia 2002, 45, 56–65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Buhl, E.S.; Jessen, N.; Pold, R.; Ledet, T.; Flyvbjerg, A.; Pedersen, S.B.; Pedersen, O.; Schmitz, O.; Lund, S. Long-Term AICAR Administration Reduces Metabolic Disturbances and Lowers Blood Pressure in Rats Displaying Features of the Insulin Resistance Syndrome. Diabetes 2002, 51, 2199–2206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Iglesias, M.A.; Ye, J.-M.; Frangioudakis, G.; Saha, A.K.; Tomas, E.; Ruderman, N.B.; Cooney, G.J.; Kraegen, E.W. AICAR Administration Causes an Apparent Enhancement of Muscle and Liver Insulin Action in Insulin-Resistant High-Fat-Fed Rats. Diabetes 2002, 51, 2886–2894. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Pold, R.; Jensen, L.S.; Jessen, N.; Buhl, E.S.; Schmitz, O.; Flyvbjerg, A.; Fujii, N.; Goodyear, L.J.; Gotfredsen, C.F.; Brand, C.L.; et al. Long-Term AICAR Administration and Exercise Prevents Diabetes in ZDF Rats. Diabetes 2005, 54, 928–934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Boon, H.; Bosselaar, M.; Praet, S.F.E.; Blaak, E.E.; Saris, W.H.M.; Wagenmakers, A.J.M.; McGee, S.L.; Tack, C.J.; Smits, P.; Hargreaves, M.; et al. Intravenous AICAR administration reduces hepatic glucose output and inhibits whole body lipolysis in type 2 diabetic patients. Diabetologia 2008, 51, 1893. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Dixon, R.; Gourzis, J.; McDermott, D.; Fujitaki, J.; Dewland, P.; Gruber, H. AICA-Riboside: Safety, Tolerance, and Pharmacokinetics of a Novel Adenosine-Regulating Agent. J. Clin. Pharmacol. 1991, 31, 342–347. [Google Scholar] [CrossRef]
  50. Dzamko, N.; Schertzer, J.D.; Ryall, J.G.; Steel, R.; Macaulay, S.L.; Wee, S.; Chen, Z.-P.; Michell, B.J.; Oakhill, J.S.; Watt, M.J.; et al. AMPK-independent pathways regulate skeletal muscle fatty acid oxidation. J. Physiol. 2008, 586, 5819–5831. [Google Scholar] [CrossRef] [PubMed]
  51. Jørgensen, S.B.; Wojtaszewski, J.F.P.; Viollet, B.; Andreelli, F.; Birk, J.B.; Hellsten, Y.; Schjerling, P.; Vaulont, S.; Neufer, P.D.; Richter, E.A.; et al. Effects of α-AMPK knockout on exercise-induced gene activation in mouse skeletal muscle. FASEB J. 2005, 19, 1146–1148. [Google Scholar] [CrossRef]
  52. Daval, M.; Diot-Dupuy, F.; Bazin, R.; Hainault, I.; Viollet, B.; Vaulont, S.; Hajduch, E.; Ferré, P.; Foufelle, F. Anti-lipolytic Action of AMP-activated Protein Kinase in Rodent Adipocytes. J. Biol. Chem. 2005, 280, 25250–25257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Mauser, M.; Hoffmeister, H.M.; Nienaber, C.; Schaper, W. Influence of ribose, adenosine, and “AICAR” on the rate of myocardial adenosine triphosphate synthesis during reperfusion after coronary artery occlusion in the dog. Circ. Res. 1985, 56, 220–230. [Google Scholar] [CrossRef] [Green Version]
  54. Gruber, H.E.; Hoffer, M.E.; McAllister, D.R.; Laikind, P.K.; Lane, T.A.; Schmid-Schoenbein, G.W.; Engler, R.L. Increased adenosine concentration in blood from ischemic myocardium by AICA riboside. Effects on flow, granulocytes, and injury. Circulation 1989, 80, 1400–1411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Russell, R.R.; Bergeron, R.; Shulman, G.I.; Young, L.H. Translocation of myocardial GLUT-4 and increased glucose uptake through activation of AMPK by AICAR. Am. J. Physiol. Circ. Physiol. 1999, 277, H643–H649. [Google Scholar] [CrossRef]
  56. Arad, M.; Seidman, C.E.; Seidman, J. AMP-Activated Protein Kinase in the Heart. Circ. Res. 2007, 100, 474–488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Ren, J.; Xu, X.; Wang, Q.; Ren, S.Y.; Dong, M.; Zhang, Y. Permissive role of AMPK and autophagy in adiponectin deficiency-accentuated myocardial injury and inflammation in endotoxemia. J. Mol. Cell. Cardiol. 2016, 93, 18–31. [Google Scholar] [CrossRef]
  58. Nam, D.H.; Kim, E.; Benham, A.; Park, H.-K.; Soibam, B.; Taffet, G.E.; Kaelber, J.T.; Suh, J.H.; Taegtmeyer, H.; Entman, M.L.; et al. Transient activation of AMPK preceding left ventricular pressure overload reduces adverse remodeling and preserves left ventricular function. FASEB J. 2019, 33, 711–721. [Google Scholar] [CrossRef]
  59. Viglino, C.; Foglia, B.; Montessuit, C. Chronic AICAR treatment prevents metabolic changes in cardiomyocytes exposed to free fatty acids. Pflügers Arch. Eur. J. Physiol. 2019, 471, 1219–1234. [Google Scholar] [CrossRef]
  60. Morrow, V.A.; Foufelle, F.; Connell, J.M.C.; Petrie, J.R.; Gould, G.W.; Salt, I.P. Direct Activation of AMP-activated Protein Kinase Stimulates Nitric-oxide Synthesis in Human Aortic Endothelial Cells. J. Biol. Chem. 2003, 278, 31629–31639. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Nagata, D.; Mogi, M.; Walsh, K. AMP-activated Protein Kinase (AMPK) Signaling in Endothelial Cells Is Essential for Angiogenesis in Response to Hypoxic Stress. J. Biol. Chem. 2003, 278, 31000–31006. [Google Scholar] [CrossRef] [Green Version]
  62. Ylikorkala, A.; Rossi, D.J.; Korsisaari, N.; Luukko, K.; Alitalo, K.; Henkemeyer, M.; Mäkelä, T.P. Vascular Abnormalities and Deregulation of VEGF in Lkb1-Deficient Mice. Science 2001, 293, 1323–1326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Page, T.; Barshop, B.; Yu, A.L.; Nyhan, W.L. Treatment of Lesch-Nyhan syndrome with AICAR. Adv. Exp. Med. Biol. 1994, 370, 353–356. [Google Scholar] [CrossRef]
  64. Bosselaar, M.; Boon, H.; van Loon, L.J.C.; van den Broek, P.H.H.; Smits, P.; Tack, C.J. Intra-arterial AICA-riboside administration induces NO-dependent vasodilation in vivo in human skeletal muscle. Am. J. Physiol. Metab. 2009, 297, E759–E766. [Google Scholar] [CrossRef]
  65. Narkar, V.A.; Downes, M.; Yu, R.T.; Embler, E.; Wang, Y.-X.; Banayo, E.; Mihaylova, M.M.; Nelson, M.C.; Zou, Y.; Juguilon, H.; et al. AMPK and PPARδ Agonists Are Exercise Mimetics. Cell 2008, 134, 405–415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Pokrywka, A.; Cholbinski, P.; Kaliszewski, P.; Kowalczyk, K.; Konczak, D.; Zembron-Lacny, A. Metabolic modulators of the exercise response: Doping control analysis of an agonist of the peroxisome proliferator-activated receptor δ (GW501516) and 5-aminoimidazole-4-carboxamide ribonucleotide (AICAR). J. Physiol. Pharmacol. Off. J. Pol. Physiol. Soc. 2014, 65, 469–476. [Google Scholar]
  67. Fan, W.; Evans, R.M. Exercise Mimetics: Impact on Health and Performance. Cell Metab. 2017, 25, 242–247. [Google Scholar] [CrossRef] [Green Version]
  68. Kjøbsted, R.; Treebak, J.T.; Fentz, J.; Lantier, L.; Viollet, B.; Birk, J.B.; Schjerling, P.; Björnholm, M.; Zierath, J.R.; Wojtaszewski, J.F.P. Prior AICAR Stimulation Increases Insulin Sensitivity in Mouse Skeletal Muscle in an AMPK-Dependent Manner. Diabetes 2015, 64, 2042–2055. [Google Scholar] [CrossRef] [Green Version]
  69. Kjøbsted, R.; Munk-Hansen, N.; Birk, J.B.; Foretz, M.; Viollet, B.; Björnholm, M.; Zierath, J.R.; Treebak, J.T.; Wojtaszewski, J.F.P. Enhanced Muscle Insulin Sensitivity After Contraction/Exercise Is Mediated by AMPK. Diabetes 2017, 66, 598–612. [Google Scholar] [CrossRef] [Green Version]
  70. Sajan, M.P.; Bandyopadhyay, G.; Miura, A.; Standaert, M.L.; Nimal, S.; Longnus, S.L.; Van Obberghen, E.; Hainault, I.; Foufelle, F.; Kahn, R.; et al. AICAR and metformin, but not exercise, increase muscle glucose transport through AMPK-, ERK-, and PDK1-dependent activation of atypical PKC. Am. J. Physiol. Metab. 2009, 298, E179–E192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Ezagouri, S.; Zwighaft, Z.; Sobel, J.; Baillieul, S.; Doutreleau, S.; Ladeuix, B.; Golik, M.; Verges, S.; Asher, G. Physiological and Molecular Dissection of Daily Variance in Exercise Capacity. Cell Metab. 2019, 30, 78–91.e4. [Google Scholar] [CrossRef]
  72. Reznick, R.M.; Zong, H.; Li, J.; Morino, K.; Moore, I.K.; Yu, H.J.; Liu, Z.-X.; Dong, J.; Mustard, K.J.; Hawley, S.A.; et al. Aging-associated reductions in AMP-activated protein kinase activity and mitochondrial biogenesis. Cell Metab. 2007, 5, 151–156. [Google Scholar] [CrossRef] [Green Version]
  73. Kobilo, T.; Guerrieri, D.; Zhang, Y.; Collica, S.C.; Becker, K.G.; van Praag, H. AMPK agonist AICAR improves cognition and motor coordination in young and aged mice. Learn. Mem. 2014, 21, 119–126. [Google Scholar] [CrossRef] [Green Version]
  74. Pauly, M.; Chabi, B.; Favier, F.B.; Vanterpool, F.; Matecki, S.; Fouret, G.; Bonafos, B.; Vernus, B.; Feillet-Coudray, C.; Coudray, C.; et al. Combined Strategies for Maintaining Skeletal Muscle Mass and Function in Aging: Myostatin Inactivation and AICAR-Associated Oxidative Metabolism Induction. J. Gerontol. Ser. A 2015, 70, 1077–1087. [Google Scholar] [CrossRef] [Green Version]
  75. Bueno Júnior, C.R.; Pantaleão, L.C.; Voltarelli, V.A.; Bozi, L.H.M.; Brum, P.C.; Zatz, M. Combined Effect of AMPK/PPAR Agonists and Exercise Training in mdx Mice Functional Performance. PLoS ONE 2012, 7, e45699. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Pauly, M.; Daussin, F.; Burelle, Y.; Li, T.; Godin, R.; Fauconnier, J.; Koechlin-Ramonatxo, C.; Hugon, G.; Lacampagne, A.; Coisy-Quivy, M.; et al. AMPK activation stimulates autophagy and ameliorates muscular dystrophy in the mdx mouse diaphragm. Am. J. Pathol. 2012, 181, 583–592. [Google Scholar] [CrossRef]
  77. Gwinn, D.M.; Shackelford, D.B.; Egan, D.F.; Mihaylova, M.M.; Mery, A.; Vasquez, D.S.; Turk, B.E.; Shaw, R.J. AMPK phosphorylation of raptor mediates a metabolic checkpoint. Mol. Cell 2008, 30, 214–226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Kazyken, D.; Magnuson, B.; Bodur, C.; Acosta-Jaquez, H.A.; Zhang, D.; Tong, X.; Barnes, T.M.; Steinl, G.K.; Patterson, N.E.; Altheim, C.H.; et al. AMPK directly activates mTORC2 to promote cell survival during acute energetic stress. Sci. Signal. 2019, 12. [Google Scholar] [CrossRef]
  79. Steinberg, G.R.; Carling, D. AMP-activated protein kinase: The current landscape for drug development. Nat. Rev. Drug Discov. 2019, 18, 527–551. [Google Scholar] [CrossRef] [PubMed]
  80. Imamura, K.; Ogura, T.; Kishimoto, A.; Kaminishi, M.; Esumi, H. Cell Cycle Regulation via p53 Phosphorylation by a 5′-AMP Activated Protein Kinase Activator, 5-Aminoimidazole- 4-Carboxamide-1-β-d-Ribofuranoside, in a Human Hepatocellular Carcinoma Cell Line. Biochem. Biophys. Res. Commun. 2001, 287, 562–567. [Google Scholar] [CrossRef]
  81. Jones, R.G.; Plas, D.R.; Kubek, S.; Buzzai, M.; Mu, J.; Xu, Y.; Birnbaum, M.J.; Thompson, C.B. AMP-Activated Protein Kinase Induces a p53-Dependent Metabolic Checkpoint. Mol. Cell 2005, 18, 283–293. [Google Scholar] [CrossRef]
  82. Rattan, R.; Giri, S.; Singh, A.K.; Singh, I. 5-Aminoimidazole-4-carboxamide-1-β-D-ribofuranoside Inhibits Cancer Cell Proliferation in Vitro and in Vivo via AMP-activated Protein Kinase. J. Biol. Chem. 2005, 280, 39582–39593. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Guo, D.; Hildebrandt, I.J.; Prins, R.M.; Soto, H.; Mazzotta, M.M.; Dang, J.; Czernin, J.; Shyy, J.Y.-J.; Watson, A.D.; Phelps, M.; et al. The AMPK agonist AICAR inhibits the growth of EGFRvIII-expressing glioblastomas by inhibiting lipogenesis. Proc. Natl. Acad. Sci. USA 2009, 106, 12932–12937. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Tang, Y.-C.; Williams, B.R.; Siegel, J.J.; Amon, A. Identification of Aneuploidy-Selective Antiproliferation Compounds. Cell 2011, 144, 499–512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Chae, H.-D.; Lee, M.-R.; Broxmeyer, H.E. 5-Aminoimidazole-4-carboxyamide Ribonucleoside Induces G1/S Arrest and Nanog Downregulation via p53 and Enhances Erythroid Differentiation. Stem Cells 2012, 30, 140–149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Vincent, E.E.; Coelho, P.P.; Blagih, J.; Griss, T.; Viollet, B.; Jones, R.G. Differential effects of AMPK agonists on cell growth and metabolism. Oncogene 2015, 34, 3627–3639. [Google Scholar] [CrossRef] [Green Version]
  87. Towler, M.C.; Hardie, D.G. AMP-Activated Protein Kinase in Metabolic Control and Insulin Signaling. Circ. Res. 2007, 100, 328–341. [Google Scholar] [CrossRef] [PubMed]
  88. Myers, R.W.; Guan, H.-P.; Ehrhart, J.; Petrov, A.; Prahalada, S.; Tozzo, E.; Yang, X.; Kurtz, M.M.; Trujillo, M.; Gonzalez Trotter, D.; et al. Systemic pan-AMPK activator MK-8722 improves glucose homeostasis but induces cardiac hypertrophy. Science 2017, 357, 507–511. [Google Scholar] [CrossRef] [Green Version]
  89. Cokorinos, E.C.; Delmore, J.; Reyes, A.R.; Albuquerque, B.; Kjøbsted, R.; Jørgensen, N.O.; Tran, J.-L.; Jatkar, A.; Cialdea, K.; Esquejo, R.M.; et al. Activation of Skeletal Muscle AMPK Promotes Glucose Disposal and Glucose Lowering in Non-human Primates and Mice. Cell Metab. 2017, 25, 1147–1159.e10. [Google Scholar] [CrossRef]
  90. Langendorf, C.G.; Ngoei, K.R.W.; Scott, J.W.; Ling, N.X.Y.; Issa, S.M.A.; Gorman, M.A.; Parker, M.W.; Sakamoto, K.; Oakhill, J.S.; Kemp, B.E. Structural basis of allosteric and synergistic activation of AMPK by furan-2-phosphonic derivative C2 binding. Nat. Commun. 2016, 7, 10912. [Google Scholar] [CrossRef]
  91. Liu, X.; Chhipa, R.R.; Pooya, S.; Wortman, M.; Yachyshin, S.; Chow, L.M.L.; Kumar, A.; Zhou, X.; Sun, Y.; Quinn, B.; et al. Discrete mechanisms of mTOR and cell cycle regulation by AMPK agonists independent of AMPK. Proc. Natl. Acad. Sci. USA 2014, 111, E435–E444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Santidrián, A.F.; González-Gironès, D.M.; Iglesias-Serret, D.; Coll-Mulet, L.; Cosialls, A.M.; de Frias, M.; Campàs, C.; González-Barca, E.; Alonso, E.; Labi, V.; et al. AICAR induces apoptosis independently of AMPK and p53 through up-regulation of the BH3-only proteins BIM and NOXA in chronic lymphocytic leukemia cells. Blood 2010, 116, 3023–3032. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Lalic, H.; Dembitz, V.; Lukinovic-Skudar, V.; Banfic, H.; Visnjic, D. 5-Aminoimidazole-4-carboxamide ribonucleoside induces differentiation of acute myeloid leukemia cells. Leuk. Lymphoma 2014, 55, 2375–2383. [Google Scholar] [CrossRef] [PubMed]
  94. Mukhopadhyay, S.; Chatterjee, A.; Kogan, D.; Patel, D.; Foster, D.A. 5-Aminoimidazole-4-carboxamide-1-β-4-ribofuranoside (AICAR) enhances the efficacy of rapamycin in human cancer cells. Cell Cycle 2015, 14, 3331–3339. [Google Scholar] [CrossRef] [Green Version]
  95. Dembitz, V.; Tomic, B.; Kodvanj, I.; Simon, J.A.; Bedalov, A.; Visnjic, D. The ribonucleoside AICAr induces differentiation of myeloid leukemia by activating the ATR/Chk1 via pyrimidine depletion. J. Biol. Chem. 2019, 294, 15257–15270. [Google Scholar] [CrossRef] [Green Version]
  96. Nyhan, W.L. Nucleotide Synthesis via Salvage Pathway. eLS 2014. [Google Scholar] [CrossRef]
  97. López, J.M.; Outtrim, E.L.; Fu, R.; Sutcliffe, D.J.; Torres, R.J.; Jinnah, H.A. Physiological levels of folic acid reveal purine alterations in Lesch-Nyhan disease. Proc. Natl. Acad. Sci. USA 2020, 117, 12071–12079. [Google Scholar] [CrossRef]
  98. Douillet, D.C.; Pinson, B.; Ceschin, J.; Hürlimann, H.C.; Saint-Marc, C.; Laporte, D.; Claverol, S.; Konrad, M.; Bonneu, M.; Daignan-Fornier, B. Metabolomics and proteomics identify the toxic form and the associated cellular binding targets of the anti-proliferative drug AICAR. J. Biol. Chem. 2019, 294, 805–815. [Google Scholar] [CrossRef] [Green Version]
  99. Ceschin, J.; Hürlimann, H.C.; Saint-Marc, C.; Albrecht, D.; Violo, T.; Moenner, M.; Daignan-Fornier, B.; Pinson, B. Disruption of Nucleotide Homeostasis by the Antiproliferative Drug 5-Aminoimidazole-4-carboxamide-1-β-d-ribofuranoside Monophosphate (AICAR). J. Biol. Chem. 2015, 290, 23947–23959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Paglia, G.; Hrafnsdóttir, S.; Magnúsdóttir, M.; Fleming, R.M.T.; Thorlacius, S.; Palsson, B.Ø.; Thiele, I. Monitoring metabolites consumption and secretion in cultured cells using ultra-performance liquid chromatography quadrupole–time of flight mass spectrometry (UPLC–Q–ToF-MS). Anal. Bioanal. Chem. 2012, 402, 1183–1198. [Google Scholar] [CrossRef] [Green Version]
  101. Bardeleben, C.; Sharma, S.; Reeve, J.R.; Bassilian, S.; Frost, P.; Hoang, B.; Shi, Y.; Lichtenstein, A. Metabolomics Identifies Pyrimidine Starvation as the Mechanism of 5-Aminoimidazole-4-Carboxamide-1-β-Riboside-Induced Apoptosis in Multiple Myeloma Cells. Mol. Cancer Ther. 2013, 12, 1310–1321. [Google Scholar] [CrossRef] [Green Version]
  102. Du, L.; Yang, F.; Fang, H.; Sun, H.; Chen, Y.; Xu, Y.; Li, H.; Zheng, L.; Zhou, B.-B.S. AICAr suppresses cell proliferation by inducing NTP and dNTP pool imbalances in acute lymphoblastic leukemia cells. FASEB J. 2019, 33, 4525–4537. [Google Scholar] [CrossRef] [PubMed]
  103. Thomas, C.B.; Meade, J.C.; Holmes, E.W. Aminoimidazole carboxamide ribonucleoside toxicity: A model for study of pyrimidine starvation. J. Cell. Physiol. 1981, 107, 335–344. [Google Scholar] [CrossRef] [PubMed]
  104. ElAzzouny, M.A.; Evans, C.R.; Burant, C.F.; Kennedy, R.T. Metabolomics Analysis Reveals that AICAR Affects Glycerolipid, Ceramide and Nucleotide Synthesis Pathways in INS-1 Cells. PLoS ONE 2015, 10, e0129029. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Asby, D.J.; Cuda, F.; Beyaert, M.; Houghton, F.D.; Cagampang, F.R.; Tavassoli, A. AMPK Activation via Modulation of De Novo Purine Biosynthesis with an Inhibitor of ATIC Homodimerization. Chem. Biol. 2015, 22, 838–848. [Google Scholar] [CrossRef] [Green Version]
  106. Agarwal, S.; Bell, C.M.; Rothbart, S.B.; Moran, R.G. AMP-activated Protein Kinase (AMPK) Control of mTORC1 Is p53- and TSC2-independent in Pemetrexed-treated Carcinoma Cells. J. Biol. Chem. 2015, 290, 27473–27486. [Google Scholar] [CrossRef] [Green Version]
  107. Beckers, A.; Organe, S.; Timmermans, L.; Vanderhoydonc, F.; Deboel, L.; Derua, R.; Waelkens, E.; Brusselmans, K.; Verhoeven, G.; Swinnen, J. V Methotrexate enhances the antianabolic and antiproliferative effects of 5-aminoimidazole-4-carboxamide riboside. Mol. Cancer Ther. 2006, 5, 2211–2217. [Google Scholar] [CrossRef] [Green Version]
  108. Pirkmajer, S.; Kulkarni, S.S.; Tom, R.Z.; Ross, F.A.; Hawley, S.A.; Hardie, D.G.; Zierath, J.R.; Chibalin, A. V Methotrexate Promotes Glucose Uptake and Lipid Oxidation in Skeletal Muscle via AMPK Activation. Diabetes 2015, 64, 360–369. [Google Scholar] [CrossRef] [Green Version]
  109. Evans, J.M.M.; Donnelly, L.A.; Emslie-Smith, A.M.; Alessi, D.R.; Morris, A.D. Metformin and reduced risk of cancer in diabetic patients. BMJ 2005, 330, 1304–1305. [Google Scholar] [CrossRef] [Green Version]
  110. Sengupta, T.K.; Leclerc, G.M.; Hsieh-Kinser, T.T.; Leclerc, G.J.; Singh, I.; Barredo, J.C. Cytotoxic effect of 5-aminoimidazole-4-carboxamide-1-β-4-ribofuranoside (AICAR) on childhood acute lymphoblastic leukemia (ALL) cells: Implication for targeted therapy. Mol. Cancer 2007, 6, 46. [Google Scholar] [CrossRef] [Green Version]
  111. Vakana, E.; Altman, J.K.; Glaser, H.; Donato, N.J.; Platanias, L.C. Antileukemic effects of AMPK activators on BCR-ABL–expressing cells. Blood 2011, 118, 6399–6402. [Google Scholar] [CrossRef] [Green Version]
  112. Dembitz, V.; Lalic, H.; Kodvanj, I.; Tomic, B.; Batinic, J.; Dubravcic, K.; Batinic, D.; Bedalov, A.; Visnjic, D. 5-aminoimidazole-4-carboxamide ribonucleoside induces differentiation in a subset of primary acute myeloid leukemia blasts. BMC Cancer 2020, 20, 1090. [Google Scholar] [CrossRef]
  113. Dembitz, V.; Lalic, H.; Visnjic, D. 5-Aminoimidazole-4-carboxamide ribonucleoside-induced autophagy flux during differentiation of monocytic leukemia cells. Cell Death Discov. 2017, 3, 17066. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. De Thé, H. Differentiation therapy revisited. Nat. Rev. Cancer 2018, 18, 117–127. [Google Scholar] [CrossRef]
  115. Kanazawa, I.; Yamaguchi, T.; Yano, S.; Yamauchi, M.; Sugimoto, T. Activation of AMP kinase and inhibition of Rho kinase induce the mineralization of osteoblastic MC3T3-E1 cells through endothelial NOS and BMP-2 expression. Am. J. Physiol. Metab. 2009, 296, E139–E146. [Google Scholar] [CrossRef] [Green Version]
  116. Zang, Y.; Yu, L.-F.; Pang, T.; Fang, L.-P.; Feng, X.; Wen, T.-Q.; Nan, F.-J.; Feng, L.-Y.; Li, J. AICAR Induces Astroglial Differentiation of Neural Stem Cells via Activating the JAK/STAT3 Pathway Independently of AMP-activated Protein Kinase. J. Biol. Chem. 2008, 283, 6201–6208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Sykes, D.B.; Kfoury, Y.S.; Mercier, F.E.; Wawer, M.J.; Law, J.M.; Haynes, M.K.; Lewis, T.A.; Schajnovitz, A.; Jain, E.; Lee, D.; et al. Inhibition of Dihydroorotate Dehydrogenase Overcomes Differentiation Blockade in Acute Myeloid Leukemia. Cell 2016, 167, 171–186.e15. [Google Scholar] [CrossRef] [Green Version]
  118. Vara-Ciruelos, D.; Russell, F.M.; Hardie, D.G. The strange case of AMPK and cancer: Dr Jekyll or Mr Hyde?. Open Biol. 2021, 9, 190099. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. AICAr structure and mechanism.
Figure 1. AICAr structure and mechanism.
Cells 10 01095 g001
Figure 2. A schematic presentation of molecular mechanisms of AICAr Akt/PKB, Akt/protein kinase B; PDK1, 3-phosphoinositide dependent protein kinase-1; PI3K, phosphatidylinositol 3-kinase; PIP3, phosphatidylinositol (3,4,5)-trisphosphate; IRS, insulin receptor substrate; TSC1/2, tuberous sclerosis complex 1 protein / tuberous sclerosis complex 2 protein; Rheb, Ras homolog enriched in the brain; mTOR, mechanistic target of rapamycin; mTORC1, mTOR complex 1; mTORC2, mTOR complex 2; Rictor, rapamycin-insensitive companion of mTOR; PRAS40, proline-rich Akt substrate 40; Raptor, regulatory-associated protein of mTOR; AMPK, adenosine monophosphate (AMP)-activated protein kinase; p53, tumor suppressor protein p53; S6K1, p70 S6 Kinase 1; OXPHOS, oxidative phosphorylation; 4E-BP1, eukaryotic translation initiation factor 4E binding protein 1; GLUT4, glucose transporter 4.
Figure 2. A schematic presentation of molecular mechanisms of AICAr Akt/PKB, Akt/protein kinase B; PDK1, 3-phosphoinositide dependent protein kinase-1; PI3K, phosphatidylinositol 3-kinase; PIP3, phosphatidylinositol (3,4,5)-trisphosphate; IRS, insulin receptor substrate; TSC1/2, tuberous sclerosis complex 1 protein / tuberous sclerosis complex 2 protein; Rheb, Ras homolog enriched in the brain; mTOR, mechanistic target of rapamycin; mTORC1, mTOR complex 1; mTORC2, mTOR complex 2; Rictor, rapamycin-insensitive companion of mTOR; PRAS40, proline-rich Akt substrate 40; Raptor, regulatory-associated protein of mTOR; AMPK, adenosine monophosphate (AMP)-activated protein kinase; p53, tumor suppressor protein p53; S6K1, p70 S6 Kinase 1; OXPHOS, oxidative phosphorylation; 4E-BP1, eukaryotic translation initiation factor 4E binding protein 1; GLUT4, glucose transporter 4.
Cells 10 01095 g002
Figure 3. Schematic representation of AICAR involvement in purine and pyrimidine synthesis. D-Ribose-5-P, ribose 5-phosphate; PRPP, 5-phosphoribosyl 1-pyrophosphate; PPAT, phosphoribosyl pyrophosphate amidotransferase; Glu, glutamate; PPi, pyrophosphate; PRA, 5-phosphoribosylamine; GAR, 5'-phosphoribosylglycinamide; N10-formyl-THF, 10-formyl tetrahydrofolate; GART, glycinamide ribonucleotide transformylase; THF, tetrahydrofolate; FGAR, 5'-phosphoribosyl-N-formylglycinamide; SAICAR, 5'-phosphoribosyl-5-aminoimidazole-4-N-succinocarboxamide; ADSL, adenylosuccinate lyase; AICAr, 5-amino-4-imidazolecarboxamide (AICA) ribonucleoside; AICAR, 5-amino-4-imidazolecarboxamide (AICA) ribonucleotide; AICART, AICAR transformylase; FAICAR, 5-formamidoimidazole-4-carboxamide ribotide; IMP, inosine monophosphate; AMP, adenosine monophosphate; GMP, guanosine monophosphate; HGPRT, hypoxanthine-guanine phosphoribosyltransferase; CPS II, carbamoyl phosphate synthetase II; ATCase, aspartate transcarbamoylase; DHO, dihydroorotase; DHODH, dihydroorotate dehydrogenase; OPRT, orotate phosphoribosyltransferase; OMP, orotidine-5'-monophosphate; OMPDC, orotidine 5'-phosphate decarboxylase (OMP decarboxylase); UMP, uridine monophosphate; CTP, cytidine triphosphate; dCTP, deoxycytidine triphosphate; dTTP, deoxythymidine triphosphate.
Figure 3. Schematic representation of AICAR involvement in purine and pyrimidine synthesis. D-Ribose-5-P, ribose 5-phosphate; PRPP, 5-phosphoribosyl 1-pyrophosphate; PPAT, phosphoribosyl pyrophosphate amidotransferase; Glu, glutamate; PPi, pyrophosphate; PRA, 5-phosphoribosylamine; GAR, 5'-phosphoribosylglycinamide; N10-formyl-THF, 10-formyl tetrahydrofolate; GART, glycinamide ribonucleotide transformylase; THF, tetrahydrofolate; FGAR, 5'-phosphoribosyl-N-formylglycinamide; SAICAR, 5'-phosphoribosyl-5-aminoimidazole-4-N-succinocarboxamide; ADSL, adenylosuccinate lyase; AICAr, 5-amino-4-imidazolecarboxamide (AICA) ribonucleoside; AICAR, 5-amino-4-imidazolecarboxamide (AICA) ribonucleotide; AICART, AICAR transformylase; FAICAR, 5-formamidoimidazole-4-carboxamide ribotide; IMP, inosine monophosphate; AMP, adenosine monophosphate; GMP, guanosine monophosphate; HGPRT, hypoxanthine-guanine phosphoribosyltransferase; CPS II, carbamoyl phosphate synthetase II; ATCase, aspartate transcarbamoylase; DHO, dihydroorotase; DHODH, dihydroorotate dehydrogenase; OPRT, orotate phosphoribosyltransferase; OMP, orotidine-5'-monophosphate; OMPDC, orotidine 5'-phosphate decarboxylase (OMP decarboxylase); UMP, uridine monophosphate; CTP, cytidine triphosphate; dCTP, deoxycytidine triphosphate; dTTP, deoxythymidine triphosphate.
Cells 10 01095 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Višnjić, D.; Lalić, H.; Dembitz, V.; Tomić, B.; Smoljo, T. AICAr, a Widely Used AMPK Activator with Important AMPK-Independent Effects: A Systematic Review. Cells 2021, 10, 1095. https://doi.org/10.3390/cells10051095

AMA Style

Višnjić D, Lalić H, Dembitz V, Tomić B, Smoljo T. AICAr, a Widely Used AMPK Activator with Important AMPK-Independent Effects: A Systematic Review. Cells. 2021; 10(5):1095. https://doi.org/10.3390/cells10051095

Chicago/Turabian Style

Višnjić, Dora, Hrvoje Lalić, Vilma Dembitz, Barbara Tomić, and Tomislav Smoljo. 2021. "AICAr, a Widely Used AMPK Activator with Important AMPK-Independent Effects: A Systematic Review" Cells 10, no. 5: 1095. https://doi.org/10.3390/cells10051095

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

Article Metrics

Back to TopTop