Skip to main content
American Journal of Physiology - Lung Cellular and Molecular Physiology logoLink to American Journal of Physiology - Lung Cellular and Molecular Physiology
. 2010 Nov 19;300(2):L151–L160. doi: 10.1152/ajplung.00170.2010

Caveolin-1: a critical regulator of lung injury

Yang Jin 1,, Seon-Jin Lee 1, Richard D Minshall 2, Augustine M K Choi 1
PMCID: PMC4380484  PMID: 21097526

Abstract

Caveolin-1 (cav-1), a 22-kDa transmembrane scaffolding protein, is the principal structural component of caveolae. Cav-1 regulates critical cell functions including proliferation, apoptosis, cell differentiation, and transcytosis via diverse signaling pathways. Abundant in almost every cell type in the lung, including type I epithelial cells, endothelial cells, smooth muscle cells, fibroblasts, macrophages, and neutrophils, cav-1 plays a crucial role in the pathogenesis of acute lung injury (ALI). ALI and its severe form, acute respiratory distress syndrome (ARDS), are responsible for significant morbidity and mortality in intensive care units, despite improvement in ventilation strategies. The pathogenesis of ARDS is still poorly understood, and therapeutic options remain limited. In this article, we summarize recent data regarding the regulation and function of cav-1 in lung biology and pathology, in particular as it relates to ALI. We further discuss the potential molecular and cellular mechanisms by which cav-1 expression contributes to ALI. Investigating the cellular functions of cav-1 may provide new insights for understanding the pathogenesis of ALI and provide novel targets for therapeutic interventions in the future.

Keywords: apoptosis, inflammation, caveolae


the purpose of this review is to explore the scientific literature supporting a role for caveolin-1 (cav-1) in the pathogenesis of acute lung injury (ALI).

Caveolae are omega-shaped invaginations of plasma membrane that were first discovered by Palade (80) and Yamada (115) in the 1950s. Caveolae reside most prominently in epithelial cells, endothelial cells, adipocytes, and fibroblasts in the lung (2, 31, 86). In the past decade, the concept has emerged that caveolae are a specialized form of lipid rafts (2, 31, 80, 86, 115). Both planar lipid rafts and caveolae provide platforms that anchor membrane proteins in the “floating sea” of lipid bilayers. The structural proteins required for caveolae formation are the caveolins. These coat proteins, i.e., caveolins, include cav-1, cav-2, and cav-3 (24, 94, 96), which facilitate membrane invagination and response to external stimuli by transducing signals that modulate cellular activity (2, 13, 31, 80, 86, 115). Cav-1 and cav-2 have been identified in the majority of tissues and organs (24, 96), whereas cav-3 is mainly are expressed in cardiac, smooth, and skeletal muscle cells (24, 97). Currently, the functional role of cav-2 remains unclear, although it is expressed and heteroligomerizes with cav-1 (24, 96, 97).

Caveolin-1

Isoforms and structure.

The cav-1 gene encodes a 178-amino acid protein in a predicted β-sheet conformation (29). Two isoforms of cav-1 (cav-1α and cav-1β) have been identified (24, 29, 88, 96, 97). The 24-kDa cav-1α (residues 1–178) and 21-kDa cav-1β isoforms (residues 32–178) are generated by alternative splicing of mRNA (47). These two isoforms differ by 31 additional amino acids present in the α-isoform at its NH2 terminus (residues 1–31). The distinct function of these two isoforms remains unclear. However, recent studies in zebrafish, using antisense morpholino oligomers, reveal nonredundant and evolutionarily conserved functions for the individual cav-1 isoforms (17, 82).

As shown in Fig. 1, the sequence of cav-1 contains a hydrophobic central domain with a hairpin-like conformation inserted in the inner leaflet of the plasma membrane (101). 14–16 Cav-1 monomers form a single cav-1 homooligomer. Both the COOH and NH2 terminus of the cav-1 monomer face the cytoplasm (101). Among the modular sequences, an important domain called the caveolin scaffolding domain (CSD), located between amino acid residues 82–101 in the NH2-terminal region adjacent to the hydrophobic membrane-insertion domain, is required for caveolin dimerization and critical in controlling the interactions between cav-1 and numerous signaling proteins. The COOH-terminal region of cav-1 contains three palmitoylated cysteine residues, which are also relevant to oligomerization (101) (Fig. 1). Studies have revealed that COOH-terminal palmitoylation is crucial for cav-1 to attach the plasma membrane (32). Another key region of the cav-1 structure is tyrosine-14 (Y14). Cav-1 is phosphorylated on tyrosine-14 in response to cellular stimuli. Recent evidence indicates that phosphorylated cav-1 (pY14) is responsible for multiple biological processes including caveolae internalization and regulation of cell signaling (30).

Fig. 1.

Fig. 1.

Structure of caveolin-1. Cav-1 contains a hydrophobic central domain with a hairpin-like conformation inserted in the inner leaflet of the plasma membrane. 14–16 Cav-1 monomers form a single cav-1 homooligomer. A simplified dimmer is illustrated here. Both the COOH and NH2 terminus of the cav-1 monomer face the cytoplasm. The caveolin scaffolding domain is located between amino acid residues 82–101 in the NH2-terminal region adjacent to the hydrophobic membrane-insertion domain. The COOH-terminal region of cav-1 contains 3 palmitoylated cysteine residues. Inset: caveolae viewed by electron microscopy (×50,000).

Cav-1 is an integral membrane protein associated with various membranous structures, including endoplasmic reticulum (ER), Golgi, and plasma membranes (Fig. 1). After synthesis, cav-1 is inserted into the ER membrane with both its NH2- and COOH-terminal sequences facing the cytoplasm. In the ER, cav-1 forms homogeneous SDS-resistant homooligomers of 7–14 cav-1 molecules (81). Cav-1 is then exported to the Golgi and transported to the cell surface to form caveolae. Remarkably, cav-1 can be recycled in reverse order from the cell plasma membrane (4, 81). Interestingly, cav-1 is also found in the cytoplasm as a soluble protein as well as in the mitochondria, secretary vesicles, cell-cell contact sites, and the nucleus (1). Four regions of cav-1 have been identified to influence its intracellular trafficking, including: 1) a conserved region between amino acids 66 and 70, which is necessary for exit from the ER, 2) a region between amino acids 71–80, which regulates incorporation of cav-1 oligomers into the Golgi apparatus, 3) a region between amino acids 91–100, and 4) COOH-terminal amino acids 134–154, which are thought to control cav-1 oligomerization and exit from the Golgi apparatus (96).

Regulation of cav-1 expression.

Cav-1 expression is regulated transcriptionally or posttranslationally. The cav-1 5′-flanking region includes three G+C-rich sites which are potential sterol regulatory elements (SREs). Additionally, the 5′-flanking region contains a CAAT sequence and a Sp1 consensus sequence (4). SRE-like elements are involved in the transcriptional response to stimuli, such as low-density lipoprotein-free cholesterol (LDL-FC) (89, 114). Other transcription factors reported to regulate cav-1 expression include the forkhead (FKHR) family of transcription factors, FOXO3a (20, 105), C-myc (89, 114), and NF-κB (104).

Downregulation of cav-1 can be achieved by promoting cav-1 degradation via lysosomal degradation pathways (9). For example, in intestinal epithelial cells, cav-1 protein level is regulated by another lipid raft protein, flotillin-1 (flot1), by preventing its lysosomal degradation (106). Cav-1 is also regulated by other lipid raft component proteins, for example, the cavins. Recent studies indicate that deletion of cavin-1 diminishes cav-1 protein expression without affecting cav-1 mRNA (36), and vice versa, deletion of cav-1 abolishes cavin-1 expression (34). These results suggest that cav-1 and cavin-1 are posttranslationally regulated by degradation and also by transcriptional regulation of mRNA levels.

Cav-1 functions.

Initially described decades ago, cav-1 is the primary protein of caveolae. Although the role of cav-2 remains unclear, the function of cav-1 has been studied extensively. Many of these functions, mentioned below, are regulated by cav-1 posttranslational modifications such as palmitoylation at the three cysteine sites in the COOH terminus and phosphorylation of NH2-terminal tyrosine Y14 and serine-80 near the CSD (48).

FORMATION OF CAVEOLAE.

Cav-1 is essential for caveolae formation. Deletion of cav-1 results in absence of caveolae (84). As expected, overexpression of cav-1 leads to an increase in the number of caveolae (56). Cav-1 is thought to be the most important structural protein required for caveolae formation, although recent data suggest that the cavins also play important roles in regulating the architecture of caveolae (33).

CONTROL OF CHOLESTEROL HOMEOSTASIS.

Cav-1 directly binds cholesterol and long-chain unsaturated fatty acids and forms membrane-associated oligomers (21). Growing data suggest that cav-1 controls the import and export of cellular cholesterol by caveolae (18). Furthermore, cav-1 coordinates lipid metabolism (19). However, cav-1 has been shown to play both proatherogenic and antiatherogenic roles, depending on the cell type studied (23). In smooth muscle cells, cav-1 suppresses cell proliferation and may have antiatherogenic effects, whereas in endothelial cells, cav-1 promotes transcytosis of LDL-cholesterol particles (42).

REGULATION OF MEMBRANE TRAFFICKING, ENDOCYTOSIS, EXOCYTOSIS, AND TRANSCYTOSIS.

Cav-1 interacts with many receptor tyrosine kinases, such as EGF receptor (EGFR) as well as nonreceptor tyrosine kinases such as Src as well as serine/threonine kinases such as PKC family members (110), which play important roles in membrane trafficking. Endocytosis, exocytosis, and transcytosis of many macromolecules via caveolae require the presence of cav-1 (3). Examples of macromolecule transport include albumin, cholera toxin, and tetanus toxin (52, 71). Among these, albumin uptake by lung endothelial cells appears to be directly involved in the pathophysiology of ALI. Endothelial cell cav-1 is required for the efficient uptake and transport of albumin from the blood to the interstitium (92).

REGULATION OF CELL SIGNALING.

Cav-1 interacts with a variety of downstream signaling molecules, including endothelial nitric oxide synthase (eNOS), heterotrimeric G proteins, nonreceptor tyrosine kinases, Src-family tyrosine kinases, and p42/44 mitogen-activated protein (MAP) kinase (10, 15, 16, 20, 26, 45, 59, 98, 121). Cav-1 anchors these signal transducers in their inactive conformation until activation by appropriate stimulation (10, 15, 16, 20, 26, 45, 59, 98, 121). Many of these signaling molecules interact with the CSD directly via the hydrophobic cav-1 binding motif (ΦxxΦxxxxΦ or ΦxΦxxxxΦ, where Φ stands for aromatic amino acids). Emerging evidence demonstrates that cav-1 functions as a negative or positive regulator of cell signaling, depending on the cell type and specific cell signaling pathway investigated. For instance, as a negative regulator, cav-1 inhibits Wnt signaling by blocking β-catenin-mediated transcription (26). Cav-1 inhibits eNOS (20, 45, 98, 121), and recombinant cav-1 blocks Neu (c-erbB2)-mediated signal transduction (16). Additionally, cav-1 inhibits signaling from EGFR, Raf-1, MEK-1, and Erk2 to the nucleus. Furthermore, cav-1 peptides derived from residues 32–95 inhibit the kinase activity of purified MEK-1 and Erk2 (15, 16, 20, 26, 27, 45, 59, 98, 121). In contrast, cav-1 positively regulates integrin-dependent signaling, Shc-mediated signaling (58, 67, 112), and the phosphoinositide 3-kinase (PI3K)/Akt pathway (55, 95, 123). Cav-1 overexpression activates phospho-Akt signaling pathways in Hela cells, in prostate cancer cells, and in MCF-7 breast cancer cells (55, 95, 85).

Lung phenotype in cav-1-deficient or cav-1-transgenic mice.

Cav-1 is abundantly expressed in lung epithelia, endothelia, and fibroblasts (113). Cav-1 knockout mice (cav-1−/− mice) exhibit significant abnormalities within the lungs (14, 83, 113). In 2001, two groups, Drab et al. (14) and Razani et al. (83), independently generated cav-1-deficient mice. These initial studies and the invaluable tool of cav-1−/− mice have enhanced our understanding of the function of cav-1 both in vitro and in vivo. Cav-1−/− mice, although lacking caveolae, are viable and fertile. Cav-1−/− mice exhibit significantly lower levels of cav-2 due to posttranscriptional regulation and increased degradation of this protein in the absence of cav-1. Cav-1−/− possess multiple defects, including impaired homeostasis of lipid, which lead to dysfunctional muscle, cardiovascular, and lipidogenic tissues (14, 83, 113). Cav-1−/− mice also have a basal phenotype characterized by hyperlipidemia.

For the remainder of this review, we will focus on the pulmonary phenotype of cav-1-deficient mice as they have served as the primary tool in cav-1-related research. The high density of caveolae in the septa of lung tissue indicated that these structures play an important role in the respiratory system. Cav-1−/− mice have a thickened alveolar wall leading to constricted and irregular alveolar spaces. Although these abnormal septa contain increased amounts of endothelial and hematopoietic cells, two characteristics of this “hypercellularity” need to be recognized. 1) While hypercellularity is a notable feature of cav-1−/− mice, these cells appear to be non-differentiated cells, given that von Willebrand factor (vWF), which marks differentiated endothelial cells, is not expressed (8, 14, 83, 113). 2) The phenotype of the cav-1−/− mice, which lacks both the cav-1α and -β isoforms, shows that only the pulmonary endothelium exhibits hyperproliferation, whereas the alveolar epithelium does not (8, 14, 83, 113). In addition to hypercellularity, recently, Le Saux et al. (53) reported a progressive increase in deposition of collagen fibrils in airways and parenchyma in cav1−/− mice (53, 54). The increased deposition of collagen fibrils may lead to significantly reduced lung compliance and increased elasticity and airway resistance. This process is associated with increased TGF-β/Smad signaling pathways involved in tropoelastin, col1 α2, and col3 α1 gene expression in lung tissues (53, 116). As a consequence of the pulmonary abnormalities, cav-1−/− mice are exercise intolerant and show early onset exhaustion during swimming tests (53, 116). Based on the above-described features, presumably, cav-1−/− mice would be more susceptible to ALI. Surprisingly, numerous recent studies in cav-1−/− mice do not fully support this hypothesis. Therefore, we will review the scientific literature exploring the function of cav-1 in the pathogenesis of ALI.

Using a different approach, Yang et al. (116) generated epithelial cell-specific cav-1 overexpressing mice via a mouse mammary tumor virus (MMTV) long terminal-repeat promoter, which is predominantly expressed in specific epithelial cells. They reported that the MMTV-cav-1(+) transgenic mice demonstrate bronchiolar epithelial hyperplasia and atypia. Additionally, the MMTV-cav-1(+) transgenic mice tend to have a greater incidence of malignancies, including lung cancer, when compared with the MMTV-cav-1(−) littermates (116).

Caveolin-1 and Lung Injury

In 1994, Lisanti et al. (58) first described the expression of caveolin proteins in lung. Since then, emerging studies have investigated the functional role of caveolins in lung diseases. ALI is a diffuse heterogeneous process characterized by diffuse alveolar cell death (DAD), acute inflammatory responses including non-cardiogenic pulmonary edema, widespread capillary leakage, cytokine release, and ROS generation leading to low lung compliance, severe hypoxia, and abnormal gas exchange (5, 41, 49, 62, 64, 69, 119). Pulmonary fibrosis develops in the later stages of ALI (6264, 69, 119). Recent studies have explored the potential roles of cav-1 in lung injury from all the above-mentioned aspects (Fig. 2). After stimuli, various cell types participate in the pathogenesis of ALI, including polymorphonuclear neutrophils (PMNs), vascular endothelial cells, lung alveolar epithelial cells, and lung fibroblasts. In the rest of this article, we will discuss current literature on the functional roles of cav-1 in each of these components involved in ALI.

Fig. 2.

Fig. 2.

Scheme of proposed roles of caveolin-1 in acute lung injury.

Cav-1 regulates acute inflammation and capillary leakage during lung injury.

Primary ALI is caused by a direct injury to the lung, such as pneumonia, ventilation-associated injury, hyperoxic injury, trauma, and contusion (5, 41, 49, 63, 64, 69, 119). Secondary ALI is caused by an indirect insult to the lung from conditions such as pancreatitis, severe sepsis, or transfusion-related ALI (5, 41, 49, 63, 64, 69, 119).

Acute inflammation has been well described in the pathological stages of ALI/ARDS, along with increased vascular permeability, fibroproliferation with hyaline membranes, epithelial apoptosis, and varying degrees of interstitial fibrosis (5, 41, 49, 6264, 69, 1119).

ENDOTHELIAL CELL CAV-1 PLAYS A CRUCIAL ROLE DURING ACUTE INFLAMMATION.

The transport of macromolecules from the blood-space to the tissue-space is regulated by endothelial cell caveolae (39, 40, 93, 99). For example, albumin is not endocytosed in cav-1−/− mouse lung endothelial cells and remains in the blood vessel lumen (99). Abundant caveolae is a distinctive feature of endothelial cells. Transmembrane water channel protein aquaporin-1 is expressed in caveolae on the lung endothelial cell surface (90). This finding supports the role of caveolae in endothelia-mediated transcellular transport of water (90). The mechanical rearrangement of caveolae/lipid rafts contribute to non-cardiogenic pulmonary edema during ALI (39, 40). Unlike the above studies which focus mainly on caveolae instead of cav-1, Sundivakkam et al. (100) studied the role of CSD in Ca2+ entry in endothelial cells. Ca2+ influx has been reported to regulate pulmonary vasoconstriction (100). CSD regulates Ca2+ store release-induced Ca2+ entry in endothelial cells, suggesting a potential role in endothelial permeability (100).

Current evidence supports a dual role of cav-1 in regulating microvascular permeability: 1) as a caveolae-associated structural protein controlling caveolar transcytosis, and 2) as a tonic inhibitor of eNOS activity, which negatively regulates paracellular permeability (99). As a result of this dual function, although the cav-1−/− mouse has lung vascular abnormalities and lung fluid balance abnormalities at baseline, upon insult, cav-1−/− mice resist ALI compared with wild-type mice. Cav-1−/− mice have improved survival following LPS challenge (60). Similar results were reported in several previous studies using LPS (28) and hyperoxia exposure (28, 44, 70). Since cav-1 is a negative regulator of eNOS, cav-1−/− mice have increased plasma NO, and, paradoxically, increased pulmonary vascular resistance (PVR) (62). Elevated PVR is attributed to pulmonary precapillary vessel remodeling, and, therefore, the elevated basal plasma NO is thought to compensate for the vascular structural abnormalities in cav-1−/− mice. Decreased pulmonary artery filling defects are found in cav-1−/− mice (62). Furthermore, less ACE MAb binding was found in the cav-1−/− mice, suggesting vascular surface area may also be reduced in cav-1−/− mice (62). Cav-1−/− mice have basal pulmonary edema, with elevated extravascular lung water, and thus this apposing tissue pressure may limit further transport and accumulation of pulmonary edema fluid from vascular damage during lung injury (unpublished observation from Dr. R. Minshall). In addition, deletion of cav-1 results in activation of eNOS and dampens Toll-like receptor 4 (TLR4) signaling, thus decreasing the innate immune response to LPS which consequently confers protection from LPS-induced inflammation (70). These results further confirmed the previous studies by Garrean et al. (28), which showed that cav-1−/− mice have a reduced inflammatory response following LPS via NF-κB-mediated pathways. Lv et al. (61) observed that overexpression of cav-1 aggravates LPS-induced inflammation in that mouse lung alveolar type 1 cells had increased inflammatory cytokine production including IL-6 and TNFα. Instead of using cav-1−/−, the “gain of function” approach used in this study further endorsed the hypothesis that cav-1 mediates LPS-induced inflammation during lung injury (61, 70). Similarly, Wang and coworkers (109) first observed cav-1 expression in murine alveolar macrophages and showed that cav-1 in murine alveolar and peritoneal macrophages regulated LPS-induced proinflammatory TNFα and IL-6 cytokine production. Deletion of cav-1 increased LPS-induced proinflammatory cytokine TNFα and IL-6 production but decreased anti-inflammatory cytokine IL-10 production. Overexpressing cav-1 in RAW264.7 cells gave rise to opposite results. Thus, this study suggests that cav-1 acts as a potent proinflammatory effector molecule in immune cells.

EVIDENCE FOR A ROLE OF CAV-1 IN PMN-MEDIATED INFLAMMATION DURING ALI.

There are two recognized phases of ALI. The initial phase is acute and is characterized by disruption of the alveolar-capillary barrier, leakage of protein-rich fluid, extensive release of cytokines, migration of neutrophils, and severe hypoxia. All of these processes likely contribute to non-cardiogenic pulmonary edema. During this phase, PMNs are thought to play a critical role. PMNs mediate generation of free radicals, damage of the vascular-alveolar interface, and increased protein-rich fluid permeability (5, 41, 49, 62, 63, 69, 119). Cav-1 contributes to the PMN-mediated inflammation, vascular injury, and non-cardiogenic pulmonary edema associated with ALI (28, 39, 40). In a recent study by Hu et al. (39), formyl-Met-Leu-Phe (FMLP) was used as a PMN activator and phorbol ester phorbol 12-myristate 13-acetate (PMA) as an alternative stimulator (39). In wild-type mice, infusion of cav-1+/+ PMNs followed by FMLP and PAF increased the pulmonary capillary filtration coefficient by 150% and lung wet-to-dry weight ratio by 50% (39). Increased PMN accumulation in lung tissue was also observed. In contrast, when cav-1−/− PMNs followed by FMLP and PAF were infused into wild-type mouse lungs, the above effects were abolished (39, 40). Furthermore, deletion of cav-1 in PMNs reduced PMN adhesion to endothelial cells, inhibited PMN chemotaxis, transendothelial migration, and reduced PMN superoxide production (28, 39, 40). In this study, the authors did not address the mortality of mice receiving infusions of cav-1+/+ PMNs vs. cav-1−/− PMNs, but their results support the hypothesis that neutrophil cav-1 is associated with the acute inflammatory responses during lung injury. An earlier study using a different approach drew a similar conclusion. Instead of infusing cav-1−/− PMNs into wild-type mice, Garrean and colleagues (28) observed marked attenuation of LPS-induced PMN sequestration and pulmonary edema in cav-1−/− mice. In this study, PMN adhesion to endothelial cells and PMN sequestration in lungs following LPS challenge was reduced in cav-1−/− mice due to reduced ICAM-1 expression (28). Adhesion of wild-type PMN to mouse lung vascular endothelial cells (MLVEC) after exposure of endothelial cells to LPS was markedly reduced in ECs cultured from cav-1−/− mice relative to WT mice. Diminished ICAM-1 expression in cav-1−/− mice paralleled the reduction in PMN binding to MLVEC cultured from cav-1−/− mice. Furthermore, lung PMN sequestration following LPS administration was reduced significantly in cav-1−/− lungs relative to wild-type mice. Garrean et al. assessed pulmonary microvascular liquid permeability by measuring the capillary filtration coefficient (Kf,c). LPS-induced elevation in Kf,c was observed in wild-type lungs but not Cav-1−/− lungs. Although the Kf,c measurement relies strongly on the vascular surface area and cav-1−/− lungs may have a reduced functional vascular surface area, as mentioned previously (62), the effect of LPS on Kf,c in the isolated perfused mouse lung was significantly reduced in cav-1−/− mice. While it is unclear how much of the dysfunctional vascular surface area is “recruited” during the pressure pulse of the Kf,c measurement, significantly greater wet-to-dry weight ratios were observed in wild-type lungs after LPS challenge compared with that measured in cav-1−/− lungs. Last, this study revealed that following administration of a relatively high dose of LPS, as much as 87% of wild-type mice died within 12 h in contrast to only 40% of cav-1−/− mice receiving the same dose of LPS. These reports demonstrate that the absence of cav-1 results in resistance to the lethal effects of LPS (28, 39, 40).

Hoetzel et al. (38) reported that in their ventilator-induced lung injury (VILI) model, in contrast to the above-mentioned studies, cav-1 confers a protective role, especially in the presence of carbon monoxide (CO). They found that cav-1−/− mice suffered significantly more severe lung injury in the VILI model compared with the wild-type mice, which was a three- to fivefold elevation in protein and cell counts in BAL fluid. This study, however, failed to show increased neutrophil recruitment in cav-1−/− mice exposed to high tidal volume ventilation. Furthermore, this study did not investigate the survival discrepancy between wild-type mice and cav-1−/− in the VILI model.

Cav-1 regulates DAD during lung injury.

CAV-1 REGULATES EPITHELIAL AND ENDOTHELIAL CELL DEATH IN LUNG INJURY.

For decades, it has been appreciated that ARDS, the severe form of ALI, is characterized by DAD (5, 41, 49, 62, 63, 69, 119). Lung epithelial cell death is a main feature of DAD. Apoptosis and other causes of cell death, including necrosis and oncosis, have been observed in ALI and attributed to a number of mechanisms (5, 41, 49, 62, 63, 69, 119). Franek et al. (22) identified apoptosis as a prominent component of the acute inflammatory response in the lung. The percentage of apoptotic cells was correlated with the severity of lung injury after hyperoxia (22). Recent studies illustrated that deletion of cav-1 protects against hyperoxia-induced cell death (43, 44, 122). Using the human bronchial epithelial cell line (Beas2B) and/or primary fibroblasts and endothelial cells isolated from the wild-type and cav-1−/− mice, Zhang et al. (122) found that deletion of cav-1 confers cytoprotection via upregulation of cytoprotective genes, such as hemeoxygenase (HO-1) and survivin (43, 44, 122). Survivin, a member of the inhibitors of apoptosis protein family, directly binds with caspase-3 and limits its activity. Thus, by regulating the level of survivin, cav-1 controls a common pathway of apoptosis in lung epithelial cells following hyperoxia exposure (43, 44, 122). However, in the above-mentioned studies, primary lung epithelial type I cells were not used, but instead, human bronchial cell lines were the major cells used for in vitro models. Thus, the above observations require further investigation using primary lung cells. Similar to survivin, deletion of cav-1 upregulates generic cytoprotective machinery (122). HO-1 confers well-known cytoprotection after hyperoxia (50, 51, 78). Compared with wild-type mice, cav-1−/− mice showed a markedly elevated HO-1 level and significant resistance to hyperoxia-induced death in vivo (43, 44, 122). Consistent with LPS/sepsis-induced lung injury models, cav-1−/− mice survived, on average, 2 days longer than wild-type mice after hyperoxia and showed less lung injury (43, 44, 122).

Cav-1 regulates bleomycin-induced lung injury and fibrosis.

The model of bleomycin-induced lung injury has been widely applied in pulmonary research. Bleomycin stimulates reactive oxygen species (ROS) production resulting in oxidative stress and pulmonary fibrosis (46, 72, 108). Bleomycin further induces apoptosis and senescence in epithelial and non-epithelial cells of the lung (46, 72, 108). Initial reports showed that cav-1 is upregulated 1 h after exposure to bleomycin, before the appearance of caspase-8, caspase-3, and caspase-9 cleavage products (46, 72, 108). Bleomycin leads to a partial translocation of cav-1 from lipid raft membrane fractions to non-raft fractions (46). Recently, additional studies suggested that bleomycin-induced injury of lung cells is accompanied by altered expression levels of cav-1 (73, 111). However, in the bleomycin-induced lung injury model, cav-1 was observed to play a greater role in epithelial cell growth arrest and senescence rather than promoting apoptosis (46, 72, 73, 108, 111). Reduction of cav-1 expression using siRNA and antisense oligonucleotides restored DNA synthesis and allowed senescent cells to re-enter the cell cycle (7). Dasari et al. (11) showed that oxidative stress induced premature senescence by stimulating cav-1 gene transcription through p38 mitogen-activated protein kinase/SP-1-mediated activation of cav-1 promoter elements (11). It is also known that cav-1 overexpression is sufficient to arrest mouse embryonic fibroblasts in the G0/G1 phase of the cell cycle, reduce their proliferative life span, and promote premature cellular senescence through activation of the p53/p21 axis (25, 107). However, despite these findings, the exact molecular mechanisms linking cav-1 to the induction of growth arrest, particularly in the G2/M phase, remain to be determined. Linge et al. (57) confirmed that bleomycin increases the expression of cav-1 in A549 cells and that deletion of cav-1 prior to bleomycin-exposure modulates p53 and p21 levels and subsequently prevents growth arrest in the G2/M phase. It was speculated that cav-1 inhibits cell proliferation by suppressing STAT3 activation, although details of these mechanisms require further investigation. Although this study, again, carries a limitation that A549 lung epithelial cell lines were used instead of primary lung epithelial cells, it supports a key role for cav-1 in modulating the cell cycle of bleomycin-damaged lung cells.

Multiple studies have demonstrated that cav-1 possesses an antifibrotic feature, led by Wang et al. (111), showing that cav-1 markedly ameliorates bleomycin-induced pulmonary fibrosis (7, 11, 12, 57, 73, 107, 108). This is consistent with the aforementioned characteristics of cav-1 in lung cell biology. Overall, cav-1 possesses antiproliferative and proapoptotic effects on all cell types including fibroblasts. In reflection, cav-1 shows an antifibrotic feature in lung fibrosis due to fibroblast apoptosis. Unfortunately, there are no survival data in cav-1−/− mice in the bleomycin-induced lung injury models.

The above reports largely rely on cell culture and/or animal studies, and thus, despite the plethora of information gleaned from these studies, there is limited information from human studies to compare with these and thus validate their relative usefulness. Wang et al. (111) and Tourkina et al. (102) reported marked reduction in cav-1 expression in lung tissues or monocytes/neutrophils from patients with idiopathic pulmonary fibrosis or scleroderma, respectively, compared with controls (102, 103, 111). Moreover, Tourkina et al. demonstrated the potential role of cav-1 as a lung-protective protein by using the cav-1 CSD peptide to treat scleroderma/fibrosis (102, 103, 111).

Future directions for assessing the role of cav-1 in ALI include: 1) determination of whether cav-1 can act as a potential biomarker for lung injury, and 2) whether development of therapeutic protocols using the cav-1 CSD peptides will be practical.

Additional cellular and molecular mechanisms by which cav-1 participates in ALI.

The above-mentioned studies provide robust evidence that cav-1 is an important regulator in the pathogenesis of ALI in various models. These reports assessed lung injury using multiple parameters including wet:dry weight ratio, morphological/histological examination, non-cardiogenic pulmonary edema, protein/cell counts in the BAL fluid, and mortality. In this section, we will discuss some additional hypotheses and evidence in support of the cellular/molecular mechanisms by which cav-1 regulates lung injury.

Does cav-1 play a role in the pathogen uptake by host cells in lung injury?

There is evidence showing that cav-1 is involved in both bacteria and virus uptake by host cells. Two recent studies on the function of cav-1 in Pseudomonas aeruginosa-induced pneumonia revealed variable results. Zaas et al. (118) found that deletion of cav-1 decreases lethality from P. aeruginosa acute pneumonia (118), whereas Gadjeva et al. (25) observed that cav-1−/− mice have increased sensitivity to P. aeruginosa infection, with an increased mortality rate. Interestingly, in the first study, Zaas et al. probably used a more cytotoxic strain of P. aeruginosa. In contrast, strains used in the second study (Gadjeva et al.) were more invasive but not cytotoxic. Furthermore, when Gadjeva et al. infected mice with the exotoxin U-positive and cytotoxic P. aeruginosa strain PA14, the previously observed survival benefits in wild-type mice diminished (25). These two studies suggest that in the model of P. aeruginosa-induced pneumonia, the cytotoxicity of the bacterial strain plays a critical role in determining the outcome observed in cav-1−/− mice.

Although the underlying mechanisms by which cav-1 regulates pulmonary infection-associated lung injury remain unclear, studies have shown evidence supporting a role of cav-1 in pathogen endocytosis and cell signaling. Eight caveolin-binding sites have been identified in severe acute respiratory syndromes (SARS) coronavirus. These caveolin-binding sites are located in replicase 1AB, spike protein, orf3 protein, and M protein (79). Furthermore, colocalization of cav-1 and orf3a (the largest unique ORF in the SARS coronavirus genome) has been confirmed (60). These studies provide evidence that cav-1 is involved in regulating pathogen-induced lung injury. However, not all viruses rely on cav-1/caveolae to enter host cells. For instance, internalization of coxsackievirus A9 into A549 lung epithelial cells does not require cav-1 but is dependent on integrin αVβ6, β2-microglobulin, dynamin, and Arf6 (35). Additionally, whether cav-1 protein or the structure of caveolae plays a role in pathogen uptake remains unclear. Delineation of the mechanisms by which cav-1/caveolae mediate pathogen uptake by host cells will assist future development of therapeutic options for infection-associated lung injury.

Does cav-1 play a role in antigen processing and immune response in lung injury?

In non-infection-associated ALI, an unspecified immune response has been suspected to trigger severe inflammation leading to DAD. Emerging evidence suggests that cav-1 is involved in this process. Ohnuma et al. (77) first showed that T cell costimulatory molecule CD26 interacts with cav-1, resulting in antigen-specific T cell activation. A series of studies further demonstrated that CD26 binds to cav-1 in antigen-presenting cells (APC). The interaction between CD26 and cav-1 induces T cell proliferation. More interestingly, blocking CD26-caveolin-1 costimulation with soluble cav-1-Ig (Cav-Ig) inhibits T cell proliferation and cytokine production in response to recall antigen (74, 75, 76). This result potentially provides a new therapeutic approach by using soluble Cav-Ig as an immunosuppressive agent. More evidence from Medina et al. (65, 66, 67) suggests that cav-1 modulates immune responses against pathogens. In their study, cav-1−/− mice showed increased production of inflammatory cytokines, chemokines, and nitric oxide, and cav-1−/− macrophages displayed increased inflammatory responses to LPS.

Does cav-1 play a role in apoptotic and/or autophagic pathways in lung injury?

Numerous studies have shown that cav-1 regulates apoptotic pathways at various stages of lung injury. Death receptor agonists including Fas ligand, TNFα, and TRAIL have all been reported to interact with or upregulate cav-1 (13, 87, 120). Cav-1 protein can be significantly induced by TNFα and TRAIL (13, 87, 120), whereas Fas has been suspected to interact with cav-1 upon stimulation by FasL (122, 123). Elevated caspase activities and an increased active form of caspase-3 are known features of hyperoxia-induced lung cell apoptosis (43, 44, 122, 50, 51, 78). Deletion of cav-1 blunts these responses after hyperoxia (43, 44, 122). Further investigations suggest that cav-1 decreases survivin, the inhibitor of apoptosis, and directly regulates apoptotic pathways (43, 44, 122). Additionally, a recent report suggests that cav-1 physically interacts with LC3B, the key autophagic marker protein, at baseline. Upon exposure to ROS, this interaction diminishes, suggesting a regulatory role of cav-1 in the process of autophagic cell death during lung injury (6a).

CONCLUSIONS

Cav-1 is involved in PMN adhesion, chemotaxis, and epithelial and endothelial cell apoptosis/senescence during lung injury. These features lead to increased alveolar damage (DAD), vascular leakage (non-cardiogenic pulmonary edema), and profound inflammation during the initial stage of ALI. However, as a potential antifibrotic protein, cav-1 may be beneficial in the late stage (fibrotic phase) of lung injury, although this remains unclear and requires further investigation. The overwhelming association of caveolin-1 in the pathology of ALI suggests that additional cell- and animal-based experiments, in combination with clinical observations, have the potential to give rise to important therapeutic tools and interventions for treating ALI/ARDS in the future.

DISCLOSURES

No conflicts of interest, financial or otherwise, are declared by the author(s).

REFERENCES

  • 1. Alonso MA, Millán J. The role of lipid rafts in signalling and membrane trafficking in T lymphocytes. J Cell Sci 114: 3957–3965, 2001. [DOI] [PubMed] [Google Scholar]
  • 2. Benlimame N, Le PU, Nabi IR. Localization of autocrine motility factor receptor to caveolae and clathrin-independent internalization of its ligand to smooth endoplasmic reticulum. Mol Biol Cell 9: 1773–1786, 1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3. Bento-Abreu A, Velasco A, Polo-Hernández E, Lillo C, Kozyraki R, Tabernero A, Medina JM. Albumin endocytosis via megalin in astrocytes is caveola- and Dab-1 dependent and is required for the synthesis of the neurotrophic factor oleic acid. J Neurochem 111: 49–60, 2009. [DOI] [PubMed] [Google Scholar]
  • 4. Bist A, Fielding PE, Fielding CJ. Two sterol regulatory element-like sequences mediate up-regulation of caveolin gene transcription in response to low density lipoprotein free cholesterol. Proc Natl Acad Sci USA 94: 10693–10698, 1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5. Buckley S, Barsky L, Driscoll B, Weinberg K, Anderson KD, Warburton D. Apoptosis and DNA damage in type 2 alveolar epithelial cells cultured from hyperoxic rats. Am J Physiol Lung Cell Mol Physiol 274: L714–L720, 1998. [DOI] [PubMed] [Google Scholar]
  • 6. Cai QC, Jiang QW, Zhao GM, Guo Q, Cao GW, Chen T. Putative caveolin-binding sites in SARS-CoV proteins. Acta Pharmacol Sin 24: 1051–1059, 2003. [PubMed] [Google Scholar]
  • 6a. Chen ZH, Lam HC, Jin Y, Kim HP, Cao J, Lee SJ, Ifedigbo E, Parameswaran H, Ryter SW, Choi AM. Autophagy protein microtubule-associated protein 1 light chain-3B (LC3B) activates extrinsic apoptosis during cigarette smoke-induced emphysema. Proc Natl Acad Sci USA 107: 18880–18885, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7. Cho KA, Ryu SJ, Park JS, Jang IS, Ahn JS, Kim KT, Park SC. Senescent phenotype can be reversed by reduction of caveolin status. J Biol Chem 278: 27789–27795, 2003. [DOI] [PubMed] [Google Scholar]
  • 8. Cho WJ, Chow AK, Schulz R, Daniel EE. Matrix metalloproteinase-2, caveolins, focal adhesion kinase and c-Kit in cells of the mouse myocardium. J Cell Mol Med 11: 1069–1086, 2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9. Conrad PA, Smart EJ, Ying YS, Anderson RG, Bloom GS. Caveolin cycles between plasma membrane caveolae and the Golgi complex by microtubule-dependent and microtubule-independent steps. J Cell Biol 131: 1421–1433, 1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10. Couet J, Sargiacomo M, Lisanti MP. Interaction of a receptor tyrosine kinase, EGF-R, with caveolins. Caveolin binding negatively regulates tyrosine and serine/threonine kinase activities. J Biol Chem 272: 30429–30438, 1997. [DOI] [PubMed] [Google Scholar]
  • 11. Dasari A, Bartholomew JN, Volonte D, Galbiati F. Oxidative stress induces premature senescence by stimulating caveolin-1 gene transcription through p38 mitogen-activated protein kinase/Sp1-mediated activation of two GC-rich promoter elements. Cancer Res 66: 10805–10814, 2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12. Del Galdo F, Lisanti MP, Jimenez SA. Caveolin-1, transforming growth factor-beta receptor internalization, and the pathogenesis of systemic sclerosis. Curr Opin Rheumatol 20: 713–719, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13. Dong F, Zhang Wold LE X, Ren Q, Zhang Z, Ren J. Endothelin-1 enhances oxidative stress, cell proliferation and reduces apoptosis in human umbilical vein endothelial cells: role of ETB receptor, NADPH oxidase and caveolin-1. Br J Pharmacol 145: 323–333, 2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14. Drab M, Verkade P, Elger M, Kasper M, Lohn M, Lauterbach B, Menne J, Lindschau C, Mende F, Luft FC, Schedl A, Haller H, Kurzchalia TV. Loss of caveolae, vascular dysfunction, and pulmonary defects in caveolin-1 gene-disrupted mice. Science 293: 2449–2452, 2001. [DOI] [PubMed] [Google Scholar]
  • 15. Engelman JA, Chu C, Lin A, Jo H, Ikezu T, Okamoto T, Kohtz DS, Lisanti MP. Caveolin-mediated regulation of signaling along the p42/44 MAP kinase cascade in vivo. A role for the caveolin-scaffolding domain. FEBS Lett 428: 205–211, 1998. [DOI] [PubMed] [Google Scholar]
  • 16. Engelman JA, Lee RJ, Karnezis A, Bearss DJ, Webster M, Siegel P, Muller WJ, Windle JJ, Pestell RG, Lisanti MP. Reciprocal regulation of neu tyrosine kinase activity and caveolin-1 protein expression in vitro and in vivo. Implications for human breast cancer. J Biol Chem 273: 20448–20455, 1998. [DOI] [PubMed] [Google Scholar]
  • 17. Fang PK, Solomon KR, Zhuang L, Qi M, McKee M, Freeman MR, Yelick PC. Caveolin-1alpha and -1beta perform nonredundant roles in early vertebrate development. Am J Pathol 169: 2209–2222, 2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18. Fernández MA, Albor C, Ingelmo-Torres M, Nixon SJ, Ferguson C, Kurzchalia T, Tebar F, Enrich C, Parton RG, Pol A. Caveolin-1 is essential for liver regeneration. Science 313: 1628–1632, 2006. [DOI] [PubMed] [Google Scholar]
  • 19. Feron O. Endothelial nitric oxide synthase expression, and its functionality. Curr Opin Clin Nutr Metab Care 2: 291–296, 1999. [DOI] [PubMed] [Google Scholar]
  • 20. Feron O, Belhassen L, Kobzik L, Smith TW, Kelly RA, Michel T. Endothelial nitric oxide synthase targeting to caveolae. Specific interactions with caveolin isoforms in cardiac myocytes and endothelial cells. J Biol Chem 271: 22810–22814, 1996. [DOI] [PubMed] [Google Scholar]
  • 21. Fielding CJ, Fielding PE. Caveolae and intracellular trafficking of cholesterol. Adv Drug Deliv Rev 49: 251–264, 2001. [DOI] [PubMed] [Google Scholar]
  • 22. Franek WR, Horowitz S, Stansberry L, Kazzaz JA, Koo HC, Li Y, Arita Y, Davis JM, Mantell AS, Scott W, Mantell LL. Hyperoxia inhibits oxidant-induced apoptosis in lung epithelial cells. J Biol Chem 276: 569–575, 2001. [DOI] [PubMed] [Google Scholar]
  • 23. Frank PG, Lisanti MP. Caveolin-1 and caveolae in atherosclerosis: differential roles in fatty streak formation and neointimal hyperplasia. Curr Opin Lipidol 15: 523–529, 2004. [DOI] [PubMed] [Google Scholar]
  • 24. Fujimoto T, Kogo H, Nomura R, Une T. Isoforms of caveolin-1 and caveolar structure. J Cell Sci 113: 3509–3517, 2000. [DOI] [PubMed] [Google Scholar]
  • 25. Gadjeva M, Paradis-Bleau C, Priebe GP, Fichorova R, Pier GB. Caveolin-1 modifies the immunity to Pseudomonas aeruginosa. J Immunol 184: 296–302, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26. Galbiati F, Volonte D, Brown AM, Weinstein DE, Ben-Ze'ev A, Pestell RG, Lisanti MP. Caveolin-1 expression inhibits Wnt/beta-catenin/Lef-1 signaling by recruiting beta-catenin to caveolae membrane domains. J Biol Chem 275: 23368–23377, 2000. [DOI] [PubMed] [Google Scholar]
  • 27. Galbiati F, Volonte D, Engelman JA, Watanabe G, Burk R, Pestell RG, Lisanti MP. Targeted downregulation of caveolin-1 is sufficient to drive cell transformation and hyperactivate the p42/44 MAP kinase cascade. EMBO J 17: 6633–6648, 1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Garrean S, Gao XP, Brovkovych V, Shimizu J, Zhao YY, Vogel SM, Malik AB. Caveolin-1 regulates NF-kappaB activation and lung inflammatory response to sepsis induced by lipopolysaccharide. J Immunol 177: 4853–4860, 2006. [DOI] [PubMed] [Google Scholar]
  • 29. Glenney JR, Jr, Soppet D. Sequence and expression of caveolin, a protein component of caveolae plasma membrane domains phosphorylated on tyrosine in Rous sarcoma virus-transformed fibroblasts. Proc Natl Acad Sci USA 89: 10517–10521, 1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30. Goetz JG, Joshi B, Lajoie P, Strugnell SS, Scudamore T, Kojic LD, Nabi IR. Concerted regulation of focal adhesion dynamics by galectin-3 and tyrosine-phosphorylated caveolin-1. J Cell Biol 180: 1261–1275, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31. Gumbleton M, Abulrob AG, Campbell L. Caveolae: an alternative membrane transport compartment. Pharm Res 17: 1035–1048, 2000. [DOI] [PubMed] [Google Scholar]
  • 32. Han JM, Kim Y, Lee JS, Lee CS, Lee BD, Ohba M, Kuroki T, Suh PG, Ryu SH. Localization of phospholipase D1 to caveolin-enriched membrane via palmitoylation: implications for epidermal growth factor signaling. Mol Biol Cell 13: 3976–3988, 2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33. Hansen CG, Bright NA, Howard G, Nichols BJ. SDPR induces membrane curvature and functions in the formation of caveolae. Nat Cell Biol 11: 807–814, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34. Hansen CG, Nichols BJ. Exploring the caves: cavins, caveolins and caveolae. Trends Cell Biol 20: 177–186, 2010. [DOI] [PubMed] [Google Scholar]
  • 35. Heikkilä O, Susi P, Stanway G, Hyypiä T. Integrin alphaVbeta6 is a high-affinity receptor for coxsackievirus A9. J Gen Virol 90: 197–204, 2009. [DOI] [PubMed] [Google Scholar]
  • 36. Hill MM, Bastiani M, Luetterforst R, Kirkham M, Kirkham A, Nixon SJ, Walser P, Abankwa D, Oorschot VM, Martin S, Hancock JF, Parton RG. PTRF-Cavin, a conserved cytoplasmic protein required for caveola formation and function. Cell 132: 113–124, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37. Hill WG, Almasri E, Ruiz WG, Apodaca G, Zeidel ML. Water and solute permeability of rat lung caveolae: high permeabilities explained by acyl chain unsaturation. Am J Physiol Cell Physiol 289: C33–C41, 2005. [DOI] [PubMed] [Google Scholar]
  • 38. Hoetzel A, Schmidt R, Vallbracht S, Goebel U, Dolinay T, Kim HP, Ifedigbo E, Ryter SW, Choi AM. Carbon monoxide prevents ventilator-induced lung injury via caveolin-1. Crit Care Med 37: 1708–1715, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39. Hu G, Vogel SM, Schwartz DE, Malik AB, Minshall RD. Intercellular adhesion molecule-1-dependent neutrophil adhesion to endothelial cells induces caveolae-mediated pulmonary vascular hyperpermeability. Circ Res 102: e120–e131, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40. Hu G, Ye RD, Dinauer MC, Malik AB, Minshall RD. Neutrophil caveolin-1 expression contributes to mechanism of lung inflammation and injury. Am J Physiol Lung Cell Mol Physiol 294: L178–L186, 2008. [DOI] [PubMed] [Google Scholar]
  • 41. Ingbar DH. Mechanisms of repair and remodeling following acute lung injury. Clin Chest Med 21: 589–616, 2000. [DOI] [PubMed] [Google Scholar]
  • 42. Ito J, Kheirollah A, Nagayasu Y, Lu R, Kato K, Yokoyama S. Apolipoprotein A-I increases association of cytosolic cholesterol and caveolin-1 with microtubule cytoskeletons in rat astrocytes. J Neurochem 97: 1034–1043, 2006. [DOI] [PubMed] [Google Scholar]
  • 43. Jin Y, Kim HP, Cao J, Zhang M, Ifedigbo E, Choi AM. Caveolin-1 regulates the secretion and cytoprotection of Cyr61 in hyperoxic cell death. FASEB J 23: 341–350, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44. Jin Y, Kim HP, Chi M, Ifedigbo E, Ryter SW, Choi AM. Deletion of caveolin-1 protects against oxidative lung injury via up-regulation of heme oxygenase-1. Am J Respir Cell Mol Biol 39: 171–179, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45. Ju H, Zou R, Venema VJ, Venema RC. Direct interaction of endothelial nitric-oxide synthase and caveolin-1 inhibits synthase activity. J Biol Chem 272: 18522–18525, 1997. [DOI] [PubMed] [Google Scholar]
  • 46. Kasper M, Barth K. Bleomycin and its role in inducing apoptosis and senescence in lung cells – modulating effects of caveolin-1. Curr Cancer Drug Targets 9: 341–353, 2009. [DOI] [PubMed] [Google Scholar]
  • 47. Kogo H, Aiba T, Fujimoto T. Cell type-specific occurrence of caveolin-1alpha and -1beta in the lung caused by expression of distinct mRNAs. J Biol Chem 279: 25574–25581, 2004. [DOI] [PubMed] [Google Scholar]
  • 48. Lee H, Woodman SE, Engelman JA, Volonté D, Galbiati F, Kaufman HL, Lublin DM, Lisanti MP. Palmitoylation of caveolin-1 at a single site (Cys-156) controls its coupling to the c-Src tyrosine kinase: targeting of dually acylated molecules (GPI-linked, transmembrane, or cytoplasmic) to caveolae effectively uncouples c-Src and caveolin-1 (TYR-14). J Biol Chem 276: 35150–35158, 2001. [DOI] [PubMed] [Google Scholar]
  • 49. Lee PJ, Alam J, Sylvester SL, Inamdar N, Otterbein L, Choi AM. Regulation of heme oxygenase-1 expression in vivo and in vitro in hyperoxic lung injury. Am J Respir Cell Mol Biol 14: 556–568, 1996. [DOI] [PubMed] [Google Scholar]
  • 50. Lee PJ, Alam J, Wiegand GW, Choi AM. Overexpression of heme oxygenase-1 in human pulmonary epithelial cells results in cell growth arrest and increased resistance to hyperoxia. Proc Natl Acad Sci USA 93: 10393–10398, 1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51. Lee PJ, Jiang BH, Chin BY, Iyer NV, Alam J, Semenza GL, Choi AM. Hypoxia-inducible factor-1 mediates transcriptional activation of the heme oxygenase-1 gene in response to hypoxia. J Biol Chem 272: 5375–5381, 1997. [PubMed] [Google Scholar]
  • 52. Le PU, Nabi IR. Distinct caveolae-mediated endocytic pathways target the Golgi apparatus and the endoplasmic reticulum. J Cell Sci 116: 1059–1071, 2003. [DOI] [PubMed] [Google Scholar]
  • 53. Le Saux O, Teeters K, Miyasato S, Choi J, Nakamatsu G, Richardson JA, Starcher B, Davis EC, Tam EK, Jourdan-Le Saux C. The role of caveolin-1 in pulmonary matrix remodeling and mechanical properties. Am J Physiol Lung Cell Mol Physiol 295: L1007–L1017, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54. Le Saux CJ, Teeters K, Miyasato SK, Hoffmann PR, Bollt O, Douet V, Shohet RV, Broide DH, Tam EK. Down-regulation of caveolin-1, an inhibitor of transforming growth factor-beta signaling, in acute allergen-induced airway remodeling. J Biol Chem 283: 5760–5768, 2008. [DOI] [PubMed] [Google Scholar]
  • 55. Li L, Ren CH, Tahir SA, Ren C, Thompson TC. Caveolin-1 maintains activated Akt in prostate cancer cells through scaffolding domain binding site interactions with and inhibition of serine/threonine protein phosphatases PP1 and PP2A. Mol Cell Biol 24: 9389–9404, 2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56. Lim JS, Choy HE, Park SC, Han JM, Jang IS, Cho KA. Caveolae-mediated entry of Salmonella typhimurium into senescent nonphagocytotic host cells. Aging Cell 9: 243–251, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57. Linge A, Weinhold K, Bläsche R, Kasper M, Barth K. Downregulation of caveolin-1 affects bleomycin-induced growth arrest and cellular senescence in A549 cells. Int J Biochem Cell Biol 39: 1964–1974, 2007. [DOI] [PubMed] [Google Scholar]
  • 58. Lisanti MP, Scherer PE, Tang Z, Sargiacomo M. Caveolae, caveolin and caveolin-rich membrane domains: a signalling hypothesis. Trends Cell Biol 4: 231–235, 1994. [DOI] [PubMed] [Google Scholar]
  • 59. Lohse MJ, Blüml K, Danner S, Krasel C. Regulators of G-protein-mediated signalling. Biochem Soc Trans 24: 975–980, 1996. [DOI] [PubMed] [Google Scholar]
  • 60. Lu Y, Liu DX, Tam JP. Lipid rafts are involved in SARS-CoV entry into Vero E6 cells. Biochem Biophys Res Commun 369: 344–349, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61. Lv XJ, Li YY, Zhang YJ, Mao M, Qian GS. Over-expression of caveolin-1 aggravate LPS-induced inflammatory response in AT-1 cells via up-regulation of cPLA2/p38 MAPK. Inflamm Res 59: 531–541, 2010. [DOI] [PubMed] [Google Scholar]
  • 62. Maniatis NA, Brovkovych V, Allen SE, John TA, Shajahan AN, Tiruppathi C, Vogel SM, Skidgel RA, Malik AB, Minshall RD. Novel mechanism of endothelial nitric oxide synthase activation mediated by caveolae internalization in endothelial cells. Circ Res 99: 870–877, 2006. [DOI] [PubMed] [Google Scholar]
  • 63. Mantell LL, Lee PJ. Signal transduction pathways in hyperoxia-induced lung cell death. Mol Genet Metab 71: 359–370, 2000. [DOI] [PubMed] [Google Scholar]
  • 64. Matute-Bello G, Frevert CW, Martin TR. Animal models of acute lung injury. Am J Physiol Lung Cell Mol Physiol 295: L379–L399, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65. Medina FA, Cohen AW, de Almeida CJ, Nagajyothi F, Braunstein VL, Teixeira MM, Tanowitz HB, Lisanti MP. Immune dysfunction in caveolin-1 null mice following infection with Trypanosoma cruzi (Tulahuen strain). Microbes Infect 9: 325–333, 2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66. Medina FA, de Almeida CJ, Dew E, Li J, Bonuccelli G, Williams TM, Cohen AW, Pestell RG, Frank PG, Tanowitz HB, Lisanti MP. Caveolin-1-deficient mice show defects in innate immunity and inflammatory immune response during Salmonella enterica serovar Typhimurium infection. Infect Immun 74: 6665–6674, 2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67. Medina FA, Williams TM, Sotgia F, Tanowitz HB, Lisanti MP. A novel role for caveolin-1 in B lymphocyte function and the development of thymus-independent immune responses. Cell Cycle 5: 1865–1871, 2006. [DOI] [PubMed] [Google Scholar]
  • 68. Mettouchi A, Klein S, Guo W, Lopez-Lago M, Lemichez E, Westwick JK, Giancotti FG. Integrin-specific activation of Rac controls progression through the G(1) phase of the cell cycle. Mol Cell 8: 115–127, 2001. [DOI] [PubMed] [Google Scholar]
  • 69. Meyrick B. Pathology of the adult respiratory distress syndrome. Crit Care Clin 2: 405–428, 1986. [PubMed] [Google Scholar]
  • 70. Mirza MK, Yuan J, Gao XP, Garrean S, Brovkovych V, Malik AB, Tiruppathi C, Zhao YY. Caveolin-1 deficiency dampens Toll-like receptor 4 signaling through eNOS activation. Am J Pathol 176: 2344–2351, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71. Montesano R, Roth J, Robert A, Orci L. Non-coated membrane invaginations are involved in binding and internalization of cholera and tetanus toxins. Nature 296: 651–653, 1982. [DOI] [PubMed] [Google Scholar]
  • 72. Mungunsukh O, Griffin AJ, Lee YH, Day RM. Bleomycin induces the extrinsic apoptotic pathway in pulmonary endothelial cells. Am J Physiol Lung Cell Mol Physiol 298: L696–L703, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73. Odajima N, Betsuyaku T, Nasuhara Y, Nishimura M. Loss of caveolin-1 in bronchiolization in lung fibrosis. J Histochem Cytochem 55: 899–909, 2007. [DOI] [PubMed] [Google Scholar]
  • 74. Ohnuma K, Uchiyama M, Hatano R, Takasawa W, Endo Y, Dang NH, Morimoto C. Blockade of CD26-mediated T cell costimulation with soluble caveolin-1-Ig fusion protein induces anergy in CD4+T cells. Biochem Biophys Res Commun 386: 327–332, 2009. [DOI] [PubMed] [Google Scholar]
  • 75. Ohnuma K, Uchiyama M, Yamochi T, Nishibashi K, Hosono O, Takahashi N, Kina S, Tanaka H, Lin X, Dang NH, Morimoto C. Caveolin-1 triggers T-cell activation via CD26 in association with CARMA1. J Biol Chem 282: 10117–10131, 2007. [DOI] [PubMed] [Google Scholar]
  • 76. Ohnuma K, Yamochi T, Uchiyama M, Nishibashi K, Iwata S, Hosono O, Kawasaki H, Tanaka H, Dang NH, Morimoto C. CD26 mediates dissociation of Tollip and IRAK-1 from caveolin-1 and induces upregulation of CD86 on antigen-presenting cells. Mol Cell Biol 25: 7743–7757, 2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77. Ohnuma K, Yamochi T, Uchiyama M, Nishibashi K, Yoshikawa N, Shimizu N, Iwata S, Tanaka H, Dang NH, Morimoto C. CD26 up-regulates expression of CD86 on antigen-presenting cells by means of caveolin-1. Proc Natl Acad Sci USA 101: 14186–14191, 2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78. Otterbein LE, Kolls JK, Mantell LL, Cook JL, Alam J, Choi AM. Exogenous administration of heme oxygenase-1 by gene transfer provides protection against hyperoxia-induced lung injury. J Clin Invest 103: 1047–1054, 1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79. Padhan K, Tanwar C, Hussain A, Hui PY, Lee MY, Cheung CY, Peiris JS, Jameel S. Severe acute respiratory syndrome coronavirus Orf3a protein interacts with caveolin. J Gen Virol 88: 3067–3077, 2007. [DOI] [PubMed] [Google Scholar]
  • 80. Palade GE. Fine structure of blood capillaries. J Appl Physiol 24: 1424, 1953. [Google Scholar]
  • 81. Parat MO, Fox PL. Oxidative stress, caveolae and caveolin-1. Subcell Biochem 37: 425–441, 2004. [DOI] [PubMed] [Google Scholar]
  • 82. Ramirez MI, Pollack L, Millien G, Cao YX, Hinds A, Williams MC. The alpha-isoform of caveolin-1 is a marker of vasculogenesis in early lung development. J Histochem Cytochem 50: 33–42, 2002. [DOI] [PubMed] [Google Scholar]
  • 83. Razani B, Combs TP, Wang XB, Frank PG, Park DS, Russell RG, Li M, Tang B, Jelicks LA, Scherer PE, Lisanti MP. Caveolin-1-deficient mice are lean, resistant to diet-induced obesity, and show hypertriglyceridemia with adipocyte abnormalities. J Biol Chem 277: 8635–8647, 2002. [DOI] [PubMed] [Google Scholar]
  • 84. Razani B, Engelman JA, Wang XB, Schubert W, Zhang XL, Marks CB, Macaluso F, Russell RG, Li M, Pestell RG, Di Vizio D, Hou H, Jr, Kneitz B, Lagaud G, Christ GJ, Edelmann W, Lisanti MP. Caveolin-1 null mice are viable but show evidence of hyperproliferative and vascular abnormalities. J Biol Chem 276: 38121–38138, 2001. [DOI] [PubMed] [Google Scholar]
  • 85. Ravid D, Chuderland D, Landsman L, Lavie Y, Reich R, Liscovitch M. Filamin A is a novel caveolin-1-dependent target in IGF-I-stimulated cancer cell migration. Exp Cell Res 314: 2762–2773, 2008. [DOI] [PubMed] [Google Scholar]
  • 86. Rothberg KG, Heuser JE, Donzell WC, Ying YS, Glenney JR, Anderson RG. Caveolin, a protein component of caveolae membrane coats. Cell 68: 673–682, 1992. [DOI] [PubMed] [Google Scholar]
  • 87. Saqr HE, Omran OM, Oblinger JL, Yates AJ. TRAIL-induced apoptosis in U-1242 MG glioma cells. J Neuropathol Exp Neurol 65: 152–161, 2006. [DOI] [PubMed] [Google Scholar]
  • 88. Scherer PE, Tang Z, Chun M, Sargiacomo M, Lodish HF, Lisanti MP. Caveolin isoforms differ in their N-terminal protein sequence and subcellular distribution. Identification and epitope mapping of an isoform-specific monoclonal antibody probe. J Biol Chem 270: 16395–16401, 1995. [DOI] [PubMed] [Google Scholar]
  • 89. Schlott T, Eiffert H, Bohne W, Landgrebe J, Brunner E, Spielbauer B, Knight B. Chlamydia trachomatis modulates expression of tumor suppressor gene caveolin-1 and oncogene C-myc in the transformation zone of non-neoplastic cervical tissue. Gynecol Oncol 98: 409–419, 2005. [DOI] [PubMed] [Google Scholar]
  • 90. Schnitzer JE, Oh P. Aquaporin-1 in plasma membrane and caveolae provides mercury-sensitive water channels across lung endothelium. Am J Physiol Heart Circ Physiol 270: H416–H422, 1996. [DOI] [PubMed] [Google Scholar]
  • 91. Schnitzer JE, Oh P, Jacobson BS, Dvorak AM. Caveolae from luminal plasmalemma of rat lung endothelium: microdomains enriched in caveolin, Ca(2+)-ATPase, and inositol trisphosphate receptor. Proc Natl Acad Sci USA 92: 1759–1763, 1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92. Schubert W, Frank PG, Razani B, Park DS, Chow CW, Lisanti MP. Caveolae-deficient endothelial cells show defects in the uptake and transport of albumin in vivo. J Biol Chem 276: 48619–48622, 2001. [DOI] [PubMed] [Google Scholar]
  • 93. Schubert W, Frank PG, Woodman SE, Hyogo H, Cohen DE, Chow CW, Lisanti MP. Microvascular hyperpermeability in caveolin-1 (-/-) knock-out mice. Treatment with a specific nitric-oxide synthase inhibitor, L-NAME, restores normal microvascular permeability in Cav-1 null mice. J Biol Chem 277: 40091–40098, 2002. [DOI] [PubMed] [Google Scholar]
  • 94. Schwab W, Galbiati F, Volonte D, Hempel U, Wenzel KW, Funk RH, Lisanti MP, Kasper M. Characterisation of caveolins from cartilage: expression of caveolin-1, -2 and -3 in chondrocytes and in alginate cell culture of the rat tibia. Histochem Cell Biol 112: 41–49, 1999. [DOI] [PubMed] [Google Scholar]
  • 95. Shack S, Wang XT, Kokkonen GC, Gorospe M, Longo DL, Holbrook NJ. Caveolin-induced activation of the phosphatidylinositol 3-kinase/Akt pathway increases arsenite cytotoxicity. Mol Cell Biol 23: 2407–2414, 2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96. Smart EJ, Ying YS, Conrad PA, Anderson RG. Caveolin moves from caveolae to the Golgi apparatus in response to cholesterol oxidation. J Cell Biol 127: 1185–1197, 1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97. Sowa G, Pypaert M, Fulton D, Sessa WC. The phosphorylation of caveolin-2 on serines 23 and 36 modulates caveolin-1-dependent caveolae formation. Proc Natl Acad Sci USA 100: 6511–6516, 2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98. Sowa G, Pypaert M, Sessa WC. Distinction between signaling mechanisms in lipid rafts vs. caveolae. Proc Natl Acad Sci USA 98: 14072–14077, 2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99. Sun Y, Hu G, Zhang X, Minshall RD. Phosphorylation of caveolin-1 regulates oxidant induced pulmonary vascular permeability via paracellular and transcellular pathways. Circ Res 105: 676–685, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100. Sundivakkam PC, Kwiatek AM, Sharma TT, Minshall RD, Malik AB, Tiruppathi C. Caveolin-1 scaffold domain interacts with TRPC1 and IP3R3 to regulate Ca2+ store release-induced Ca2+ entry in endothelial cells. Am J Physiol Cell Physiol 296: C403–C413, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101. Thomas CM, Smart EJ. Caveolae structure and function. J Cell Mol Med 12: 796–809, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102. Tourkina E, Richard M, Gööz P, Bonner M, Pannu J, Harley R, Bernatchez PN, Sessa WC, Silver RM, Hoffman S. Antifibrotic properties of caveolin-1 scaffolding domain in vitro and in vivo. Am J Physiol Lung Cell Mol Physiol 294: L843–L861, 2008. [DOI] [PubMed] [Google Scholar]
  • 103. Tourkina E, Richard M, Oates J, Hofbauer A, Bonner M, Gööz P, Visconti R, Zhang J, Znoyko S, Hatfield CM, Silver RM, Hoffman S. Caveolin-1 regulates leucocyte behaviour in fibrotic lung disease. Ann Rheum Dis 69: 1220–1226, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104. Tiruppathi C, Shimizu J, Miyawaki-Shimizu K, Vogel S, Bair AM, Minshall RD, Predescu D, Malik AB. Role of NF-κB dependent caveolin-1 expression in the menchanism of increased endothelial permeability induced by LPS. J Biol Chem 283: 4210–4218, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105. van den Heuvel AP, Schulze A, Burgering BM. Direct control of caveolin-1 expression by FOXO transcription factors. Biochem J 385: 795–802, 2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106. Vassilieva EV, Ivanov AI, Nusrat A. Flotillin-1 stabilizes caveolin-1 in intestinal epithelial cells. Biochem Biophys Res Commun 379: 460–465, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107. 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 13: 2502–2517, 2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108. Wallach-Dayan SB, Izbicki G, Cohen PY, Gerstl-Golan R, Fine A, Breuer R. Bleomycin initiates apoptosis of lung epithelial cells by ROS but not by Fas/FasL pathway. Am J Physiol Lung Cell Mol Physiol 290: L790–L796, 2006. [DOI] [PubMed] [Google Scholar]
  • 109. Wang XM, Kim HP, Song R, Choi AM. Caveolin-1 confers antiinflammatory effects in murine macrophages via the MKK3/p38 MAPK pathway. Am J Respir Cell Mol Biol 34: 434–442, 2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110. Wang XQ, Yan Q, Sun P, Liu JW, Go L, McDaniel SM, Paller AS. Suppression of epidermal growth factor receptor signaling by protein kinase C-alpha activation requires CD82, caveolin-1, and ganglioside. Cancer Res 67: 9986–9995, 2007. [DOI] [PubMed] [Google Scholar]
  • 111. Wang XM, Zhang Y, Kim HP, Zhou Z, Feghali-Bostwick CA, Liu F, Ifedigbo E, Xu X, Oury TD, Kaminski N, Choi AM. Caveolin-1: a critical regulator of lung fibrosis in idiopathic pulmonary fibrosis. J Exp Med 203: 2895–2906, 2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112. Wary KK, Mariotti A, Zurzolo C, Giancotti FG. A requirement for caveolin-1 and associated kinase Fyn in integrin signaling and anchorage-dependent cell growth. Cell 94: 625–634, 1998. [DOI] [PubMed] [Google Scholar]
  • 113. Williams TM, Medina F, Badano I, Hazan RB, Hutchinson J, Muller WJ, Chopra NG, Scherer PE, Pestell RG, Lisanti MP. Caveolin-1 gene disruption promotes mammary tumorigenesis and dramatically enhances lung metastasis in vivo. Role of Cav-1 in cell invasiveness and matrix metalloproteinase (MMP-2/9) secretion. J Biol Chem 279: 51630–51646, 2004. [DOI] [PubMed] [Google Scholar]
  • 114. Xie Z, Zeng X, Waldman T, Glazer RI. Transformation of mammary epithelial cells by 3-phosphoinositide-dependent protein kinase-1 activates beta-catenin and c-Myc, and down-regulates caveolin-1. Cancer Res 63: 5370–5375, 2003. [PubMed] [Google Scholar]
  • 115. Yamada E. The fine structure of the gall bladder epithelium of the mouse. J Biophys Biochem Cytol 1: 445–458, 1955. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116. Yang G, Park S, Cao G, Goltsov A, Ren C, Truong LD, Demayo F, Thompson TC. MMTV promoter-regulated caveolin-1 overexpression yields defective parenchymal epithelia in multiple exocrine organs of transgenic mice. Exp Mol Pathol 89: 9–19, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117. Yang G, Timme TL, Naruishi K, Fujita T, Fattah el MA, Cao G, Rajagopalan K, Troung LD, Thompson TC. Mice with cav-1 gene disruption have benign stromal lesions and compromised epithelial differentiation. Exp Mol Pathol 84: 131–140, 2008. [DOI] [PubMed] [Google Scholar]
  • 118. Zaas DW, Swan ZD, Brown BJ, Li G, Randell SH, Degan S, Sunday ME, Wright JR, Abraham SN. Counteracting signaling activities in lipid rafts associated with the invasion of lung epithelial cells by Pseudomonas aeruginosa. J Biol Chem 284: 9955–9964, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119. Zaher TE, Miller EJ, Morrow DM, Javdan M, Mantell LL. Hyperoxia-induced signal transduction pathways in pulmonary epithelial cells. Free Radic Biol Med 42: 897–908, 2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120. Zhao X, Liu Y, Ma Q, Wang X, Jin H, Mehrpour M, Chen Q. Caveolin-1 negatively regulates TRAIL-induced apoptosis in human hepatocarcinoma cells. Biochem Biophys Res Commun 378: 21–26, 2009. [DOI] [PubMed] [Google Scholar]
  • 121. Zhao YY, Malik AB. A novel insight into the mechanism of pulmonary hypertension involving caveolin-1 deficiency and endothelial nitric oxide synthase activation. Trends Cardiovasc Med 19: 238–242, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122. Zhang M, Lin L, Lee SJ, Mo L, Cao J, Ifedigbo E, Jin Y. Deletion of caveolin-1 protects hyperoxia-induced apoptosis via survivin-mediated pathways. Am J Physiol Lung Cell Mol Physiol 297: L945–L953, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123. Zundel W, Swiersz LM, Giaccia A. Caveolin 1-mediated regulation of receptor tyrosine kinase-associated phosphatidylinositol 3-kinase activity by ceramide. Mol Cell Biol 20: 1507–1514, 2000. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from American Journal of Physiology - Lung Cellular and Molecular Physiology are provided here courtesy of American Physiological Society

RESOURCES