Abstract
Free full text
Chromogranin A and its derived peptides: potential regulators of cholesterol homeostasis
Associated Data
Abstract
Chromogranin A (CHGA), a member of the granin family of proteins, has been an attractive therapeutic target and candidate biomarker for several cardiovascular, neurological, and inflammatory disorders. The prominence of CHGA stems from the pleiotropic roles of several bioactive peptides (e.g., catestatin, pancreastatin, vasostatins) generated by its proteolytic cleavage and by their wide anatomical distribution. These peptides are emerging as novel modulators of cardiometabolic diseases that are often linked to high blood cholesterol levels. However, their impact on cholesterol homeostasis is poorly understood. The dynamic nature of cholesterol and its multitudinous roles in almost every aspect of normal body function makes it an integral component of metabolic physiology. A tightly regulated coordination of cholesterol homeostasis is imperative for proper functioning of cellular and metabolic processes. The deregulation of cholesterol levels can result in several pathophysiological states. Although studies till date suggest regulatory roles for CHGA and its derived peptides on cholesterol levels, the mechanisms by which this is achieved still remain unclear. This review aims to aggregate and consolidate the available evidence linking CHGA with cholesterol homeostasis in health and disease. In addition, we also look at common molecular regulatory factors (viz., transcription factors and microRNAs) which could govern the expression of CHGA and genes involved in cholesterol homeostasis under basal and pathological conditions. In order to gain further insights into the pathways mediating cholesterol regulation by CHGA/its derived peptides, a few prospective signaling pathways are explored, which could act as primers for future studies.
Introduction
Chromogranin A (CHGA), belonging to the granin family of acidic, soluble glycoproteins stored in secretory granules and predominantly expressed in the neuroendocrine system, has garnered considerable attention owing to its multitude of physiological and pathophysiological roles [1–5]. At the intracellular level, CHGA plays a crucial role in the biogenesis of secretory vesicles [3, 6–8], binds with calcium [9–11], catecholamines and proteins involved in the regulatory secretory pathway [12–14]. At the extracellular level, CHGA acts as a prohormone giving rise to a number of biologically active peptides by proteolytic cleavage [15]. In view of their pleiotropic roles in several aspects of cellular and metabolic physiology, the CHGA-derived peptides are attractive therapeutic/drug target molecules [16–22].
A number of studies in the literature indicated that CHGA and some of its peptides might regulate circulating cholesterol levels, but any attempt to consolidate this link has not yet been made. Since cholesterol homeostasis is vital for various cellular/physiological processes, understanding the regulation of cholesterol by CHGA would have important implications for several aspects of metabolism. This review aims to dissect the link between CHGA and cholesterol homeostasis, with a focus on the CHGA-derived peptides catestatin (CST) [23–26], pancreastatin (PST) [27–30], vasostatins (VS-I and II) [31–34]. We discuss studies which chronicle the interplay of CHGA and cholesterol pathways, roles of the CHGA-derived peptides in mediating cholesterol homeostasis in basal and cholesterol-dysregulated conditions. We also look at a few intersecting signaling cascades, which provide compelling evidence of the pathways through which CHGA/CHGA-derived peptides could govern cholesterol homeostasis.
Cholesterol: a multi-faceted sterol
Cholesterol has enthralled the scientific community since its isolation from gallstones in the eighteenth century, and a testament to its remarkable significance are the thirteen Nobel Prizes awarded to scientists who have worked extensively on elucidating the several aspects of cholesterol structure and biosynthesis [35]. Cholesterol plays pivotal roles in membrane biology, signal transduction and acts as a precursor to several classes of steroid hormones (progestogens, estrogens, androgens, glucocorticoids and mineralocorticoids), oxysterols and bile acids [36]. The significance of cholesterol in physiology is further underscored by the numerous pathophysiological conditions which arise as a consequence of perturbed cholesterol homeostasis.
Cholesterol and membrane function
Glycerophospholipids, sphingolipids and cholesterol constitute the majority of the membrane bilayer, which not only help in stabilizing membrane proteins but also provide compatible environments in each organelle to facilitate their function. Cholesterol and sphingolipids are more enriched in the plasma membrane when compared to the Golgi and endoplasmic reticulum (ER), which renders the plasma membrane impermeable to the passage of small molecules [37, 38]. Although synthesized in the ER, the counter gradient accumulation of cholesterol in the plasma membrane involves specific transport between the different cellular membrane compartments at specific interfaces called membrane contact sites [39]. Cholesterol associates with the sphingolipid molecules in the membrane bilayer to form liquid-ordered microdomains called lipid rafts which regulate interactions between membrane components, thus altering the conformation of raft-associated proteins and consequently their activity, in addition to modulating membrane trafficking, signaling cascades and host–pathogen interactions [40, 41]. In addition, cholesterol modulates membrane rigidity and thickness [42, 43].
Role of cholesterol in signal transduction
Cholesterol seems to play roles in modulating channel receptors function such as the nicotinic acetylcholinergic receptor (nAChR), G-protein coupled receptor (GPCR) and tyrosine kinase signaling. Cholesterol also regulates Hedgehog signaling and influences membrane receptor function by either directly binding to the receptor and altering its conformation or indirectly by modifying the biophysical properties of the membrane [44–46].
Cholesterol and the secretory pathway
The secretory pathway directs the synthesis, maintenance and release of secretory proteins in a constitutive or regulated manner, in response to secretagogues [47]. The journey of a secretory protein begins in the lumen of the endoplasmic reticulum (ER), where it is subjected to post-translational modifications. The ER also regulates the efficiency of secretion by modulating protein folding [48]. The protein is then transported through the three distinct sub-compartments of the Golgi apparatus (cis, medial and trans) to the trans-Golgi network (TGN), from where it is sorted to its proper destination, either the plasma membrane or the endosomal/lysosomal system [47]. Proteins could then either enter the constitutive pathway and travel to the plasma membrane or are stored in granules and released from the cell upon external stimuli (regulatory pathway). The acidic milieu of the TGN is believed to facilitate aggregation of the proteins destined for the regulatory pathway, which are then packed into immature secretory granules at the TGN. These granules bud off from the TGN and undergo maturation during their transport towards the cell periphery. The secretory material gets concentrated by ten to hundred folds along the pathway [49–52].
Cholesterol plays crucial roles at various stages of the secretory pathway. It is an important constituent of the Golgi network and secretory vesicle membrane in endocrine/neuroendocrine cells (~25–65 mol%) [53–56]. Cholesterol depletion impairs binding of the prohormones, sorting and packaging of proopiomelanocortin (POMC) into secretory vesicles at the TGN while elevation of cellular cholesterol levels promotes efficient formation of chromogranin B (CHGB)-containing granules at the TGN in AtT-20 cells. Moreover, cholesterol seems to play a role in the formation of both constitutive and regulated secretory vesicles [54, 57]. Cholesterol also affects the integrity of SNARE (soluble N-ethylmaleimide-sensitive-factor attachment protein receptor) clusters, thus modulating exocytic release [58]. Cholesterol was observed to have an effect on granule fusion kinetics and exocytic events in the neuroendocrine cell line PC12 as well [59]. Cholesterol, by its virtue of forming ordered microdomains with sphingolipids in the plasma membrane, is believed to be involved in defining exocytosis sites and may promote membrane fusion by regulating membrane curvature [60].
Cholesterol also plays key roles in granin-mediated effects on secretory granules. The binding of secretogranin III (SgIII) to chromogranin A (CHGA) is essential for its targeting to secretory granules [61]. Interestingly, SgIII requires cholesterol to bind to the membrane of secretory granules and anchors CHGA to the membrane, which then mediates condensation/aggregation of pro-hormones in the secretory granule [62]. Thus, formation of dense-core aggregate-containing secretory granules requires molecules which could interact with both cholesterol and the core aggregates. SgIII acts as a bridge between the core aggregate and the cholesterol-rich microdomains which facilitates crucial processes in secretory granule biogenesis [63].
Cholesterol homeostasis
Circulatory forms of cholesterol
Lipid metabolism is regulated by lipoproteins. These are macromolecular, heterogeneous complexes which shield insoluble lipids from the aqueous plasma. These are made up of an inner dense hydrophobic core of both free and esterified forms of cholesterol, triacylglycerides, and an outer layer of hydrophilic moieties of apolipoproteins, phospholipids and free cholesterol [64]. Based on size, density, lipid and protein composition, plasma lipoproteins can be divided into seven classes: chylomicrons, chylomicron remnants, VLDL (very low-density lipoprotein), IDL (intermediate-density lipoprotein), LDL (low-density lipoprotein), HDL (high-density lipoprotein) and Lp(a) (Lipoprotein (a)) (Table (Table1)1) [65, 66].
Table 1
Lipoprotein | Size (nm) | Density (g/ml) | Major lipids | Major apolipoproteins | Pathophysiological implications |
---|---|---|---|---|---|
Chylomicrons | 75–1200 | <0.930 | Triglycerides | Apo B-48; Apo C-I, C-II, C-III; Apo E; Apo A-I, A-II, A-IV, A-V | Transport of dietary triglycerides and cholesterol, pro-atherogenic |
Chylomicron remnants | 30–80 | 0.930–1.006 | Triglycerides, Cholesterol | Apo B-48, Apo E | Pro-atherogenic |
VLDL | 30–80 | 0.930–1.006 | Triglycerides | Apo B-100; Apo E; Apo C-I, C-II, C-III; Apo A-V | Pro-atherogenic |
IDL | 25–35 | 1.006–1.019 | Triglycerides, Cholesterol | Apo B-100, Apo E, Apo C | Pro-atherogenic |
LDL | 18–25 | 1.019–1.063 | Cholesterol | Apo B-100 | Hypertriglyceridemia, type 2 diabetes, obesity, inflammation, pro-atherogenic |
HDL | 5–12 | 1.063–1.210 | Cholesterol, Phospholipids | Apo A-I, A-II, A-IV, A-V; Apo C-I, C-II, C-III; Apo E | Anti-oxidant, antithrombotic, anti-apoptotic, anti-inflammatory, antiatherogenic |
Lp (a) | ~30 | 1.055–1.085 | Cholesterol | Apo B-100; Apo (a) | Pro-atherogenic |
Cholesterol homeostasis involves several processes including biosynthesis, absorption, storage by the process of esterification, and degradation by efflux and conversion to bile, which are depicted schematically in Fig. 1.
Biosynthesis
Typically, all cells and tissues can synthesize cholesterol, but the major sites of synthesis are the liver, intestine and reproductive organs [67]. It is synthesized in the ER through a series of 30 enzymatic reactions. It all begins with the condensation of acetate molecules by successive reversible reactions to first form acetoacetyl-CoA, followed by 3-hydroxy-3-methylglutaryl-CoA (HMG-CoA). The subsequent reaction is the rate-limiting step of the cholesterol biosynthesis pathway, catalyzed by HMG-CoA reductase (HMGCR), which reduces HMG-CoA to mevalonate. Further reactions, like condensation and cyclization, result in the formation of the cholesterol molecule [68]. The protein, which is considered as an important rate-limiting enzyme next to HMGCR, is squalene epoxidase (SQLE).
Absorption
Cholesterol uptake is regulated directly by NPC1L1 transporter (highly expressed in the apical surfaces of hepatocytes and enterocytes) and indirectly by LDL-receptor (LDLR, expressed in the basolateral surface of most cells) [69]. NPC1L1 is a glycosylated membrane protein, which is the primary moderator of cholesterol absorption. Under steady-state conditions, NPC1L1 resides in the endocytic recycling compartment (ERC) and upon cholesterol availability, it translocates to the plasma membrane. The N-terminal domain of NPC1L1 binds with cholesterol, which leads to a conformational change in the structure of the protein. This results in the dissociation of NPC1L1-bound cholesterol from the membrane to be packed within endocytic vesicles in a clathrin-dependent manner. These vesicles migrate towards the ERC, from where they reach the ER, by an unknown mechanism [69, 70]. The membrane glycoprotein LDLR binds with LDL from blood and this complex gets internalized via a clathrin-dependent endocytic pathway, involving ARH and DAB2 [71, 72]. Multiple LDL-LDLR-containing vesicles fuse together to form endosomes harbouring a pH of 6.5, leading to LDLR dissociation and return to the plasma membrane. The remaining LDL-containing vesicles fuse with the lysosomal compartment to undergo hydrolysis to yield cholesterol [35].
LDLR levels at the plasma membrane are regulated by the secretory serine protease PCSK9, which is highly expressed in liver and small intestine. Upon activation, it undergoes a conformational change which enables it to bind with the epidermal growth factor A (EGF-A) domain of LDLR [73]. The PCSK9-LDLR complex gets internalized via a clathrin-mediated process and gets transported to the endosomes and then to the lysosomes, where it gets eventually degraded. The acidic pH of endosomes facilitates the interaction of PCSK9 with the ligand-binding domain of LDLR, thereby disrupting the process of recycling back to the plasma membrane [69]. This negative regulation elevates plasma LDL levels leading to hypercholesterolemia, which recently led to the development of anti-PCSK9 therapies [74].
Storage
Accumulation of intracellular free cholesterol above physiological levels leads to cytotoxicity [75]. Hence, it is stored in the form of cytoplasmic lipid droplets by the process of esterification. This is mediated by the enzyme acyl-CoA:cholesterol acyltransferase (ACAT), which is an ER membrane protein. Its two homologs, ACAT1 and ACAT2 are differentially expressed in mammals. The highest expression of ACAT1 is found in macrophages, while ACAT2 is mainly localized to hepatocytes and enterocytes in humans [76]. ACAT1 plays a major role in the formation of foam cells in macrophages, while ACAT2 is involved in cholesterol absorption. Both these isozymes work together in chylomicron and VLDL assembly/synthesis [77].
Trafficking
Intracellular cholesterol transport across various sub-cellular membranes takes place because of both de novo synthesis from ER and its uptake from extracellular milieu. It can involve both vesicular and non-vesicular pathways.
Once synthesized in the ER, cholesterol can be transported to the plasma membrane or different compartments with the help of specialized lipid-binding proteins. These include members of steroidogenic acute regulatory protein-related lipid transfer (START) domain family and oxysterol-binding protein (OSBP)-related proteins (ORPs) [78]. In humans, STARD1, STARD3 and STARD4 play an important role in cholesterol transport. In steroidogenic tissues, STARD1 transports cholesterol from outer to inner mitochondrial membrane. STARD3 facilitates transport from ER to endosomes, and then to plasma membrane, while STARD4 mediates non-vesicular transport from plasma membrane to ERC [79]. ORP2 is involved in assisting movement of lipids across lipid droplets [80]; ORP9L and OSBP1 are involved in sterol transfer between ER and Golgi [69, 81].
LDLR internalizes extracellular cholesterol (with APOB or APOE proteins) which is transported to endosomes. Efflux from late endosomes is facilitated by the membrane spanning protein, NPC1 and the soluble protein NPC2. Two more proteins containing sterol-binding domains include MLN64 and ORP1, which bind to cholesterol and 25-hydroxycholesterol, respectively; these proteins regulate membrane trafficking and subcellular distribution of late endosomes. Rab GTPases including Rab8, Rab11, Rab7 and Rab9 are also involved in this sterol-dependent endosomal transport system [76].
Efflux
The selective transport of cholesterol from peripheral tissues by HDL to liver for its excretion in the form of bile and feces is called “reverse cholesterol transport.” Most of the peripheral tissues cannot catabolize cholesterol and thus require special transporters for its efflux to the acceptor, HDL. The lipid-poor ApoA-I protein accepts cholesterol from both liver and macrophages via ABCA1 transporter. This is the nascent HDL particle. With the help of LCAT, it acquires cholesterol released from the ABCG1 transporter (highly expressed in macrophages) and via ABCG5 and ABCG8 (widely expressed in the luminal surface of intestine and bile duct). The resultant mature HDL can unload cholesterol back to the liver by two processes: SR-BI transporter (direct process); or CETP-mediated (indirect process) [69, 82].
Degradation
The formation of bile is the most important regulated process for the degradation of cholesterol which occurs via two pathways: classical pathway (neutral) and alterative (acidic pathway). The major enzymes catalyzing these pathways include the ER-specific enzyme cholesterol 7α hydroxylase (CYP7A1) for the neutral pathway and the mitochondria-specific sterol 27-hydroxylase (CYP27A) for the acidic pathway [76].
Transcriptional and post-transcriptional mechanisms regulating cholesterol homeostasis are depicted in Table Table22.
Table 2
Regulator | Pathway | Target | Function of the target | Mechanism of regulation | |
---|---|---|---|---|---|
Transcriptional regulation | |||||
SCAP | Biosynthesis | SREBP2 | Master transcriptional regulator | SCAP interacts with SREBP2 during sterol depletion for transcriptional activation of target gene [69] | |
SREBP2 | Absorption | NPC1L1 | Key mediator of cholesterol absorption | Positive regulation [69] | |
LXRs | Efflux | ABCA1, ABCG5, ABCG8 | Transporters for cholesterol efflux | Positive regulation [69] | |
IFN-γ | Esterification | ACAT1 | Esterification enzyme | Activation of P1 promoter of ACAT1 [69] | |
Estrogen | Metabolism | LXRα/RXRα | Nuclear receptor | Sp1 mediated transcriptional activation [278] | |
PPAR-α and PPAR-γ | Efflux | ABCA1 | Transporter for cholesterol efflux | Transcriptional activation via LXRα [279] | |
FXR | HDL metabolism | ApoA-I | Major apolipoprotein in HDL | Downregulation of ApoA-I promoter activity in the presence of bile or bile agonists [280] | |
FOXO3 | Biosynthesis, Absorption | SREBP2 | Master transcriptional regulator | Downregulation of Srebp2 and its target genes like PCSK9 [281] | |
Nrf1 | Efflux | LXR | Nuclear receptor | De-repression of LXR pathway in the presence of cholesterol [69] | |
Post-transcriptional regulation | |||||
INSIG | Biosynthesis | SREBP2, HMGCR | Transcriptional regulator; rate limiting enzyme of cholesterol biosynthesis | SCAP recruits INSIG during cholesterol repletion to retain SCAP-SREBP2-INSIG complex within ER; Sterol levels can induce binding of INSI-E3 ligase to HMGCR for its proteasomal degradation [69] | |
Hsp90 | Biosynthesis | SREBP2 | Master transcriptional regulator | Proteolytic activation of SREBP2 by stabilizing the SCAP-SREBP2 complex in ER[69] | |
AMPK | Biosynthesis | HMGCR | Rate limiting enzyme of cholesterol biosynthesis | AMPK blocks cholesterol biosynthesis [199] | |
PP2A | Biosynthesis | AMPK, HMGCR | Kinase; rate limiting enzyme of cholesterol biosynthesis | Dephosphorylation [282] | |
IDOL | Absorption | LDLR | Receptor for LDL, VLDL | Proteasomal degradation [69] | |
PCSK9 | Absorption | LDLR | Receptor for LDL, VLDL | Lysosomal degradation [69] | |
Estrogen | Degradation | CYP7A1 | Bile synthesis | Increased enzyme activity of CYP7A1 [278] |
Perturbed cholesterol homeostasis: pathophysiological implications
Cardiovascular diseases
Aberrations in the cholesterol pathway trigger several metabolic, cardiovascular, and neurological disorders. The most prominent consequence of dysregulated cholesterol homeostasis is familial hypercholesterolemia (FH), which is characterized by elevations in blood cholesterol levels and increased incidence of myocardial infarctions (MI) [35]. Early studies delineated the predominant causative genetic region to span the LDLR locus, wherein mutations impairing LDLR function, thus impacting LDL uptake and turnover, were observed in FH patients [35, 83]. Subsequent studies documented mutations in the APOB, LDLRAP1, PCSK9, LIPA, ABCG5 and ABCG8 genes in FH patients [83, 84]. FH predisposes the individual to an enhanced risk (~3–13 times) of atherosclerotic cardiovascular diseases (CVDs) [84].
Elevated plasma total cholesterol and LDL cholesterol levels, but also lower HDL cholesterol (HDL-c) levels, are major risk factors for CVDs [85]. Increased number of small LDL particles are associated with a higher risk of ischemic heart disease [86], whereas apo-B/apoA-I levels were observed to be a good predictor of MI risk [87]. Elevated non-HDL cholesterol levels were attributed to~3.9 million global deaths from ischemic heart disease and ischemic stroke in 2017 [88]. Moreover, multiple lines of evidence suggest involvement of disrupted cholesterol homeostasis in the pathophysiology of metabolic syndrome [89], a key risk factor of CVDs.
Elucidation of the genetic basis of CAD and MI identified loci associated with LDL metabolism and other aspects of cholesterol homeostasis [90–95]. Investigation of statins (HMGCR inhibitors), CETP inhibitors and PCSK9 inhibitors showed promising potential to improve cardiovascular outcomes [96].
Other diseases resulting from perturbation of cholesterol homeostasis include: the rare autosomal recessive sitosterolemia (defects in the ABCG5-ABCG8 locus characterized by high levels of plant sterols in the plasma and tissues, due to augmented intestinal absorption and diminished bile secretion, resulting in heightened risk for cardiovascular diseases, xanthomas and hematologic abnormalities [97]); autosomal recessive Tangier’s disease (genetic mutations in the cholesterol transporter ABCA1 resulting in neurological abnormalities, enlarged spleen and tonsils, deposition of cholesterol esters in reticuloendothelial cells and varying degrees of atherosclerosis [98]).
Neurological disorders
SNPs in the cholesterol homeostasis pathway are associated with several neurological/neurodegenerative diseases and inflammatory diseases which are depicted in Fig. 2. The brain accounts for a substantial proportion of the total cholesterol in humans and the salient roles of cholesterol in brain development and neuronal function are well known. The importance of cholesterol in the central nervous system is further underscored by findings of disrupted cholesterol homeostasis in several neurodegenerative diseases such as Alzheimer’s disease, Parkinson’s disease, Huntington’s disease, and autism spectrum disorders [99–101]. Mutations in the NPC1 and NPC2 genes cause an autosomal recessive condition: Niemann-Pick type C disease, which is characterized by defects in cholesterol transport in the lysosomes and subsequent cell accumulation, resulting in neurodegeneration and hepatosplenomegaly [102]. Intriguingly, dysregulated cholesterol homeostasis is also suggested to play a role in mood disorders like anxiety, depression and suicidal behavior [103].
Hepatic diseases
Several studies have demonstrated elevated free cholesterol levels, accumulation of hepatic cholesterol, presumably due to diminished efflux and catabolism, in addition to impaired esterification, in non-alcoholic fatty liver disease (NAFLD) and non-alcoholic steatohepatitis (NASH). The escalated cholesterol levels incite a strong inflammatory response in the liver cells, promoting fibrosis and liver damage [104, 105]. LIPA deficiency could result in the infantile-onset Wolman disease (WD) (characterized by build-up of lysosomal cholesteryl esters in several organs) and the adult-onset cholesteryl ester storage disease (CESD), an autosomal recessive condition characterized by accumulation of lysosomal cholesterol and triglycerides, with liver abnormalities, premature atherosclerosis and cardiovascular disease [106, 107].
COVID-19
Obesity has been documented as the second risk factor after age for severity in SARS-CoV-2 infection, highlighting the importance of lipid metabolism in the disease [108]. An accumulating body of evidence expound the role of cholesterol in the pathophysiology of the recent global pandemic, COVID-19. A reduction in total cholesterol (TC), HDL and LDL cholesterol levels were observed in a Wuhan COVID-19 patient cohort compared to controls with statistically significant decreases in the LDL and TC levels across increasing degrees of disease severity. These observations need however to be considered cautiously as these modifications may result from liver damage especially in severe and elderly patients. Moreover, LDL, HDL and TC levels negatively correlated with CRP, whereas LDL and TC levels exhibited significant negative correlations with IL-6 levels, suggesting an interplay of lipid metabolism and inflammatory cascades in the development of COVID-19 progression [109]. Interestingly, the levels of cholesterol, precursor sterols and the oxysterol 27-hydroxycholesterol (27-OHC) are diminished in patients and 27-OHC exhibits antiviral effects against SARS-CoV-2 [110]. Another retrospective study reported a lower risk of mortality in COVID-19 patients who were on statin medications, which could be attributed to the immunomodulatory properties of statins [111]. A comprehensive review by Schmidt et al. enumerates the salient features of statins and their enormous potential in devising therapeutic strategies [112]. Intriguingly, SREBP-2 and PCSK9 mRNA levels were enhanced in more severe COVID-19 cases. Moreover, SREBP-2 activity reflected the degree of disease severity and mediated the inflammatory responses in these patients, further corroborated by the suppression of inflammatory cytokines and the NLRP3 inflammasome expression in an infectious disease mouse model [113]. Of note, cholesterol modulated viral fusion by murine coronavirus [114], and a recent study demonstrated that cholesterol binds to the SARS-CoV-2 spike protein and the HDL receptor SR-B1 facilitated viral entry [115]. Put together, these studies and others highlight the relevance of cholesterol homeostasis in COVID-19 and the need for more studies to enable development of appropriate therapeutic strategies.
Chromogranin A and its derived peptides
The tale of the chromaffin cells originated in the mid-nineteenth century with the discovery of catecholamine release from adrenomedullary cells, and subsequent breakthrough studies unearthed evidence of the localization of catecholamines and ATP in chromaffin granules, along with the granins, neuropeptides, biogenic amines, and growth factors [116]. The granins are acidic, heat-stable, soluble, calcium-binding proteins that help stabilize the contents of the granules to an osmotically inert form until discharge. The most prominent member of the granin family is CHGA, whose roles in secretory granule biogenesis and regulated secretory pathway are becoming understood. For instance, CHGA was recently shown to promote granule biogenesis through the interaction with the lipid phosphatidic acid at the Golgi leading to changes in membrane topology [117]. In addition, CHGA is also believed to regulate cytosolic calcium levels through their interactions with IP3R [118]. Chga−/− mice are hypertensive and exhibit a reduction in secretory granule number [119]. The physiological relevance of CHGA can be attributed to the presence of multiple dibasic sites which facilitate proteolytic processing resulting in the generation of several bioactive peptides with diverse autocrine, paracrine, and endocrine effects including VS-I and VS-II (human CHGA1-76, and CHGA1-115, respectively), vasodilators; PST (human CHGA250-301), a dysglycemic peptide; and CST (human CHGA352-372), a master regulator of cardiovascular functions with anti-hypertensive, cardioprotective and pro-insulinic effects [1–3, 120] (Fig. 3). The roles of these peptides in physiology and metabolism, except GE25, whose physiological significance is poorly understood, are listed in Table Table3.3. The roles of CST, PST and VS-I and VS-II in cholesterol homeostasis are discussed in greater detail in subsequent sections.
Table 3
CHGA-derived peptide | Function |
---|---|
Pancreastatin (PST) | Inhibits glucose-stimulated insulin release [27] Inhibits insulin-mediated glucose uptake [161, 283] Pro-gluconeogenesis and pro-glycogenolysis [30, 284] Pro-inflammatory [210] |
Catestatin (CST) | Catecholamine release inhibitory [285] Anti-hypertensive [119] Anti-adrenergic [234] Vasodilatory [288] Promotes insulin sensitivity [267] Anti-inflammatory [267] |
Vasostatins (VS-I and VS-II) | Vasoconstriction-inhibitory [32] Anti-adrenergic [237] Anti-inflammatory [141] Anti-microbial [293] Regulation of cell adhesion [294] |
Parastatin (PRST) | Inhibits parathyroid hormone secretion [295] |
Serpinin | β-adrenergic agonist [239] Cardioprotective [296] |
Chromofungin | Anti-fungal [297] Cardioprotective [298] |
Chromacin | Anti-microbial [299] |
WE14 | Modulates histamine release from mast cells [300] Autoantigen for diabetogenic CD4+ T cells [301] |
Chromogranin A: a potential master player in cholesterol homeostasis?
Microarray analysis of the adrenal glands of blood pressure high (BPH) mice exhibited~1.5-fold and~3.1-fold overexpression of Chga and Hmgcr respectively, suggesting enhanced catecholamine storage and dysregulated cholesterol metabolism in hypertensive mice [121]. Upregulation of both these genes in the hypertensive state suggests an interplay of these two pathways in hypertension: both genes could be under common regulatory mechanisms in the hypertensive state or CHGA and derived peptides could regulate Hmgcr expression or both. This seemingly causal link is reinforced by the observation that~93% of the genes involved in the cholesterol biosynthesis pathway are downregulated in the adrenal glands of Chga−/− mice whereas~73% of the cholesterol biosynthesis genes are upregulated in the liver of these mice [122]. Consistently, the expression of Hmgcr is downregulated by~2.6-fold in the adrenal glands while that is upregulated by~1.6-fold in the liver with concomitant reduction of adrenal cholesterol and elevated hepatic cholesterol levels in Chga−/− mice. This finding strongly suggests the impact of CHGA and/or its derived peptides on cholesterol homeostasis in pathological conditions.
Furthermore, a closer look at the primary structure of the CHGA protein reveals the presence of several potential cholesterol-binding motifs which are found to occur in the transmembrane domains of receptors and ion channels [123, 124]. These motifs, which were discovered using the Fuzzpro application in the EMBOSS suite, are outlined in Fig. 4. Putative motifs were identified in the vasostatin/chromofungin region (45–54 aa) and the parastatin (391–401 aa) peptides, implicating potential roles for these peptides in cholesterol homeostasis. Interestingly, the motifs, especially the critical residues, are well-conserved across primates and artiodactylae, and to a lesser extent in the rodents. The presence of multiple cholesterol-binding motifs is a promising premise to propel elaborate investigations on the binding of cholesterol to CHGA and its functional ramifications.
In this direction, we discuss studies which support the potential impact of CHGA-derived peptides on cholesterol levels in basal and pathological states below.
Correlations between the plasma/serum levels of CHGA-derived peptides and cholesterol in basal and pathophysiological conditions
Plasma CST levels were observed to be higher in untreated hypertensive patients when compared to controls, although the significance was lost upon adjustments for age, gender, height and weight. However, in hypertensive patients, CST levels correlated with age and HDL cholesterol. When the patients were stratified according to their HDL-c levels, plasma CST levels increased gradually with increasing HDL-c levels among groups. Interestingly, the group with the lowest range of HDL-c levels had a higher proportion of males, larger waist circumference and higher metabolic syndrome score, suggesting that CST impacts lipid and metabolic profiles in hypertension [125]. However, in male adults with obstructive sleep apnea (OSA), which is a major cause of secondary hypertension [126], serum CST levels were higher when compared to controls, and exhibited significant negative correlation with HDL-c [127]. Interestingly, plasma CST levels were lower in children with moderate-to-severe OSA, when compared to controls and children with mild OSA. No correlations between CST levels and cholesterol levels were observed in this cohort [128]. The difference in the trend observed in the OSA patients could reflect the effect of age and gender-related hormones, which could have an influence on circulating CST levels [127]. These studies also reflect the interplay between CST and cholesterol metabolism, which seems to be altered in different hypertensive contexts, thus implicating a multitude of factors mediating the CST-cholesterol pathway interaction.
Effects of genetic variations in the CHGA locus on cholesterol levels in basal and pathophysiological conditions
Although modest there is some evidence on the link between the variants in the CHGA locus and cholesterol levels. Figure 5 depicts the associations of several SNPs in the CHGA locus with total/HDL/LDL cholesterol levels from different genome-wide association studies. Most of the promoter SNPs seem to be associated with lower total cholesterol except for rs9658629 (A-1018 T) and rs9658638 (C-57 T), which seem to contribute to higher cholesterol levels. Interestingly, the A-1018 T and C-57 T variants exhibited enhanced affinities to the binding of the c-Rel transcription factor at basal and TNF-α induced conditions [129]. The subsequent alteration in the expression of CHGA could play roles in higher cholesterol levels.
SNPs in the vasostatin region exhibit associations with decreased HDL-c levels, while the PST variant rs9658664 (G297S) was associated with elevated HDL-c and ratio of total cholesterol to HDL levels (Fig. 5). Of note, carriers of this variant exhibited higher total cholesterol levels than the wild-type carriers in an Indian population [130]. Interestingly, two synonymous SNPs in the PST region seem to enhance the risk for higher LDL and total cholesterol (Fig. 5). A variant in the serpinin region appears to be associated with higher LDL cholesterol levels (Fig. 5), thus suggesting metabolic roles for this peptide, which are hitherto unknown.
As shown in Fig. 5, a few of the SNPs are in linkage disequilibrium (LD). Globally, the cholesterol-related SNPs in the CHGA region which exhibited strong LD are the promoter variants rs9658629 and rs9658638; the coding variant rs9658655 (E246D in the mature protein) and the intron variant rs9658656; and the intronic variants rs9658656 and rs4900163. Knowledge of SNPs in LD throws more light on the effect of the genetic architecture of the CHGA locus on cholesterol levels and could potentially aid in development of suitable screening strategies.
Roles of CHGA and its derived peptides in dysregulated cholesterol homeostasis conditions
CHGA levels are altered in several cardiovascular diseases and associated risk factors. CHGA levels are elevated in both essential and secondary hypertension [131, 132], while CST levels were diminished in normotensive offspring of hypertensive patients, suggesting that CST could act as an indicator of risk of developing hypertension [133]. The pathophysiological significance of CHGA is further underscored by the observation that Chga−/− mice are hypertensive and the administration of CST diminished blood pressure levels substantially [119]. Although there is a report of an elevation in PST levels after glucose-loading in hypertensive subjects (normoinsulinaemic/hyperinsulinaemic/glucose-intolerant), the role of PST in hypertension has not been explored properly [134]. PST levels are elevated in diabetes [28, 135–137] and plasma CHGA levels were also observed to be higher in Type 2 diabetes mellitus (T2DM) patients [138]. Moreover, the roles played by CHGA-derived peptides in different aspects of diabetes, such as glucose metabolism, insulin resistance, inflammation, etc., are well known, and discussed in subsequent sections.
Atherosclerosis and coronary artery disease
CHGA is expressed in the medial layer of coronary arteries from CAD and non-CAD subjects. CST is expressed in the endothelium and part of medial layer in non-CAD patients and higher expression was observed in the intima and atheromatous plaques in stenotic arteries from CAD patients. However, plasma CST levels were lower in CAD patients, and decreased with the increase in severity of vessel disease [139, 140]. CST was observed to play a protective role in atherosclerosis. In vitro, CST treatment exerted anti-inflammatory effects and suppressed foam cell formation induced by ox-LDL. Moreover, CST suppressed ACAT-1 expression and promoted ABCA1 expression in cultured macrophages [139], suggesting that CST could modulate cholesterol homeostasis in atherosclerotic conditions by downregulating cholesterol esterification and promoting efflux. Administration of CST in ApoE−/− mice, an animal model of atherosclerosis, diminished plaque size and reduced infiltration by immune cells. However, the mice treated with saline or CST did not exhibit any differences in the total cholesterol levels [139]. CST suppressed endothelial cell dysfunction, a hallmark feature of atherosclerosis, and the anti-inflammatory role of CST seemed to be mediated via ACE2 as evidenced by the mitigation of the anti-inflammatory effects of CST upon inhibition of ACE2. Remarkably, administration of DX600 (an ACE2 inhibitor) partially reversed the anti-atherogenic effects of CST in ApoE−/− mice [140], implying that CST exerts its protective effects via ACE2, and it would be interesting to see if this notion holds true in other pathological conditions involving the RAAS (Renin–Angiotensin–Aldosterone System) pathway.
Like CST, plasma levels of VS-II were lower in CAD patients as compared to controls and diminished in diabetic CAD patients as opposed to non-diabetic CAD patients [141]. However, plasma VS-I and total CHGA levels were higher in diabetic patients with carotid artery atherosclerosis than non-diabetic patients, and they correlated with carotid artery maximum stenosis [142]. The expression of VS-I and VS-II in atherosclerotic arteries were observed to be lesser when compared to controls, while CHGA protein levels were higher in atherosclerotic aorta [141], raising the possibility of reduced processing of CHGA into its anti-inflammatory peptides in diseased conditions. Indeed, lower conversion of CHGA to CST was observed in heart failure [143], suggesting impaired proteolytic processing in pathological states. Both VS-I and VS-II exerted anti-adhesive and anti-inflammatory roles in human arterial endothelial cells [141], while VS-I administration ameliorated vascular inflammation/remodeling, macrophage infiltration and reduced the size of atherosclerotic lesions in ApoE−/− mice, though there were no differences in cholesterol levels. In addition, like CST, VS-I suppressed ox-LDL-induced foam cell formation by modulating the expression of ACAT-1 and ABCA1 in THP-1 monocyte-derived macrophages. Interestingly, VS-II repressed AngII-mediated migration of human aortic smooth muscle cells [144], suggesting that the anti-atherogenic effect of VS-I could involve the RAAS pathway, similar to CST. Future studies in this direction are required to delineate the link between CHGA/derived peptides and the RAAS pathway.
Studies suggest plasma CHGA levels to be a good prognostic marker in patients with MI and CAD, which exhibited association with higher long-term mortality and worse prognosis [145–147]. However, the molecular mechanisms are yet to be investigated. There are inconsistent reports on plasma CST levels in CAD [148–151], which make it difficult to draw solid conclusions, and these could be attributed to smaller sample size in most of these studies and different time points of plasma collection. However, higher plasma CST levels seem to be coincident with more severe outcomes and elevated LDL-c levels in MI patients [152, 153], suggesting that CST might have roles in regulating LDL-c levels in MI. Patients with chronic heart failure (HF) and previous MI had lower levels of VS-II and HDL-cholesterol compared to healthy controls, and VS-II levels correlated positively with left ventricle ejection fraction (LVEF), negatively with stage of HF and inflammatory markers (TNF-α and hsCRP). Moreover, VS-II had a cardio-protective effect and alleviated myocardial remodeling and inflammation in rats with ischemic heart failure. Interestingly, VS-II treatment led to an increase in the protein levels of ACE2 and decrease in the protein levels of ACE [154], further implicating the role of CHGA-derived peptides in modulation of the RAAS system in pathophysiological conditions.
Diabetes and insulin resistance
The association of diabetes with cardiovascular diseases is well established, with two- to three-fold higher incidence of cardiovascular complications in diabetics than the non-diabetic group [155]. As mentioned previously, dysregulated cholesterol homeostasis in metabolic diseases is a key risk factor in CVDs. In an interesting study, insulin stimulated the expression of Liver-X Receptor-α (LXRα), an important regulator of cholesterol metabolism in a time-dependent manner. Gene targeting studies using the LXRα/β−/−mice revealed downregulation in insulin-mediated induction of SREBP-1c, a key transcriptional regulator of cholesterol metabolism and many other lipogenic genes, thereby implying insulin’s effect on LXR-induced genes [156]. Insulin’s stimulatory role on SREBP-1c [157] expression could be annulled by PST, which downregulated SREBP-1c expression in Chga knockout mice [30].
The hexosamine biosynthetic pathway (HBP) has an important significance in diabetic cardiomyopathy, insulin resistance and β-cell failure [158, 159]. Interestingly, through HBP, hyperinsulinemia increased glucose flux and O-linked glycosylation of Sp1, which activated HMGCR and SREBP1 in mouse adipocytes [158]. Further, hyperinsulinemia due to increased HBP-mediated glycation resulted in dysregulation of F-actin which is essential for GLUT4 translocation [158, 160]. Of note, PST inhibited insulin-stimulated GLUT4 translocation in rat adipocytes [161] while CST treatment enabled GLUT4 translocation to the plasma membrane in rat cardiomyocytes by inducing Akt phosphorylation [162].
Common regulatory factors governing the expression of CHGA and genes involved in cholesterol homeostasis
This section focuses on common regulatory factors, viz. transcription factors (TF) and miRNAs, which govern the expression of both CHGA and genes involved in cholesterol homeostasis, in order to gain insights on the similar regulatory cues which could mediate their expression in different pathophysiological states.
Transcription factors
Many transcription factors that regulate CHGA expression under basal and pathophysiological conditions, viz., CREB, Sp1, Egr1, Olf/EBF, LEF1, COUP-II-TF, c-Rel [163–175]; are observed to modulate the expression of several cholesterol homeostatic genes, and the major transcriptional regulators of cholesterol homeostasis, viz., SREBPs, LXRs and PPARs, as outlined in Table Table44.
Table 4
CHGA-regulatory TF | Role in cholesterol homoeostasis |
---|---|
CREB | Co-activator to regulate the transcription of HMGCR [302] Perturbation of cholesterol homeostatic genes were observed in the brain of CREB/CREM mutant mice [303] Sterol-independent regulatory element (SIRE) in LDLR promoter which exhibited interactions with CREB [304] Co-regulates SREBP expression in transcriptional activation of major cholesterol homeostatic genes along with Sp1 [302, 305, 306] Modulates hepatic lipid homeostasis by repressing PPARγ [307] |
Egr1 | Mediates transcriptional activation of LDLR promoter [308] Egr1−/− mice exhibit diminished serum cholesterol levels and was observed to regulate several cholesterol biosynthetic genes in the liver [309] Negative regulation of SREBP-1a transcriptional activation [310] |
Sp1 | Upregulates human LCAT promoter activity [311] Modulates HMGCR, SQLE and sterol-mediated activation of Dhcr7 [312–314] Interplay among Sp1, SREBP, LXRα, NF-Y governs insulin-mediated hepatic activation of SREBP-1C [315] |
COUP-TFII | Important for activation of CYP7A1 [316, 317] Regulates LXR-mediated activation of genes involved in RCT in skeletal muscle cells [318] Synergistically activates murine lipoprotein lipase promoter with PPAR2/RXRα [319] |
In order to understand the intersecting transcriptional regulatory factors in CHGA and cholesterol homeostasis, common motifs in the promoter regions of CHGA and the genes involved in various aspects of cholesterol homeostasis, such as biosynthesis, trafficking, storage and metabolism, were identified using the MEME Suite [176]. The putative transcription factors binding to these motifs were identified using in silico tools such as JASPAR [177], LASAGNA [178], PROMO [179] and P-Match [180]. Table Table55 lists the common motifs, their consensus sequences and the putative TFs that bind to them. In addition, Pscan [181] was used to identify the over-represented TF binding motifs in the promoters of CHGA and cholesterol homeostatic genes from the JASPAR and Transfac databases. A few of these TFs were also found to be harbored in cholesterol quantitative trait loci (QTLs) in the rat genome (Fig. 6), indicative of their roles in the maintenance of cholesterol levels. The interplay of these transcription factors indicates that similar interactions could also play significant roles in mediating the expression of CHGA and cholesterol homeostatic genes.
Table 5
Motif (MEME) | Consensus sequence | Putative TFs |
---|---|---|
AATCCCAGCACTTTG | Elk-1, Evi-1, STATx, STAT3, Nkx3-1, Nkx3-2, MZF1, ZEB1, c-Ets, GATA2 | |
AGGCTGGAGTGCAGT | ZID, NFIC, ZNF354C, TEAD1 |
Enriched TF-binding sites (Pscan) | |||
---|---|---|---|
TF | Z-score | P-value | Bonferroni adjusted P- value |
COUP-TF | 4.3893 | 5.45E-06 | 0.0015 |
HNF-4 | 4.20739 | 1.24E-05 | 0.0034 |
HNF-4alpha1 | 3.93519 | 4.02E-05 | 0.0113 |
Nr2f6 | 4.50711 | 3.16E-06 | 0.0018 |
NFYA | 4.42712 | 4.64E-06 | 0.0026 |
NFYB | 3.99162 | 3.20E-05 | 0.0185 |
Rxra | 3.97277 | 3.43E-05 | 0.0198 |
RXRB | 3.80249 | 6.94E-05 | 0.0401 |
*Motifs were identified in the promoters of CHGA and the genes involved in cholesterol homeostasis using the MEME program [176] and the putative transcription factors (TFs) binding to these motifs were predicted by four in silico tools (JASPAR [177], LASAGNA [178], PROMO [179] and P-Match [180]. TFs predicted by at least two tools are listed. In addition, enriched TF-binding sites in the promoters of CHGA and cholesterol homeostatic genes were analyzed using Pscan [181]. The Z-score, unadjusted P-values and Bonferroni adjusted P-values are shown. The genes involved in different aspects of cholesterol homeostasis in the KEGG and Reactome databases were considered
miRNAs
Very little is known about the post-transcriptional regulation of CHGA. A polymorphism in the 3’-UTR of rat Chga (G+174 T; G in normotensive WKY and T in the hypertensive SHR) altered the free energy for the binding of miR-22 to the Chga 3’-UTR, that is, miR-22 reduced the luciferase signal of the SHR 3’-UTR construct more than WKY. Administration of miR-22 antagomir in the SHR resulted in a decline in the blood pressure [182]. Similarly, the C+87 T in the human CHGA 3’-UTR altered the affinity for miR-107, which was evident by greater inhibition of CHGA expression in constructs harboring the T allele as compared to the C allele. Moreover, miR-107 antagomir diminished blood pressure in mice [183].
Do these miRNAs have any role in cholesterol homeostasis? It was observed that cholesterol levels were higher in wild-type mice fed with high-fat (HF) diet than in miR-22 KO mice following the same diet, suggesting that miR-22 is linked to dyslipidemia [184]. QTL data from the Rat Genome Database showed that the Mir22 gene is localized in the rat serum cholesterol QTL Scl47 (LOD score: 3.6) in chromosome 10, suggesting key roles for miR-22 in cholesterol homeostasis.
In order to identify the targets of miR-22 and miR-107 in cholesterol homeostasis, validated interactions of these two miRNAs with the genes involved in cholesterol homeostasis were mined from Mirpath [185] and miRWalk 2.0 [186], and indeed, several genes involved in the various aspects of cholesterol homeostasis appear to be under the regulation of miR-22 and miR-107 (Fig. 7A). Interestingly, in addition to these genes, the major regulators of cholesterol homeostasis, INSIG1, SCAP and PPARA seem to be modulated by these two miRNAs. Moreover, in silico prediction analysis of the miR-22 and miR-107 targets using the miRWalk 2.0 and RNAhybrid [187] tools reveal that LRP2, MVD and NSDHL are putative targets of these miRNAs, further underscoring the role of these miRNAs in cholesterol homeostasis.
There is a vast body of literature which catalogue the involvement of miRNAs in cholesterol homeostasis. Figure 7B depicts the putative miRNAs involved in the maintenance of cholesterol homeostasis which could regulate CHGA expression at the post-transcriptional level. miR-148a modulated LDLR expression and cholesterol uptake in the liver. Moreover, elevated HDL-c levels were observed in mice treated with LNA (Locked Nucleic Acid) miR-148a, and a concomitant decrease in ABCA1 hepatic protein levels, implicating the role of miR-148a in mediating hepatic LDL-c clearance [188]. miR-10b expression was upregulated in the arterial walls with advanced atherosclerotic plaques in ApoE−/−mice, and downregulated macrophage ABCA1 expression in these mice. Moreover, free-cholesterol induced apoptotic macrophages (FC-AM) induced miR-10b expression [189]. A major player in the maintenance of cholesterol homeostasis, miR-33a, is encoded within the SREBF2 gene, and its expression is modulated by dietary cholesterol. The pathophysiological significance of miR-33a is further substantiated by its dysregulation in rodent models of hypercholesterolemia / atherosclerosis. Moreover, its role in regulating key genes involved in cholesterol transport and HDL metabolism is well-characterized [190–193]. Notably, there is evidence of miR-33a targeting bile acid metabolism and secretion [194, 195] in addition to key metabolic players like AMPK [196], a significant modulator of cholesterol metabolism. An investigation of the aforementioned miRNAs in the concerted regulation of CHGA and cholesterol homeostasis would broaden our understanding of the post-transcriptional regulatory mechanisms of these genes for instance in conditions of perturbed cholesterol homeostasis.
Post-translational modifications
It has been well-established that post-translational modifications including O-glycosylation and phosphorylation play an important regulatory function in CHGA processing and in the potential interactions of the cleaved peptides [197]. For instance, glycans made of N-acetyl-galactosamine, galactose and sialic acid have been found on residues Thr163, Thr165, and Thr233 in the core domain of hCHGA present in urine samples [198]. On the same line, hCHGA phosphorylation in the middle and C-terminal part occur on residues Ser200, Ser253, and Ser315 [198]. It is thus likely that the O-glycosylation and phosphorylation of CHGA may affect the interaction of CHGA and fragments with cholesterol, but this obviously remains to be established experimentally.
Regulatory mechanisms of CST/PST/VS on cholesterol homeostasis: intersecting signaling cascades
Although several lines of evidence suggest that CHGA and its derived peptides exert their influence on cholesterol levels, the signaling pathway(s) through which this regulation is accomplished remains obscure. In this section, we discuss a few signaling cascades common to CHGA action and cholesterol homeostasis regulation which merit consideration for further studies (Fig. 8).
AMPK pathway
Adenosine monophosphate kinase (AMPK) is a cellular energy sensor which is activated by phosphorylation in response to altered AMP: ATP ratio, i.e., where ATP consumption exceeds ATP production, in conditions such as hypoxia, ischemia, heat shock, nutritional stress, etc. AMPK plays crucial roles in regulating glucose and lipid metabolism by direct action on metabolic enzymes or indirect transcriptional control of key regulators of these pathways. AMPK inactivates anabolic pathways like cholesterol, fatty acid, triglyceride, and protein synthesis; whereas it upregulates ATP-generating catabolic pathways like glycolysis and fatty acid oxidation [199–202]. Phosphorylation of HMGCR by AMPK (at Ser 871 in rodents and Ser 872 in humans) inhibits its activity with a concomitant inhibition of cholesterol biosynthesis [203–205].
Leptin signaling
Interestingly, both CST and PST are known to modulate AMPK signaling. CST treatment of adipose tissue explants from Chga knockout (KO) mice stimulated phosphorylation of AMPK [206]. Moreover, leptin levels were higher and leptin signaling was depressed in Chga KO mice. CST treatment lowered plasma leptin levels and resensitized the leptin receptor [206]. However, CST enhanced AMPK signaling in concert with leptin in DIO mice and enhanced leptin-mediated AMPK phosphorylation in ob/ob mice, suggesting that other CHGA-derived peptides may downregulate CST-mediated AMPK signaling [206]. Leptin treatment lowered the LDL/HDL1 fraction in ob/ob mice and altered hepatic expression of a few genes involved in cholesterol homeostasis: downregulation of ACAT1 and HMGCR; upregulation of APOA1 [207]. SCAP, HMGCR, PLC, MAPK and PI3K were involved in the leptin-induced increase in cholesterol synthesis in control and hypercholesterolemic monocytes [208]. Injection of leptin into intestinal loops increased intestinal cholesterol levels and decreased plasma HDL cholesterol levels. Acute leptin treatment lowered plasma cholesterol levels in chow-fed mice while chronic leptin treatment diminished plasma cholesterol and elevated intestinal and hepatic cholesterol levels [209].
GRP78
Muscle and white adipose tissues of Chga KO-DIO mice exhibited reduced AMPK phosphorylation, but not the liver [210]. PST supplementation did not alter the pattern of AMPK phosphorylation in any of these tissues, suggesting that PST does not influence AMPK signaling. However, PSTi8 stimulated AMPK phosphorylation and augmented the ATP/AMP ratio in dexamethasone-treated hepatic cells and liver tissues of mice treated with dexamethasone (DEX). This stimulatory effect was observed to be mediated by enhancing the ATPase activity of GRP78, another known stimulator of AMPK signaling [211, 212]. Hepatic GRP78 expression was higher in Chga-KO mice as compared to the wild-type. Moreover, PST was observed to interact with GRP78 and inhibit its ATPase activity [213]; whereas the PST inhibitor, PSTi8 competed with PST to bind to GRP78 and enhanced its ATPase activity [214]. Glucose-related protein 78 (GRP78), also known as BiP/HSP5a, is a major player in the unfolded protein response (UPR) in response to ER stress [215]. ER stress enhanced SREBP-2 cleavage and expression of cholesterol biosynthetic genes [216, 217]. Of note, GRP78 ameliorates hepatic expression of ABCA1, ABCG5, ABCG8 and SR-BI and biliary cholesterol levels in db/db mice [218]. Of note, knocking down the CST precursor (CHGA) elevated GRP78 levels while CST treatment reduced GRP78 protein levels in cardiomyocytes during ER stress in ischemia/reperfusion injury [219].
These studies demonstrate the crosstalk among AMPK, GRP78 and leptin signaling in cholesterol homeostasis, and the impact of CST and PST on AMPK signaling via different pathways. An interesting line of study in this direction would be to validate the modulation of HMGCR expression/cholesterol homeostasis by CST/PST via AMPK in a leptin/GRP78 dependent/independent manner.
PPARα and PPAR pathway
Peroxisome proliferator-activated receptor (PPAR) belongs to the nuclear receptor superfamily and plays salient roles in lipid metabolism and glucose homeostasis. These transcription factors heterodimerize with the retinoic acid receptor and regulate expression of several target genes. There are three forms of PPAR: PPARα, PPARδ and PPAR [220, 221].
CST treatment elevated Pparα expression in the liver of Chga-KO mice while Ppar levels were unaltered [206]. Although PST supplementation did not alter the expression levels of Pparα in Chga KO-DIO mice, PST treatment depressed insulin-mediated elevations in Ppar and Pparα expression in cultured adipocytes, which was reversed by PSTi8 treatment [214]. Moreover, administration of PSTi8 reduced Ppar and Pparα expression in DEX-treated mice [210, 211], thus modulating lipid and glucose homeostasis. It is evident from these reports that knowledge on the impact of CST/PST on PPAR and PPARα remains incomplete and given the involvement of these factors in the maintenance of cholesterol homeostasis, more studies are required to delineate the mechanisms of CST/PST regulation of cholesterol metabolism via these transcription factors.
β-adrenergic signaling
The effects of catecholamines on fat and lipid metabolism are well known [222]. Although the exact molecular mechanisms of adrenergic signaling in cholesterol homeostasis are poorly understood, speculations are rife in literature. The β-adrenergic receptor agonist isoproterenol inhibited sterol synthesis while the β-adrenergic antagonist propranolol abolished the epinephrine-induced suppression of sterol synthesis in human mononuclear leukocytes [223, 224]. This effect seemed to be mediated by the β2 isoform. In addition, β2-adrenergic receptor polymorphisms exhibited associations with LDL cholesterol levels [225]. Several other studies have shown the beneficial action of β2 agonism by lowering LDL-c levels. These effects could be explained by the enhancing action of β2 receptors on intracellular cAMP levels [226]. Of note, cAMP is a regulator of LDLR and SR-BI expression [227, 228], promoted cholesterol efflux in macrophages [229], by modulating ABCA1 expression [230], stimulated cholesteryl ester hydrolysis and clearance [231]. Another speculation is the decline of LCAT activity by β-adrenergic receptor agonists, which could account for the increase in HDL content upon these treatments [232].
The role of CHGA-derived peptides in the modulation of β-adrenergic signaling is well-documented. CST abrogated the response to adrenergic stimulation in the frog heart [233] and rodent papillary muscles [234]. Notably, the G364S variant peptide exhibited diminished potency to generate NO levels and inhibit isoproterenol-stimulated activation of ERK due to altered interactions with the ADRB2 when compared with the wild-type [235]. The treatment of primary neonatal cardiomyocytes with β2 antagonist abolished the CST-induced cardioprotective effects [236]. Vasostatin was also observed to block isoproterenol-stimulated positive inotropy in the rodent heart [237, 238]. Recently, our group reported differential catecholamine secretion from neuronal cell lines expressing either PST or its naturally occurring variant PST-Ser [137]. Serpinin and pGlu-serpinin act as myocardial β1-adrenergic like agonists, and interestingly also upregulate cAMP levels in cardiac tissue extracts [239]. As the metabolic roles of serpinin are poorly understood, its putative action on cholesterol homeostasis via the adrenergic pathway would make a novel line of study.
Nicotinic acetylcholine receptor pathway
The nicotinic acetylcholine receptor (nAChR) harbors a high affinity cholesterol-binding CARC motif, a well-established domain for several membrane receptors [240]. Cholesterol plays an important role in the homeostasis of nAChR levels, which not just affects its structure and dynamics, but also dictates its endocytic pathway upon internalization [241]. With cholesterol being a chief regulator of nAChR, the receptor itself regulates cholesterol homeostasis and several metabolic pathways.
Treatment of human placental villus trophoblastic cells with nicotine reduced the expression of cholesterol regulatory genes ABCA1, ABCG1, SR-B1, LXRα, LXRβ by stimulating nAChR [242]. Conversely, in another study, nicotine stimulated the expression of ABCA1 in THP-1 cells via nAChR [243]. As CST is a potent endogenous inhibitor of the nicotinic acetylcholine receptor [244], it may be possible that it relays its effects on cholesterol homeostasis via nAChR pathway. However, it is essential to delineate the molecular mechanisms to depict a conclusion, considering the tissue-specific effects of nAChR. In ApoE−/− mice, nicotine treatment activated α7nAChR by exerting its effects via JAK2/STAT3 and Akt signaling pathways. This resulted in mast cell activation and release of proinflammatory cytokines, thereby leading to atherogenesis in ApoE−/− mice [245]. As previously mentioned, PST downregulates AKT phosphorylation, while CST upregulates Stat3 and AKT phosphorylation which converges these peptides to an interesting mechanism in atherogenesis via nAChR. In adipocytes, nicotine activated AMPK via nACHR7α by mediating ROS levels [246].
Akt signaling
An important signaling cascade involved in cell growth, survival and metabolism is the PI3K-Akt pathway, which has implications in glucose homeostasis, insulin signaling, lipid metabolism, etc. With respect to cholesterol homeostasis, Akt was found to negatively influence ABCA1 function by suppressing cholesterol efflux to ApoA-I via mTORC1 [247]. Akt was also observed to be involved in the ER-Golgi transport of the SREBP/SCAP, an important cholesterol regulator [248]. The association of Akt phosphorylation with ischemia/reperfusion (IR) injury was observed to be cardioprotective [249]. Another line of evidence emerges from the unfolded protein response (UPR) which gets activated during I/R injury. UPR elevates the expression of ER chaperone Grp78 which protects cardiomyocytes from ROS accumulation by Akt activation [250].
The association of CHGA-derived peptides with Akt pathway is well established. CST stimulates Akt phosphorylation in rat cardiomyocytes against I/R injury [251]. It also reduces apoptosis of cardiomyocytes induced by oxidative stress during I/R injury by activating β-adrenergic receptor and increasing Akt phosphorylation along the reperfusion injury salvage kinase (RISK) pathway [236]. Pre-treatment of rat cardiomyocytes with CST was found to stimulate muscarinic acetylcholine M2 receptor and thus activate PI3K/Akt pathway. This resulted in reducing ER-stress-induced cell apoptosis under I/R conditions [219].
mTORC1 pathway
The mammalian target receptor complex 1 (mTORC1) can activate SREBPs at several steps like processing, trafficking, and transcription [252]. Insulin signaling activates mTOR (mTORC1) via Akt phosphorylation. The activated mTOR phosphorylates USP20, a deubiquitilase, which facilitates its interaction with the E3 ligase gp78, in order to stabilize HMGCR for cholesterol biosynthesis [253]. Thus, Akt phosphorylation becomes necessary for the mTOR pathway to stimulate HMGCR. Interestingly, PST treatment in Chga KO mice reduced Akt phosphorylation in hepatic tissues in the presence of insulin. Hence, PST might deregulate this mTOR pathway with its anti-insulin-like effects [30]. Opposing to this, CST’s activity on enhanced Akt phosphorylation, as cited in the previous section might probably restore and enhance HMGCR activity.
Inflammation
Inflammation is the protective response of our immune system against alarming stimuli like pathogens, oxidative stress, injury, toxins, etc. If this process gets dysregulated, it leads to a sustained hyperactive state of immune cells resulting in chronic inflammation, a common factor in several metabolic syndromes. The link between inflammation with cholesterol metabolism and homeostasis can be studied in metabolic disorders like dyslipidemia, atherosclerosis, diabetes, obesity, etc.
Acute phase responses
Inflammation triggers a series of local and distant systemic changes to maintain homeostasis, which comprise acute phase responses (APR). During this, the levels of inflammatory cytokines (TNF-α, IL-1β, etc.) increase and changes in lipid metabolism are observed. These include increased serum triglyceride (TG) levels, decreased HDL cholesterol levels, changes in LDL oxidation, etc. [254]. In Hep 3B cells, TNF-α and IL-1 were found to decrease the levels of LXRα, RXRα, PPARα and PPARγ, including its target SREBP1c and CPT1α [255]. These genes are important regulators of several cholesterol homeostasis genes like HMGCR, ABCG5, ABCG8, ABCA1, CETP, cholesterol 7α hydroxylase, SR-BI, etc. [254, 256]. The downregulation of cholesterol transporters under inflammation could possibly explain the higher levels of intracellular cholesterol and lower serum ApoA-I and HDL. The role of CHGA and its peptides in lipid metabolism could be interpreted from the fact that the Chga KO mice showed augmented levels of both PPARγ and SREBP1c, whereas these levels were conversely reduced upon PST treatment. The same trend was observed with SREBP1c protein levels which were elevated in Chga KO mice but significantly reduced in the presence of PST [30].
Atherosclerosis
Atherosclerosis is associated with upregulation of adhesion molecules and recruitment of immune cells which takes up oxidized and modified lipids to form foam cells resulting in plaque formation. In a recent study, the vasoprotective roles of CST in atherosclerosis were studied. In human umbilical vein endothelial cells (HUVECs), CST attenuated vascular cell adhesion molecule-1 (VCAM1), intercellular adhesion molecule-1 (ICAM1) and TNFα, which were upregulated by LPS. There was a significant reduction in oxidized LDL-induced foam cell formation associated with ABCA1 upregulation in macrophages and ACAT1 downregulation in human macrophages. CST administration in ApoE−/− mice retarded the atherosclerotic lesions by 40% in the entire surface area of aorta with reduced macrophage/monocyte infiltration [139]. PCSK9 inhibition was found to reduce the risk of venous thromboembolism. It was also found that PCSK9 gene silencing suppressed atherosclerosis by decreasing vascular inflammation via TLR4/NFKB pathway in hyperlipidemia-induced ApoE−/− mice [257].
Toll-like receptor-mediated signaling
Toll-like receptors (TLRs) are an important class of pattern recognition receptors that are highly expressed in monocytes. Both TLR2 and TLR4 were found to be elevated in the endothelium of atherosclerotic lesions [258] and are also implicated as potent inflammatory inducers in atherothrombotic diseases [259]. The potential role of CST as an anti-inflammatory molecule in atherothrombotic conditions could be deduced from a study in which CST treatment of human pulmonary artery endothelial cells (HPAECs) showed attenuated thrombin-induced endothelial inflammation by inhibiting TLR4-p38 signaling [260]. Interestingly, there exists a crosstalk between TLRs and LXR. Ligands for TLR3 and TLR4 have been reported to inhibit transcriptional activity of LXR via interferon regulatory factor 3 (IRF3) [261]. Thus, CST could rescue anti-inflammatory role of LXR via ABCA1 [262] by inhibiting TLR4 signaling.
Insulin signaling
Cholesterol homeostasis and cholesterol efflux mediated by both ABCA1 and ABCG1 play critical roles in pancreatic β-cell function. Conditional knockout of both Abca1 and Abcg1 in mice resulted in impaired glucose-stimulated insulin secretion, glucose intolerance, increased adipose tissue and systemic inflammation [263]. One of the consequences of metabolic inflammation is impaired Akt signaling leading to insulin resistance. As a result, downstream targets of Akt, FOXO, SREBP, GLUT4, etc., are also downregulated, which play major roles in glucose and lipid metabolism [264]. Interestingly, PST inhibits insulin-stimulated GLUT4 translocation in rat adipocytes [161]. This effect of PST is similar to the inflammatory cytokine TNFα which downregulates GLUT4 in mouse adipocytes [265]. In Chga KO mice, PST suppressed insulin signaling by downregulating Akt phosphorylation [266].
Involvement of iNos
The mRNA levels of the pro-inflammatory gene iNos were downregulated in the adipose tissue of Chga KO-DIO mice compared to WT-DIO, and supplementation of PST was observed to upregulate iNos levels [206]. CST downregulated Nos2 levels in monocytes and macrophages which infiltrated the liver in DIO mice [267]. Another study reported elevated levels of total cholesterol in iNOS-KO mice, in addition to increased hepatic cholesterol levels. Moreover, in these mice, genes involved in cholesterol homeostasis (hepatic HMGCR, SREBP-2, PPARα, PPARγ, LDLR, ABCG5, ABCG8) were upregulated compared to the wild-type mice [268]. These reports suggest the involvement of iNOS in the effects of PST/CST on cholesterol metabolism.
The renin–angiotensin–aldosterone system
The putative effects of CHGA-derived peptides on RAAS have been discussed above. Remarkably, this pathway could also potentially act as a link between CHGA and cholesterol homeostasis, especially in the context of dysregulated cholesterol states. Impairment of the RAAS pathway has been implicated in atherosclerosis, ischemic disease, and congestive heart failure. It is fascinating to note that hypercholesterolemia regulates several components of the RAAS pathway [269–272]. However, the RAAS pathway has also been observed to influence cholesterol homeostasis. AngII injection increased atherosclerotic lesion size in ApoE−/− mice, increased cholesterol biosynthesis in macrophages, which was reduced by treatment with ACE inhibitor or AT1 receptor antagonist or HMGCR inhibitor. AngII also elevated HMGCR expression [273]. AngII enhanced LDL oxidation and lipid peroxidation in macrophages [274]. AngII was found to bind to LDL and modified its oxidative and aggregation properties in addition to exhibiting enhanced uptake by the macrophage scavenger receptor [275]. AngII injection increased ox-LDL uptake by mouse peritoneal macrophages and human coronary artery endothelial cells, which could contribute to accelerated atherosclerosis [276, 277]. It would make an interesting study to see if the anti-atherogenic effects of CST/VS-I/VS-II are mediated by a shift to the protective axis of the RAAS pathway.
Conclusions and future perspectives
The salient roles of cholesterol in diverse aspects of physiology and metabolism makes it an indispensable molecule for normal body function. The wide spectrum of disorders arising as a consequence of perturbed cholesterol homeostasis and their high incidences in the general population calls for identification of novel regulators which can modulate the key aspects of cholesterol homeostasis. The pleiotropic nature of CHGA and its derived peptides makes them attractive candidates to investigate their potential as novel regulators of cholesterol homeostasis. Indeed, multiple studies suggest the involvement of CHGA/CST/PST/VS-I/VS-II in governing cholesterol levels in pathological conditions. Furthermore, genome-wide association studies also suggest the putative roles for the serpinin peptides in this regard, and the identification of putative cholesterol binding motifs in the CHGA protein further strengthens the CHGA-cholesterol link. The observation of common transcription factors and miRNAs suggest similar regulatory mechanisms in play which could dictate their expression in health and disease. The association between CHGA and cholesterol homeostasis is further compounded by several intersecting pathways, through which CHGA-derived peptides could potentially mediate cholesterol levels. Of note, there is substantial evidence supporting the involvement of the AMPK and the RAAS pathways in connecting CHGA and cholesterol homeostasis, which warrants further investigation. Studies focusing on the direct/indirect effects of CHGA and its derived peptides on cholesterol biosynthesis and/or on the expression of associated genes will also provide valuable insights into devising check points of cholesterol levels in pathophysiological conditions. This opens up exciting avenues for future studies, which could bridge the gap in our understanding of the dynamic field of cholesterol research and facilitate development of innovative therapeutic strategies to combat cholesterol deregulated conditions.
Acknowledgements
We acknowledge all the researchers who contributed to the areas of Chromogranin A and cholesterol research. All studies could not be cited due to space constraints. DRI and JV would like to thank Dr. Abrar Ali Khan (IIT Madras) for his valuable inputs.
Author contributions
DRI and JV performed literature search, data analysis and drafted the manuscript. ET and NV provided critical inputs and edited the manuscript. NRM conceptualized and edited the manuscript. All authors approved the final version of the manuscript.
Funding
NRM received research grants from the Department of Biotechnology (BT/PR25796/GET/119/98/2017; BT/PR23017/MED/30/1838/2017), Department of Science and Technology (SR/SO/HS-084/2013A) and Council of Scientific and Industrial Research (37/1564/12-EMR-II), Government of India. DRI and JV are thankful to Ministry of Human Resource Development (MHRD), Government of India for research fellowships. NV acknowledges support from the Agence Nationale pour la Recherche (ANR-19-CE44-0019 and ANR-22-CE44-0029).
Availability of data and material
The datasets analyzed during the current study are available from the corresponding author on reasonable request.
Declarations
The authors have no relevant financial or non-financial interests to disclose.
Not applicable.
Not applicable.
Footnotes
The original online version of this article was revised: to update the entries in Table 1, 5th column heading is incorrect as “lipoproteins” and should have been reads as “apolipoproteins”. In entry Chylomicrons under the Pathophysiological implications column the term dietarytriglycerides should read as “dietary triglycerides”. In last entry LPA the value for size was incorrect as 30 and should have been reads as ~30.
Publisher's Note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Dhanya R. Iyer and Janani Venkatraman have contributed equally to this article.
Change history
10/12/2023
Error in the Table 1 has been updated.
Change history
11/6/2023
A Correction to this paper has been published: 10.1007/s00018-023-04990-7
Contributor Information
Nicolas Vitale, Email: rf.artsinu@nelativ.
Nitish R. Mahapatra, Email: ni.ca.mtii@artapahamn.
References
Articles from Cellular and Molecular Life Sciences: CMLS are provided here courtesy of Springer
Full text links
Read article at publisher's site: https://doi.org/10.1007/s00018-023-04908-3
Read article for free, from open access legal sources, via Unpaywall: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC11072126
Citations & impact
This article has not been cited yet.
Impact metrics
Alternative metrics
Discover the attention surrounding your research
https://www.altmetric.com/details/153432038
Data
Data behind the article
This data has been text mined from the article, or deposited into data resources.
SNPs (Showing 6 of 6)
- (2 citations) dbSNP - rs9658638
- (2 citations) dbSNP - rs9658629
- (2 citations) dbSNP - rs9658656
- (1 citation) dbSNP - rs9658664
- (1 citation) dbSNP - rs4900163
- (1 citation) dbSNP - rs9658655
Show less
Similar Articles
To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.
Global metabolic consequences of the chromogranin A-null model of hypertension: transcriptomic detection, pathway identification, and experimental verification.
Physiol Genomics, 40(3):195-207, 01 Dec 2009
Cited by: 13 articles | PMID: 19952279 | PMCID: PMC2825767
Chromogranin A-derived peptides pancreastatin and catestatin: emerging therapeutic target for diabetes.
Amino Acids, 55(5):549-561, 13 Mar 2023
Cited by: 4 articles | PMID: 36914766
Review
Chromogranin-A and its derived peptides and their pharmacological effects during intestinal inflammation.
Biochem Pharmacol, 152:315-326, 13 Apr 2018
Cited by: 20 articles | PMID: 29656116
Review
Granin-derived peptides.
Prog Neurobiol, 154:37-61, 22 Apr 2017
Cited by: 32 articles | PMID: 28442394
Review
Funding
Funders who supported this work.
Agence Nationale pour la Recherche (1)
Grant ID: ANR-19-CE44-0019