Next Article in Journal
Molecular Cytological Analysis and Specific Marker Development in Wheat-Psathyrostachys huashanica Keng 3Ns Additional Line with Elongated Glume
Previous Article in Journal
The Solvation Effect of C=O Group of Cyclic Anhydrides in Solution
Previous Article in Special Issue
Repositioning the Role of Tumor Necrosis Factor-Related Apoptosis-Inducing Ligand (TRAIL) on the TRAIL to the Development of Diabetes Mellitus: An Update of Experimental and Clinical Evidence
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Prognostic, Diagnostic, and Therapeutic Potential of TRAIL Signalling in Cardiovascular Diseases

1
Heart Research Institute, The University of Sydney, Sydney 2042, Australia
2
Royal Prince Alfred Hospital, Sydney 2006, Australia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(7), 6725; https://doi.org/10.3390/ijms24076725
Submission received: 28 February 2023 / Revised: 29 March 2023 / Accepted: 29 March 2023 / Published: 4 April 2023
(This article belongs to the Special Issue The Role of TRAIL in Human Health and Metabolic Disorders)

Abstract

:
TNF-related apoptosis-inducing ligand (TRAIL) was originally discovered, almost 20 years ago, for its ability to kill cancer cells. More recent evidence has described pleiotropic functions, particularly in the cardiovascular system. There is potential for TRAIL concentrations in the circulation to act as prognostic and/or diagnostic factors for cardiovascular diseases (CVD). Pre-clinical studies also describe the therapeutic capacity for TRAIL signals, particularly in the context of atherosclerotic disease and diseases of the myocardium. Because diabetes mellitus significantly contributes to the progression and pathogenesis of CVDs, in this review we highlight recent evidence for the prognostic, diagnostic, and therapeutic potential of TRAIL signals in CVDs, and where relevant, the impact of diabetes mellitus. A greater understanding of how TRAIL signals regulate cardiovascular protection and pathology may offer new diagnostic and therapeutic avenues for patients suffering from CVDs.

1. Introduction

Cardiovascular disease (CVD) is an umbrella term for a group of disorders related to the heart and blood vessels and is the leading cause of death worldwide. Since 2019, >500 million cases of CVD were reported globally, with 18.6 million CVD-associated deaths [1]. Acute events caused by CVD include heart attack, stroke, and gangrene, which is a direct result of atherosclerosis, where blood vessels become blocked due to a fibro-fatty plaque that reduces nutrients and oxygen to the heart, brain, and lower limbs. Damage to the blood vessels can also promote heart failure, affecting the heart’s ability to pump blood around the body. Alarmingly, metabolic derangements such as diabetes mellitus and obesity significantly impact and contribute to the prevalence of CVDs. For example, the risk of ischaemic CVDs including myocardial infarction (MI) and stroke is increased by more than 50% with type-2 diabetes; diabetes and obesity increase the risk of heart failure by 112% and 65%, respectively [2,3]. Greater comprehension of CVD development, pathogenesis and early detection, and impact of metabolic diseases is critical for finding new strategies to reduce this burden.
A member of the TNF family of cytokines, TNF-related apoptosis-inducing ligand (TRAIL) was first discovered for its ability to kill human cancer cells upon ligation with its signalling receptors, namely, death receptor-4 and -5 (DR4 and DR5), without affecting normal cells [4,5]. Mice only have one death receptor, mDR5 (homologous to both human receptors), and similar to its actions in humans, TRAIL binding to mDR5 reduced tumours [6]. Because of these findings, TRAIL was hailed a potential cancer therapeutic; however, clinical trials in patients showed little survival benefit [7]. It is now clear that TRAIL signals have pleiotropic effects, like most TNF cytokines. In addition to apoptosis, TRAIL can stimulate necroptosis [8] and autophagy [9], as well as cell survival processes such as proliferation, migration, and differentiation [10,11,12,13,14]. Human decoy receptors (DcR) for TRAIL have also been identified and include DcR1, DcR2, and the soluble receptor osteoprotegrin (OPG). Upon binding TRAIL, they inhibit the induction of apoptosis [15], yet how these receptors impact non-apoptotic function(s) of TRAIL is not clear.
TRAIL signalling in CVD has gained considerable interest, and the evidence in atherosclerotic disease mostly points towards a protective role, particularly since reduced circulating TRAIL levels are associated with increased cardiovascular events and mortality [16,17]. TRAIL’s role in the heart is conflicting. In some circumstances it may protect, while in others it may contribute to pathogenesis. Furthermore, it is unclear as to whether TRAIL is a risk factor or a risk marker in CVDs [18]. This review summarises our current understanding of TRAIL signals in atherosclerotic vascular disease as well as conditions of the heart. A better understanding of how TRAIL signalling regulates cardiovascular protection and pathology may offer new diagnostic and therapeutic avenues for patients suffering from CVDs.

2. Atherosclerosis and Vessel Diseases

Atherosclerosis is the pathological process underlying coronary heart disease (e.g., coronary artery disease (CAD)), stroke, and peripheral artery disease (PAD). It is a condition where vascular smooth muscle cells (VSMCs), inflammatory cells (e.g., monocytes, macrophages), lipids, cholesterol, and cellular waste accumulate, producing a thickened plaque in the arterial wall. Because the endothelium has anti-inflammatory, anti-thrombotic, and anti-atherosclerotic properties, one of the earliest indications for atherogenesis is endothelial dysfunction, with diabetes mellitus precipitating this process [19]. As the lesion grows, the plaque can become vulnerable and rupture. Plaque rupture is the major complication and cause of CVD mortality and morbidity. Below we have summarised our current knowledge of TRAIL biology in the context of atherosclerotic disease, specifically focusing on findings implicating a prognostic/diagnostic and therapeutic role (Table 1 and Table 2; Figure 1 and Figure 2). Where relevant, the impact of diabetes mellitus is also highlighted.

2.1. Diagnostic and Prognostic Potential

(i) CAD—The occlusion of blood vessels to the heart by atherosclerotic lesions leads to reduced blood flow to the myocardium, leading to MI and potentially heart failure. Several reports showed that circulating TRAIL levels were lower in patients diagnosed with CAD vs. those undiagnosed [20,21,22,23], associating with oxidative stress [67] and disease severity [16,17]. Interestingly, monocyte Trail mRNA was reduced in CAD, concomitant with circulating levels from the same patients, implicating monocytes as a significant source of TRAIL in healthy circulation and compromised in CAD [23]. The Canakinumab Anti-inflammatory Thrombosis Outcomes Study identified IL-18 and IL-6 as associated with residual inflammatory risk following IL-1β inhibition in patients [89]. A negative association between IL-18 and TRAIL, but not IL-6, was identified with CAD [23], and colchicine anti-inflammatory treatment increased plasma TRAIL levels in these patients [61], supporting TRAIL’s anti-inflammatory role. Because diabetes mellitus is a common comorbidity of atherosclerosis, lower levels of circulating TRAIL were also reported in type-2 diabetes [90,91]; however, TRAIL levels did not reflect early stages of atherosclerosis (as a measure of carotid artery intima-media thickness) in these patients [92].
The prognostic potential of TRAIL has been reported. In the InCHIANTI study, a population-based study of aging, almost 1300 individuals >65 years of age were selected randomly, and blood was sampled at baseline and at 3 and 6 years later. A strong and independent association between low TRAIL levels and all-cause mortality over a period of 6 years was found in participants with pre-existing CVDs (including and not limited to heart failure, MI, stroke, and PAD) [16]. In fact, the authors reported that participants with TRAIL levels <84.5 pg/mL at baseline had 2–3-fold greater risk of cardiovascular mortality [16]. A similar finding was observed in a second smaller prospective study; TRAIL serum levels were a strong predicter of death [17]. These findings suggest that measuring TRAIL levels in patients with CVD has prognostic value and could be considered as a measure of cardiovascular risk.
TRAIL receptors have also been linked to CAD. For example, a recent biomarker study retrospectively examined two clinical trials and identified DR5 as one of 18 proteins increased with CAD [24], suggesting that it may act as a potential diagnostic marker. DR5 may also have prognostic value since higher levels are associated with all-cause mortality in acute MI patients, along with growth development factor-15, and in combination with established risk factors predicted survival with 88% accuracy [25]. Increased DR5 levels may also have prognostic potential in chronic kidney disease. Indeed, DR5 was identified as a predictor of cardiovascular mortality, as well as a predictor of MI and heart failure readmission [26,27], a finding also observed in diabetes [28].
Several studies suggest that increased concentrations of OPG are associated with increased cardiovascular risk. For example, the Tromsø Study, a large population-based cohort study of 6265 participants, found high levels associated with increased risk of MI, stroke, total mortality and mortality of ischaemic heart disease, stroke, and non-vascular conditions [29]. These were independent of cardiovascular risk factors [29]. OPG levels may be independently associated with traditional cardiovascular risk factors, including diabetes [30,31]. High OPG levels were also reflective of disease severity and predicted cardiovascular events and all-cause mortality in CAD patients [30,33]; however, how this impacts TRAIL signalling is unclear. The OPG/TRAIL ratio may have prognostic potential; higher OPG/TRAIL ratios were observed in CAD [21] and predicted all-cause cardiovascular mortality in patients with renal failure [32]. In contrast, a single OPG measurement was deemed insufficient to diagnose CAD in patients with angina [33].
(ii) Stroke—Similar to coronary heart disease, studies report lower levels of circulating TRAIL in patients that had a stroke vs. healthy individuals [34,35,36,37,38,39], and in some cases (but not all) it was associated with stroke severity [34,35,37,39]. The two stroke subtypes involving active atherosclerosis are large artery atherosclerosis (LAA), often in the carotid artery, and small artery occlusion (SAO). Conflicting evidence exists as to whether TRAIL levels differ significantly between subtypes, some reporting no change [34,36]; however, one study identified that circulating TRAIL levels were lower in patients with SAO [38]; low levels persisted for at least three months after the onset of stroke. Peripheral blood mononuclear cells (PBMCs) were also assessed for TRAIL expression from patients with ischaemic stroke, and while serum TRAIL levels were reduced, this did not reflect PBMC mRNA expression, which was increased at the time of admission [36]. The impact of PBMC-derived TRAIL vs. serum TRAIL in stroke is still unclear. DR5 and OPG levels may also be relevant. For example, DR5 is increased in the circulation of symptomatic patients with carotid plaque, associated with increased DR5 plaque protein when compared to asymptomatic patients [40], suggesting that levels may reflect the severity of disease. DR5 and OPG levels were also associated with stroke, and both were elevated in LAA patients compared to controls [35]. Furthermore, increased OPG levels (assessed at the time of admission) predicted poorer prognosis and mortality of patients who suffered an ischaemic stroke [41,42].
(iii) PAD—There is limited and conflicting data as to whether TRAIL could act as a prognostic marker factor for PAD. Serum TRAIL levels were reduced in type-2 diabetic nephropathy patients with foot ulcers vs. those without [43], a finding also observed in diabetic PAD patients alone [44]. In contrast, O’Sullivan et al. reported higher levels of TRAIL in PAD [45]. DR4 and DR5 measurements from PAD patients have not been described; however, increased levels of circulating OPG were evident in PAD and type-2 diabetic PAD patients [44,45]. Further, the OPG/TRAIL ratio was described to be higher in both cohorts [44]. More work is needed to understand the contribution of TRAIL and its receptors to PAD.

2.2. Therapeutic Potential for Atherosclerotic Disease—Teachings from In Vitro and Pre-Clinical Studies

In vitro cell studies and pre-clinical animal models have shown that TRAIL plays important role(s) in the development of atherosclerosis, either protecting or contributing to pathogenesis (Figure 2). In response to peri-vascular cuff injury, Trail−/− mice had reduced neointimal hyperplasia compared to Trail+/+ mice, and recombinant TRAIL delivery recovered neointimal thickening [10], a finding supported by in vitro studies using human VSMCs [10,14,81]. These findings suggest that TRAIL may contribute to the development of early atherosclerosis. Because TRAIL promotes VSMC migration into the plaque, this process may also contribute to plaque stability in advanced lesions, reducing the incidence of rupture [70]. Indeed, atherosclerotic Trail−/−Apoe−/− mice developed a larger [23,65,69], more macrophage-rich plaque of unstable phenotype with reduced VSMC and collagen content [23,65,66,69]. Mice lacking TRAIL had greater vascular oxidative stress [67], inflammation [23,65,68], and endothelial dysfunction [67] compared to the control. Importantly, metabolic derangements were observed; Trail−/−Apoe−/− mice developed features of type-2 diabetes [65,68]. Trail−/− mice or neutralisation of TRAIL in mice resulted in increased susceptibility to streptozotocin-induced diabetes or high fat diet-induced insulin resistance [63,64]. TRAIL’s expression in the vessel wall is controlled by insulin. We showed that chronic exposure of human VSMCs to insulin suppressed TRAIL gene expression, promoting apoptosis [80]. TRAIL expression was also downregulated in vessels of diabetic rats [93]. Furthermore, microarray analysis identified Glut1, a glucose transporter, as a pathological gene upregulated in aortic tissues of Trail−/− mice [61]. These pre-clinical findings support TRAIL’s involvement in metabolic CVDs and provide insight into the impact of TRAIL suppression in people. Given that TRAIL can regulate the vascular system in diabetes [94], a greater understanding of TRAIL signalling in the progression of diabetic CVD is needed.
To understand why TRAIL was suppressed in CAD, we found that elevated levels of IL-18 repressed TRAIL transcription and gene expression in healthy human monocytes by inhibiting NFκB’s ability to bind the TRAIL promoter [23]. Indeed, macrophages lacking TRAIL were more inflammatory, less effective in their ability to efferocytose, showed impaired cholesterol handling, and had reduced migratory ability [23], which are hallmarks of dysfunctional macrophages in lesions, accelerating atherosclerosis [23,95]. In contrast, exogenous TRAIL pre-treatment increased lipid uptake and foam cell formation and contributed to macrophage apoptosis [82]. The exogenous delivery of TRAIL in pre-clinical models of atherosclerosis has been described and for the most part shows promising therapeutic potential. Administration of TRAIL protein, TRAIL gene therapy, or TRAIL bone marrow transplantation attenuated atherosclerosis development, reduced macrophage content in the vessel wall, and reduced inflammation in diabetic Apoe−/− or Trail−/−Apoe−/− mice [23,70].
As described earlier, the vascular endothelium is critical for the maintenance of cardiovascular homeostasis. Although reports indicate that TRAIL can stimulate apoptosis of endothelial cells (ECs) [86], the majority of findings reports on increased EC survival processes, particularly at physiological concentrations. For example, diabetes-induced endothelial dysfunction was improved by TRAIL, in part via its ability to increase endothelial nitric oxide synthase (eNOS) production [73]. TRAIL also prevented high glucose-induced apoptosis of ECs and protected against angiotensin II (AngII)- or TNF-α-induced oxidative stress, which are pro-atherogenic conditions [67,73,83]. Furthermore, TRAIL reduced AngII-induced endothelial reactivity and monocyte adhesion, improving endothelial integrity [67]. These findings further support TRAIL’s therapeutic potential in the vasculature.
In addition to their multiple functions in maintaining vascular homeostasis is the regenerative capacity of ECs and their ability to develop new blood vessels by angiogenesis, an essential process that is upregulated during ischemia to increase blood perfusion. However, in CVDs, endogenous angiogenic processes are impaired, contributing to vascular insufficiency. In vitro, TRAIL stimulated EC proliferation, migration, and differentiation [13,62,84], processes important for the formation of vascular tubules. We used the hindlimb ischemia model of PAD; Trail−/− mice had impaired limb movement and increased limb necrosis associated with markedly reduced (~70%) capillary density in limb tissues [62]. Viral TRAIL gene therapy dramatically improved limb blood perfusion and vascularisation mediated by NOX4-inducible eNOS phosphorylation and generation of nitric oxide [62], a key factor regulating vessel patency. This effect may be a consequence of TRAIL interacting with its receptor mDR5 [62], although the exact contribution of TRAIL receptor(s) to ischemia-induced angiogenesis is unknown. In contrast, TRAIL stimulated apoptosis in the human brain endothelial cell line, hCMEC/d3 [85], suggesting organ- and cell-specific effects. Indeed, amyloid-beta (Aβ), associated with neurodegeneration and known to accumulate after stroke or in cerebral ischemia [96,97], interacted with DR4 and DR5, triggering the activation of caspase-8 and mitochondrial pathways for apoptosis in these cells [98].
The impact of TRAIL receptors in atherosclerosis in pre-clinical studies is less clear. Brachiocephalic arteries of Apoe−/− mice express OPG [71]; its expression was associated with lesions that are unstable [99]. However, Opg−/−Apoe−/− mice had increased lesion area, with a 40% reduction in plaque cellularity compared to Apoe−/− mice [72]. The exogenous treatment of VSMCs promoted survival in vitro, supporting this finding. Furthermore, the increased lesion size was a result of calcification and extracellular matrix deposition [72]. The authors suggested that reduced MMP-9 activity could contribute to increased matrix deposition in Opg−/−Apoe−/− plaque [72]. These findings imply that OPG has a complex role in atherogenesis, and more work is needed to understand its contribution and whether it can be targeted for its therapeutic potential.

3. TRAIL Signalling in the Myocardium

Injury to the myocardium due to multiple factors, e.g., ischemia, atherosclerosis, infections, etc., can result in heart failure, a progressive disease that impacts the heart’s ability to adequately pump blood around the body, manifesting impaired cardiac function, disturbed electrical activity, and abnormal tissue architecture. Diabetes mellitus increases the risk of heart failure and may also contribute to the progression of cardiomyopathy and atrial fibrillation [100,101], two related pathologies. Below is a summary of the current findings examining the diagnostic/biomarker potential of TRAIL signals in the myocardium (Table 1 and Figure 3). We also describe in vitro and pre-clinical findings that show the therapeutic potential of TRAIL signals (Table 2 and Figure 4), and where possible, we describe these in the context of metabolic disease.

3.1. Diagnostic and Prognostic Potential

(i) Heart Failure—Some studies have linked circulating levels of TRAIL or its receptors to heart failure. For example, a strong inverse association of all-cause mortality was observed in advanced heart failure with TRAIL, whereas higher levels of TRAIL reflected better prognoses [46]. In support, another study identified TRAIL as one of five multi-biomarkers that could predict patient mortality [47]. Furthermore, a negative association between TRAIL levels and all-cause mortality and hospitalisation was identified in patients with preserved ejection fraction, whereas a positive association with circulating DR5 was found [48]. In contrast, a prospective observational study showed no difference in TRAIL levels in heart failure patients undergoing cardiac resynchronisation therapy, and TRAIL levels did not predict mortality [49]. Circulating DR5 is increased in heart failure patients with worse left ventricular ejection fraction and diastolic function but positively associating with the incidence of disease [48,50]. The decoy receptor OPG has also been linked to chronic heart failure with increased levels observed in patients [51], associated with adverse outcomes within 2 years [52]. OPG levels were also shown to be a significant predictor of mortality [53]. These findings suggest that TRAIL and TRAIL receptors may act as a potential biomarker in heart failure as well as predict patient outcomes and mortality; however, more studies are needed to confirm these.
(ii) Cardiomyopathies—Non-ischaemic dilated cardiomyopathy is a condition that causes hypertrophy of the ventricles, effecting myocardial contractility and reducing the ejection fraction, leading to heart failure if left untreated. Plasma TRAIL levels were upregulated in patients with non-ischaemic dilated cardiomyopathy and positively correlated with left ventricular end-diastolic diameter, whereas OPG levels remained unchanged [54]. TRAIL levels were also altered in patients with Chagas cardiomyopathy, an inflammatory disease caused by the protozoan Trypanosoma cruzi, which can progress to dilated cardiomyopathy and heart failure. Patients suffering from severe Chagus had elevated levels of circulating TRAIL, correlated with left ventricular ejection fraction and left ventricular diastolic diameter [55]. Whether TRAIL and its receptors act as diagnostic or prognostic factors in cardiomyopathy is unclear.
(iii) Atrial fibrillation (AF) is a common atrial arrhythmia that can be paroxysmal (~1 week), persistent (>1 week), or permanent and can increase the risk of stroke and heart failure. A prospective observational study identified circulating TRAIL levels to be decreased in patients with successful ablation of AF [56]. Conversely, low levels of circulating TRAIL were evident in acute onset AF, and they were increased following sinus rhythm maintenance [58]. Another study found no differences in plasma TRAIL levels observed in patients with or without AF recurrence; however, when the transcardiac gradient was measured, TRAIL levels were reduced, revealing this gradient to be a negative predictor for AF recurrence [57]. Like TRAIL, DR5 levels are reduced in AF; however, no links have been described between AF and sinus rhythm [59]. As for decoy receptors, very little information exists. There is some evidence linking OPG to AF; OPG expression was increased in samples of the right atrial appendage from persistent and paroxysmal AF patients vs. normal controls and sinus rhythm patients [60]. Whether TRAIL and its receptors can act as prognostic factors or biomarkers requires further elucidation.

3.2. Therapeutic Potential for Diseases of the Heart—Teachings from In Vitro and Pre-Clinical Studies

TRAIL and its receptors are expressed in normal and diseased human and rodent hearts at varying levels [51,74,102,103], although the impact of TRAIL signals in the heart is not fully elucidated. Apoptosis and proliferation play key roles under normal and pathogenic conditions, but it is not clear if TRAIL play a protective or detrimental role here. Administration of TRAIL or a small molecule DR5 agonist (bioymifi) to cardiomyocytes in vitro did not induce apoptosis or affect cell viability, but it altered cardiomyocyte structure, promoting hypertrophy in an ERK1/2-dependent manner [75]. Similarly, the administration of recombinant TRAIL or adenoviral TRAIL resulted in a significant reduction in cardiac fibrosis and apoptosis compared to control diabetic animals [76], and MD5-1 (agonistic mDR5 mAb) treatment to wildtype mice resulted in increased heart weight and cardiomyocyte area, in part through the activation of the epidermal growth factor receptor [75]. Increased ventricular fractional shortening was also observed with DR5 activation [75]. OPG may also associate with cardiomyocyte hypertrophy, since hearts from Apoe−/− mice had increased cardiomyocyte diameter associated with increased OPG protein [77]. Further, OPG delivery to spontaneously hypertensive rats resulted in enlarged cardiomyocytes and fibroblasts, with OPG regulating cardiac and fibrosis-related proteins [78]. These findings suggest that the activation of TRAIL signals via its signalling receptors may regulate structural changes in the heart under physiological conditions and in conditions of diabetic cardiomyopathy. Whether the actions of OPG are TRAIL-dependent or -independent is unclear.
Other studies report opposing actions of TRAIL in the heart. For example, TRAIL stimulated, whereas neutralising DR5 inhibited, the stretch-induced apoptosis of cardiomyocytes [87]. The increased expression of DR4/DR5 was also observed in doxorubicin-treated human cardiomyocytes, associated with spontaneous apoptosis [88]. Apoptotic cell death is increased in heart failure and may contribute to unfavourable left ventricular remodelling [104]. Blocking DR5 using a soluble immunoglobulin fusion protein (sDR5-fc) in a heart failure model that prevents cardiac cell death and inflammation, preserves ejection fraction and fractional shortening, reduces fibrosis, and prevents ventricular wall thinning, findings observed in rodents, pigs, and monkeys [79]. Silencing DR5 in an MI model in rats also reduced myocardial damage and infarct size, and it reduced the cardiac expression of apoptotic mediators [74]. These imply that under certain conditions, the activation of TRAIL signals in the heart could be detrimental, and a blockade of TRAIL signalling may be used as a potential therapeutic. More research is needed to fully comprehend the diverse roles of TRAIL and its receptors in cardiac function under normal and pathological conditions.

4. Conclusions

Targeting the TRAIL pathway in CVDs holds great prognostic and diagnostic potential. TRAIL concentrations are suppressed, whereas TRAIL receptor levels are increased in people with CVDs, which are associated with cardiovascular risk. Pre-clinical models have also identified that TRAIL signals also play a role in disease protection or progression, offering new therapeutic possibilities for the treatment of CVDs. Promoting TRAIL-receptor activation in atherosclerotic disease could be beneficial; drugs already in use to activate TRAIL signals in clinical trials in cancer could be repurposed or modified for atherosclerosis. On the other hand, novel therapeutics aimed at blocking TRAIL signalling in the myocardium could improve heart failure. More research is needed to fully comprehend the role of TRAIL and its receptors in atherosclerotic vessel diseases and the myocardium.

Author Contributions

Conceptualisation, E.K., M.S.P., S.P., S.P.C. and M.M.K.; writing draft and figures, E.K., M.S.P., S.P.C. and M.M.K.; final review and editing, E.K., M.S.P., S.P., S.P.C. and M.M.K. All authors have read and agreed to the published version of the manuscript.

Funding

E.K. is supported by the Heart Research Institute New Zealand Pathways Program; M.S.P. is supported by an Australian Government Research Program Training Scholarship; M.M.K. and S.P.C. are supported by grants from the Australian National Health and Medical Research Council (APP1188218) and the Heart Research Institute.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Roth, G.A.; Mensah, G.A.; Johnson, C.O.; Addolorato, G.; Ammirati, E.; Baddour, L.M.; Barengo, N.C.; Beaton, A.Z.; Benjamin, E.J.; Benziger, C.P.; et al. Global Burden of Cardiovascular Diseases and Risk Factors, 1990–2019: Update From the GBD 2019 Study. J. Am. Coll. Cardiol. 2020, 76, 2982–3021. [Google Scholar] [CrossRef] [PubMed]
  2. Einarson, T.R.; Acs, A.; Ludwig, C.; Panton, U.H. Prevalence of cardiovascular disease in type 2 diabetes: A systematic literature review of scientific evidence from across the world in 2007–2017. Cardiovasc. Diabetol. 2018, 17, 83. [Google Scholar] [CrossRef] [Green Version]
  3. Powell-Wiley, T.M.; Poirier, P.; Burke, L.E.; Després, J.-P.; Gordon-Larsen, P.; Lavie, C.J.; Lear, S.A.; Ndumele, C.E.; Neeland, I.J.; Sanders, P.; et al. Obesity and Cardiovascular Disease: A Scientific Statement From the American Heart Association. Circulation 2021, 143, e984–e1010. [Google Scholar] [CrossRef]
  4. Pitti, R.M.; Marsters, S.A.; Ruppert, S.; Donahue, C.J.; Moore, A.; Ashkenazi, A. Induction of apoptosis by Apo-2 ligand, a new member of the tumor necrosis factor cytokine family. J. Biol. Chem. 1996, 271, 12687–12690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Wiley, S.R.; Schooley, K.; Smolak, P.J.; Din, W.S.; Huang, C.P.; Nicholl, J.K.; Sutherland, G.R.; Smith, T.D.; Rauch, C.; Smith, C.A.; et al. Identification and characterization of a new member of the TNF family that induces apoptosis. Immunity 1995, 3, 673–682. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Grosse-Wilde, A.; Kemp, C.J. Metastasis suppressor function of tumor necrosis factor-related apoptosis-inducing ligand-R in mice: Implications for TRAIL-based therapy in humans? Cancer Res. 2008, 68, 6035–6037. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Snajdauf, M.; Havlova, K.; Vachtenheim, J.; Ozaniak, A.; Lischke, R.; Bartunkova, J.; Smrz, D.; Strizova, Z. The TRAIL in the Treatment of Human Cancer: An Update on Clinical Trials. Front. Mol. Biosci. 2021, 8, 628332. [Google Scholar] [CrossRef]
  8. Jouan-Lanhouet, S.; Arshad, M.I.; Piquet-Pellorce, C.; Martin-Chouly, C.; Le Moigne-Muller, G.; Van Herreweghe, F.; Takahashi, N.; Sergent, O.; Lagadic-Gossmann, D.; Vandenabeele, P.; et al. TRAIL induces necroptosis involving RIPK1/RIPK3-dependent PARP-1 activation. Cell Death Differ. 2012, 19, 2003–2014. [Google Scholar] [CrossRef] [Green Version]
  9. Chen, Y.; Zhou, X.; Qiao, J.; Bao, A. Autophagy is a regulator of TRAIL-induced apoptosis in NSCLC A549 cells. J. Cell Commun. Signal. 2017, 11, 219–226. [Google Scholar] [CrossRef] [Green Version]
  10. Chan, J.; Prado-Lourenco, L.; Khachigian, L.M.; Bennett, M.R.; Bartolo, B.A.D.; Kavurma, M.M. TRAIL Promotes VSMC Proliferation and Neointima Formation in a FGF-2–, Sp1 Phosphorylation–, and NFκB-Dependent Manner. Circ. Res. 2010, 106, 1061–1071. [Google Scholar] [CrossRef] [Green Version]
  11. Secchiero, P.; Gonelli, A.; Carnevale, E.; Milani, D.; Pandolfi, A.; Zella, D.; Zauli, G. TRAIL Promotes the Survival and Proliferation of Primary Human Vascular Endothelial Cells by Activating the Akt and ERK Pathways. Circulation 2003, 107, 2250–2256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Kahraman, S.; Yilmaz, O.; Altunbas, H.A.; Dirice, E.; Sanlioglu, A.D. TRAIL induces proliferation in rodent pancreatic beta cells via AKT activation. J. Mol. Endocrinol. 2021, 66, 325–338. [Google Scholar] [CrossRef] [PubMed]
  13. Cartland, S.P.; Genner, S.W.; Zahoor, A.; Kavurma, M.M. Comparative Evaluation of TRAIL, FGF-2 and VEGF-A-Induced Angiogenesis In Vitro and In Vivo. Int. J. Mol. Sci. 2016, 17, 2025. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Kavurma, M.M.; Schoppet, M.; Bobryshev, Y.V.; Khachigian, L.M.; Bennett, M.R. TRAIL Stimulates Proliferation of Vascular Smooth Muscle Cells via Activation of NF-κB and Induction of Insulin-like Growth Factor-1 Receptor. J. Biol. Chem. 2008, 283, 7754–7762. [Google Scholar] [CrossRef] [Green Version]
  15. Pan, G.; Ni, J.; Wei, Y.F.; Yu, G.; Gentz, R.; Dixit, V.M. An antagonist decoy receptor and a death domain-containing receptor for TRAIL. Science 1997, 277, 815–818. [Google Scholar] [CrossRef]
  16. Volpato, S.; Ferrucci, L.; Secchiero, P.; Corallini, F.; Zuliani, G.; Fellin, R.; Guralnik, J.M.; Bandinelli, S.; Zauli, G. Association of tumor necrosis factor-related apoptosis-inducing ligand with total and cardiovascular mortality in older adults. Atherosclerosis 2010, 215, 452–458. [Google Scholar] [CrossRef] [Green Version]
  17. Osmancik, P.; Teringova, E.; Tousek, P.; Paulu, P.; Widimsky, P. Prognostic value of TNF-related apoptosis inducing ligand (TRAIL) in acute coronary syndrome patients. PLoS ONE 2013, 8, e53860. [Google Scholar] [CrossRef] [Green Version]
  18. Kakareko, K.; Rydzewska-Rosołowska, A.; Zbroch, E.; Hryszko, T. TRAIL and Cardiovascular Disease—A Risk Factor or Risk Marker: A Systematic Review. J. Clin. Med. 2021, 10, 1252. [Google Scholar] [CrossRef]
  19. Hadi, H.A.; Suwaidi, J.A. Endothelial dysfunction in diabetes mellitus. Vasc. Health Risk Manag. 2007, 3, 853–876. [Google Scholar]
  20. Mori, K.; Ikari, Y.; Jono, S.; Shioi, A.; Ishimura, E.; Emoto, M.; Inaba, M.; Hara, K.; Nishizawa, Y. Association of serum TRAIL level with coronary artery disease. Thromb. Res. 2009, 125, 322–325. [Google Scholar] [CrossRef]
  21. Shaker, O.G.; El-Shehaby, A.; Nabih, M. Possible Role of Osteoprotegerin and Tumor Necrosis Factor-Related Apoptosis-Inducing Ligand as Markers of Plaque Instability in Coronary Artery Disease. Angiology 2010, 61, 756–762. [Google Scholar] [CrossRef] [PubMed]
  22. Ajala, O.; Zhang, Y.; Gupta, A.; Bon, J.; Sciurba, F.; Chandra, D. Decreased serum TRAIL is associated with increased mortality in smokers with comorbid emphysema and coronary artery disease. Respir. Med. 2018, 145, 21–27. [Google Scholar] [CrossRef]
  23. Cartland, S.P.; Genner, S.W.; Martínez, G.J.; Robertson, S.; Kockx, M.; Lin, R.C.Y.; O’Sullivan, J.F.; Koay, Y.C.; Manuneedhi Cholan, P.; Kebede, M.A.; et al. TRAIL-Expressing Monocyte/Macrophages Are Critical for Reducing Inflammation and Atherosclerosis. iScience 2019, 12, 41–52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Wallentin, L.; Eriksson, N.; Olszowka, M.; Grammer, T.B.; Hagström, E.; Held, C.; Kleber, M.E.; Koenig, W.; März, W.; Stewart, R.A.H.; et al. Plasma proteins associated with cardiovascular death in patients with chronic coronary heart disease: A retrospective study. PLoS Med. 2021, 18, e1003513. [Google Scholar] [CrossRef]
  25. Skau, E.; Henriksen, E.; Wagner, P.; Hedberg, P.; Siegbahn, A.; Leppert, J. GDF-15 and TRAIL-R2 are powerful predictors of long-term mortality in patients with acute myocardial infarction. Eur. J. Prev. Cardiol. 2017, 24, 1576–1583. [Google Scholar] [CrossRef]
  26. Feldreich, T.; Nowak, C.; Fall, T.; Carlsson, A.C.; Carrero, J.-J.; Ripsweden, J.; Qureshi, A.R.; Heimbürger, O.; Barany, P.; Stenvinkel, P.; et al. Circulating proteins as predictors of cardiovascular mortality in end-stage renal disease. J. Nephrol. 2019, 32, 111–119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Edfors, R.; Lindhagen, L.; Spaak, J.; Evans, M.; Andell, P.; Baron, T.; Mörtberg, J.; Rezeli, M.; Salzinger, B.; Lundman, P.; et al. Use of proteomics to identify biomarkers associated with chronic kidney disease and long-term outcomes in patients with myocardial infarction. J. Intern. Med. 2020, 288, 581–592. [Google Scholar] [CrossRef]
  28. Ferreira, J.P.; Sharma, A.; Mehta, C.; Bakris, G.; Rossignol, P.; White, W.B.; Zannad, F. Multi-proteomic approach to predict specific cardiovascular events in patients with diabetes and myocardial infarction: Findings from the EXAMINE trial. Clin. Res. Cardiol. 2021, 110, 1006–1019. [Google Scholar] [CrossRef]
  29. Vik, A.; Mathiesen, E.B.; Brox, J.; Wilsgaard, T.; Njolstad, I.; Jorgensen, L.; Hansen, J.B. Serum osteoprotegerin is a predictor for incident cardiovascular disease and mortality in a general population: The Tromso Study. J. Thromb. Haemost. 2011, 9, 638–644. [Google Scholar] [CrossRef]
  30. Bjerre, M.; Hilden, J.; Winkel, P.; Jensen, G.B.; Kjøller, E.; Sajadieh, A.; Kastrup, J.; Kolmos, H.J.; Larsson, A.; Ärnlöv, J.; et al. Serum osteoprotegerin as a long-term predictor for patients with stable coronary artery disease and its association with diabetes and statin treatment: A CLARICOR trial 10-year follow-up substudy. Atherosclerosis 2020, 301, 8–14. [Google Scholar] [CrossRef]
  31. Mogelvang, R.; Pedersen, S.H.; Flyvbjerg, A.; Bjerre, M.; Iversen, A.Z.; Galatius, S.; Frystyk, J.; Jensen, J.S. Comparison of osteoprotegerin to traditional atherosclerotic risk factors and high-sensitivity C-reactive protein for diagnosis of atherosclerosis. Am. J. Cardiol. 2012, 109, 515–520. [Google Scholar] [CrossRef]
  32. Kuzniewski, M.; Fedak, D.; Dumnicka, P.; Stepien, E.; Kusnierz-Cabala, B.; Cwynar, M.; Sulowicz, W. Osteoprotegerin and osteoprotegerin/TRAIL ratio are associated with cardiovascular dysfunction and mortality among patients with renal failure. Adv. Med. Sci. 2016, 61, 269–275. [Google Scholar] [CrossRef]
  33. Hosbond, S.E.; Diederichsen, A.C.P.; Saaby, L.; Rasmussen, L.M.; Lambrechtsen, J.; Munkholm, H.; Sand, N.P.R.; Gerke, O.; Poulsen, T.S.; Mickley, H. Can osteoprotegerin be used to identify the presence and severity of coronary artery disease in different clinical settings? Atherosclerosis 2014, 236, 230–236. [Google Scholar] [CrossRef]
  34. Kang, Y.H.; Park, M.-G.; Noh, K.-H.; Park, H.R.; Lee, H.W.; Son, S.M.; Park, K.-P. Low serum TNF-related apoptosis-inducing ligand (TRAIL) levels are associated with acute ischemic stroke severity. Atherosclerosis 2015, 240, 228–233. [Google Scholar] [CrossRef]
  35. Pan, X.; Pang, M.; Ma, A.; Wang, K.; Zhang, Z.; Zhong, Q.; Yang, S. Association of TRAIL and Its Receptors with Large-Artery Atherosclerotic Stroke. PLoS ONE 2015, 10, e0136414. [Google Scholar] [CrossRef] [Green Version]
  36. Tufekci, K.U.; Vurgun, U.; Yigitaslan, O.; Keskinoglu, P.; Yaka, E.; Kutluk, K.; Genc, S. Follow-up Analysis of Serum TNF-Related Apoptosis-Inducing Ligand Protein and mRNA Expression in Peripheral Blood Mononuclear Cells from Patients with Ischemic Stroke. Front. Neurol. 2018, 9, 102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Altintas Kadirhan, O.; Kucukdagli, O.T.; Gulen, B. The effectiveness of serum S100B, TRAIL, and adropin levels in predicting clinical outcome, final infarct core, and stroke subtypes of acute ischemic stroke patients. Biomédica 2022, 42, 55–63. [Google Scholar] [CrossRef]
  38. Stanne, T.M.; Angerfors, A.; Andersson, B.; Brännmark, C.; Holmegaard, L.; Jern, C. Longitudinal Study Reveals Long-Term Proinflammatory Proteomic Signature After Ischemic Stroke Across Subtypes. Stroke 2022, 53, 2847–2858. [Google Scholar] [CrossRef] [PubMed]
  39. Mihalovic, M.; Mikulenka, P.; Línková, H.; Neuberg, M.; Štětkářová, I.; Peisker, T.; Lauer, D.; Tousek, P. Tumor Necrosis Factor-Related Apoptosis-Inducing Ligand (TRAIL) in Patients after Acute Stroke: Relation to Stroke Severity, Myocardial Injury, and Impact on Prognosis. J. Clin. Med. 2022, 11, 2552. [Google Scholar] [CrossRef] [PubMed]
  40. Gonçalves, I.; Singh, P.; Tengryd, C.; Cavalera, M.; Yao Mattisson, I.; Nitulescu, M.; Flor Persson, A.; Volkov, P.; Engström, G.; Orho-Melander, M.; et al. sTRAIL-R2 (Soluble TNF [Tumor Necrosis Factor]-Related Apoptosis-Inducing Ligand Receptor 2) a Marker of Plaque Cell Apoptosis and Cardiovascular Events. Stroke 2019, 50, 1989–1996. [Google Scholar] [CrossRef]
  41. Wajda, J.; Świat, M.; Owczarek, A.J.; Holecki, M.; Duława, J.; Brzozowska, A.; Olszanecka-Glinianowicz, M.; Chudek, J. Osteoprotegerin Assessment Improves Prediction of Mortality in Stroke Patients. J. Stroke Cerebrovasc. Dis. 2019, 28, 1160–1167. [Google Scholar] [CrossRef] [PubMed]
  42. Jensen, J.K.; Ueland, T.; Atar, D.; Gullestad, L.; Mickley, H.; Aukrust, P.; Januzzi, J.L. Osteoprotegerin concentrations and prognosis in acute ischaemic stroke. J. Intern. Med. 2010, 267, 410–417. [Google Scholar] [CrossRef]
  43. Arık, H.O.; Yalcin, A.D.; Gumuslu, S.; Genç, G.E.; Turan, A.; Sanlioglu, A.D. Association of circulating sTRAIL and high-sensitivity CRP with type 2 diabetic nephropathy and foot ulcers. Med. Sci. Monit. 2013, 19, 712–715. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Moon, A.R.; Park, Y.; Chang, J.H.; Lee, S.S. Inverse regulation of serum osteoprotegerin and tumor necrosis factor-related apoptosis-inducing ligand levels in patients with leg lesional vascular calcification: An observational study. Medicine 2019, 98, e14489. [Google Scholar] [CrossRef]
  45. O’Sullivan, E.P.; Ashley, D.T.; Davenport, C.; Kelly, J.; Devlin, N.; Crowley, R.; Leahy, A.L.; Kelly, C.J.; Agha, A.; Thompson, C.J.; et al. Osteoprotegerin is higher in peripheral arterial disease regardless of glycaemic status. Thromb. Res. 2010, 126, e423–e427. [Google Scholar] [CrossRef]
  46. Niessner, A.; Hohensinner, P.J.; Rychli, K.; Neuhold, S.; Zorn, G.; Richter, B.; Hulsmann, M.; Berger, R.; Mortl, D.; Huber, K.; et al. Prognostic value of apoptosis markers in advanced heart failure patients. Eur. Heart J. 2009, 30, 789–796. [Google Scholar] [CrossRef] [Green Version]
  47. Richter, B.; Koller, L.; Hohensinner, P.J.; Zorn, G.; Brekalo, M.; Berger, R.; Mörtl, D.; Maurer, G.; Pacher, R.; Huber, K.; et al. A multi-biomarker risk score improves prediction of long-term mortality in patients with advanced heart failure. Int. J. Cardiol. 2013, 168, 1251–1257. [Google Scholar] [CrossRef]
  48. Hage, C.; Michaelsson, E.; Linde, C.; Donal, E.; Daubert, J.C.; Gan, L.M.; Lund, L.H. Inflammatory Biomarkers Predict Heart Failure Severity and Prognosis in Patients With Heart Failure With Preserved Ejection Fraction: A Holistic Proteomic Approach. Circ. Cardiovasc. Genet. 2017, 10, e001633. [Google Scholar] [CrossRef] [Green Version]
  49. Osmancik, P.; Herman, D.; Stros, P.; Linkova, H.; Vondrak, K.; Paskova, E. Changes and prognostic impact of apoptotic and inflammatory cytokines in patients treated with cardiac resynchronization therapy. Cardiology 2013, 124, 190–198. [Google Scholar] [CrossRef] [PubMed]
  50. Stenemo, M.; Nowak, C.; Byberg, L.; Sundström, J.; Giedraitis, V.; Lind, L.; Ingelsson, E.; Fall, T.; Ärnlöv, J. Circulating proteins as predictors of incident heart failure in the elderly: Circulating proteins as predictors of incident heart failure. Eur. J. Heart Fail. 2018, 20, 55–62. [Google Scholar] [CrossRef]
  51. Ueland, T.; Yndestad, A.; Oie, E.; Florholmen, G.; Halvorsen, B.; Froland, S.S.; Simonsen, S.; Christensen, G.; Gullestad, L.; Aukrust, P. Dysregulated osteoprotegerin/RANK ligand/RANK axis in clinical and experimental heart failure. Circulation 2005, 111, 2461–2468. [Google Scholar] [CrossRef] [Green Version]
  52. Ueland, T.; Jemtland, R.; Godang, K.; Kjekshus, J.; Hognestad, A.; Omland, T.; Squire, I.B.; Gullestad, L.; Bollerslev, J.; Dickstein, K.; et al. Prognostic value of osteoprotegerin in heart failure after acute myocardial infarction. J. Am. Coll. Cardiol. 2004, 44, 1970–1976. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Roysland, R.; Masson, S.; Omland, T.; Milani, V.; Bjerre, M.; Flyvbjerg, A.; Di Tano, G.; Misuraca, G.; Maggioni, A.P.; Tognoni, G.; et al. Prognostic value of osteoprotegerin in chronic heart failure: The GISSI-HF trial. Am. Heart J. 2010, 160, 286–293. [Google Scholar] [CrossRef]
  54. Schoppet, M.; Ruppert, V.; Hofbauer, L.C.; Henser, S.; Al-Fakhri, N.; Christ, M.; Pankuweit, S.; Maisch, B. TNF-related apoptosis-inducing ligand and its decoy receptor osteoprotegerin in nonischemic dilated cardiomyopathy. Biochem. Biophys. Res. Commun. 2005, 338, 1745–1750. [Google Scholar] [CrossRef] [PubMed]
  55. Lula, J.F.; Rocha, M.O.d.C.; Nunes, M.d.C.P.; Ribeiro, A.L.P.; Teixeira, M.M.; Bahia, M.T.; Talvani, A. Plasma concentrations of tumour necrosis factor-alpha, tumour necrosis factor-related apoptosis-inducing ligand, and FasLigand/CD95L in patients with Chagas cardiomyopathy correlate with left ventricular dysfunction. Eur. J. Heart Fail. 2009, 11, 825–831. [Google Scholar] [CrossRef] [PubMed]
  56. Osmancik, P.; Peroutka, Z.; Budera, P.; Herman, D.; Stros, P.; Straka, Z.; Vondrak, K. Decreased Apoptosis following Successful Ablation of Atrial Fibrillation. Cardiology 2010, 116, 302–307. [Google Scholar] [CrossRef]
  57. Deftereos, S.; Giannopoulos, G.; Kossyvakis, C.; Raisakis, K.; Angelidis, C.; Efremidis, M.; Panagopoulou, V.; Kaoukis, A.; Theodorakis, A.; Toli, K.; et al. Association of post-cardioversion transcardiac concentration gradient of soluble tumor necrosis factor-related apoptosis-inducing ligand (sTRAIL) and inflammatory biomarkers to atrial fibrillation recurrence. Clin. Biochem. 2013, 46, 1020–1025. [Google Scholar] [CrossRef]
  58. Rewiuk, K.; Grodzicki, T. Osteoprotegerin and TRAIL in Acute Onset of Atrial Fibrillation. Biomed. Res. Int. 2015, 2015, 259843. [Google Scholar] [CrossRef] [Green Version]
  59. Chua, W.; Purmah, Y.; Cardoso, V.R.; Gkoutos, G.V.; Tull, S.P.; Neculau, G.; Thomas, M.R.; Kotecha, D.; Lip, G.Y.H.; Kirchhof, P.; et al. Data-driven discovery and validation of circulating blood-based biomarkers associated with prevalent atrial fibrillation. Eur. Heart J. 2019, 40, 1268–1276. [Google Scholar] [CrossRef] [Green Version]
  60. Cao, H.; Li, Q.; Li, M.; Od, R.; Wu, Z.; Zhou, Q.; Cao, B.; Chen, B.; Chen, Y.; Wang, D. Osteoprotegerin/RANK/RANKL axis and atrial remodeling in mitral valvular patients with atrial fibrillation. Int. J. Cardiol. 2013, 166, 702–708. [Google Scholar] [CrossRef]
  61. Cartland, S.P.; Lin, R.C.Y.; Genner, S.; Patil, M.S.; Martínez, G.J.; Barraclough, J.Y.; Gloss, B.; Misra, A.; Patel, S.; Kavurma, M.M. Vascular transcriptome landscape of Trail−/− mice: Implications and therapeutic strategies for diabetic vascular disease. FASEB J. 2020, 34, 9547–9562. [Google Scholar] [CrossRef]
  62. Di Bartolo, B.A.; Cartland, S.P.; Prado-Lourenco, L.; Griffith, T.S.; Gentile, C.; Ravindran, J.; Azahri, N.S.; Thai, T.; Yeung, A.W.; Thomas, S.R.; et al. Tumor Necrosis Factor-Related Apoptosis-Inducing Ligand (TRAIL) Promotes Angiogenesis and Ischemia-Induced Neovascularization Via NADPH Oxidase 4 (NOX4) and Nitric Oxide-Dependent Mechanisms. J. Am. Heart Assoc. 2015, 4, e002527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Cartland, S.P.; Harith, H.H.; Genner, S.W.; Dang, L.; Cogger, V.C.; Vellozzi, M.; Di Bartolo, B.A.; Thomas, S.R.; Adams, L.A.; Kavurma, M.M. Non-alcoholic fatty liver disease, vascular inflammation and insulin resistance are exacerbated by TRAIL deletion in mice. Sci. Rep. 2017, 7, 1898. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Lamhamedi-Cherradi, S.E.; Zheng, S.; Tisch, R.M.; Chen, Y.H. Critical roles of tumor necrosis factor-related apoptosis-inducing ligand in type 1 diabetes. Diabetes 2003, 52, 2274–2278. [Google Scholar] [CrossRef] [Green Version]
  65. Di Bartolo, B.A.; Chan, J.; Bennett, M.R.; Cartland, S.; Bao, S.; Tuch, B.E.; Kavurma, M.M. TNF-related apoptosis-inducing ligand (TRAIL) protects against diabetes and atherosclerosis in Apoe−/− mice. Diabetologia 2011, 54, 3157–3167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Watt, V.; Chamberlain, J.; Steiner, T.; Francis, S.; Crossman, D. TRAIL attenuates the development of atherosclerosis in apolipoprotein E deficient mice. Atherosclerosis 2011, 215, 348–354. [Google Scholar] [CrossRef] [PubMed]
  67. Manuneedhi Cholan, P.; Cartland, S.P.; Dang, L.; Rayner, B.S.; Patel, S.; Thomas, S.R.; Kavurma, M.M. TRAIL protects against endothelial dysfunction in vivo and inhibits angiotensin-II-induced oxidative stress in vascular endothelial cells in vitro. Free Radic. Biol. Med. 2018, 126, 341–349. [Google Scholar] [CrossRef]
  68. Cartland, S.P.; Erlich, J.H.; Kavurma, M.M. TRAIL Deficiency Contributes to Diabetic Nephropathy in Fat-Fed ApoE−/− Mice. PLoS ONE 2014, 9, e92952. [Google Scholar] [CrossRef] [Green Version]
  69. Di Bartolo, B.A.; Cartland, S.P.; Harith, H.H.; Bobryshev, Y.V.; Schoppet, M.; Kavurma, M.M. TRAIL-deficiency accelerates vascular calcification in atherosclerosis via modulation of RANKL. PLoS ONE 2013, 8, e74211. [Google Scholar] [CrossRef]
  70. Secchiero, P.; Candido, R.; Corallini, F.; Zacchigna, S.; Toffoli, B.; Rimondi, E.; Fabris, B.; Giacca, M.; Zauli, G. Systemic tumor necrosis factor-related apoptosis-inducing ligand delivery shows antiatherosclerotic activity in apolipoprotein E-null diabetic mice. Circulation 2006, 114, 1522–1530. [Google Scholar] [CrossRef] [Green Version]
  71. Rattazzi, M.; Bennett, B.J.; Bea, F.; Kirk, E.A.; Ricks, J.L.; Speer, M.; Schwartz, S.M.; Giachelli, C.M.; Rosenfeld, M.E. Calcification of advanced atherosclerotic lesions in the innominate arteries of ApoE-deficient mice: Potential role of chondrocyte-like cells. Arterioscler. Thromb. Vasc. Biol. 2005, 25, 1420–1425. [Google Scholar] [CrossRef] [Green Version]
  72. Bennett, B.J.; Scatena, M.; Kirk, E.A.; Rattazzi, M.; Varon, R.M.; Averill, M.; Schwartz, S.M.; Giachelli, C.M.; Rosenfeld, M.E. Osteoprotegerin inactivation accelerates advanced atherosclerotic lesion progression and calcification in older ApoE−/− mice. Arterioscler. Thromb. Vasc. Biol. 2006, 26, 2117–2124. [Google Scholar] [CrossRef] [Green Version]
  73. Liu, M.; Xiang, G.; Lu, J.; Xiang, L.; Dong, J.; Mei, W. TRAIL protects against endothelium injury in diabetes via Akt-eNOS signaling. Atherosclerosis 2014, 237, 718–724. [Google Scholar] [CrossRef] [PubMed]
  74. Wang, M.; Wei, Y.; Wang, X.; Ma, F.; Zhu, W.; Chen, X.; Zhong, X.; Li, S.; Zhang, J.; Liu, G.; et al. TRAIL inhibition by soluble death receptor 5 protects against acute myocardial infarction in rats. Heart Vessels 2022, 38, 448–458. [Google Scholar] [CrossRef] [PubMed]
  75. Tanner, M.A.; Thomas, T.P.; Grisanti, L.A. Death receptor 5 contributes to cardiomyocyte hypertrophy through epidermal growth factor receptor transactivation. J. Mol. Cell. Cardiol. 2019, 136, 1–14. [Google Scholar] [CrossRef]
  76. Toffoli, B.; Bernardi, S.; Candido, R.; Zacchigna, S.; Fabris, B.; Secchiero, P. TRAIL shows potential cardioprotective activity. Investig. New Drugs 2012, 30, 1257–1260. [Google Scholar] [CrossRef]
  77. Toffoli, B.; Fabris, B.; Bartelloni, G.; Bossi, F.; Bernardi, S. Dyslipidemia and Diabetes Increase the OPG/TRAIL Ratio in the Cardiovascular System. Mediat. Inflamm. 2016, 2016, 6529728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Shen, A.; Hou, X.; Yang, D.; Liu, T.; Zheng, D.; Deng, L.; Zhou, T. Role of osteoprotegerin and its gene polymorphisms in the occurrence of left ventricular hypertrophy in essential hypertensive patients. Medicine 2014, 93, e154. [Google Scholar] [CrossRef]
  79. Wang, Y.; Zhang, H.; Wang, Z.; Wei, Y.; Wang, M.; Liu, M.; Wang, X.; Jiang, Y.; Shi, G.; Zhao, D.; et al. Blocking the death checkpoint protein TRAIL improves cardiac function after myocardial infarction in monkeys, pigs, and rats. Sci. Transl. Med. 2020, 12, 1. [Google Scholar] [CrossRef]
  80. Harith, H.H.; Di Bartolo, B.A.; Cartland, S.P.; Genner, S.; Kavurma, M.M. Insulin promotes vascular smooth muscle cell proliferation and apoptosis via differential regulation of tumor necrosis factor-related apoptosis-inducing ligand. J. Diabetes 2016, 8, 568–578. [Google Scholar] [CrossRef]
  81. Azahri, N.S.M.; Di Bartolo, B.A.; Khachigian, L.M.; Kavurma, M.M. Sp1, acetylated histone-3 and p300 regulate TRAIL transcription: Mechanisms of PDGF-BB-mediated VSMC proliferation and migration. J. Cell. Biochem. 2012, 113, 2597–2606. [Google Scholar] [CrossRef]
  82. Liu, F.F.; Wu, X.; Zhang, Y.; Wang, Y.; Jiang, F. TRAIL/DR5 signaling promotes macrophage foam cell formation by modulating scavenger receptor expression. PLoS ONE 2014, 9, e87059. [Google Scholar] [CrossRef] [PubMed]
  83. Forde, H.; Harper, E.; Rochfort, K.D.; Wallace, R.G.; Davenport, C.; Smith, D.; Cummins, P.M. TRAIL inhibits oxidative stress in human aortic endothelial cells exposed to pro-inflammatory stimuli. Physiol. Rep. 2020, 8, e14612. [Google Scholar] [CrossRef]
  84. Secchiero, P.; Gonelli, A.; Carnevale, E.; Corallini, F.; Rizzardi, C.; Zacchigna, S.; Melato, M.; Zauli, G. Evidence for a proangiogenic activity of TNF-related apoptosis-inducing ligand. Neoplasia 2004, 6, 364–373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Chen, P.-L.; Easton, A.S. Evidence that tumor necrosis factor-related apoptosis inducing ligand (TRAIL) inhibits angiogenesis by inducing vascular endothelial cell apoptosis. Biochem. Biophys. Res. Commun. 2010, 391, 936–941. [Google Scholar] [CrossRef]
  86. Li, J.H.; Kirkiles-Smith, N.C.; McNiff, J.M.; Pober, J.S. TRAIL Induces Apoptosis and Inflammatory Gene Expression in Human Endothelial Cells 1. J. Immunol. 2003, 171, 1526–1533. [Google Scholar] [CrossRef] [Green Version]
  87. Liao, X.; Wang, X.; Gu, Y.; Chen, Q.; Chen, L.Y. Involvement of death receptor signaling in mechanical stretch-induced cardiomyocyte apoptosis. Life Sci. 2005, 77, 160–174. [Google Scholar] [CrossRef]
  88. Zhao, L.; Zhang, B. Doxorubicin induces cardiotoxicity through upregulation of death receptors mediated apoptosis in cardiomyocytes. Sci. Rep. 2017, 7, 44735. [Google Scholar] [CrossRef] [Green Version]
  89. Ridker, P.M.; MacFadyen, J.G.; Thuren, T.; Libby, P. Residual inflammatory risk associated with interleukin-18 and interleukin-6 after successful interleukin-1beta inhibition with canakinumab: Further rationale for the development of targeted anti-cytokine therapies for the treatment of atherothrombosis. Eur. Heart J. 2020, 41, 2153–2163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Xiang, G.; Zhang, J.; Ling, Y.; Zhao, L. Circulating level of TRAIL concentration is positively associated with endothelial function and increased by diabetic therapy in the newly diagnosed type 2 diabetic patients. Clin. Endocrinol. 2014, 80, 228–234. [Google Scholar] [CrossRef]
  91. Bisgin, A.; Yalcin, A.D.; Gorczynski, R.M. Circulating soluble tumor necrosis factor related apoptosis inducing-ligand (TRAIL) is decreased in type-2 newly diagnosed, non-drug using diabetic patients. Diabetes Res. Clin. Pract. 2012, 96, e84–e86. [Google Scholar] [CrossRef]
  92. Kawano, N.; Mori, K.; Emoto, M.; Lee, E.; Kobayashi, I.; Yamazaki, Y.; Urata, H.; Morioka, T.; Koyama, H.; Shoji, T.; et al. Association of serum TRAIL levels with atherosclerosis in patients with type 2 diabetes mellitus. Diabetes Res. Clin. Pract. 2011, 91, 316–320. [Google Scholar] [CrossRef] [PubMed]
  93. Corallini, F.; Celeghini, C.; Rizzardi, C.; Pandolfi, A.; Di Silvestre, S.; Vaccarezza, M.; Zauli, G. Insulin down-regulates TRAIL expression in vascular smooth muscle cells both in vivo and in vitro. J. Cell. Physiol. 2007, 212, 89–95. [Google Scholar] [CrossRef]
  94. Koliaki, C.; Katsilambros, N. Repositioning the Role of Tumor Necrosis Factor-Related Apoptosis-Inducing Ligand (TRAIL) on the TRAIL to the Development of Diabetes Mellitus: An Update of Experimental and Clinical Evidence. Int. J. Mol. Sci. 2022, 23, 3225. [Google Scholar] [CrossRef]
  95. Kavurma, M.M.; Rayner, K.J.; Karunakaran, D. The walking dead: Macrophage inflammation and death in atherosclerosis. Curr. Opin. Lipidol. 2017, 28, 91–98. [Google Scholar] [CrossRef] [Green Version]
  96. Goulay, R.; Mena Romo, L.; Hol, E.M.; Dijkhuizen, R.M. From Stroke to Dementia: A Comprehensive Review Exposing Tight Interactions Between Stroke and Amyloid-β Formation. Transl. Stroke Res. 2020, 11, 601–614. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Ouyang, F.; Jiang, Z.; Chen, X.; Chen, Y.; Wei, J.; Xing, S.; Zhang, J.; Fan, Y.; Zeng, J. Is Cerebral Amyloid-β Deposition Related to Post-stroke Cognitive Impairment? Transl. Stroke Res. 2021, 12, 946–957. [Google Scholar] [CrossRef] [PubMed]
  98. Fossati, S.; Ghiso, J.; Rostagno, A. TRAIL death receptors DR4 and DR5 mediate cerebral microvascular endothelial cell apoptosis induced by oligomeric Alzheimer’s Aβ. Cell Death Dis. 2012, 3, e321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Sandberg, W.J.; Yndestad, A.; Oie, E.; Smith, C.; Ueland, T.; Ovchinnikova, O.; Robertson, A.K.; Muller, F.; Semb, A.G.; Scholz, H.; et al. Enhanced T-cell expression of RANK ligand in acute coronary syndrome: Possible role in plaque destabilization. Arterioscler. Thromb. Vasc. Biol. 2006, 26, 857–863. [Google Scholar] [CrossRef] [Green Version]
  100. Oktay, A.A.; Akturk, H.K.; Paul, T.K.; O’Keefe, J.H.; Ventura, H.O.; Koch, C.A.; Lavie, C.J. Diabetes, Cardiomyopathy, and Heart Failure. In Endotext; Feingold, K.R., Anawalt, B., Blackman, M.R., Boyce, A., Chrousos, G., Corpas, E., de Herder, W.W., Dhatariya, K., Hofland, J., Dungan, K., et al., Eds.; MDtext: South Dartmouth, MA, USA, 2000. [Google Scholar]
  101. Wang, A.; Green, J.B.; Halperin, J.L.; Piccini, J.P., Sr. Atrial Fibrillation and Diabetes Mellitus: JACC Review Topic of the Week. J. Am. Coll. Cardiol. 2019, 74, 1107–1115. [Google Scholar] [CrossRef]
  102. Dzimiri, N.P.D.; Afrane, B.P.D.; Canver, C.C.M.D. Preferential Existence of Death-Inducing Proteins in the Human Cardiomyopathic Left Ventricle. J. Surg. Res. 2007, 142, 227–232. [Google Scholar] [CrossRef] [PubMed]
  103. Spierings, D.C.; de Vries, E.G.; Vellenga, E.; van den Heuvel, F.A.; Koornstra, J.J.; Wesseling, J.; Hollema, H.; de Jong, S. Tissue Distribution of the Death Ligand TRAIL and Its Receptors. J. Histochem. Cytochem. 2004, 52, 821–831. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Abbate, A.; Biondi-Zoccai, G.G.; Bussani, R.; Dobrina, A.; Camilot, D.; Feroce, F.; Rossiello, R.; Baldi, F.; Silvestri, F.; Biasucci, L.M.; et al. Increased myocardial apoptosis in patients with unfavorable left ventricular remodeling and early symptomatic post-infarction heart failure. J. Am. Coll. Cardiol. 2003, 41, 753–760. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Summary of circulating TRAIL and TRAIL receptor levels in atherosclerotic vascular diseases including coronary artery disease (CAD), stroke, and peripheral artery disease (PAD); s, soluble; ↑, increased; ↓, decreased. Details are within the text.
Figure 1. Summary of circulating TRAIL and TRAIL receptor levels in atherosclerotic vascular diseases including coronary artery disease (CAD), stroke, and peripheral artery disease (PAD); s, soluble; ↑, increased; ↓, decreased. Details are within the text.
Ijms 24 06725 g001
Figure 2. Therapeutic potential of TRAIL signals in atherosclerotic disease. Activating TRAIL signalling receptors in the vessel wall improve multiple features of atherosclerosis and diabetic-atherosclerotic disease. EC, endothelial cell; VSMC, vascular smooth muscle cells; ↑, increased; ↓, decreased. Details are within the text.
Figure 2. Therapeutic potential of TRAIL signals in atherosclerotic disease. Activating TRAIL signalling receptors in the vessel wall improve multiple features of atherosclerosis and diabetic-atherosclerotic disease. EC, endothelial cell; VSMC, vascular smooth muscle cells; ↑, increased; ↓, decreased. Details are within the text.
Ijms 24 06725 g002
Figure 3. Summary of soluble (s) concentrations and expression of TRAIL and TRAIL receptors in diseased myocardium; s, soluble; ↑, increased; ↓, decreased. Details are within the text.
Figure 3. Summary of soluble (s) concentrations and expression of TRAIL and TRAIL receptors in diseased myocardium; s, soluble; ↑, increased; ↓, decreased. Details are within the text.
Ijms 24 06725 g003
Figure 4. Therapeutic potential of TRAIL signals in the myocardium. Blocking TRAIL’s interaction with its signalling receptor(s) in pre-clinical models of heart failure and MI improve multiple functions of the myocardium. In contrast, activating TRAIL signals in the diabetic heart protects against cardiac fibrosis and apoptosis; ↑, increased; ↓, decreased. Details within the text.
Figure 4. Therapeutic potential of TRAIL signals in the myocardium. Blocking TRAIL’s interaction with its signalling receptor(s) in pre-clinical models of heart failure and MI improve multiple functions of the myocardium. In contrast, activating TRAIL signals in the diabetic heart protects against cardiac fibrosis and apoptosis; ↑, increased; ↓, decreased. Details within the text.
Ijms 24 06725 g004
Table 1. Summary of clinical findings and evidence of TRAIL signalling in CVDs.
Table 1. Summary of clinical findings and evidence of TRAIL signalling in CVDs.
DiseaseProteinFindingReference
CADTRAILDecreased circulating TRAIL in patients with CAD[20,21,22,23]
Negative association of circulating TRAIL with disease severity[16,17]
Reduced expression of TRAIL on monocytes from CAD patients[23]
DR5Increased circulating DR5 identified as a potential prognostic factor in all-cause mortality in MI patients, cardiovascular mortality, and MI and heart failure readmission in chronic kidney disease and diabetes patients[24,25,26,27,28]
OPGPositive association of circulating OPG with risk of all-cause and cardiovascular event-associated mortality[29,30]
Increased OPG associated with cardiovascular risk factors, e.g., diabetes[30,31]
Increased OPG a predictor for all-cause mortality in patients with renal failure[32]
Single OPG measurement insufficient to diagnose CAD in patients with angina[33]
StrokeTRAILDecreased circulating TRAIL in stroke patients[34,35,36,37,38,39]
Circulating levels associated with stroke severity[34,35,37,39]
Increased expression of TRAIL on monocytes with reduced circulating TRAIL at stroke onset[36]
DR5Increased DR5 in carotid plaques and circulation of symptomatic patients[40]
Elevated in LAA stroke[35]
OPGElevated in LAA stroke[35]
High levels at time of admission predicts poorer prognosis and mortality in ischemic stroke[41,42]
PADTRAILCirculating TRAIL reduced in patient with diabetic complication, i.e., foot ulcers and PAD[43,44]
Increased circulating TRAIL in patients with PAD and diabetes compared to diabetes alone[45]
OPGIncreased in PAD and diabetic PAD; associated with decreased TRAIL[44]
Increased in PAD and diabetic PAD; associated with higher TRAIL[45]
Heart failureTRAILInverse association of circulating TRAIL and all-cause mortality and hospitalisation[46,47,48]
TRAIL did not predict mortality in heart failure patients undergoing cardiac resynchronisation therapy[49]
DR5Positive correlation between plasma DR5 and HF incidence, preserved ejection fraction and left ventricular ejection fraction[48,50]
OPGPositive association with circulating OPG and prediction of adverse outcomes and mortality[51,52,53]
CardiomyopathyTRAILIncreased systemic TRAIL in dilated cardiomyopathy patients[54]
Positive association with circulating TRAIL and left ventricular ejection fraction and left ventricular diastolic diameter[55]
OPGIncreased in the myocardium of dilated cardiomyopathy patients, but no systemic change[54]
Atrial fibrillationTRAILReduced circulating TRAIL with successful ablation of AF[56]
Reduced trans cardiac gradient of TRAIL with AF recurrence[57]
Not useful in predicting the return to sinus rhythm[58]
DR5Inverse association of DR5 with AF, but no difference in concentration between patients in sinus rhythm and in AF[59]
OPGIdentified an increasing gradient of atrial expression of OPG with increasing degrees of AF[60]
Not useful in predicting the return to sinus rhythm[58]
Footnote: CAD, coronary artery disease; MI, myocardial infarction; LAA, large artery atherosclerosis; PAD, peripheral artery disease; HF, heart failure; AF, atrial fibrillation.
Table 2. Pre-clinical findings implicating TRAIL signals in cardiovascular disease.
Table 2. Pre-clinical findings implicating TRAIL signals in cardiovascular disease.
ModelAnimal/Cell TypeModel/TreatmentFindingReference
In vivoTrail−/− miceChowUpregulation of glucose transporter Glut1 in aortic tissue by microarray[61]
Peri-vascular cuff; intimal thickeningReduced intimal thickening; recombinant TRAIL recovered the neointima after cuff placement in Trail−/− mice; TRAIL stimulates VSMC proliferation and migration in vivo[10]
HLIReduced vascularisation after HLI; TRAIL gene therapy; improved limb perfusion and vascularisation[62]
Western dietInsulin resistance; increased vascular inflammation[63]
STZ-induced
diabetes
Increased susceptibility to STZ-induced diabetes[64]
NOD miceCY-induced diabetesNeutralising TRAIL by soluble TRAIL-R enhanced CY-induced diabetes[64]
Trail−/−Apoe−/− miceAtherosclerosis; Western dietDeveloped larger, macrophage-rich plaques of unstable phenotype (thin cap, large necrotic core, reduced VSMC and collagen content); developed features of type-2 diabetes[65]
Atherosclerosis;
cholate free Western diet
Developed larger atherosclerotic plaque[66]
Atherosclerosis; Western diet;
bone marrow
transplant
TRAIL-expressing bone marrow attenuated atherosclerosis; reduced inflammation[23]
Atherosclerosis; Western dietIncreased vascular oxidative stress; increased aortic endothelial dysfunction[67]
Atherosclerosis; Western dietIncreased inflammation; diabetic nephropathy[68]
Atherosclerosis; Western dietIncreased plaque; calcification[69]
Apoe−/− miceSTZ-induced diabetesAttenuation of atherosclerotic plaque with recombinant TRAIL or adenoviral TRAIL; reduced plaque macrophage content[70]
ChowOPG expressed in the brachiocephalic arteries, associated with chondrocyte-like cells[71]
Opg−/−Apoe−/− miceAtherosclerosis; chowIncreased atherosclerosis and calcification; reduced plaque cellularity[72]
RatsSTZ-induced diabetesEndothelial dysfunction was attenuated with recombinant TRAIL treatment[73]
Acute myocardial infarctionSoluble DR5 reduced infarct size, myocardial damage, and expression of apoptotic mediators[74]
C57BL/6 miceMD5-1 antibody and Bioymifi (small molecule DR5 agonist)DR5 activation increased heart weight, cardiac hypertrophy, left ventricular ejection fraction, and fractional shortening[75]
Apoe−/− miceSTZ-induced diabetes;Recombinant TRAIL and AAV TRAIL reduced cardiac fibrosis and apoptosis in diabetes[76]
STZ-induced diabetesIncreased OPG expression associated with cardiomyocyte hypertrophy[77]
Spontaneously hypertensive ratsRecombinant OPGIncreased left ventricular weight[78]
Rats, pigs and monkeysMyocardial infarctionDR5 inhibition reduced infarct size, cardiomyocyte death, and fibrosis and prevented ventricular wall thinning; preserved ejection fraction and fractional shortening[79]
In vitroVSMCRecombinant TRAILIncreases proliferation and migration in human aortic VSMCs[10]
Recombinant TRAILIncreases proliferation and migration via activation of insulin-like growth factor in human aortic VSMCs[14]
InsulinChronic insulin suppresses TRAIL expression and promotes apoptosis in human aortic VSMCs[80]
Recombinant PDGFBIncreases proliferation and migration via induction of TRAIL transcription and gene expression in human aortic VSMCs[81]
Monocyte/
macrophage
Recombinant IL-18Suppressed TRAIL gene expression and transcription via blocking NFκβ binding to the TRAIL promoter in human monocytes[23]
Recombinant TRAILIncreased lipid uptake and foam cell formation; macrophage apoptosis in RAW264.7 and THP-1 cells[82]
Basal, LPS and acLDLTrail−/− bone marrow-derived macrophages were more inflammatory and had a reduced ability to efferocytose; had impaired cholesterol and impaired ability to migrate compared to Trail+/+ bone marrow-derived macrophages[23]
ECsRecombinant TRAILTRAIL treatment inhibited TNFα/hyperglycaemia-induced inflammation and ROS production in HAECs [83]
TRAIL inhibited high glucose-induced ROS and cell death, in part via Akt and eNOS phosphorylation in HUVEC[73]
TRAIL protected against AngII-induced oxidative stress; reduced AngII-induced monocyte adhesion and improved EC integrity by redistributing VE-cadherin expression to the cell surface[67]
Increased HMEC-1 proliferation, migration, and tubule formation via NOX-4-inducible eNOS phosphorylation and nitric oxide production[62]
Increased HMEC-1 proliferation, migration, and tubule formation[13]
Increased HUVEC migration, invasion, and tubule formation[84]
Increased apoptosis of HMEC/d3 cells[85]
Increased HUVEC apoptosis[86]
CardiomyocytesRecombinant TRAIL and BioymifiDR5 activation via EGFR increased ERK1/2 phosphorylation for hypertrophy; cell death or viability not affected[75]
AAV-OPG vector and OPG siRNAOPG increased cell surface size and expression of hypertrophy proteins in rat cardiomyocytes[78]
Recombinant TRAIL and soluble DR5 (sDR5)TRAIL increased, while sDR5 neutralised stretch-induced apoptosis in rat cardiomyocytes[87]
DoxorubicinIncreased DR4 and DR5 mRNA/protein associating with enhanced TRAIL-induced apoptosis in human induced pluripotent stem cell-derived cardiomyocytes[88]
Footnote: AAV, adeno-associated virus; acLDL, acetylated low density lipoprotein; AngII, angiotensin II; CY, cyclophosphamide; EGFR, epidermal growth factor receptor; eNOS, endothelial nitric oxide synthase; ERK1/2, extracellular signal-regulated kinase 1/2; HAEC, human aortic endothelial cell; HLI, hindlimb ischemia; HMEC-1, human microvascular endothelial cell-1; HUVEC, human umbilical vein endothelial cell; IL-18, interleukin-18; LPS, lipopolysaccharide; NFκβ, nuclear factor κβ; NOX-4, NADPH oxidase-4; PDGFB, platelet-derived growth factor-B; ROS, reactive oxygen species; STZ, streptozotocin; VSMC, vascular smooth muscle cells.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kelland, E.; Patil, M.S.; Patel, S.; Cartland, S.P.; Kavurma, M.M. The Prognostic, Diagnostic, and Therapeutic Potential of TRAIL Signalling in Cardiovascular Diseases. Int. J. Mol. Sci. 2023, 24, 6725. https://doi.org/10.3390/ijms24076725

AMA Style

Kelland E, Patil MS, Patel S, Cartland SP, Kavurma MM. The Prognostic, Diagnostic, and Therapeutic Potential of TRAIL Signalling in Cardiovascular Diseases. International Journal of Molecular Sciences. 2023; 24(7):6725. https://doi.org/10.3390/ijms24076725

Chicago/Turabian Style

Kelland, Elaina, Manisha S. Patil, Sanjay Patel, Siân P. Cartland, and Mary M. Kavurma. 2023. "The Prognostic, Diagnostic, and Therapeutic Potential of TRAIL Signalling in Cardiovascular Diseases" International Journal of Molecular Sciences 24, no. 7: 6725. https://doi.org/10.3390/ijms24076725

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