Skip to main content
British Journal of Pharmacology logoLink to British Journal of Pharmacology
. 2015 Jan 13;172(8):2114–2125. doi: 10.1111/bph.13009

Caveolins in cardioprotection – translatability and mechanisms

Jan M Schilling 1,2, David M Roth 1,2, Hemal H Patel 1,2,
PMCID: PMC4386985  PMID: 25377989

Abstract

Translation of preclinical treatments for ischaemia-reperfusion injury into clinical therapies has been limited by a number of factors. This review will focus on a single mode of cardiac protection related to a membrane scaffolding protein, caveolin, which regulates protective signalling as well as myocyte ultrastructure in the setting of ischaemic stress. Factors that have limited the clinical translation of protection will be considered specifically in terms of signalling and structural defects. The potential of caveolin to overcome barriers to protection with the ultimate hope of clinical translation will be discussed.

Tables of Links

TARGETS
Enzymesa Transportersb
Akt ANT, adenine nucleotide transporter
eNOS GLUT4, glucose transporter 4
ERK Catalytic receptorc
F0F1 ATP synthase Insulin receptor
Glycogen synthase kinase-3β Ion channelsd
PI3K KATP channels, Kir6.2
PKC GPCRe
Src kinase Adenosine A1 receptor
LIGANDS
Resveratrol
TNFα

These Tables list key protein targets and ligands in this article which are hyperlinked to corresponding entries in http://www.guidetopharmacology.org, the common portal for data from the IUPHAR/BPS Guide to PHARMACOLOGY (Pawson et al., 2014) and are permanently archived in the Concise Guide to PHARMACOLOGY 2013/14 (a,b,c,d,cAlexander et al., 2013a,b,c,d,e,,,,).

Introduction

Myocardial infarction is a major cause of death in the United States. One of the most beneficial experimental interventions to produce cardiac protection is termed ischaemic preconditioning (IPC) where sublethal ischaemia protects from subsequent lethal injury (Murry et al., 1986). Defining molecular events regulating IPC has potential implications for therapeutic strategies for ischaemic heart disease and broader implications for cardiac hypertrophy, heart failure, diseases that secondarily lead to cardiac dysfunction (e.g. diabetes, hypertension) and other conditions where the balance between cell death and survival is critical. Despite a considerable amount of data describing preconditioning signalling, a precise pharmacological target and therefore a specific therapeutic agent remains elusive. This review will focus on limitations of protective interventions in the heart and a novel observation that caveolae, lipid-rich membrane microdomains enriched in caveolin-scaffolding proteins, may be a means to break down these barriers to the translation of experimental cardiac protection to clinical practice.

What is the state of understanding of mechanisms leading to IPC?

Following the seminal discovery of IPC by Murry, Jennings and Reimer in 1986 (Murry et al., 1986), two parallel ideas have developed over the last ∼25 years to account for IPC and provide a path for therapeutic development. One is the signalling hypothesis which proposes molecules converge to change the biochemistry and metabolism of the cell to affect protection and the other is the structural hypothesis in which IPC provides a physical, structural resiliency to the cardiac myocyte. The ratio of published papers is skewed 100-fold in favour of investigations focused on the signalling hypothesis.

Many studies have described IPC as a promiscuous stimulus that involves the initiation of many shared and interconnected signalling pathways that ultimately converge upon the mitochondria to cause cell protection and survival (Hausenloy et al., 2005; Hausenloy and Yellon, 2006). The signal transduction pathways involve sequential triggers, mediators and end effects that culminate in modulation of mitochondrial function (Juhaszova et al., 2004; Hausenloy and Yellon, 2006). Three main mitochondrial phenomenon – (i) opening of mitochondrial ATP-sensitive potassium (KATP) channels; (ii) generation of a small burst of reactive oxygen species (ROS) and (iii) maintenance of the mitochondrial permeability transition pore (mPTP) – have all been linked to the cardiac protective effects of IPC (Auchampach et al., 1992; Gross and Auchampach, 1992; VandenHoek et al., 1998; 2000,; Becker et al., 1999; Yao et al., 1999; Pain et al., 2000).

Following their initial discovery of IPC, Murry et al. showed in 1990 that IPC delays ultrastructural myocardial damage (i.e. to membrane and mitochondria) during subsequent lethal ischaemia (Murry et al., 1990). This led to exploration of the ‘structural protective’ hypothesis where IPC was shown to preserve membrane integrity and mitochondrial structure (Moolman et al., 1995; Armstrong et al., 2001). Although limited in number, some proposed effectors of this structural preservation have been postulated: the mPTP (Fang et al., 2008), dystrophin and spectrin (Armstrong et al., 2001; Kido et al., 2004), ROS (Miyamae et al., 2002) and KATP channels (Geshi et al., 1998). This concept was expanded to include preservation of mitochondrial function and structure by IPC and postconditioning in heart and liver (Zhong et al., 2000; Giovanardi et al., 2009; Penna et al., 2009; Quarrie et al., 2012). What remained unknown was whether IPC prevented injury by directly modulating structure or if signalling events initiated by IPC activated membrane repair processes that helped to maintain membrane integrity during periods of stress. The latter possibility would suggest that the molecular signalling and structural protection afforded by IPC could be linked through some common factor. Preservation of cellular ultrastructure is a tightly regulated process that has at its core molecular signalling leading to membrane repair and dynamics in the sarcolemma and mitochondria (Donaldson et al., 2009; Hausenloy and Yellon, 2010; Cao et al., 2011; Gottlieb and Gustafsson, 2011; Mousley et al., 2012). Finding a molecule that bridges signalling to preservation of cellular ultrastructure may provide a novel therapeutic target to protect against ischaemia-reperfusion injury.

Caveolae and caveolins: a bridge that provides a unifying feature to cardiac protection

Caveolae, or ‘little caves’, are cholesterol and sphingolipid-enriched invaginations of the plasma membrane (Palade, 1953) and are considered to be a subset of lipid rafts (Pike, 2003). Caveolins, the structural proteins essential for caveolae formation, are present in three isoforms (Chun et al., 1994; Parton et al., 1997). Caveolins have a 20 amino acid scaffolding domain (caveolin-scaffolding domain, CSD) that anchors and regulates proteins (Sargiacomo et al., 1995; Feron and Balligand, 2006). Canonically, caveolin (Cav)-1 and Cav-2 were thought to be expressed in many cell types, while Cav-3 was found primarily in striated (skeletal and cardiac) muscle and certain smooth muscle cells (Song et al., 1996). Such concepts are currently being challenged with the identification and description of the structural importance for Cav-3 in non-muscle cells (Niesman et al., 2013) and the identification of a functional consequence of caveolin localized to a variety of cellular compartments (Head and Insel, 2007; Fridolfsson et al., 2014).

The expression of caveolin isoforms in the heart has been hotly debated. It is accepted that cardiac myocytes express Cav-3, the muscle-specific isoform (Song et al., 1996; Tang et al., 1996), and that other cell types in the heart express Cav-1 and Cav-2. Recent studies have provided evidence for the existence of and a signalling role for Cav-1 in cardiac myocytes with respect to ischaemia-reperfusion injury and maintenance of cardiac gap junctions (Patel et al., 2007; Yang et al., 2014). It was previously thought that Cav-1 and Cav-2 form hetero-oligomers and Cav-3 forms homo-oligomers (Tang et al., 1996; Scherer et al., 1997) but more recent data indicate co-expression and interaction of Cav-2 and Cav-3 in neonatal cardiac myocytes (Rybin et al., 2003) and interaction of Cav-1, Cav-2 and Cav-3 in adult cardiac myocytes (Hagiwara et al., 2002; Head et al., 2006). Other findings show that cell type-specific environments may regulate the interaction of caveolins. Thus, in fibroblasts, Cav-1 and Cav-2, but not Cav-3, interact, whereas in myoblasts, all three caveolin isoforms co-immunoprecipitate (Capozza et al., 2005) and in microglia Cav-1 and Cav-3 have very distinct roles depending on the metabolic and structural state of the cell (Niesman et al., 2013). Studies of mice with knockout (KO) or transgenic overexpression of caveolins demonstrate that the expression of caveolin is both necessary and sufficient for the formation of caveolae. Recent studies have identified another protein, cavin, as a key component of caveolae although studies on the physiological and pathophysiological role of cavin in the heart are limited (Vinten et al., 2005; Hill et al., 2008; Liu and Pilch, 2008a).

Regulation of caveolae and caveolins is complex and may help explain the general role of this structural protein in regulation of a wide range of physiological cellular functions. Protein–protein interactions as a function of charge, size and/or steric factors may contribute to the localization of proteins within the caveolae (Yamabhai and Anderson, 2002; Nichols, 2003; Pike, 2005). In addition, being lipid-rich microdomains distinct from surrounding membranes, caveolae may facilitate lipid–protein interaction (Rothberg et al., 1992; Park et al., 2004). Lipid modification of proteins, in particular palmitoylation and myristoylation, contribute, for example, to the localization of G-protein signalling components in raft/caveolae domains (Ratajczak et al., 2003; Razzaq et al., 2004; Rodgers et al., 2005; Kim et al., 2006). As noted earlier, the CSD, a hydrophobic region in the cytoplasmic amino terminal tail that interacts with protein ‘partners’ through hydrophobic interactions, has been proposed as a critical region by which signalling proteins interact with caveolins (Chini and Parenti, 2004; Becher and McIlhinney, 2005), although this notion has been recently challenged (Collins et al., 2012). Finally, caveolins can undergo post-translational modification that may be critical to regulating not only signalling proteins, but also the response to pathophysiology. Caveolins contain three C-terminal cysteine residues that contain putative palmitoylation sites (Cys133, Cys144 and Cys156) (Dietzen et al., 1995). Cav-1 undergoes phosphorylation (at Tyr14) (Rothberg et al., 1992; Li et al., 1996) and Src-mediated phosphorylation alters the properties of Cav-1, including its interaction with extracellular matrix proteins (Grande-Garcia et al., 2007; Grande-Garcia and Del Pozo, 2008) and has other potential physiological functions (Patel et al., 2007; Yang et al., 2014). Although Cav-3 has a putative phosphorylation site, no studies have been published to identify this site or its functional significance. However, Cav-3 has been reported to be sumoylated which affects receptor desensitization (Fuhs and Insel, 2011). Thus, caveolae and caveolins appear to have a variety of ways to regulate cell function.

Cav-3 has been identified in the sarcolemmal membrane, transverse tubules (T-tubules), the I-band/A-band interface and localized with ryanodine receptors in myocytes (Ralston and Ploug, 1999; Scriven et al., 2005). Caveolins are involved in many cellular processes including vesicular transport, cholesterol and calcium homeostasis (Fujimoto et al., 1992; Fujimoto, 1993; Scriven et al., 2002; Jones et al., 2004; Peng et al., 2004), signal transduction (Lisanti et al., 1994; Steinberg and Brunton, 2001; Cohen et al., 2004; Williams and Lisanti, 2004) and have been recently detected in the mitochondria (Li et al., 2001; Fridolfsson et al., 2012). Caveolins function as chaperones and scaffolds, recruiting signalling molecules to caveolae to provide direct temporal and spatial regulation of signal transduction (Shaul and Anderson, 1998; Williams and Lisanti, 2004). Caveolins can inhibit activity of signalling proteins by interaction of the CSD with a caveolin binding motif present in many proteins found in caveolae including endothelial NOS (eNOS) and ERK1/2 (Engelman et al., 1998; Feron et al., 1998; Kamoun et al., 2006). Alternatively, caveolins can promote signalling via enhanced receptor-effector coupling or enhanced receptor affinity when caveolins are up-regulated or overexpressed (Feron and Balligand, 2006; Raikar et al., 2006). This has led to the concept of a ‘caveolar paradox’ in which caveolins may produce direct allosteric inhibition of molecules such as eNOS under basal conditions but facilitate increased signalling upon agonist stimulation through compartmentation (Feron and Kelly, 2001; Feron and Balligand, 2006).

An emerging concept suggests that signalling molecules exist as multiprotein complexes, ‘signalosomes’, continuously forming and dissociating under basal or stimulated conditions (Feron and Balligand, 2006). Caveolins are thought to play an integral role in the dynamics of these multiprotein complexes. Specifically, in regard to signalling molecules involved in cardiac protection, many GPCRs including opioid (Head et al., 2005) and adenosine receptors (Lasley et al., 2000) localize to caveolae and co-immunoprecipitate with caveolins. Additionally, many of the signalling molecules involved in cardiac protection, including the Gα subunit of heterotrimeric G-proteins, Src kinases, PI3K, eNOS, PKC isoforms and ERK are known to bind with the scaffolding domain of caveolin and be regulated by caveolin (Krajewska and Maslowska, 2004; Ballard-Croft et al., 2006). Caveolin is known to be a key component and activator of PI3K/Akt signalling (Fecchi et al., 2006), a pro-cell survival pathway that plays a significant role in preconditioning in the heart.

Initial evidence implicating a role for caveolin in cardiac protection was confirmed by infusion of the CSD peptide of Cav-1 into ischaemic/reperfused hearts which resulted in recovery of cardiac function (Young et al., 2001). It was later shown that ischaemia/reperfusion injury activates p42/44 and p38 MAPK, redistributes Cav-3 and down-regulates expression of Cav-1 (Ballard-Croft et al., 2006). The critical links between caveolar structure, caveolin protein and cardiac protection have emerged from a series of studies conducted by our group showing that physical disruption of caveolae negated protection in adult cardiac myocytes (Patel et al., 2006), loss of protection in Cav-1 and Cav-3 KO mice (Patel et al., 2007; Horikawa et al., 2008; Tsutsumi et al., 2010a,b,) and restoration of the preconditioning phenotype by cardiac-specific overexpression of Cav-3 (Tsutsumi et al., 2008).

Why is clinical translation so difficult and is caveolin a potential solution?

Since the original discovery of IPC nearly three decades ago, numerous molecular mechanisms and a host of therapeutic targets have been identified but not clinically translated. The root cause of this problem in clinical translation is likely to be complex. Simply put, the problem derives from the many complicating factors leading to a disconnection between the robustness of preclinical models, which are almost always performed in healthy young animals with no ongoing pharmacological treatments, or the various modifiers of protection, including age, sex, existing disease and drug treatment, that are present in patients. Could these complicating factors be somehow limited? The remainder of this review will explore the role of caveolin as a key feature to rescue protective pathways in pathophysiological settings.

Ageing

Why does the aged heart have decreased ischaemic tolerance?

As the population ages, there is an increasing challenge to preserve organ function in the face of disease and the well-known age-related changes in organ function and functional reserve. Age is the most important predictor of mortality in patients with ischaemic heart disease (Boersma et al., 2000). Consistent with this clinical result, aged human atrial myocytes are not protected from ischaemic insults (Mio et al., 2014). Studies with preclinical models also reveal an increased sensitivity and decreased tolerance to ischaemia-reperfusion injury in the aged heart (Headrick et al., 2003; Willems et al., 2005). Mechanisms that underlie this age-related deficit are not clear but are postulated to involve abnormalities in cellular signalling and mitochondrial function (Tani et al., 2001; Lesnefsky et al., 2006; Peart et al., 2007) although other mechanisms, such as dysfunctional calcium homeostasis (Swynghedauw et al., 1995), have been suggested.

Ageing also results in remodelling of mitochondria in terms of lipid content and membrane integrity (Pepe, 2005), defects in respiratory chain components (Lesnefsky et al., 2001) and increased oxidant stress (Hagen, 2003) which ultimately lead to reduced capacity to exclude calcium, generate ATP and limit injury mediated by ROS (Jahangir et al., 2001; Lakatta and Sollott, 2002; Lesnefsky and Hoppel, 2008). All of these factors may affect mPTP function. The molecular composition of the mPTP is controversial. Potential components of the mPTP have included the adenine nucleotide transporter (ANT) on the inner mitochondrial membrane and the voltage-dependent anion channel on the outer membrane, although genetic inactivation studies have suggested that neither of these components is necessary for mitochondrial permeability transition to occur (Bernardi, 2013). Cyclophilin D in the mitochondrial matrix also is thought to play a role in response to stress and the formation of the mPTP (Javadov and Karmazyn, 2007). A benzodiazepine receptor, hexokinase, and creatine kinase have also been proposed as regulators of the pore. Recent work has suggested the F0F1 ATP synthase forms a channel with properties similar to the functioning mPTP (Bernardi, 2013). It is unclear if cardiac protective agents act by inhibiting the opening of a preformed mPTP complex, a particular subunit of the complex, or the assembly or organization of the complex. Work in a model of protection has shown that increased phosphorylation of glycogen synthase kinase (GSK)-3β reduces the affinity of the ANT for cyclophilin D, suggesting that assembly of the complex is targeted by protective signals to limit mPTP opening (Nishihara et al., 2007). Importantly, regulation of the pore is diminished with age (Jahangir et al., 2001; Seo et al., 2008), but the precise mechanism of modulation is unknown (Di Lisa and Bernardi, 2005). Age-related changes may involve, as with cellular signalling, altered organization and function of the mPTP leading to inefficiency and dysfunction. Data from Drosophila indicate that the only protein that shows age-associated increases in carbonyl modifications (an index of oxidative injury) is ANT, a change that results in loss of mPTP function, which is accelerated by pro-oxidant stimuli (Yan and Sohal, 1998).

The function of mitochondria is intimately connected to mitochondrial dynamics. Mitochondria are in equilibrium between fusion and fission events to maintain their morphology and function. When fusion is inhibited, mitochondria become fragmented resulting in reduced glucose oxidation, respiration and loss of mitochondrial membrane potential (Olichon et al., 2003; Griparic et al., 2004). When fission is inhibited, mitochondria become tubular and elongated (Stojanovski et al., 2004). Fission is important for segregating irreversibly damaged mitochondria targeted for degradation. Excessive fission and lack of fusion result in loss of mitochondrial DNA, increased generation of ROS and loss of the mitochondrial network (Yaffe, 1999). Aged hearts have fewer mitochondria suggesting defects in fusion/fission.

Is caveolin a therapeutic target for reduced ischaemic tolerance with ageing?

The discussion thus far leads to two separate possibilities: (i) there is a potential ageing deficit that leads to an altered cardiac phenotype with age or (ii) the cellular environment created by ageing limits normal processes. Importantly, these are not mutually exclusive. Treatments aimed at restoring ischaemic tolerance in the aged myocardium must address the ‘ageing deficit’ and/or recreate a ‘young environment’ to alter not only cellular signalling but also restore dysfunctional mitochondria. From an experimental perspective, the only known external intervention to extend life in a number of species and reduce disease risk associated with ageing in primates and humans is caloric restriction (CR), an idea first conceptualized in 1935 (McCay et al., 1935). CR is likely to activate or deactivate a number of pathways to extend lifespan and to enhance protective and repair processes. These pathways include mitochondrial function, dynamics and autophagy (Masoro, 2009). Interestingly, two reports suggest that CR prevents an age-related decline in Cav-1 expression in hepatic sinusoids (Jamieson et al., 2007) and maternal CR elevates message for caveolin in the fetal cardiac left ventricle (Han et al., 2004). Could these findings somehow indicate a role for caveolin in longevity?

Little is known regarding caveolin expression and ageing. Early studies of ageing in isolated senescent cells showed increases in caveolin expression (Volonte et al., 2002; Cho and Park, 2005). However, in such studies, the concept of ‘ageing’ is contrived, as it is dependent on passage number. The concept that senescence is equivalent to ageing is flawed, as senescence is defined as an inability of cells to divide. Cardiac myocytes contain high levels of caveolin and are senescent cells by definition (they do not undergo significant cell division) but not necessarily aged. Therefore, the role of caveolin in ageing must be considered from a cell-type and organ-specific perspective. Animal studies reveal organ-specific patterns of changes in caveolin expression with age. Importantly, a decrease in the expression of cardiac Cav-3 (Kawabe et al., 2001) is observed as a function of age, a result we confirm in our preliminary data. Ageing results in dissociation of Cav-1 and Cav-3 from membrane caveolae (Ratajczak et al., 2003). Cav-3 KO mice develop a progressive cardiomyopathy (Woodman et al., 2002) and are also resistant to cardiac protective stimuli (Horikawa et al., 2008). Cav-1-deficient mice show reduced lifespan and increased cardiac dysfunction (Park et al., 2003) and are resistant to cardiac protective stimuli (Patel et al., 2007). We have recently shown that ischaemic tolerance is reduced in human atrial tissue (Peart et al., 2014). This observation was paralleled with the observation that Cav-3 is decreased in aged, compared with young, mouse hearts. We, furthermore, have indications that Cav-3 decreases with age in human hearts (unpublished data).

Caveolae are dynamic entities that form and dissipate in response to various stimuli (Tsutsumi et al., 2008) and serve as a clathrin-independent mechanism for the endocytosis of plasma membrane constituents. Caveolae-mediated endocytosis facilitates transport of vesicles to other cellular regions and across the cell (transcytosis) (Mukherjee et al., 2006; Ge et al., 2008). Co-expression of flotillins 1 and 2 in caveolae enhances the accumulation of intracellular vesicles (Frick et al., 2007). The fate of such vesicles is unknown. Recent data indicate that caveolae forms contacts with other cellular compartments to communicate membrane-derived signals to other organelles and regions of the cells. For example, smooth muscle cells have ‘nanocontacts’ between caveolae and the endoplasmic reticulum (Gherghiceanu and Popescu, 2007). Caveolins are found in cells and intracellular regions lacking caveolae, suggesting roles for caveolins in non-sarcolemmal locations (Head and Insel, 2007; Fridolfsson et al., 2014). We have recently shown that there is a stress-adaptive transfer of caveolin to mitochondria which is facilitated by IPC that leads to protection of the heart from ischaemia-reperfusion injury and that this is a generalized protective pathway active in cancer and Caenorhabditis elegans and involves the activation of GPCR signalling and survival kinases (Fridolfsson et al., 2012; Wang et al., 2014). It is possible that a loss of caveolin expression with age affects the ability of the membrane not only to house and regulate survival kinases but also limits the ability of the cell to modulate mitochondrial function during stress.

Diabetes

Why is the diabetic heart dysfunctional?

According to the American Diabetes Association in the United States, there are nearly 26 million individuals, adults and children, with diabetes. In addition, there may be as many as 79 million individuals who are prediabetic. In 2007, diabetes was listed as the underlying cause of >70 000 deaths and a contributing factor of an additional 160 000 deaths. Those aged 65 years or older represent an ever-growing population facing the consequences of diabetes. In 2004, the most recent year for which statistics are available, heart disease was noted in nearly 70% of diabetes-related deaths among people 65 years or older and adults with diabetes have heart disease mortality rates that are two to four times higher than adults without diabetes.

Controversy exists as to whether cardiac events associated with diabetes are a consequence of underlying coronary artery disease and hypertension. Growing evidence suggests that diabetes results in altered cardiac structure and function independent of vascular pathology, supporting the existence of a ‘diabetic cardiomyopathy’. Diabetes in animal models results in both diastolic (i.e. prolongation of relaxation and increased left ventricular end diastolic pressure) (Joffe et al., 1999) and systolic (i.e. heart rate, systolic BP and fractional shortening) (Joffe et al., 1999) dysfunction and such findings are also observed in humans (Poirier et al., 2001). Structural changes also have been observed in the diabetic heart that include perivascular and interstitial fibrosis, possibly as a result of replacement of myocyte loss, altered mitochondrial structure and altered cardiac ultrastructure (Eto et al., 1987; Warley et al., 1995; Mizushige et al., 2000). The molecular mechanisms proposed for diabetic cardiomyopathy are diverse and may include impaired calcium handling, altered substrate supply and utilization, altered energy generation with mitochondrial dysfunction, altered ion channel function, myocyte apoptosis, endothelial dysfunction, cardiac insulin resistance and activation of the renin-angiotensin system (Zhang and Chen, 2012). Additionally, diabetic hearts are refractory to protective interventions that limit ischaemia-reperfusion injury, suggesting major defects in survival kinase signalling (Balakumar and Sharma, 2012a). Such findings suggest that diabetic cardiomyopathy is a complex disease that is manifested with many cellular alterations that may or may not have a common control point of regulation that can be targeted therapeutically.

Are caveolins potential regulators of diabetes?

Cav-3 KO mice have a variety of deleterious phenotypes, such as muscle degeneration (Hagiwara et al., 2000), insulin resistance (Oshikawa et al., 2004) and progressive cardiomyopathy with age (Woodman et al., 2002). Knockdown of Cav-1 in adipocytes results in loss of insulin receptor signalling as a result of decreased insulin receptor and glucose transporter 4 (GLUT4) expression (Gonzalez-Munoz et al., 2009). Although hearts of Cav-1 and Cav-3 KO mice develop cardiomyopathy, they appear to have normal substrate utilization (Augustus et al., 2008). In H9C2 cardiomyoblasts, Cav-1 knockdown has been shown to inhibit signalling by insulin-like growth factors (Salani et al., 2008) and insulin signalling directly coupled to Akt and glucose transport (Ha and Pak, 2005). Importantly, Cav-3 was a positive regulator of insulin signalling (Yamamoto et al., 1998) and caveolin gene transfer to the liver improved glucose metabolism in diabetic mice (Otsu et al., 2009). Recently, our group has shown that Cav-3 overexpression in the heart leads to enhanced Akt phosphorylation that results in protection of the heart from ischaemia-reperfusion injury (Tsutsumi et al., 2008). Other studies reveal that compounds such as resveratrol, which are polyphenols shown to have lifespan-expanding properties (Frojdo et al., 2008), also recruit GLUT4 to caveolae and up-regulate Akt signalling in the setting of type I diabetes (Penumathsa et al., 2008). Caveolae also are major regulators of calcium storage and influx which may be an added cellular regulatory feature important to limiting diabetic cardiomyopathy (Shaul and Anderson, 1998).

Diabetes results in altered cardiac mitochondrial function with respect to complex activity, generation of ATP and activation of the mPTP, a key feature leading to cellular apoptosis (Oliveira et al., 2003; Boudina et al., 2007). Recent evidence suggests that caveolin-deficient stromal cells have compromised mitochondrial function (Pavlides et al., 2010) and mitochondria from Cav-1-KO fibroblasts accumulate cholesterol and have severe dysfunction; such cells adapt poorly to nutrient starvation and are predisposed to apoptosis (Bosch et al., 2011). Loss of caveolin leads to altered mitochondrial function in adipose tissue, suggesting a link between caveolin and metabolism (Wernstedt Asterholm et al., 2012).

Caveolins also may play a role in pathologies associated with diabetes including metabolic syndrome, as recently reviewed by Zhang (2014). Specifically, the association between GLUT4 transporters and caveolin plays an important role in the development of insulin resistance (Kabayama et al., 2007; Liu et al., 2008b). Clinical studies utilizing caveolin as a marker or protein of interest in the setting of insulin resistance are rare and primarily address insulin resistance in the context of caveolinopathies (Mendez-Gimenez et al., 2014). Some authors argue in favour of the importance of the caveolar structure, rather than the loss of either Cav-1 or Cav-3 (Mendez-Gimenez et al., 2014). In a translational approach, Cav-1 polymorphisms have been linked to insulin resistance and hypertension in Caucasian and Hispanic patients (Pojoga et al., 2011). In diabetes mellitus patients that underwent flow-mediated dilation of coronary arterioles during heart surgery, membrane localized Cav-1 was significantly reduced. This reduction was attributed to peroxynitrite, which contributes to microvascular dysfunction in diabetes mellitus (Cassuto et al., 2014).

A major confounding factor in the translation of protective strategies to patients with diabetes is that many pharmacological agents that patients are prescribed may negate cardioprotection. Most diabetic patients receiving oral medications will be taking a sulfonylurea that blocks KATP channels, which in the pancreas increases insulin secretion but, in the heart, the same drug results in the attenuation of cardiac protection (Gross and Auchampach, 1992). Most diabetics and elderly patients are also on statins to maintain low blood cholesterol. Although statins have been shown to have many effects that have potential to protect the heart in specific settings, there is growing concerns that diabetics and individual with other pathophysiologies may not benefit as much as previously thought (Gullestad et al., 2007; Drummond et al., 2010; Schilling et al., 2014). Reduction of caveolin through statin treatment in endothelial cells could affect protection indirectly, through modifying eNOS signalling (Balakumar et al., 2012b). Conversely, one could argue that statin inhibition of the cholesterol pathway and a consequent decrease in Cav-3 in cardiac myocytes could result in impaired survival kinase signalling. The proof of this concept in the heart, specifically under long-term treatment, is still under investigation, while some effects of statin treatment on Cav-1 and the development of diabetes in preclinical models have been recently reviewed (Brault et al., 2014).

In both the ageing and the diabetic heart, the two central features of the pathology are loss of effective signalling networks and compromised ultrastructure, the two perquisites Murry, Jennings and Reimer described early on, as being critical to the induction of IPC. Central to this dysfunction appears to be the loss of caveolin in the heart.

Caveolins in myopathies

Cav-3 interacts with signalling molecules involved in cardiac hypertrophy, remodelling and the progression of heart failure (Fujita et al., 2001; Krajewska and Maslowska, 2004). Cav-3 KO mice exhibit reduced cardiac function and cardiomyopathy (Woodman et al., 2002). These results suggest a potential role for Cav-3 in heart failure. Expression of cardiac Cav-3 is changed in models of heart failure and patients with cardiomyopathy, although there are inconsistencies in the findings (Hare et al., 2000; Damy et al., 2004; Hayashi et al., 2004; Ruiz-Hurtado et al., 2007). In the heart, mutations of Cav-3 can lead to familial hypertrophic cardiomyopathy (Hayashi et al., 2004) as well as arrhythmias such as the congenital long-QT syndrome (Vatta et al., 2006; Balijepalli and Kamp, 2008). In the models and patients examined, the variability of the results may be due to species differences or the stage of heart failure development and ventricular dysfunction. Recently, Feiner et al. (2011) reported reduced levels of Cav-3 in two well-established models of heart failure in mice, overexpression of the adenosine A1 receptor or TNFα. They found significantly reduced levels of Cav-3 protein and mRNA in the mice with heart failure and showed a significant correlation between the reduced levels of Cav-3 and reduced cardiac function. In addition, these investigators found a significant correlation between the reduced levels of Cav-3 in failing human heart samples and the levels of the sarcoplasmic-endoplasmic reticulum calcium ATPase, a marker of heart failure. Our group has shown that cardiac myocyte-specific overexpression of Cav-3 limits the hypertrophic response to transverse aortic constriction and improves survival (Horikawa et al., 2011). Such data indicate a role for Cav-3 as a therapeutic protein in heart failure. Cav-3 also plays a role in muscular dystrophies (Woodman et al., 2004) including limb girdle muscular dystrophy type 1C (Angelini, 2004). The pathogenesis involved in these diseases involves a failure to traffic Cav-3 from the Golgi network to the plasma membrane (Woodman et al., 2004).

Translational approaches to increasing caveolin

From the data presented, it is evident that heart-specific decreases in caveolin are detrimental to the heart, whereas up-regulation of caveolin may be beneficial to the heart. Our laboratory is currently developing a gene therapy-based approach to overexpress caveolin in a cell type-specific manner using selective promoters and regulatory elements. Although this approach has clinical potential, translation is likely to be far in the future. Therefore, we need to find other natural means to increase caveolin expression. In one study on 14 male pentathlon athletes, exercise increased Cav-1, Cav-3, GLUT4 and the insulin receptor-β in samples from the vastus lateralis muscle after a 1500 m swim trial (Kim et al., 2009). In another experiment, the manipulation of pre-exercise muscle glycogen storage was assessed. Here, an increase in the baseline levels of Cav-1 after recovery from initial glycogen depletion exercise was noted (Roepstorff et al., 2004). Furthermore, in an exercise countermeasure during 12 weeks of bed rest, exercise altered NOS2/Cav-3 co-immunostaining patterns in vastus lateralis and soleus myofibres (Rudnick et al., 2004). Additionally, in an intensive care unit model in rats, immobilization results in distinct alterations in gene expression and down-regulation of Cav-3 expression (Llano-Diez et al., 2011).

Conclusion

Caveolae and caveolins are comparatively new players in a relatively saturated field of ischaemia-reperfusion injury. Given the data provided here, it should be clear that there is a central role for caveolin expression in the protection of the heart, and potentially other organs, from ischaemia-reperfusion injury and cell stress in general. It is intriguing that conditions in which protection is lost show marked loss of caveolin expression coupled to decreased survival kinase signalling and dysfunctional myocyte ultrastructure. Mice with cardiac specific overexpression have dramatic cardiac stress adaptation in a variety of disease settings and provide hope that caveolin may serve as a critical mediator and potential therapeutic target to provide protection from ischaemia-reperfusion injury in humans.

Acknowledgments

This work was supported by grants from the National Institutes of Health HL091071 (H. H. P.), HL107200 (H. H. P.), HL066941 (D. M. R. and H. H. P.) and HL115933 (D. M. R.), and VA Merit BX001963 (H. H. P.) and BX000783 (D. M. R.).

Glossary

ANT

adenine nucleotide transporter

Cav

caveolin

CR

caloric restriction

CSD

caveolin-scaffolding domain

eNOS

endothelial NOS

GLUT4

glucose transporter 4

IPC

ischaemic preconditioning

KO

knockout

mPTP

mitochondrial permeability transition pore

ROS

reactive oxygen species

Conflict of interest

There are no competing interests or disclosures.

References

  1. Alexander SPH, Benson HE, Faccenda E, Pawson AJ, Sharman JL. Spedding M, et al. The Concise Guide to PHARMACOLOGY 2013/14: Enzymes. Br J Pharmacol. 2013a;170:1797–1867. doi: 10.1111/bph.12451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Alexander SPH, Benson HE, Faccenda E, Pawson AJ, Sharman JL. Spedding M, et al. The Concise Guide to PHARMACOLOGY 2013/14: Transporters. Br J Pharmacol. 2013b;170:1706–1796. doi: 10.1111/bph.12450. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Alexander SPH, Benson HE, Faccenda E, Pawson AJ, Sharman JL. Spedding M, et al. The Concise Guide to PHARMACOLOGY 2013/14: Catalytic Receptors. Br J Pharmacol. 2013c;170:1676–1705. doi: 10.1111/bph.12449. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Alexander SPH, Benson HE, Faccenda E, Pawson AJ, Sharman JL, Catterall WA, et al. The Concise Guide to PHARMACOLOGY 2013/14: Ion Channels. Br J Pharmacol. 2013d;170:1607–1651. doi: 10.1111/bph.12447. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Alexander SPH, Benson HE, Faccenda E, Pawson AJ, Sharman JL. Spedding M, et al. The Concise Guide to PHARMACOLOGY 2013/14: G Protein-Coupled Receptors. Br J Pharmacol. 2013e;170:1459–1581. doi: 10.1111/bph.12445. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Angelini C. Limb-girdle muscular dystrophies: heterogeneity of clinical phenotypes and pathogenetic mechanisms. Acta Myol. 2004;23:130–136. [PubMed] [Google Scholar]
  7. Armstrong SC, Latham CA, Shivell CL, Ganote CE. Ischemic loss of sarcolemmal dystrophin and spectrin: correlation with myocardial injury. J Mol Cell Cardiol. 2001;33:1165–1179. doi: 10.1006/jmcc.2001.1380. [DOI] [PubMed] [Google Scholar]
  8. Auchampach JA, Grover GJ, Gross GJ. Blockade of ischaemic preconditioning in dogs by the novel ATP dependent potassium channel antagonist sodium 5-hydroxydecanoate. Cardiovasc Res. 1992;26:1054–1062. doi: 10.1093/cvr/26.11.1054. [DOI] [PubMed] [Google Scholar]
  9. Augustus AS, Buchanan J, Addya S, Rengo G, Pestell RG, Fortina P, et al. Substrate uptake and metabolism are preserved in hypertrophic caveolin-3 knockout hearts. Am J Physiol Heart Circ Physiol. 2008;295:H657–H666. doi: 10.1152/ajpheart.00387.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Balakumar P, Sharma NK. Healing the diabetic heart: does myocardial preconditioning work? Cell Signal. 2012a;24:53–59. doi: 10.1016/j.cellsig.2011.09.007. [DOI] [PubMed] [Google Scholar]
  11. Balakumar P, Kathuria S, Taneja G, Kalra S, Mahadevan N. Is targeting eNOS a key mechanistic insight of cardiovascular defensive potentials of statins? J Mol Cell Cardiol. 2012b;52:83–92. doi: 10.1016/j.yjmcc.2011.09.014. [DOI] [PubMed] [Google Scholar]
  12. Balijepalli RC, Kamp TJ. Caveolae, ion channels and cardiac arrhythmias. Prog Biophys Mol Biol. 2008;98:149–160. doi: 10.1016/j.pbiomolbio.2009.01.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Ballard-Croft C, Locklar AC, Kristo G, Lasley RD. Regional myocardial ischemia induced activation of MAPKs is associated with subcellular redistribution of caveolin and cholesterol. Am J Physiol Heart Circ Physiol. 2006;291:H658–H667. doi: 10.1152/ajpheart.01354.2005. [DOI] [PubMed] [Google Scholar]
  14. Becher A, McIlhinney RA. Consequences of lipid raft association on G-protein-coupled receptor function. Biochem Soc Symp. 2005;72:151–164. doi: 10.1042/bss0720151. [DOI] [PubMed] [Google Scholar]
  15. Becker LB, VandenHoek TL, Shao Z-H, Li C-Q, Schumacker PT. Generation of superoxide in cardiomyocytes during ischemia before reperfusion. Am J Physiol Heart Circ Physiol. 1999;277:H2240–H2246. doi: 10.1152/ajpheart.1999.277.6.H2240. [DOI] [PubMed] [Google Scholar]
  16. Bernardi P. The mitochondrial permeability transition pore: a mystery solved? Front Physiol. 2013;4:95. doi: 10.3389/fphys.2013.00095. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Boersma E, Pieper KS, Steyerberg EW, Wilcox RG, Chang WC, Lee KL, et al. Predictors of outcome in patients with acute coronary syndromes without persistent ST-segment elevation. Results from an international trial of 9461 patients. The PURSUIT Investigators. Circulation. 2000;101:2557–2567. doi: 10.1161/01.cir.101.22.2557. [DOI] [PubMed] [Google Scholar]
  18. Bosch M, Mari M, Herms A, Fernandez A, Fajardo A, Kassan A, et al. Caveolin-1 deficiency causes cholesterol-dependent mitochondrial dysfunction and apoptotic susceptibility. Curr Biol. 2011;21:681–686. doi: 10.1016/j.cub.2011.03.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Boudina S, Sena S, Theobald H, Sheng X, Wright JJ, Hu XX, et al. Mitochondrial energetics in the heart in obesity-related diabetes: direct evidence for increased uncoupled respiration and activation of uncoupling proteins. Diabetes. 2007;56:2457–2466. doi: 10.2337/db07-0481. [DOI] [PubMed] [Google Scholar]
  20. Brault M, Ray J, Gomez YH, Mantzoros CS, Daskalopoulou SS. Statin treatment and new-onset diabetes: a review of proposed mechanisms. Metabolism. 2014;63:735–745. doi: 10.1016/j.metabol.2014.02.014. [DOI] [PubMed] [Google Scholar]
  21. Cao CM, Zhang Y, Weisleder N, Ferrante C, Wang X, Lv F, et al. MG53 constitutes a primary determinant of cardiac ischemic preconditioning. Circulation. 2011;121:2565–2574. doi: 10.1161/CIRCULATIONAHA.110.954628. [DOI] [PubMed] [Google Scholar]
  22. Capozza F, Cohen AW, Cheung MW, Sotgia F, Schubert W, Battista M, et al. Muscle-specific interaction of caveolin isoforms: differential complex formation between caveolins in fibroblastic vs. muscle cells. Am J Physiol Cell Physiol. 2005;288:C677–C691. doi: 10.1152/ajpcell.00232.2004. [DOI] [PubMed] [Google Scholar]
  23. Cassuto J, Dou H, Czikora I, Szabo A, Patel VS, Kamath V, et al. Peroxynitrite disrupts endothelial caveolae leading to eNOS uncoupling and diminished flow-mediated dilation in coronary arterioles of diabetic patients. Diabetes. 2014;63:1381–1393. doi: 10.2337/db13-0577. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Chini B, Parenti M. G-protein coupled receptors in lipid rafts and caveolae: how, when and why do they go there? J Mol Endocrinol. 2004;32:325–338. doi: 10.1677/jme.0.0320325. [DOI] [PubMed] [Google Scholar]
  25. Cho KA, Park SC. Caveolin-1 as a prime modulator of aging: a new modality for phenotypic restoration? Mech Ageing Dev. 2005;126:105–110. doi: 10.1016/j.mad.2004.09.029. [DOI] [PubMed] [Google Scholar]
  26. Chun M, Liyanage UK, Lisanti MP, Lodish HF. Signal transduction of a G protein-coupled receptor in caveolae: colocalization of endothelin and its receptor with caveolin. Proc Natl Acad Sci U S A. 1994;91:11728–11732. doi: 10.1073/pnas.91.24.11728. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Cohen AW, Hnasko R, Schubert W, Lisanti MP. Role of caveolae and caveolins in health and disease. Physiol Rev. 2004;84:1341–1379. doi: 10.1152/physrev.00046.2003. [DOI] [PubMed] [Google Scholar]
  28. Collins BM, Davis MJ, Hancock JF, Parton RG. Structure-based reassessment of the caveolin signaling model: do caveolae regulate signaling through caveolin-protein interactions? Dev Cell. 2012;23:11–20. doi: 10.1016/j.devcel.2012.06.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Damy T, Ratajczak P, Shah AM, Camors E, Marty I, Hasenfuss G, et al. Increased neuronal nitric oxide synthase-derived NO production in the failing human heart. Lancet. 2004;363:1365–1367. doi: 10.1016/S0140-6736(04)16048-0. [DOI] [PubMed] [Google Scholar]
  30. Di Lisa F, Bernardi P. Mitochondrial function and myocardial aging. A critical analysis of the role of permeability transition. Cardiovasc Res. 2005;66:222–232. doi: 10.1016/j.cardiores.2005.02.009. [DOI] [PubMed] [Google Scholar]
  31. Dietzen DJ, Hastings WR, Lublin DM. Caveolin is palmitoylated on multiple cysteine residues. Palmitoylation is not necessary for localization of caveolin to caveolae. J Biol Chem. 1995;270:6838–6842. doi: 10.1074/jbc.270.12.6838. [DOI] [PubMed] [Google Scholar]
  32. Donaldson JG, Porat-Shliom N, Cohen LA. Clathrin-independent endocytosis: a unique platform for cell signaling and PM remodeling. Cell Signal. 2009;21:1–6. doi: 10.1016/j.cellsig.2008.06.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Drummond JC, Head BP, Patel PM. Statins might contribute to postoperative delirium. Anesthesiology. 2010;113:500–501. doi: 10.1097/ALN.0b013e3181e33bf7. [DOI] [PubMed] [Google Scholar]
  34. Engelman JA, Chu C, Lin A, Jo H, Ikezu T, Okamoto T, et al. Caveolin-mediated regulation of signaling along the p42/44 MAP kinase cascade in vivo. A role for the caveolin-scaffolding domain. FEBS Lett. 1998;428:205–211. doi: 10.1016/s0014-5793(98)00470-0. [DOI] [PubMed] [Google Scholar]
  35. Eto M, Watanabe K, Sekiguchi M, Iwashima Y, Morikawa A, Oshima E, et al. Metabolic and morphological changes of the heart in Chinese hamsters (CHAD strain) with spontaneous long-term diabetes. Diabetes Res Clin Pract. 1987;3:297–305. doi: 10.1016/s0168-8227(87)80054-2. [DOI] [PubMed] [Google Scholar]
  36. Fang J, Wu L, Chen L. Postconditioning attenuates cardiocyte ultrastructure injury and apoptosis by blocking mitochondrial permeability transition in rats. Acta Cardiol. 2008;63:377–387. doi: 10.2143/AC.63.3.1020316. [DOI] [PubMed] [Google Scholar]
  37. Fecchi K, Volonte D, Hezel MP, Schmeck K, Galbiati F. Spatial and temporal regulation of GLUT4 translocation by flotillin-1 and caveolin-3 in skeletal muscle cells. FASEB J. 2006;20:705–707. doi: 10.1096/fj.05-4661fje. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Feiner EC, Chung P, Jasmin JF, Zhang J, Whitaker-Menezes D, Myers V, et al. Left ventricular dysfunction in murine models of heart failure and in failing human heart is associated with a selective decrease in the expression of caveolin-3. J Card Fail. 2011;17:253–263. doi: 10.1016/j.cardfail.2010.10.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Feron O, Balligand JL. Caveolins and the regulation of endothelial nitric oxide synthase in the heart. Cardiovasc Res. 2006;69:788–797. doi: 10.1016/j.cardiores.2005.12.014. [DOI] [PubMed] [Google Scholar]
  40. Feron O, Kelly RA. The caveolar paradox: suppressing, inducing, and terminating eNOS signaling. Circ Res. 2001;88:129–131. doi: 10.1161/01.res.88.2.129. [DOI] [PubMed] [Google Scholar]
  41. Feron O, Saldana F, Michel JB, Michel T. The endothelial nitric-oxide synthase-caveolin regulatory cycle. J Biol Chem. 1998;273:3125–3128. doi: 10.1074/jbc.273.6.3125. [DOI] [PubMed] [Google Scholar]
  42. Frick M, Bright NA, Riento K, Bray A, Merrified C, Nichols BJ. Coassembly of flotillins induces formation of membrane microdomains, membrane curvature, and vesicle budding. Curr Biol. 2007;17:1151–1156. doi: 10.1016/j.cub.2007.05.078. [DOI] [PubMed] [Google Scholar]
  43. Fridolfsson HN, Kawaraguchi Y, Ali SS, Panneerselvam M, Niesman IR, Finley JC, et al. Mitochondria-localized caveolin in adaptation to cellular stress and injury. FASEB J. 2012;26:4637–4649. doi: 10.1096/fj.12-215798. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Fridolfsson HN, Roth DM, Insel PA, Patel HH. Regulation of intracellular signaling and function by caveolin. FASEB J. 2014;28:3823–3831. doi: 10.1096/fj.14-252320. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Frojdo S, Durand C, Pirola L. Metabolic effects of resveratrol in mammals – a link between improved insulin action and aging. Curr Aging Sci. 2008;1:145–151. doi: 10.2174/1874609810801030145. [DOI] [PubMed] [Google Scholar]
  46. Fuhs SR, Insel PA. Caveolin-3 undergoes SUMOylation by the SUMO E3 ligase PIASy: sumoylation affects G-protein-coupled receptor desensitization. J Biol Chem. 2011;286:14830–14841. doi: 10.1074/jbc.M110.214270. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Fujimoto T. Calcium pump of the plasma membrane is localized in caveolae. J Cell Biol. 1993;120:1147–1157. doi: 10.1083/jcb.120.5.1147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Fujimoto T, Nakade S, Miyawaki A, Mikoshiba K, Ogawa K. Localization of inositol 1,4,5-trisphosphate receptor-like protein in plasmalemmal caveolae. J Cell Biol. 1992;119:1507–1513. doi: 10.1083/jcb.119.6.1507. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Fujita T, Toya Y, Iwatsubo K, Onda T, Kimura K, Umemura S, et al. Accumulation of molecules involved in alpha1-adrenergic signal within caveolae: caveolin expression and the development of cardiac hypertrophy. Cardiovasc Res. 2001;51:709–716. doi: 10.1016/s0008-6363(01)00348-0. [DOI] [PubMed] [Google Scholar]
  50. Ge S, Song L, Serwanski DR, Kuziel WA, Pachter JS. Transcellular transport of CCL2 across brain microvascular endothelial cells. J Neurochem. 2008;104:1219–1232. doi: 10.1111/j.1471-4159.2007.05056.x. [DOI] [PubMed] [Google Scholar]
  51. Geshi E, Ishioka H, Nomizo A, Nakatani M, Katagiri T. Biochemical and ultrastructural evaluations of the effect of ischemic preconditioning on ischemic myocardial injury – role of the adenosine triphosphate-sensitive potassium channel. Jpn Circ J. 1998;62:915–924. doi: 10.1253/jcj.62.915. [DOI] [PubMed] [Google Scholar]
  52. Gherghiceanu M, Popescu LM. Electron microscope tomography: further demonstration of nanocontacts between caveolae and smooth muscle sarcoplasmic reticulum. J Cell Mol Med. 2007;11:1416–1418. doi: 10.1111/j.1582-4934.2007.00166.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Giovanardi RO, Rhoden EL, Cerski CT, Salvador M, Kalil AN. Ischemic preconditioning protects the pig liver by preserving the mitochondrial structure and downregulating caspase-3 activity. J Invest Surg. 2009;22:88–97. doi: 10.1080/08941930802712995. [DOI] [PubMed] [Google Scholar]
  54. Gonzalez-Munoz E, Lopez-Iglesias C, Calvo M, Palacin M, Zorzano A, Camps M. Caveolin-1 loss of function accelerates glucose transporter 4 and insulin receptor degradation in 3T3-L1 adipocytes. Endocrinology. 2009;150:3493–3502. doi: 10.1210/en.2008-1520. [DOI] [PubMed] [Google Scholar]
  55. Gottlieb RA, Gustafsson AB. Mitochondrial turnover in the heart. Biochim Biophys Acta. 2011;1813:1295–1301. doi: 10.1016/j.bbamcr.2010.11.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Grande-Garcia A, Del Pozo MA. Caveolin-1 in cell polarization and directional migration. Eur J Cell Biol. 2008;87:641–647. doi: 10.1016/j.ejcb.2008.02.001. [DOI] [PubMed] [Google Scholar]
  57. Grande-Garcia A, Echarri A, de Rooij J, Alderson NB, Waterman-Storer CM, Valdivielso JM, et al. Caveolin-1 regulates cell polarization and directional migration through Src kinase and Rho GTPases. J Cell Biol. 2007;177:683–694. doi: 10.1083/jcb.200701006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Griparic L, van der Wel NN, Orozco IJ, Peters PJ, van der Bliek AM. Loss of the intermembrane space protein Mgm1/OPA1 induces swelling and localized constrictions along the lengths of mitochondria. J Biol Chem. 2004;279:18792–18798. doi: 10.1074/jbc.M400920200. [DOI] [PubMed] [Google Scholar]
  59. Gross GJ, Auchampach JA. Blockade of ATP-sensitive potassium channels prevents myocardial preconditioning in dogs. Circ Res. 1992;70:223–233. doi: 10.1161/01.res.70.2.223. [DOI] [PubMed] [Google Scholar]
  60. Gullestad L, Oie E, Ueland T, Yndestad A, Aukrust P. The role of statins in heart failure. Fundam Clin Pharmacol. 2007;21:35–40. doi: 10.1111/j.1472-8206.2007.00538.x. [DOI] [PubMed] [Google Scholar]
  61. Ha H, Pak Y. Modulation of the caveolin-3 and Akt status in caveolae by insulin resistance in H9c2 cardiomyoblasts. Exp Mol Med. 2005;37:169–178. doi: 10.1038/emm.2005.23. [DOI] [PubMed] [Google Scholar]
  62. Hagen TM. Oxidative stress, redox imbalance, and the aging process. Antioxid Redox Signal. 2003;5:503–506. doi: 10.1089/152308603770310149. [DOI] [PubMed] [Google Scholar]
  63. Hagiwara Y, Sasaoka T, Araishi K, Imamura M, Yorifuji H, Nonaka I, et al. Caveolin-3 deficiency causes muscle degeneration in mice. Hum Mol Genet. 2000;9:3047–3054. doi: 10.1093/hmg/9.20.3047. [DOI] [PubMed] [Google Scholar]
  64. Hagiwara Y, Nishina Y, Yorifuji H, Kikuchi T. Immunolocalization of caveolin-1 and caveolin-3 in monkey skeletal, cardiac and uterine smooth muscles. Cell Struct Funct. 2002;27:375–382. doi: 10.1247/csf.27.375. [DOI] [PubMed] [Google Scholar]
  65. Han HC, Austin KJ, Nathanielsz PW, Ford SP, Nijland MJ, Hansen TR. Maternal nutrient restriction alters gene expression in the ovine fetal heart. J Physiol. 2004;558:111–121. doi: 10.1113/jphysiol.2004.061697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Hare JM, Lofthouse RA, Juang GJ, Colman L, Ricker KM, Kim B, et al. Contribution of caveolin protein abundance to augmented nitric oxide signaling in conscious dogs with pacing-induced heart failure. Circ Res. 2000;86:1085–1092. doi: 10.1161/01.res.86.10.1085. [DOI] [PubMed] [Google Scholar]
  67. Hausenloy DJ, Yellon DM. Survival kinases in ischemic preconditioning and postconditioning. Cardiovasc Res. 2006;70:240–253. doi: 10.1016/j.cardiores.2006.01.017. [DOI] [PubMed] [Google Scholar]
  68. Hausenloy DJ, Yellon DM. Cell membrane repair as a mechanism for ischemic preconditioning? Circulation. 2010;121:2547–2549. doi: 10.1161/CIRCULATIONAHA.110.958462. [DOI] [PubMed] [Google Scholar]
  69. Hausenloy DJ, Tsang A, Yellon DM. The reperfusion injury salvage kinase pathway: a common target for both ischemic preconditioning and postconditioning. Trends Cardiovasc Med. 2005;15:69–75. doi: 10.1016/j.tcm.2005.03.001. [DOI] [PubMed] [Google Scholar]
  70. Hayashi T, Arimura T, Ueda K, Shibata H, Hohda S, Takahashi M, et al. Identification and functional analysis of a caveolin-3 mutation associated with familial hypertrophic cardiomyopathy. Biochem Biophys Res Commun. 2004;313:178–184. doi: 10.1016/j.bbrc.2003.11.101. [DOI] [PubMed] [Google Scholar]
  71. Head BP, Insel PA. Do caveolins regulate cells by actions outside of caveolae? Trends Cell Biol. 2007;17:51–57. doi: 10.1016/j.tcb.2006.11.008. [DOI] [PubMed] [Google Scholar]
  72. Head BP, Patel HH, Roth DM, Lai NC, Niesman IR, Farquhar MG, et al. G-protein-coupled receptor signaling components localize in both sarcolemmal and intracellular caveolin-3-associated microdomains in adult cardiac myocytes. J Biol Chem. 2005;280:31036–31044. doi: 10.1074/jbc.M502540200. [DOI] [PubMed] [Google Scholar]
  73. Head BP, Patel HH, Roth DM, Murray F, Swaney JS, Niesman IR, et al. Microtubules and actin microfilaments regulate lipid raft/caveolae localization of adenylyl cyclase signaling components. J Biol Chem. 2006;281:26391–26399. doi: 10.1074/jbc.M602577200. [DOI] [PubMed] [Google Scholar]
  74. Headrick JP, Willems L, Ashton KJ, Holmgren K, Peart J, Matherne GP. Ischaemic tolerance in aged mouse myocardium: the role of adenosine and effects of A1 adenosine receptor overexpression. J Physiol. 2003;549:823–833. doi: 10.1113/jphysiol.2003.041541. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Hill MM, Bastiani M, Luetterforst R, Kirkham M, Kirkham A, Nixon SJ, et al. PTRF-Cavin, a conserved cytoplasmic protein required for caveola formation and function. Cell. 2008;132:113–124. doi: 10.1016/j.cell.2007.11.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Horikawa YT, Patel HH, Tsutsumi YM, Jennings MM, Kidd MW, Hagiwara Y, et al. Caveolin-3 expression and caveolae are required for isoflurane-induced cardiac protection from hypoxia and ischemia/reperfusion injury. J Mol Cell Cardiol. 2008;44:123–130. doi: 10.1016/j.yjmcc.2007.10.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Horikawa YT, Panneerselvam M, Kawaraguchi Y, Tsutsumi YM, Ali SS, Balijepalli RC, et al. Cardiac-specific overexpression of caveolin-3 attenuates cardiac hypertrophy and increases natriuretic Peptide expression and signaling. J Am Coll Cardiol. 2011;57:2273–2283. doi: 10.1016/j.jacc.2010.12.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Jahangir A, Ozcan C, Holmuhamedov EL, Terzic A. Increased calcium vulnerability of senescent cardiac mitochondria: protective role for a mitochondrial potassium channel opener. Mech Ageing Dev. 2001;122:1073–1086. doi: 10.1016/s0047-6374(01)00242-1. [DOI] [PubMed] [Google Scholar]
  79. Jamieson HA, Hilmer SN, Cogger VC, Warren A, Cheluvappa R, Abernethy DR, et al. Caloric restriction reduces age-related pseudocapillarization of the hepatic sinusoid. Exp Gerontol. 2007;42:374–378. doi: 10.1016/j.exger.2006.11.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Javadov S, Karmazyn M. Mitochondrial permeability transition pore opening as an endpoint to initiate cell death and as a putative target for cardioprotection. Cell Physiol Biochem. 2007;20:1–22. doi: 10.1159/000103747. [DOI] [PubMed] [Google Scholar]
  81. Joffe II, Travers KE, Perreault-Micale CL, Hampton T, Katz SE, Morgan JP, et al. Abnormal cardiac function in the streptozotocin-induced non-insulin-dependent diabetic rat: noninvasive assessment with doppler echocardiography and contribution of the nitric oxide pathway. J Am Coll Cardiol. 1999;34:2111–2119. doi: 10.1016/s0735-1097(99)00436-2. [DOI] [PubMed] [Google Scholar]
  82. Jones KA, Jiang X, Yamamoto Y, Yeung RS. Tuberin is a component of lipid rafts and mediates caveolin-1 localization: role of TSC2 in post-Golgi transport. Exp Cell Res. 2004;295:512–524. doi: 10.1016/j.yexcr.2004.01.022. [DOI] [PubMed] [Google Scholar]
  83. Juhaszova M, Zorov DB, Kim SH, Pepe S, Fu Q, Fishbein KW, et al. Glycogen synthase kinase-3beta mediates convergence of protection signaling to inhibit the mitochondrial permeability transition pore. J Clin Invest. 2004;113:1535–1549. doi: 10.1172/JCI19906. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Kabayama K, Sato T, Saito K, Loberto N, Prinetti A, Sonnino S, et al. Dissociation of the insulin receptor and caveolin-1 complex by ganglioside GM3 in the state of insulin resistance. Proc Natl Acad Sci U S A. 2007;104:13678–13683. doi: 10.1073/pnas.0703650104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Kamoun WS, Karaa A, Kresge N, Merkel SM, Korneszczuk K, Clemens MG. LPS inhibits endothelin-1-induced endothelial NOS activation in hepatic sinusoidal cells through a negative feedback involving caveolin-1. Hepatology. 2006;43:182–190. doi: 10.1002/hep.20940. [DOI] [PubMed] [Google Scholar]
  86. Kawabe JI, Grant BS, Yamamoto M, Schwencke C, Okumura S, Ishikawa Y. Changes in caveolin subtype protein expression in aging rat organs. Mol Cell Endocrinol. 2001;176:91–95. doi: 10.1016/s0303-7207(01)00472-5. [DOI] [PubMed] [Google Scholar]
  87. Kido M, Otani H, Kyoi S, Sumida T, Fujiwara H, Okada T, et al. Ischemic preconditioning-mediated restoration of membrane dystrophin during reperfusion correlates with protection against contraction-induced myocardial injury. Am J Physiol Heart Circ Physiol. 2004;287:H81–H90. doi: 10.1152/ajpheart.01140.2003. [DOI] [PubMed] [Google Scholar]
  88. Kim HA, Kim KH, Lee RA. Expression of caveolin-1 is correlated with Akt-1 in colorectal cancer tissues. Exp Mol Pathol. 2006;80:165–170. doi: 10.1016/j.yexmp.2005.09.001. [DOI] [PubMed] [Google Scholar]
  89. Kim HS, Kim HJ, Kim YS, Park SC, Harris R, Kim CK. Caveolin, GLUT4 and insulin receptor protein content in human arm and leg muscles. Eur J Appl Physiol. 2009;106:173–179. doi: 10.1007/s00421-009-1001-1. [DOI] [PubMed] [Google Scholar]
  90. Krajewska WM, Maslowska I. Caveolins: structure and function in signal transduction. Cell Mol Biol Lett. 2004;9:195–220. [PubMed] [Google Scholar]
  91. Lakatta EG, Sollott SJ. The ‘heartbreak’ of older age. Mol Interv. 2002;2:431–446. doi: 10.1124/mi.2.7.431. [DOI] [PubMed] [Google Scholar]
  92. Lasley R, Prakash N, Uittenbogaard A, Smart E. Activated cardiac adenosine A1 receptors translocate out of caveolae. J Biol Chem. 2000;275:4417–4421. doi: 10.1074/jbc.275.6.4417. [DOI] [PubMed] [Google Scholar]
  93. Lesnefsky EJ, Hoppel CL. Cardiolipin as an oxidative target in cardiac mitochondria in the aged rat. Biochim Biophys Acta. 2008;1777:1020–1027. doi: 10.1016/j.bbabio.2008.05.444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Lesnefsky EJ, Gudz TI, Migita CT, Ikeda-Saito M, Hassan MO, Turkaly PJ, et al. Ischemic injury to mitochondrial electron transport in the aging heart: damage to the iron-sulfur protein subunit of electron transport complex III. Arch Biochem Biophys. 2001;385:117–128. doi: 10.1006/abbi.2000.2066. [DOI] [PubMed] [Google Scholar]
  95. Lesnefsky EJ, He D, Moghaddas S, Hoppel CL. Reversal of mitochondrial defects before ischemia protects the aged heart. FASEB J. 2006;20:1543–1545. doi: 10.1096/fj.05-4535fje. [DOI] [PubMed] [Google Scholar]
  96. Li S, Seitz R, Lisanti MP. Phosphorylation of caveolin by src tyrosine kinases. The alpha-isoform of caveolin is selectively phosphorylated by v-Src in vivo. J Biol Chem. 1996;271:3863–3868. [PubMed] [Google Scholar]
  97. Li WP, Liu P, Pilcher BK, Anderson RG. Cell-specific targeting of caveolin-1 to caveolae, secretory vesicles, cytoplasm or mitochondria. J Cell Sci. 2001;114:1397–1408. doi: 10.1242/jcs.114.7.1397. [DOI] [PubMed] [Google Scholar]
  98. Lisanti MP, Scherer PE, Tang Z, Sargiacomo M. Caveolae, caveolin and caveolin-rich membrane domains: a signalling hypothesis. Trends Cell Biol. 1994;4:231–235. doi: 10.1016/0962-8924(94)90114-7. [DOI] [PubMed] [Google Scholar]
  99. Liu L, Pilch PF. A critical role of cavin (polymerase I and transcript release factor) in caveolae formation and organization. J Biol Chem. 2008a;283:4314–4322. doi: 10.1074/jbc.M707890200. [DOI] [PubMed] [Google Scholar]
  100. Liu L, Brown D, McKee M, Lebrasseur NK, Yang D, Albrecht KH, et al. Deletion of Cavin/PTRF causes global loss of caveolae, dyslipidemia, and glucose intolerance. Cell Metab. 2008b;8:310–317. doi: 10.1016/j.cmet.2008.07.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Llano-Diez M, Gustafson AM, Olsson C, Goransson H, Larsson L. Muscle wasting and the temporal gene expression pattern in a novel rat intensive care unit model. BMC Genomics. 2011;12:602. doi: 10.1186/1471-2164-12-602. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Masoro EJ. Caloric restriction-induced life extension of rats and mice: a critique of proposed mechanisms. Biochim Biophys Acta. 2009;1790:1040–1048. doi: 10.1016/j.bbagen.2009.02.011. [DOI] [PubMed] [Google Scholar]
  103. McCay C, Crowell M, Maynard L. The effect of retarded growth upon the length of life and upon the ultimate body size. J Nutr. 1935;10:63–79. [PubMed] [Google Scholar]
  104. Mendez-Gimenez L, Rodriguez A, Balaguer I, Fruhbeck G. Role of aquaglyceroporins and caveolins in energy and metabolic homeostasis. Mol Cell Endocrinol. 2014;397:78–92. doi: 10.1016/j.mce.2014.06.017. [DOI] [PubMed] [Google Scholar]
  105. Mio Y, Bienengraeber MW, Marinovic J, Gutterman DD, Rakic M, Bosnjak ZJ, et al. Age-related attenuation of isoflurane preconditioning in human atrial cardiomyocytes: roles for mitochondrial respiration and sarcolemmal adenosine triphosphate-sensitive potassium channel activity. Anesthesiology. 2008;108:612–620. doi: 10.1097/ALN.0b013e318167af2d. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Miyamae M, Fujiwara H, Tanaka M, Yokota R, Takemura G, Itoh S, et al. Oxygen radicals mediate ultrastructural and metabolic protection of preconditioning in vivo in pig hearts. Exp Clin Cardiol. 2002;7:173–179. [PMC free article] [PubMed] [Google Scholar]
  107. Mizushige K, Yao L, Noma T, Kiyomoto H, Yu Y, Hosomi N, et al. Alteration in left ventricular diastolic filling and accumulation of myocardial collagen at insulin-resistant prediabetic stage of a type II diabetic rat model. Circulation. 2000;101:899–907. doi: 10.1161/01.cir.101.8.899. [DOI] [PubMed] [Google Scholar]
  108. Moolman JA, Genade S, Winterbach R, Harper IS, Williams K, Lochner A. Preconditioning with a single short episode of global ischemia in the isolated working rat heart: effect on structure, mechanical function, and energy metabolism for various durations of sustained global ischemia. Cardiovasc Drugs Ther. 1995;9:103–115. doi: 10.1007/BF00877750. [DOI] [PubMed] [Google Scholar]
  109. Mousley CJ, Yuan P, Gaur NA, Trettin KD, Nile AH, Deminoff SJ, et al. A sterol-binding protein integrates endosomal lipid metabolism with TOR signaling and nitrogen sensing. Cell. 2012;148:702–715. doi: 10.1016/j.cell.2011.12.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Mukherjee S, Tessema M, Wandinger-Ness A. Vesicular trafficking of tyrosine kinase receptors and associated proteins in the regulation of signaling and vascular function. Circ Res. 2006;98:743–756. doi: 10.1161/01.RES.0000214545.99387.e3. [DOI] [PubMed] [Google Scholar]
  111. Murry CE, Jennings RB, Reimer KA. Preconditioning with ischemia: a delay of lethal cell injury in ischemic myocardium. Circulation. 1986;74:1124–1136. doi: 10.1161/01.cir.74.5.1124. [DOI] [PubMed] [Google Scholar]
  112. Murry CE, Richard VJ, Reimer KA, Jennings RB. Ischemic preconditioning slows energy metabolism and delays ultrastructural damage during a sustained ischemic episode. Circ Res. 1990;66:913–931. doi: 10.1161/01.res.66.4.913. [DOI] [PubMed] [Google Scholar]
  113. Nichols B. Caveosomes and endocytosis of lipid rafts. J Cell Sci. 2003;116:4707–4714. doi: 10.1242/jcs.00840. [DOI] [PubMed] [Google Scholar]
  114. Niesman IR, Zemke N, Fridolfsson HN, Haushalter KJ, Levy K, Grove A, et al. Caveolin isoform switching as a molecular, structural, and metabolic regulator of microglia. Mol Cell Neurosci. 2013;56:283–297. doi: 10.1016/j.mcn.2013.07.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Nishihara M, Miura T, Miki T, Tanno M, Yano T, Naitoh K, et al. Modulation of the mitochondrial permeability transition pore complex in GSK-3beta-mediated myocardial protection. J Mol Cell Cardiol. 2007;43:564–570. doi: 10.1016/j.yjmcc.2007.08.010. [DOI] [PubMed] [Google Scholar]
  116. Olichon A, Baricault L, Gas N, Guillou E, Valette A, Belenguer P, et al. Loss of OPA1 perturbates the mitochondrial inner membrane structure and integrity, leading to cytochrome c release and apoptosis. J Biol Chem. 2003;278:7743–7746. doi: 10.1074/jbc.C200677200. [DOI] [PubMed] [Google Scholar]
  117. Oliveira PJ, Seica R, Coxito PM, Rolo AP, Palmeira CM, Santos MS, et al. Enhanced permeability transition explains the reduced calcium uptake in cardiac mitochondria from streptozotocin-induced diabetic rats. FEBS Lett. 2003;554:511–514. doi: 10.1016/s0014-5793(03)01233-x. [DOI] [PubMed] [Google Scholar]
  118. Oshikawa J, Otsu K, Toya Y, Tsunematsu T, Hankins R, Kawabe J, et al. Insulin resistance in skeletal muscles of caveolin-3-null mice. Proc Natl Acad Sci U S A. 2004;101:12670–12675. doi: 10.1073/pnas.0402053101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Otsu K, Toya Y, Oshikawa J, Kurotani R, Yazawa T, Sato M, et al. Caveolin gene transfer improves glucose metabolism in diabetic mice. Am J Physiol Cell Physiol. 2009;298:C450–C456. doi: 10.1152/ajpcell.00077.2009. [DOI] [PubMed] [Google Scholar]
  120. Pain T, Yang X-M, Critz SD, Yue Y, Nakano A, Liu GS, et al. Opening of mitochondrial KATP channels triggers the preconditioned state by generating free radicals. Circ Res. 2000;87:460–466. doi: 10.1161/01.res.87.6.460. [DOI] [PubMed] [Google Scholar]
  121. Palade G. Fine structure of blood capillaries. J Appl Phys. 1953;24:1424. [Google Scholar]
  122. Park DS, Cohen AW, Frank PG, Razani B, Lee H, Williams TM, et al. Caveolin-1 null (-/-) mice show dramatic reductions in life span. Biochemistry. 2003;42:15124–15131. doi: 10.1021/bi0356348. [DOI] [PubMed] [Google Scholar]
  123. Park SC, Cho KA, Jang IS, Kim KT, Ryu SJ. Functional efficiency of the senescent cells: replace or restore? Ann N Y Acad Sci. 2004;1019:309–316. doi: 10.1196/annals.1297.052. [DOI] [PubMed] [Google Scholar]
  124. Parton RG, Way M, Zorzi N, Stang E. Caveolin-3 associates with developing T-tubules during muscle differentiation. J Cell Biol. 1997;136:137–154. doi: 10.1083/jcb.136.1.137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Patel HH, Head BP, Petersen HN, Niesman IR, Huang D, Gross GJ, et al. Protection of adult rat cardiac myocytes from ischemic cell death: role of caveolar microdomains and delta-opioid receptors. Am J Physiol Heart Circ Physiol. 2006;291:H344–H350. doi: 10.1152/ajpheart.01100.2005. [DOI] [PubMed] [Google Scholar]
  126. Patel HH, Tsutsumi YM, Head BP, Niesman IR, Jennings M, Horikawa Y, et al. Mechanisms of cardiac protection from ischemia/reperfusion injury: a role for caveolae and caveolin-1. FASEB J. 2007;21:1565–1574. doi: 10.1096/fj.06-7719com. [DOI] [PubMed] [Google Scholar]
  127. Pavlides S, Tsirigos A, Vera I, Flomenberg N, Frank PG, Casimiro MC, et al. Loss of stromal caveolin-1 leads to oxidative stress, mimics hypoxia and drives inflammation in the tumor microenvironment, conferring the ‘reverse Warburg effect’: a transcriptional informatics analysis with validation. Cell Cycle. 2010;9:2201–2219. doi: 10.4161/cc.9.11.11848. [DOI] [PubMed] [Google Scholar]
  128. Pawson AJ, Sharman JL, Benson HE, Faccenda E, Alexander SP, Buneman OP, et al. NC-IUPHAR. The IUPHAR/BPS Guide to PHARMACOLOGY: an expert-driven knowledge base of drug targets and their ligands. Nucleic Acids Res. 2014;42(Database Issue):D1098–1106. doi: 10.1093/nar/gkt1143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Peart JN, Gross ER, Headrick JP, Gross GJ. Impaired p38 MAPK/HSP27 signaling underlies aging-related failure in opioid-mediated cardioprotection. J Mol Cell Cardiol. 2007;42:972–980. doi: 10.1016/j.yjmcc.2007.02.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Peart JN, Pepe S, Reichelt ME, Beckett N, See Hoe L, Ozberk V, et al. Dysfunctional survival-signaling and stress-intolerance in aged murine and human myocardium. Exp Gerontol. 2014;50:72–81. doi: 10.1016/j.exger.2013.11.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Peng Y, Akmentin W, Connelly MA, Lund-Katz S, Phillips MC, Williams DL. Scavenger receptor BI (SR-BI) clustered on microvillar extensions suggests that this plasma membrane domain is a way station for cholesterol trafficking between cells and high-density lipoprotein. Mol Biol Cell. 2004;15:384–396. doi: 10.1091/mbc.E03-06-0445. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Penna C, Perrelli MG, Raimondo S, Tullio F, Merlino A, Moro F, et al. Postconditioning induces an anti-apoptotic effect and preserves mitochondrial integrity in isolated rat hearts. Biochim Biophys Acta. 2009;1787:794–801. doi: 10.1016/j.bbabio.2009.03.013. [DOI] [PubMed] [Google Scholar]
  133. Penumathsa SV, Thirunavukkarasu M, Zhan L, Maulik G, Menon VP, Bagchi D, et al. Resveratrol enhances GLUT-4 translocation to the caveolar lipid raft fractions through AMPK/Akt/eNOS signalling pathway in diabetic myocardium. J Cell Mol Med. 2008;12:2350–2361. doi: 10.1111/j.1582-4934.2008.00251.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Pepe S. Effect of dietary polyunsaturated fatty acids on age-related changes in cardiac mitochondrial membranes. Exp Gerontol. 2005;40:751–758. doi: 10.1016/j.exger.2005.03.013. [DOI] [PubMed] [Google Scholar]
  135. Pike LJ. Lipid rafts: bringing order to chaos. J Lipid Res. 2003;44:655–667. doi: 10.1194/jlr.R200021-JLR200. [DOI] [PubMed] [Google Scholar]
  136. Pike LJ. Growth factor receptors, lipid rafts and caveolae: an evolving story. Biochim Biophys Acta. 2005;1746:260–273. doi: 10.1016/j.bbamcr.2005.05.005. [DOI] [PubMed] [Google Scholar]
  137. Poirier P, Bogaty P, Garneau C, Marois L, Dumesnil JG. Diastolic dysfunction in normotensive men with well-controlled type 2 diabetes: importance of maneuvers in echocardiographic screening for preclinical diabetic cardiomyopathy. Diabetes Care. 2001;24:5–10. doi: 10.2337/diacare.24.1.5. [DOI] [PubMed] [Google Scholar]
  138. Pojoga LH, Underwood PC, Goodarzi MO, Williams JS, Adler GK, Jeunemaitre X, et al. Variants of the caveolin-1 gene: a translational investigation linking insulin resistance and hypertension. J Clin Endocrinol Metab. 2011;96:E1288–E1292. doi: 10.1210/jc.2010-2738. [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Quarrie R, Lee DS, Steinbaugh G, Cramer B, Erdahl W, Pfeiffer DR, et al. Ischemic preconditioning preserves mitochondrial membrane potential and limits reactive oxygen species production. J Surg Res. 2012;178:8–17. doi: 10.1016/j.jss.2012.05.090. [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Raikar LS, Vallejo J, Lloyd PG, Hardin CD. Overexpression of caveolin-1 results in increased plasma membrane targeting of glycolytic enzymes: the structural basis for a membrane associated metabolic compartment. J Cell Biochem. 2006;98:861–871. doi: 10.1002/jcb.20732. [DOI] [PubMed] [Google Scholar]
  141. Ralston E, Ploug T. Caveolin-3 is associated with the T-tubules of mature skeletal muscle fibers. Exp Cell Res. 1999;246:510–515. doi: 10.1006/excr.1998.4305. [DOI] [PubMed] [Google Scholar]
  142. Ratajczak P, Damy T, Heymes C, Oliviero P, Marotte F, Robidel E, et al. Caveolin-1 and -3 dissociations from caveolae to cytosol in the heart during aging and after myocardial infarction in rat. Cardiovasc Res. 2003;57:358–369. doi: 10.1016/s0008-6363(02)00660-0. [DOI] [PubMed] [Google Scholar]
  143. Razzaq TM, Ozegbe P, Jury EC, Sembi P, Blackwell NM, Kabouridis PS. Regulation of T-cell receptor signalling by membrane microdomains. Immunology. 2004;113:413–426. doi: 10.1111/j.1365-2567.2004.01998.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Rodgers W, Farris D, Mishra S. Merging complexes: properties of membrane raft assembly during lymphocyte signaling. Trends Immunol. 2005;26:97–103. doi: 10.1016/j.it.2004.11.016. [DOI] [PubMed] [Google Scholar]
  145. Roepstorff C, Vistisen B, Roepstorff K, Kiens B. Regulation of plasma long-chain fatty acid oxidation in relation to uptake in human skeletal muscle during exercise. Am J Physiol Endocrinol Metab. 2004;287:E696–E705. doi: 10.1152/ajpendo.00001.2004. [DOI] [PubMed] [Google Scholar]
  146. Rothberg KG, Heuser JE, Donzell WC, Ying YS, Glenney JR, Anderson RG. Caveolin, a protein component of caveolae membrane coats. Cell. 1992;68:673–682. doi: 10.1016/0092-8674(92)90143-z. [DOI] [PubMed] [Google Scholar]
  147. Rudnick J, Puttmann B, Tesch PA, Alkner B, Schoser BG, Salanova M, et al. Differential expression of nitric oxide synthases (NOS 1–3) in human skeletal muscle following exercise countermeasure during 12 weeks of bed rest. FASEB J. 2004;18:1228–1230. doi: 10.1096/fj.03-0792fje. [DOI] [PubMed] [Google Scholar]
  148. Ruiz-Hurtado G, Fernandez-Velasco M, Mourelle M, Delgado C. LA419, a novel nitric oxide donor, prevents pathological cardiac remodeling in pressure-overloaded rats via endothelial nitric oxide synthase pathway regulation. Hypertension. 2007;50:1049–1056. doi: 10.1161/HYPERTENSIONAHA.107.093666. [DOI] [PubMed] [Google Scholar]
  149. Rybin VO, Grabham PW, Elouardighi H, Steinberg SF. Caveolae-associated proteins in cardiomyocytes: caveolin-2 expression and interactions with caveolin-3. Am J Physiol Heart Circ Physiol. 2003;285:H325–H332. doi: 10.1152/ajpheart.00946.2002. [DOI] [PubMed] [Google Scholar]
  150. Salani B, Briatore L, Garibaldi S, Cordera R, Maggi D. Caveolin-1 down-regulation inhibits insulin-like growth factor-I receptor signal transduction in H9C2 rat cardiomyoblasts. Endocrinology. 2008;149:461–465. doi: 10.1210/en.2007-0312. [DOI] [PubMed] [Google Scholar]
  151. Sargiacomo M, Scherer PE, Tang Z, Kübler E, Song KS, Sanders MC, et al. Oligomeric structure of caveolin: implications for caveolae membrane organization. Proc Natl Acad Sci U S A. 1995;92:9407–9411. doi: 10.1073/pnas.92.20.9407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Scherer PE, Lewis RY, Volonte D, Engelman JA, Galbiati F, Couet J, et al. Cell-type and tissue-specific expression of caveolin-2. Caveolins 1 and 2 co-localize and form a stable hetero-oligomeric complex in vivo. J Biol Chem. 1997;272:29337–29346. doi: 10.1074/jbc.272.46.29337. [DOI] [PubMed] [Google Scholar]
  153. Schilling JM, Cui W, Godoy JC, Risbrough VB, Niesman IR, Roth DM, et al. Long-term atorvastatin treatment leads to alterations in behavior, cognition, and hippocampal biochemistry. Behav Brain Res. 2014;267:6–11. doi: 10.1016/j.bbr.2014.03.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Scriven DR, Klimek A, Lee KL, Moore ED. The molecular architecture of calcium microdomains in rat cardiomyocytes. Ann N Y Acad Sci. 2002;976:488–499. doi: 10.1111/j.1749-6632.2002.tb04783.x. [DOI] [PubMed] [Google Scholar]
  155. Scriven DR, Klimek A, Asghari P, Bellve K, Moore ED. Caveolin-3 is adjacent to a group of extradyadic ryanodine receptors. Biophys J. 2005;89:1893–1901. doi: 10.1529/biophysj.105.064212. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Seo AY, Xu J, Servais S, Hofer T, Marzetti E, Wohlgemuth SE, et al. Mitochondrial iron accumulation with age and functional consequences. Aging Cell. 2008;7:706–716. doi: 10.1111/j.1474-9726.2008.00418.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  157. Shaul PW, Anderson RG. Role of plasmalemmal caveolae in signal transduction. Am J Physiol Lung Cell Mol Physiol. 1998;275:L843–L851. doi: 10.1152/ajplung.1998.275.5.L843. [DOI] [PubMed] [Google Scholar]
  158. Song KS, Scherer PE, Tang Z, Okamoto T, Li S, Chafel M, et al. Expression of caveolin-3 in skeletal, cardiac, and smooth muscle cells. Caveolin-3 is a component of the sarcolemma and co-fractionates with dystrophin and dystrophin-associated glycoproteins. J Biol Chem. 1996;271:15160–15165. doi: 10.1074/jbc.271.25.15160. [DOI] [PubMed] [Google Scholar]
  159. Steinberg SF, Brunton LL. Compartmentation of g protein-coupled signaling pathways in cardiac myocytes. Annu Rev Pharmacol Toxicol. 2001;41:751–773. doi: 10.1146/annurev.pharmtox.41.1.751. [DOI] [PubMed] [Google Scholar]
  160. Stojanovski D, Koutsopoulos OS, Okamoto K, Ryan MT. Levels of human Fis1 at the mitochondrial outer membrane regulate mitochondrial morphology. J Cell Sci. 2004;117:1201–1210. doi: 10.1242/jcs.01058. [DOI] [PubMed] [Google Scholar]
  161. Swynghedauw B, Besse S, Assayag P, Carre F, Chevalier B, Charlemagne D, et al. Molecular and cellular biology of the senescent hypertrophied and failing heart. Am J Cardiol. 1995;76:2D–7D. doi: 10.1016/s0002-9149(99)80484-6. [DOI] [PubMed] [Google Scholar]
  162. Tang Z, Scherer PE, Okamoto T, Song K, Chu C, Kohtz DS, et al. Molecular cloning of caveolin-3, a novel member of the caveolin gene family expressed predominantly in muscle. J Biol Chem. 1996;271:2255–2261. doi: 10.1074/jbc.271.4.2255. [DOI] [PubMed] [Google Scholar]
  163. Tani M, Honma Y, Hasegawa H, Tamaki K. Direct activation of mitochondrial KATP channels mimics preconditioning but protein kinase C activation is less effective in middle-aged rat hearts. Cardiovasc Res. 2001;49:56–68. doi: 10.1016/s0008-6363(00)00240-6. [DOI] [PubMed] [Google Scholar]
  164. Tsutsumi YM, Horikawa YT, Jennings MM, Kidd MW, Niesman IR, Yokoyama U, et al. Cardiac-specific overexpression of caveolin-3 induces endogenous cardiac protection by mimicking ischemic preconditioning. Circulation. 2008;118:1979–1988. doi: 10.1161/CIRCULATIONAHA.108.788331. [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Tsutsumi YM, Kawaraguchi Y, Niesman IR, Patel HH, Roth DM. Opioid-induced preconditioning is dependent on caveolin-3 expression. Anesth Analg. 2010a;111:1117–1121. doi: 10.1213/ANE.0b013e3181f3351a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Tsutsumi YM, Kawaraguchi Y, Horikawa YT, Niesman IR, Kidd MW, Chin-Lee B, et al. Role of caveolin-3 and glucose transporter-4 in isoflurane-induced delayed cardiac protection. Anesthesiology. 2010b;112:1136–1145. doi: 10.1097/ALN.0b013e3181d3d624. [DOI] [PMC free article] [PubMed] [Google Scholar]
  167. VandenHoek TL, Becker LB, Shao Z, Li C, Schumacker PT. Reactive oxygen species released from mitochondria during brief hypoxia induce preconditioning in cardiomyocytes. J Biol Chem. 1998;273:18092–18098. doi: 10.1074/jbc.273.29.18092. [DOI] [PubMed] [Google Scholar]
  168. VandenHoek TL, Becker LB, Shao Z-H, Li C-Q, Schumacker PT. Preconditioning in cardiomyocytes protects by attenuating oxidant stress at reperfusion. Circ Res. 2000;86:541–548. doi: 10.1161/01.res.86.5.541. [DOI] [PubMed] [Google Scholar]
  169. Vatta M, Ackerman MJ, Ye B, Makielski JC, Ughanze EE, Taylor EW, et al. Mutant caveolin-3 induces persistent late sodium current and is associated with long-QT syndrome. Circulation. 2006;114:2104–2112. doi: 10.1161/CIRCULATIONAHA.106.635268. [DOI] [PubMed] [Google Scholar]
  170. Vinten J, Johnsen AH, Roepstorff P, Harpoth J, Tranum-Jensen J. Identification of a major protein on the cytosolic face of caveolae. Biochim Biophys Acta. 2005;1717:34–40. doi: 10.1016/j.bbamem.2005.09.013. [DOI] [PubMed] [Google Scholar]
  171. Volonte D, Zhang K, Lisanti MP, Galbiati F. Expression of caveolin-1 induces premature cellular senescence in primary cultures of murine fibroblasts. Mol Biol Cell. 2002;13:2502–2517. doi: 10.1091/mbc.01-11-0529. [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Wang J, Schilling JM, Niesman IR, Headrick JP, Finley JC, Kwan E, et al. Cardioprotective trafficking of caveolin to mitochondria is Gi-protein dependent. Anesthesiology. 2014;121:538–548. doi: 10.1097/ALN.0000000000000295. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Warley A, Powell JM, Skepper JN. Capillary surface area is reduced and tissue thickness from capillaries to myocytes is increased in the left ventricle of streptozotocin-diabetic rats. Diabetologia. 1995;38:413–421. doi: 10.1007/BF00410278. [DOI] [PubMed] [Google Scholar]
  174. Wernstedt Asterholm I, Mundy DI, Weng J, Anderson RG, Scherer PE. Altered mitochondrial function and metabolic inflexibility associated with loss of caveolin-1. Cell Metab. 2012;15:171–185. doi: 10.1016/j.cmet.2012.01.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  175. Willems L, Zatta A, Holmgren K, Ashton KJ, Headrick JP. Age-related changes in ischemic tolerance in male and female mouse hearts. J Mol Cell Cardiol. 2005;38:245–256. doi: 10.1016/j.yjmcc.2004.09.014. [DOI] [PubMed] [Google Scholar]
  176. Williams TM, Lisanti MP. The caveolin proteins. Genome Biol. 2004;5:214. doi: 10.1186/gb-2004-5-3-214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Woodman SE, Park DS, Cohen AW, Cheung M, Chandra M, Shirani J, et al. Caveolin-3 knock-out mice develop a progressive cardiomyopathy and show hyperactivation of the p42/44 MAP kinase cascade. J Biol Chem. 2002;277:38988–38997. doi: 10.1074/jbc.M205511200. [DOI] [PubMed] [Google Scholar]
  178. Woodman SE, Sotgia F, Galbiati F, Minetti C, Lisanti MP. Caveolinopathies: mutations in caveolin-3 cause four distinct autosomal dominant muscle diseases. Neurology. 2004;62:538–543. doi: 10.1212/wnl.62.4.538. [DOI] [PubMed] [Google Scholar]
  179. Yaffe MP. The machinery of mitochondrial inheritance and behavior. Science. 1999;283:1493–1497. doi: 10.1126/science.283.5407.1493. [DOI] [PubMed] [Google Scholar]
  180. Yamabhai M, Anderson RG. Second cysteine-rich region of epidermal growth factor receptor contains targeting information for caveolae/rafts. J Biol Chem. 2002;277:24843–24846. doi: 10.1074/jbc.C200277200. [DOI] [PubMed] [Google Scholar]
  181. Yamamoto M, Toya Y, Schwencke C, Lisanti MP, Myers MG, Jr, Ishikawa Y. Caveolin is an activator of insulin receptor signaling. J Biol Chem. 1998;273:26962–26968. doi: 10.1074/jbc.273.41.26962. [DOI] [PubMed] [Google Scholar]
  182. Yan LJ, Sohal RS. Mitochondrial adenine nucleotide translocase is modified oxidatively during aging. Proc Natl Acad Sci U S A. 1998;95:12896–12901. doi: 10.1073/pnas.95.22.12896. [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Yang KC, Rutledge CA, Mao M, Bakhshi FR, Xie A, Liu H, et al. Caveolin-1 modulates cardiac gap junction homeostasis and arrhythmogenecity by regulating cSrc tyrosine kinase. Circ Arrhythm Electrophysiol. 2014;7:701–710. doi: 10.1161/CIRCEP.113.001394. [DOI] [PMC free article] [PubMed] [Google Scholar]
  184. Yao Z, Tong J, Tan X, Li C, Shao Z, Kim WC, et al. Role of reactive oxygen species in acetylcholine-induced preconditioning in cardiomyocytes. Am J Physiol Heart Circ Physiol. 1999;277:H2504–H2509. doi: 10.1152/ajpheart.1999.277.6.H2504. [DOI] [PubMed] [Google Scholar]
  185. Young LH, Ikeda Y, Lefer AM. Caveolin-1 peptide exerts cardioprotective effects in myocardial ischemia-reperfusion via nitric oxide mechanism. Am J Physiol Heart Circ Physiol. 2001;280:H2489–H2495. doi: 10.1152/ajpheart.2001.280.6.H2489. [DOI] [PubMed] [Google Scholar]
  186. Zhang WZ. An association of metabolic syndrome constellation with cellular membrane caveolae. Pathobiol Aging Age Relat Dis. 2014;4:23866. doi: 10.3402/pba.v4.23866. . doi: 10.3402/pba.v4.23866; eCollection 2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  187. Zhang X, Chen C. A new insight of mechanisms, diagnosis and treatment of diabetic cardiomyopathy. Endocrine. 2012;41:398–409. doi: 10.1007/s12020-012-9623-1. [DOI] [PubMed] [Google Scholar]
  188. Zhong N, Zhang Y, Zhu HF, Zhou ZN. Intermittent hypoxia exposure prevents mtDNA deletion and mitochondrial structure damage produced by ischemia/reperfusion injury. Sheng Li Xue Bao. 2000;52:375–380. [PubMed] [Google Scholar]

Articles from British Journal of Pharmacology are provided here courtesy of The British Pharmacological Society

RESOURCES