Skip to main content
American Journal of Physiology - Lung Cellular and Molecular Physiology logoLink to American Journal of Physiology - Lung Cellular and Molecular Physiology
. 2015 Jul 17;309(6):L507–L525. doi: 10.1152/ajplung.00139.2015

Lost after translation: insights from pulmonary surfactant for understanding the role of alveolar epithelial dysfunction and cellular quality control in fibrotic lung disease

Surafel Mulugeta 1,, Shin-Ichi Nureki 2, Michael F Beers 1
PMCID: PMC4572416  PMID: 26186947

Abstract

Dating back nearly 35 years ago to the Witschi hypothesis, epithelial cell dysfunction and abnormal wound healing have reemerged as central concepts in the pathophysiology of idiopathic pulmonary fibrosis (IPF) in adults and in interstitial lung disease in children. Alveolar type 2 (AT2) cells represent a metabolically active compartment in the distal air spaces responsible for pulmonary surfactant biosynthesis and function as a progenitor population required for maintenance of alveolar integrity. Rare mutations in surfactant system components have provided new clues to understanding broader questions regarding the role of AT2 cell dysfunction in the pathophysiology of fibrotic lung diseases. Drawing on data generated from a variety of model systems expressing disease-related surfactant component mutations [surfactant proteins A and C (SP-A and SP-C); the lipid transporter ABCA3], this review will examine the concept of epithelial dysfunction in fibrotic lung disease, provide an update on AT2 cell and surfactant biology, summarize cellular responses to mutant surfactant components [including endoplasmic reticulum (ER) stress, mitochondrial dysfunction, and intrinsic apoptosis], and examine quality control pathways (unfolded protein response, the ubiquitin-proteasome system, macroautophagy) that can be utilized to restore AT2 homeostasis. This integrated response and its derangement will be placed in the context of cell stress and quality control signatures found in patients with familial or sporadic IPF as well as non-surfactant-related AT2 cell dysfunction syndromes associated with a fibrotic lung phenotype. Finally, the need for targeted therapeutic strategies for pulmonary fibrosis that address epithelial ER stress, its downstream signaling, and cell quality control are discussed.

Keywords: alveolar type 2 cells, cell quality control, epithelial ER stress, fibrotic lung disease, surfactant proteins


by the very nature of its juxtaposition with the outside world, the epithelium of the distal lung, and specifically the alveolar type 2 (AT2) cell, in the course of performing the biosynthetic, metabolic, and repair/regenerative functions critical to maintenance of alveolar homeostasis, must also protect itself from significant stress imparted by its constant exposure to mechanical, metabolic, and environmental factors. To address this considerable threat to protein synthesis, macromolecular turnover, and maintenance of organelle mass and integrity, eukaryotic cells and higher organisms have evolved a complex network of cellular quality control pathways designed to improve, modify, compartmentalize, and/or remove abnormal molecular and organellar substrates that accumulate during the course of “daily life.” Disruption of this network inevitably results in the accumulation of misfolded proteins/aggregates, complex macromolecules, and/or dysfunctional organelles with elaboration of multiple signaling cascades that can drive disease initiation/progression through the triggering of inflammation, cell death, and organ dysfunction. Although cellular quality control and its derangement have been shown to play an important role in disease pathogenesis in other organ systems such as brain, liver, and pancreas, its contribution to pulmonary disease is only emerging.

Pulmonary fibrosis represents a heterogeneous group of postnatal interstitial lung disorders characterized by destruction of pulmonary architecture caused by an abnormal wound repair response that ultimately leads to scar formation, organ malfunction, disruption of gas exchange, and respiratory failure. In adults, idiopathic pulmonary fibrosis (IPF), the most common subtype of the larger family of idiopathic interstitial pneumonias, is a diffuse parenchymal lung disease of unknown etiology that affects over 5 million people worldwide and typically results in a need for lung transplantation or in death within 2–5 years of diagnosis (81). In the United States, the annual incidence of IPF appears to be rising and is estimated at 5–16/100,000 individuals with a prevalence of 13–20/100,000, is more common in men, and increases significantly with age. IPF is characterized histologically by a pattern of usual interstitial pneumonia (UIP), in which the peripheral subpleural parenchyma shows evidence of fibroblastic foci, traction bronchiectasis, and microscopic honeycombing lined by hyperplastic AT2 cells interspersed with areas of normal or nearly normal lung tissue. Familial forms of pulmonary fibrosis (FPF) can also present in children and are part of the larger spectrum of childhood interstitial lung disease (chILD) (38; 88). The incompletely defined pathogenesis of IPF (and chILD) has been a major obstacle in developing effective therapies capable of stabilizing or improving lung function for these disorders.

Although research into the details of pathways responsible for epithelial cell dysfunction in IPF is still in its infancy, the pulmonary surfactant system has provided an important platform and experimental substrates to study the role aberrant quality control, cell stress, and cytotoxicity in these events. In this review, we describe developments in the understanding of the molecular mechanisms underlying childhood and adult interstitial lung diseases (ILDs) from the viewpoint of disease-related mutations in components of the surfactant system [surfactant protein C (SP-C); surfactant protein A (SP-A), ATP binding cassette class A3 (ABCA3) lipid transporter] and the cellular responses to their expression that could modulate disease pathogenesis. Data on the elicited quality control events and signaling pathways leveraged from a variety of in vitro and in vivo models expressing these mutant isoforms will be examined and placed in the context of translational data from patients showing similar molecular signatures in both familial and sporadic IPF and in preclinical and patient data from other non-surfactant-related AT2 cell dysfunction syndromes associated with a fibrotic lung phenotype.

Alveolar Epithelial Injury and Fibrotic Lung Remodeling: the Witschi Hypothesis Reborn

The health of the distal lung relies on the proper function of AT2 cells, which play a critical role in the production and maintenance of pulmonary surfactant, contribute to overall barrier function, perform xenobiotic metabolism, and participate in lung repair serving as a progenitor population following lung injury to replace damaged AT1 and AT2 cells (10, 16). The cuboidal-shaped AT2 cells comprise only 3–5% of the alveolar surface area while the remainder is occupied by juxtaposed AT1 cells that serve to facilitate gas exchange. Both AT1 and AT2 cells have important roles in ion transport in the distal lung (101). Nevertheless, AT2 cells constitute 60% of alveolar epithelial cells and 10–15% of all lung cells (36), rendering them a critical component in alveolar homeostasis.

Over 35 years ago, Haschek and Witschi published a theory on the pathogenesis of lung fibrosis that challenged the then contemporaneous view of lung fibrosis as an “inflammatory disease” (56, 165). Although chronic inflammation as a key to its pathogenesis had persisted in some circles, it is not a prominent feature in biopsies of IPF patients (144). Coupled with the low efficacy or even harmful effects of immunomodulatory therapies shown in many clinical trials involving these patients (62a, 67, 132, 133), a paradigm shift developed away from the so-called “alveolitis” theory of the 1970s and 1980s to again suggest a pivotal role for alveolar epithelial cells (AEC) and specifically AT2 cell dysfunction in the development of a fibrotic lung phenotype. Although not primary drivers of the fibrotic lesion, the exact roles of inflammation and the local immune milieu in IPF pathogenesis remain to be defined.

In this “epithelial injury/abnormal wound repair” model, rechampioned in 2001 (144), IPF development and progression were proposed to be initiated and sustained by microfoci of repeated cycles of AEC injury occur within a larger dysfunctional repair process reminiscent of aberrant wound healing described in other organs such as skin, kidney, and liver (18, 184). More recently, the “postmodern” link between intrinsic epithelial cell dysfunction and IPF/ILD postulated by Witschi has been firmly established based on multiple lines of experimental evidence and clinical correlations:

1) Monogenetic diseases such as Hermansky-Pudlak syndrome (HPS) are characterized by dysfunctional AT2 cells and ILD (114). Mutations in nine different HPS genes disrupt intracellular cargo protein trafficking leading to dysfunctional lysosome-related organelles (LROs) in various organs. Lamellar bodies (LBs), the LROs of AT2 cells, show marked increase in size due to overaccumulation of phospholipids. Two HPS genetic subgroups, HPS1 and HPS4, manifest a common lung phenotype characterized by the premature development of a UIP-like fibrotic pattern in the fourth decade of life (114). Some pathological features can be recapitulated in mouse models of HPS, where AT2 cells were shown to be important effector cells elaborating proinflammatory cytokines such as monocyte chemoattractant protein-1 (MCP-1) (5, 191, 192).

2) Depletion of AT2 cells induces spontaneous fibrosis in mice. In a transgenic mouse model using the SP-C gene promoter to express a diphtheria toxin (DT) receptor exclusively in AT2 cells, administration of DT in these mice induced AT2 cell death and increased lung collagen deposition.

3) AEC apoptosis is paramount in mouse models of lung fibrosis generated by single or repetitive exposures to bleomycin. Inhibition of apoptosis in these models by use of caspase inhibitors, Fas/FasL pathway blockade, or genetic deletion of proapoptotic Bcl2 proteins attenuates fibrosis (164).

4) It has been noted that aberrantly activated lung epithelial cells produce virtually all the mediators responsible for fibroblast migration, proliferation, and activation with the subsequent exaggerated extracellular matrix accumulation and destruction of the lung parenchyma (144, 187).

5) Some of the most compelling evidence for the role of AT2 cells in these processes is derived from studies linking patients with UIP or chILD histopathology with mutations in AT2 cell-restricted genes involved in the production of pulmonary surfactant. To date, more than 60 SP-C (SFTPC), and 150 ABCA3 mutations, as well as two mutations in the surfactant protein A2 gene (SFTPA2) have been identified in affected children and adults (25, 116, 175, 178). As will be detailed below, signatures of protein aggregation, endoplasmic reticulum (ER) stress, proinflammatory/profibrotic cytokine elaboration, altered macroautophagy, and apoptosis defined by model systems expressing SFTPC, SFTPA, and ABCA3 mutations are also present in the lungs and AT2 epithelia of sporadic and familial IPF patients (3, 23, 83, 91).

Thus substantial evidence exists that a vulnerable and/or dysfunctional AT2 epithelium is a pivotal player in aberrant injury/repair responses occurring in IPF and other forms of fibrotic lung remodeling.

AT2 CELLS AND THE BIOSYNTHETIC CHALLENGE OF SURFACTANT

The alveolar gas exchange surface is coated with a thin film of surface active agent (= surfactant), representing a complex heterogeneous mixture of primarily lipids (90% by weight) and some protein that serves to promote alveolar stability by reducing surface tension at the air-liquid interface along the epithelial lining layer (129). In addition to lipids [mainly phosphatidylcholine (PC) with one (lyso-PC) or two (DPPC) palmitic acid side chains], biochemical analysis of surfactant has identified four unique proteins designated surfactant proteins: SP-A, SP-B, SP-C, and SP-D (131). A large volume of literature has demonstrated that the surface tension-reducing function of surfactant stems from the interaction of phospholipids and the two low-molecular-weight hydrophobic proteins, SP-B and SP-C (reviewed in Ref. 180). The relatively hydrophilic and more abundant oligomeric proteins, SP-A and SP-D, are members of the collectin family of C-type lectins that share distinct collagen-like and globular, carbohydrate-binding domains. Although SP-A and SP-D do not have a direct function in the surface tension activity of surfactant, they play an essential role in innate lung host defense (reviewed in Ref. 183). Numerous studies have shown that, under pathological conditions, SP-A and SP-D can each undergo a variety of posttranslational modifications (such as oxidation, nitration, S-nitrosylation) that can lead to loss of function (50, 53, 102, 196).

AT2 cells synthesize, secrete, and recycle all components of surfactant (Fig. 1). Compared with neuroendocrine and many secretory epithelia (exocrine and endocrine) in other tissues, both functionally and morphologically, the AT2 cell is somewhat unique. The diversity of cellular cargo that must be managed (phospholipids, hydrophobic proteins, multimeric hydrophilic proteins, cytokines) creates a significant burden on cell quality control and metabolic demands. Ultrastructurally, AT2 cells lack a dense secretory granule; but instead, transmission electron microscopy (EM) reveals the presence of cytoplasmic, osmiophilic, lamellated organelles (i.e., LBs) long recognized as the storage organelle from which surfactant is released into the alveolar lumen (9). Biochemically, LBs resemble other LROs in that they are acidic, express lysosomal markers (CD63; LAMP-1), but also contain surfactant lipids, SP-B, and SP-C as well as a fraction of intracellular SP-A (121). LBs release these contents into the alveolus via regulated exocytosis regulated by a variety of signaling pathways (reviewed in Ref. 134). In contrast much of the synthesized SP-A and all SP-D is secreted via a non-LB, constitutive pathway. In addition to biosynthesis, AT2 cells participate in the uptake, catabolism, and reutilization of surfactant. Protein and lipid components are endocytosed via clathrin-dependent and -independent mechanisms, routed through the central vacuolar system [early endosomes (EE), late endosomes/multivesicular bodies (LE/MVB)] and then sorted to lysosomes for degradation or to LB for resecretion.

Fig. 1.

Fig. 1.

Schematic representation of structure, biosynthetic processing, and life cycle of 3 surfactant system components associated with interstitial lung disease in humans by the alveolar type 2 cell. Biosynthesis: along with phospholipids, surfactant protein (SP)-B, and SP-D, AT2 cells synthesize SP-C, SP-A, and ABCA3. SP-C: the human SFTPC gene is transcribed and translated to a 197-amino acid (21-kDa) proprotein (proSP-C21). ProSP-C21 is translocated to the endoplasmic reticulum (ER), is sorted in the Golgi, and enters the regulated secretory pathway where it is processed by 4 endoproteolytic cleavages of NH2 and COOH propeptide domains as it transits through small or sorting vesicles (SV), multivesicular bodies (MVB), and composite bodies (CB) to lamellar bodies (LB), the site of the final cleavage. Inset, left: topological representation of the bitopic proSP-C showing its type II orientation. The 4 domains of proSP-C include NH2-terminal targeting domain harboring lysine 6 ubiquitination site (yellow oval) and Nedd4-2 binding “PY” targeting motif (black rectangle); mature domain (depicted in blue) containing juxtamembrane palmitoylation sites (pink bars) and valine-rich α-helical transmembrane hydrophobic region; linker domain, site for the common I73T mutant variant (yellow pentagon); BRICHOS domain with the exon4 deletion region (depicted in gray), and L188Q mutation (yellow triangle). Conserved cysteine residues (C121 and C189) form the primary disulfide-bond (gray bar). ABCA3: initially routed to post-Golgi sorting vesicles, ABCA3 is trafficked via the MVB/CB network to LBs and plasma membrane. ABCA3 undergoes a posttranslational proteolytic cleavage within the proximal NH2-terminal region at distal post-Golgi compartments. Inset, bottom right: predicted topological structure of the ABCA3 transporter. ABCA3 comprises 12 putative membrane-spanning helices where the NH2- and COOH-terminal domains and the 2 ABC binding cassettes (ABC1 and ABC2) face the cytosol. The NH2 targeting signal for Golgi exit (yellow region) and the 2 glycosylation sites within the first luminal loop (orange) are shown. ECD, extracellular domain. SP-A: the human SFTPA gene gives rise to multiple mRNA transcripts that assemble into an 18-mer native structure made up from 6 SP-A trimers. The assembled form of SP-A is secreted by both regulated and constitutive pathways. Posttranslational modifications of SP-A include signal peptide cleavage prior to entering the ER, hydroxylation of proline residues, and N-linked glycosylation. Inset, top right: structural representation of a native SP-A. The primary structure of SP-A protein consists of 4 domains: a short amino-terminal domain, a collagenous domain, a neck domain, and a carbohydrate recognition domain (CRD). Six of the SP-A trimers composed of 2 SP-A1 molecules (depicted in green) and 1 SP-A2 molecule (depicted in black) give rise to the mature SP-A 18-mer. A proline residue in the collagenous domain of SP-A primary structure provides a flexible kink that causes the trimers to bend outward in different directions, giving the native SP-A 18-mer a “flower bouquet” appearance. Secretion: surfactant phospholipid, processed SP-C, SP-B, and SP-A contained in the LB are released by regulated exocytosis; ABCA3 remains in the LB membrane and is recycled. A portion of SP-A and SP-D is secreted constitutively. Reuptake/degradation/recycling: SP-A, SP-C, together with other surfactant proteins and lipids, are taken up by AT2 cells via endocytosis and are either targeted to the lysosomes for degradation or recycled back to LBs via endosomes and MVBs.

SP-C: a Unique Secretory Product

Of the four surfactant proteins, only SP-C is synthesized exclusively by the AT2 cell. The 2.4-kb gene encoding human SFTPC is located on chromosome 8p and is organized into six exons (I through V coding, VI untranslated) and five introns, which produce a 0.9-kb mRNA encoding either a 191- or 197-amino acid 21-kDa proprotein (proSP-C21). The protein form isolated and sequenced from lung lavage fractions (SP-C3.7) is a lipid-avid peptide composed of 33–35 highly conserved amino acids containing a high content of Val, Ile, and Leu (∼60–65% of the primary sequence), which in aqueous solution self-aggregates, adopting β-sheet conformation and forming amyloid fibrils. Thus SP-C represents a structurally and functionally challenging substrate for the AT2 cell in which the proSP-C21 propeptide is trafficked through the regulated secretory pathway as an integral type II bitopic transmembrane (TM) protein (Ncytosol/Clumen), undergoing four endoproteolytic cleavages of its flanking NH2 and COOH propeptides to yield the mature, biophysically active 3.7-kDa form found in secreted surfactant (Fig. 1B) (reviewed in Ref. 13). These posttranslational events as well as the behavior of disease-causing proSP-C mutants are influenced (directly and indirectly) by four functional domains and structural motifs contained within proSP-C21 (Fig. 1B, inset):

1) NH2 targeting domain (PY motif).

A series of mutagenesis studies identified and characterized a tryptophan-tryptophan (WW) domain-interacting PPxY (PY) motif within the NH2 propeptide that directs post-ER/Golgi proSP-C trafficking to distal compartments (34, 66, 84). The PPDY motif specifically interacts with two members of an E3 ubiquitin ligase family, Nedd4 and Nedd4-2 (Nedd4-L), with Nedd4-2 exerting the dominant effect on proSP-C transport via catalysis of monoubiquitination (a well-known signal for lysosomal targeting of proteins) of proSP-C21 at lysine residue 6 of the cytosolic NH2 propeptide (34).

2) Mature SP-C (TM) domain.

In phospholipid-rich environments the mature SP-C domain forms a very stable and rigid α-helix and functions as a noncleavable signal anchor peptide for proSP-C21, facilitating its translocation to the ER (33, 75). Additionally, a conserved pair of juxtamembrane basic amino acids, lysine (K) and arginine (R) (at positions 34 and 35 of human proSP-C), play an essential role in the determination of the propeptide's type II TM topology (75, 109).

3) The BRICHOS domain.

The distal proSP-C21 COOH-terminus (residues 94–197), known as the BRICHOS domain, shares a high degree of structural homology with domains identified in over 300 proteins (59). The term BRICHOS was derived from the original disease-associated family members including BRI2, linked to familial British and Danish Dementia; Chondromodulin-1, related to chondrosarcoma; and SP-C (139). The structural similarity of these three is punctuated by the fact that each exist as integral type II TM proteins comprised of a cytosolic region, a TM segment, a linker domain, and BRICHOS domain containing a pair of conserved cysteine residues The human SFTPC encodes four cysteines in the BRICHOS domain. Cysteines 121 and 189 are positionally conserved across species and form the primary folding function. Johannson’s group has also modeled a second disulfide bridge between cysteine 120 and 148 (see Refs. 18, 181).

For proSP-C, mutation of either one of these cysteine residues results initially in proximal (ER) retention and eventual formation of aggregates of unprocessed polyubiquitinated proprotein in vitro (69). Functionally, in addition to promoting proSP-C folding, the BRICHOS domain appears to have an intramolecular chaperone-like activity safeguarding the metastable, β-sheet-prone, mature SP-C domain from amyloid fibril formation (51, 65, 154, 182). Amyloid deposits composed of mature SP-C have also been observed in lung tissue samples from some ILD patients with mutations in the BRICHOS domain (181).

4) Linker domain.

As the name signifies, the proximal COOH-terminal propeptide (residues 59–93) is linearly positioned between the mature SP-C (TM) and BRICHOS domains and through formation of β-hairpin structure thereby facilitates docking of the BRICHOS domain to the TM segment. Importantly, mutations in the linker domain do not induce cytosolic aggregation of proSP-C but instead result in its mistrafficking and accumulation in plasma membrane, early endosomes, and LE/MVB (15, 58) (discussed in detail later).

Surfactant Protein A: a Multimeric Glycoprotein AT2 Product

SP-A, the most abundant surfactant protein, is a luminal, multifunctional sialoglycoprotein of Mr 28–36,000 that contains a COOH-terminal C-type lectin motif, a triple helical collagen domain, and carbohydrate recognition domain (CRD) (183) (Fig. 1, inset). In humans, there are two genes (SFTPA1 and SFTPA2), 4.5 kb each, located on chromosome 10. The amino acid differences that distinguish SP-A1 from SP-A2 are in general conservative and located mainly in the collagen-like domain with only one shown to affect protein structure. Like SP-D, SP-A is synthesized by airway cells, including nonciliated (club/Clara) cells. Although it has been localized to extrapulmonary sites, including salivary glands, lacrimal glands, and the female urogenital tract, the exact physiological relevance of its expression in these organs remains to be defined (reviewed in Ref. 78).

The biosynthesis of SP-A differs markedly from SP-C (and SP-B). In contrast to proteolytic processing of a larger proprotein, posttranslational processing of the SP-A primary translation product consists of NH2 signal peptide cleavage, N-linked glycosylation, and proline hydroxylation followed by trimeric association of monomers via coiled-coil bundling of α-helices in the neck thought to be composed of two SP-A1 and one SP-A2 monomer(s). The completed oligomeric structure of SP-A represents a higher order 18-mer assembled from six trimeric subunits linked by interchain disulfide bonding (Fig. 1, inset).

Metabolic labeling studies have demonstrated that the intracellular trafficking and secretion of nascent SP-A may involve both a LB-dependent (42) and a LB-independent (constitutive) pathway (122). Since alveolar SP-A can also be taken up via endocytosis and incorporated into LBs it has been difficult to accurately gauge the contribution of each pathway. Beyond the presence of a signal peptide and N-linked glycosylation, structural motifs and signals regulating trafficking of SP-A are not defined.

ABCA3: a Transporter Critical to AT2 Cell Homeostasis

A key component in the formation of LBs is the ABCA3 glycoprotein, a member of the ABC superfamily TM transporter proteins that use ATP energy to drive various substrates, ranging from small ions (e.g., ABCC7 (CFTR) = Cl) to large molecules (e.g., ABCA1 = cholesterol) across the plasma and intracellular membranes. ABCA3 belongs to a subclass (class A) of ABC transporters that are involved in a variety of lipid translocation events in multiple cell types, including macrophages, epithelial cells, and neurons, with each localized in distinct subcellular compartments (71).

Human ABCA3 has been mapped to chromosome 16p13.3 and encodes a 1,704-amino acid protein (35). Although ABCA3 mRNA is detected in many tissues, the ABCA3 message is highly expressed in AT2 cells and ABCA3 protein localized in the LB-limiting membrane (110, 186). ABCA3 has been shown in vitro to transport phosphatidylcholine, phosphatidylglycerol, sphingomyelin, and cholesterol into lysosomes of model cell line systems (8, 30, 31, 41, 110). Functionally homozygous null mutations of ABCA3 described in neonates with respiratory failure are associated with the presence of abnormal electron dense bodies on transmission EM, further demonstrating its importance as one of the critical regulators of LB biogenesis and lung surfactant metabolism (8, 25, 31).

The structure of ABCA3 is schematically illustrated in Fig. 1. To date studies have identified two functional domains in the molecule: 1) xLxxKN targeting motif: nearly all ABCA subfamily of transporters share significant homology in their NH2-terminal domain, including a highly conserved motif (xLxxKN) (71) that we have shown to be essential for directing these proteins to post-Golgi compartments (12). Since each ABCA transporter has a unique subcellular destination and function, subsequent sorting of ABCA3 to LB requires additional unidentified signal(s). 2) N-linked glycosylation: site-directed mutagenesis of two asparagine residues within the first NH2-terminal luminal loop of the transporter impairs protein stability and disrupts its anterograde trafficking, resulting in proteasomal degradation and cell stress (14).

To summarize, AT2 cells represent a metabolically active cellular component of the distal lung responsible for the biosynthesis, assembly, and complex intracellular trafficking of a diverse and challenging set of cargo, resident proteins, and organelles that often mix in multiple subcellular compartments of the secretory, endocytic, and degradative (lysosomal) pathways.

CELLULAR QUALITY CONTROL

In 2001, Nogee and colleagues (117) first reported a mutation in the COOH-terminal encoding region of SFTPC that resulted in deletion of exon 4 and its 37 amino acids (SP-CΔexon4) in both an infant and mother with ILD. Shortly after, a second report described a different SFTPC mutation resulting in substitution of glutamine for leucine at position 188 of the SP-C propeptide (SP-CL188Q) in a large kindred of whom the biopsies of 11 adults showed IPF/UIP patterns and nonspecific interstitial pneumonia (NSIP) patterns were present in three children. Using a variety of model systems, we and others subsequently demonstrated that expression of SP-CΔexon4 or SP-CL188Q in vitro resulted in intracellular SP-C protein aggregation and cell death by apoptosis. Importantly, all affected patients with SFTPC-related ILD reported to date have been heterozygous for the mutant allele, suggesting a toxic gain of function mechanism.

To preserve homeostasis in the face of a variety of exogenous and endogenous insults that challenge the integrity of their molecular and organelle components, eukaryotic cells have developed a quality control network integrating various combinations of the following processes: 1) sensing mechanisms designed to detect the presence of unwanted proteins, nucleic acids, macromolecules, and organelles; 2) effector pathways composed of chaperone/cochaperone/enzyme systems designed to improve macromolecular integrity (e.g., protein folding), facilitate normal trafficking, or repair damaged organelles and DNA-based structures (e.g., telomeres); and 3) degradative components such as the ubiquitin-proteasome system (UPS) and autophagosome-lysosomal-endosomal pathways (principally macroautophagy will be considered here) acting synergistically to restore protein/organelle integrity and remove defective or unwanted substrates. In addition, signaling pathways emanating from many of these component pieces serve to modulate the overall composition of the quality control network.

Quality Control Lessons for the Lung from Other Organs

At the NHLBI workshop “Malformed Protein Structure and Proteostasis in Lung Diseases,” the role that disorders of protein folding and degradation play in the pathobiology of lung disorders and their associated complications was highlighted (7). By extension, recent discoveries have also implicated organelle dysfunction and turnover, in particular in mitochondria or in the endosomal-lysosomal system, in the pathogenesis of some age-dependent (e.g., IPF), age-accelerated (e.g., HPS-induced fibrosis), and environmentally modified (e.g., tobacco) lung disorders, pointing to a critical role for the need to edit and remove damaged organelles to maintain lung health. A major conclusion from this workshop was that research into the mechanisms and pathways underlying protein (and organelle) quality control in the lung and the coupling of alterations in quality control to lung disease has been largely underdeveloped.

By contrast, much of the progress made in our understanding the role of quality control pathways in disease has been derived from studies of a seemingly diverse set of chronic, degenerative disorders observed in a variety of organ systems including the central nervous system (CNS), liver, and pancreas. Huntington's and Alzheimer's diseases represent progressive neurodegenerative disorders associated with expression or altered processing of a mutant TM protein (analogous to proSP-C) that undergoes a fatal conformational rearrangement that unmasks an intrinsic ability to self-associate (aggregate), induce aberrant cell signaling/cytotoxicity, and/or be deposited in nonnative subcellular or tissue compartments. Similarly, liver disease from expression of the mutant Z allele of alpha-1-antitrypsin (Z-A1AT) is associated with ER retention and cytosolic aggregation of Z-A1AT in hepatocytes leading to their eventual death. The sheer magnitude of biosynthetic capacity for insulin and other hormones in islet cells of the endocrine pancreas requires an equally robust quality control system, which appears to fail in models of Type 2 diabetes. Furthermore, disorders such as Parkinson's disease have supplied the recognition that protein “conformational diseases” are often accompanied by mitochondrial dysfunction and, like all organelles, must be “recycled,” requiring an additional layer of cell quality control (130). Finally, lipid storage disorders such as Niemann-Pick C (NPC) disease that are marked by severe neurodegeneration and lung disease have been shown to acquire a defective cell quality control phenotype (principally impaired macroautophagy) (140).

Together, from basic and translational studies of these diseases using these abnormal protein isoforms as substrates, the cellular responses to misfolded proteins and dysfunctional organelles have been dissected out in the context of cellular quality control.

Protein Quality Control (Proteostasis)

Mutant protein accumulation within the ER and cytosol is countered by a complex mechanism aimed at restoring protein homeostasis or “proteostasis” (6). Proteostasis refers to the control of concentration, formation, binding interactions, quaternary structure, and location of individual proteins making up the proteome by readapting the innate biology of the cell, often through transcriptional and translational changes. The proteostasis network comprises processes controlling protein synthesis, folding, trafficking, aggregation, and degradation and is regulated by the integration of inputs from numerous cell sensing and signaling pathways. Failure of proteostasis to cope with misfolding-prone proteins, aging, or metabolic/environmental stress can trigger or exacerbate conformational disease.

The major components of a prototypical quality control response to aberrant conformers include the following:

The unfolded protein response.

As an initial response, highly specific signaling pathways, evolved to ensure that ER protein-folding capacity is not overwhelmed, are collectively termed the unfolded protein response (UPR) (reviewed in detail in Ref. 73). The UPR pathways are governed by three ER TM protein sensors: PKR-like endoplasmic reticulum kinase (PERK), activating transcription factor 6 (ATF6), and inositol-requiring enzyme 1 (IRE-1) (illustrated in Fig. 2). Their activation by abnormally folded proteins initiate one or more of three signaling cascades with two objectives: 1) generation, translocation, and binding of select transcription factors (XBP-1, ATF-6p50, ATF4) to nuclear ER-stress-responsive elements to expand ER refolding capacity by upregulating a variety of chaperones (e.g., Bip, GRP94, calnexin, protein disulfide isomerase); and 2) through phosphorylated eif2α to promote translation attenuation at the level of the ribosome. Importantly, molecular signatures generated by these events can be used to detect UPR activation in cells and in tissues: IRE1 activation is heralded by appearance of spliced XBP1 mRNA or XBP1 protein; PERK activation can be confirmed by detection of phosphor-eIF2α, CHOP/GADD153, and ATF4; ATF6 cleavage fragments confirm engagement of this pathway. The relative contributions of each UPR pathway to the proteostatic response are protein substrate and cell specific.

Fig. 2.

Fig. 2.

Quality control machinery utilized by cells in response to the expression of misfolded or unassembled proteins and dysfunctional organelles. Unfolded protein response (UPR): 1 or more of the 3 ER proximal sensors for misfolded proteins, IRE1, PERK, and ATF6, are activated by misfolded SP-C to initiate the downstream signaling of the UPR. These sensors are typically maintained in an inactive state as they are bound to the molecular chaperone, Bip/GRP78. The presence of increasing load of misfolded SP-C proteins within the ER requires the availability of Bip/GRP78 to bind to the misfolding proteins and consequently Bip/GRP78 dissociates from the sensors sites to meet this requirement and to allow 3 downstream signaling pathways to activate the following. 1) ATF6: when Bip is released from ATF6p90, this transmembrane sensor is first translocated to Golgi and subjected to cleavage by S1 and S2 protease generating a soluble cytosolic form (ATF6p50) that traffics to the nucleus where it binds to ER-stress-responsive elements (ERSE) to upregulate chaperone production. 2) IRE1, in the absence of Bip binding, undergoes phosphodimerization and acquires specific endonuclease activity for cytosolic splicing of XBP1 mRNA. The translated XBP1, a transcription factor, also binds to ERSEs in the nucleus. 3) PERK activation results in both phosphorylation of cytosolic eIF2α leading to a generalized translation repression plus selective upregulation of transcription factors such as CHOP/GADD153 and ATF4 to promote a specific subset of gene expression involved in expanding proteostatic capacity. ER-associated protein degradation (ERAD) involves a multistep process consisting of recognition of misfolded cargo by ER-resident proteins [e.g., EDEM, protein disulfide isomerase (PDI), Bip/Grp78] followed by their retrotranslocation out of the ER through translocons and subsequent polyubiquitination mediated by cytosolic E1/E2/E3 ubiquitin ligases. Targeting of ubiquitin-decorated coformers to the 26S proteasome results in eventual degradation. Autophagy: cytosolic macroaggregates are targeted for degradation by autophagy. Activation of a cascade of upstream events (not depicted) mediated by a complex composed of Beclin 1, class III PI3K, and other components initiates isolation membrane assembly and the eventual formation of a phagophore. The Atg5-Atg12-Atg16L complex and atg8 (LC3-I) phosphatidylethanolamine (PE) conjugate recruit and through elongation envelop cytosolic ubiquitinated cargos including protein aggregates and mitochondria, leading to the formation of autophagosome. Subsequently, the autophagosome fuses with lysosomes, resulting in an acidified and functional autophagolysosome (“autolysosome”), which also contains lysosome-derived acid hydrolases, to promote degradation of internalized cytosolic content. Aggresome formation: as a last coping mechanism, failure of UPR/ubiquitin-proteasome system (UPS)/autophagy network to control mutant protein levels results in transport of aggregate forms of misfolded protein in a microtubule-dependent manner to the microtubule organizing center (MTOC) near the nucleus, resulting in the formation of aggresomes.

ER-associated protein degradation.

If the UPR is unsuccessful at restoring normal protein conformation, misfolded or unassembled protein targets are removed from the ER by retrotranslocation and ultimately destroyed by the ubiquitin-proteasome system (UPS) depicted in Fig. 2 (107, 170). The proteasome is the main protein degradation machinery of the cell and degrades ∼90% of all cellular proteins. Covalent attachment of one or more K48-linked polyubiquitin chain(s) to a lysine residue(s) represents the primary targeting signal for direction to 26S proteasomes and then degraded into small peptides. In addition to a role in protein quality control, protein degradation by the proteasome is essential for diverse cellular functions such as apoptosis, transcription, cell signaling, and immune responses (reviewed in Ref. 119, 141). The proteasome itself is composed of a cylinder-shaped catalytic core particle, the 20S proteasome, which can be capped at either end by several different 19S regulator complexes when ATP/Mg2+ is present (48). The 20S proteasome alone participates but is inefficient at protein degradation therefore activation mediated by these 19S regulatory particles is crucial and can contribute to differential proteasome responses to physiological or pathological conditions (reviewed in Ref. 106).

The 26S proteasome used for ER-associated protein degradation (ERAD) is subject to failure. In a variety of systems, overexpression of misfolded protein isomers and conformers can actually inhibit global UPS function, leading to the cytosolic accumulation of both the primary mutant as well as bystander proteins (17). In addition, environmental factors such as cigarette smoke can directly impair proteasome activity in lung cells (without affecting proteasome expression), resulting in accumulation of polyubiquitinated proteins (168).

Autophagy.

The process of macroautophagy (hereafter referred to as autophagy) was originally described as a cellular response to fasting starvation. However, an important emerging concept is that autophagy also functions as a second major degradation pathway for the targeted removal of long-lived, aggregation-prone proteins (e.g., huntingtin, α-synuclein, and neuroserpins) (82, 89). An extensive discussion of the autophagic process is beyond the scope of this review; the reader is referred to several excellent recent reviews (32, 94, 113). In brief, as illustrated in Fig. 2, autophagy is a dynamic process controlled by at least 30 known autophagy (Atg) genes whose assembly and turnover enclose cytosolic materials and organelles by generation of isolation membranes (phagophores) that mature to form double membrane-bound vesicles (“autophagosomes”) that are readily identified by EM. Subsequently, the autophagosome containing captured cargo fuses with the lysosome for degradation of the contents. The entire process is highly regulated from inputs from a variety of sensing and signaling mechanisms including mammalian target of rapamycin (mTOR1), Class III phosphoinositide 3-kinases (PI3K), and several mTOR-independent pathways. The selectivity of autophagy for misfolded protein is dependent on the recognition of ubiquitinated aggregates by ubiquitin-binding receptors such as p62/sequestosome-1 (p62/SQSTM1) (125).

Various reports have confirmed the existence of a link between ERAD/UPS and autophagy in the clearance of misfolded proteins (28, 124), since perturbations in the flux through either pathway can affect the activity of the other system (85). Evidence also exists for cross talk between the UPR and autophagy mediated by several pathways including an eIF2α/ATF4 axis and JNK1 (105, 171).

Aggresomes, JUNQ, and IPOD.

Current consensus holds that direct cellular toxicity from aberrant mutant protein conformers is associated with nonnative soluble protein oligomers and that the formation of large aggregates is instead cytoprotective. To limit toxicity from misfolded proteins accumulating in the cytosol or nucleus that escape degradation by either the UPS and/or autophagy, mammalian cells contain an additional cellular quality control-based defense that involves the recognition, segregation, and active compartmentalization of ubiquitylated protein oligomers/microaggregates in a microtubule-dependent manner to pericentriolar, membrane-free, vimentin-enwrapped perinuclear structures termed aggresomes (68). Work by several groups has shown that, in addition to protein aggregates, Atgs and lysosomes can also be directed to aggresomes in a process dependent on the microtubule deactylase HDAC6, suggesting that aggresomes may function as a delivery and concentration system to improve the efficiency of autophagic degradation (63, 74).

Recently, our understanding of cellular aggregate deposition sites has been further expanded by the discovery of two other distinct types of deposition sites that can concurrently exist within a single cell (70). Like aggresomes, one site is located in close proximity to the nucleus and functions as a quality control center that attracts chaperones and proteasomes to either refold or degrade misfolded aggregated proteins. This type of deposition site, which contains ubiquitylated misfolded proteins, was termed “JUxta Nuclear Quality control compartment” (JUNQ) and is a dynamic structure that rapidly exchanges proteins with the cytosol. The other aggregate deposition site, located near cytosolic vacuoles, was termed “insoluble protein deposit” (IPOD) and sequesters immobile substrates that have no ubiquitin signal. The interrelationships of aggresomes, JUNQ, and IPOD to each other and to the UPS and macroautophagy in mammalian cells remain to be established.

Consequences of Failed Protein Quality Control: ER Stress, Aggregation, and Apoptosis

If the integrated proteostatic program (combined efforts of UPR, ERAD, and autophagy) fails to control mutant protein levels, a series of deleterious consequences for the cell can ensue. Illustrated in Fig. 3, these include the following events:

Fig. 3.

Fig. 3.

Pathways to AT2 cell injury/dysfunction defined by using mutant proSP-C isoforms as prototype substrates. The expression of proSP-C mutants can produce 1 of 2 profiles of aberrant cellular responses. Misfolding BRICHOS proSP-C (Group A1): ER stress marks the state in which the UPR can no longer maintain ER homeostasis because of overwhelming expression of misfolded SP-C, leading to a cascade of cellular disruption and injury: 1) Dysfunction of ERAD due to an overloaded and defective UPS. This event can be further exacerbated by second hits such as cigarette smoke and viruses. 2) ER stress-dependent recruitment of TRAF2 by IRE1 to activate JNK that promotes upregulation and release of cytokines. 3) Induction of apoptotic pathways by ER stress-induced activation of either the PERK/EIF2α/ATF4 network to trigger CHOP activation, IRE1/TRAF2 activation of caspase 4/12, and/or cytochrome c release from dysfunctional mitochondria. All 3 intersect downstream at the activation of caspase 3. 4) A failure of aggresomal compartmentalization or autophagy also contributes to the eventual accumulation of toxic macroaggregates. Mistrafficked Linker (non-BRICHOS) proSP-C (Group B): mutations in the COOH-terminus outside BRICHOS domain of proSP-C reported to date appear to be initially mistargeted to the plasma membrane via a constitutive pathway. Because of proper processing failure, proSP-C in the plasma membrane can be reinternalized and trafficked through early endosomes to late endosomes (LE)/MVB. The enhanced presence of aberrantly trafficked and misprocessed proSP-C in these compartments leads to a functional disruption of normal endosome/amphisome/lysosome turnover within the autophagic machinery that ultimately results in a distal block in autophagy. Such a block leads to the accumulation of abnormally large autophagic vacuoles containing undegraded, organellar, and proteinaceous debris populated with autophagy markers including LC3 and P62, as well as parkin, a protein critical for autophagy-dependent removal of dysfunctional mitochondria (i.e., mitophagy). *Potential sites for organelle disruption by the expression of mistrafficked proSP-C.

1) ER stress.

In addition to its normal adaptive signaling described previously, prolonged or severe overload of the ER with client proteins can activate “alarm signals” that overlap with some of the same signaling transduction cascades associated with innate immunity (185, 194). Termed “ER stress,” UPR sensors (particularly IRE1) can then activate downstream MAP kinases such as JNK to promote the elaboration of proinflammatory cytokines and growth factors from affected cells (115, 166). The proof of principle that the balance of UPR/ER stress can determine cytoprotection/cytotoxicity has been demonstrated repeatedly whereby overexpression or knockdown of Grp78/Bip can modulate both ER stress markers and susceptibility to cell death in vitro and in vivo (72, 179).

2) Intrinsic apoptosis.

In mammalian cells, ER stress can also induce “intrinsic apoptosis” via one or more pathways (80, 185). One involves the cleavage and activation of the ER-membrane-localized caspase-4/12 (60). In another, a complex generated by IRE1 and TNF-receptor-associated factor 2 (TRAF2) recruited to the ER membrane signals downstream effectors such as JNK (62). In some systems, PERK via an ATF4-mediated pathway and/or IRE1 via XBP-1 can upregulate C/EBP-homologous protein (CHOP), a transcriptional repressor of expression of prosurvival Bcl2 proteins. Additionally, an ER-mitochondria death-signaling cascade mediated through a cytochrome c/apoptotic proteasome activating factor 1 (APAF1)/caspase 9 apoptosome mechanism has been described in a variety of model systems (52). Further downstream integrative death signaling can involve activated caspase cascades that vary among cell types, by pathway and by inciting substrate. As examples, numerous reports have linked ER stress-mediated apoptosis of pancreatic beta cells to a caspase 4/12-mediated activation of caspase 3 (reviewed in Ref. 123). Similarly, amyloid-beta (Aβ) also induces apoptosis via activation of caspases 4/12 and then caspase 3 in neuroblastoma cell lines (61); however, neurons expressing misfolded PolyQ72 huntingtin undergo caspase 8-mediated apoptotic events (138).

3) Protein aggregation.

Best exemplified by chronic neurodegenerative disease, the intracellular or extracellular accumulation of an abnormal protein isoform (such as Aβ amyloid precursor protein, α-synuclein, or polyglutamine huntingtin-1) with inherent instability can lead to time-dependent neuronal cell toxicity and the histopathological appearance of inclusion bodies, amyloid fibrils, and protein macroaggregates in the cytosol or within the nucleus (29, 76, 135, 137). As will be discussed in detail subsequently, given the propensity of SP-C3.7 to form β-amyloid structures (51), it is not surprising that expression of some SFTPC mutations results in generation of SDS-insoluble proSP-C macroaggregates in vitro (111, 112) as well as intracellular amyloid formation in vitro and in vivo (182).

4) Mitochondrial dysfunction.

ER stress can induce mitochondrial dysfunction through both physical and functional communications between mitochondria and the ER network, which can further exacerbate apoptotic events (reviewed in Ref. 169).

Quality Control for Dysfunctional Organelles: Mitophagy

In addition to the utilization of macroautophagy for selective removal of polyubiquitinated protein aggregates (aggrephagy), autophagy can target and remove specific subcellular components (selective autophagy) including invading pathogens (xenophagy), lipids (lipophagy), and dysfunctional cellular organelles such as mitochondria or ER. Degradation of dysfunctional mitochondria, termed mitophagy, is a remarkably specific process. When a single mitochondrion is uncoupled via photoirradiation, this and only this mitochondrion will be degraded (79). Additionally, global disruption of mitochondrial function in cells treated by using an uncoupler such as carbonyl cyanide 3-chlorophenylhydrazone results in selective mitochondrial degradation, leaving other organelles intact.

The process of mitophagy requires the coordination of cytosolic factors and signals assembled at the outer mitochondrial membrane (reviewed in Refs. 189, 190). A key step in the initiation of the process involves PTEN-induced putative kinase 1 (PINK1), stabilized in the outer mitochondrial membrane in response to lowered TM potential (Δψ), which then recruits the E3 ligase parkin, resulting in K-63-linked polyubiquitin chains of a variety of mitochondrial protein substrates such as mitofusins (Mfn1; mfn2). The subsequent recognition and sequestration of the ubiquitin decorated mitochondrion utilizes much of the same autophagy machinery commissioned for misfolded protein removal. For example, the ubiquitin-binding adaptor p62/SQM1, which recruits ubiquitinated protein cargo into autophagosomes by binding to LC3, also accumulates on mitochondria following their parkin-mediated ubiquitination and facilitates their engulfment by the phagophore/autophagosome.

Aging, Cellular Quality Control, and Parenchymal Lung Disease

The effects of aging on cellular quality control capacity have been well documented, particularly in the CNS where perturbations to proteostasis and mitochondrial function are hallmarks of age-related neurodegenerative disorders (108). An age-dependent decrease in proteasome activity has also been observed in other tissues including skin, muscle, heart, liver, kidney, and eye, each with a variety of consequences (reviewed in Ref. 95). Aging has been shown to markedly downregulate autophagic activity (136). Given their key role in processing cellular constituents, either to the lysosome for degradation or to other cellular compartments for immune activation, autophagy and the UPS represent important cytoprotective strategies against aging, especially for postmitotic cells. As proof of concept, expansion of either proteasome activity or autophagy in the CNS extends lifespan and protects organisms from symptoms of proteotoxic stress (96).

Aging is a known risk factor for IPF (81). Many of the hallmarks of aging (e.g., genomic instability, telomere attrition, epigenetic alterations, and cellular senescence) have been proposed as essential mechanisms for the development of IPF; however, these disturbances are not restricted to IPF (reviewed in Ref. 145). Telomerase mutations have been described in IPF families but also in chronic obstructive pulmonary disease (4, 150). A recent report has described enhanced sensitivity of the alveolar stem cell function of AT2 cells from mice with short telomere syndromes subjected to exogenous “second hits” providing a potential mechanism for their role in aberrant injury/repair processes in IPF pathogenesis (2).

In addition to these pathways, several lines of evidence link age-dependent alterations in cell quality control, AT2 cell function, and susceptibility to fibrosis. Aged mice are susceptible to fibrosis from herpesvirus infection with increased expression of ER stress markers (161). In a preliminary report, lung tissue from aged mice displays decreased proteasomal activity and alterations in 26S proteasome assembly (27). Recently, Bueno et al. (23) have found that AT2 cells from the lungs of both IPF patients and aged wild-type mice exhibit marked accumulation of dysmorphic and dysfunctional mitochondria. This was associated with both an upregulation of ER stress markers and a downregulation of PINK1. Moreover, young PINK1-deficient mice had premature development of similarly dysmorphic, dysfunctional mitochondria in AT2 cells and were vulnerable to apoptosis and development of lung fibrosis in a viral challenge model (23). Interestingly, in a transgenic mouse model of atg5 deficiency, a 90% decrease in autophagy was well tolerated in young adult mice but resulted in alveolar septal thickening and altered lung mechanics in aged animals, consistent with accumulation of damage over time (54). Collectively, these data suggest that aging has marked effects on both the proteostasis network and autophagy in the lung epithelia and that IPF, like many neurodegenerative conditions, may reflect an acquired and potentially aged-related defect in one or more arms of cellular quality control.

SURFACTANT MUTATIONS, EPITHELIAL DYSFUNCTION, AND IPF PATHOGENESIS

The three components of the surfactant system associated with ILD opportunistically represent distinct substrates (a bitopic integral membrane protein, a soluble luminal protein, and a polytopic membrane transporter) with which to assess cellular responses and aberrant signaling present in the distal lung epithelia in IPF. As the first of these to be described, SFTPC represents a starting point to understand AT2 cell dysfunction in both sporadic IPF and FPF. Since the initial descriptions of the clinical proSP-C mutants, SP-CΔexon4 and SP-CL188Q, nearly 15 years ago, over 60 additional SFTPC mutations have been reported in the literature in association with parenchymal lung disease (summarized in Table 1). These can be grouped on the basis of their functional behavior and domain location within the proSP-C sequence. The potential targets and various processes that have been revealed by their expression in cells, preclinical animal models, and patients are schematically illustrated in Fig. 3.

Table 1.

Summary of reported phenotypic features for surfactant component mutations

Mutation (Domain) Clinical Diagnosis Lung Phenotype (in vivo) Subcellular Localization Trafficking Cellular Responses (in vitro) References
SFTPA2
F198S (CRD)
G231V (CRD)
Familial pulmonary fibrosis Total BAL [SP-A] Normal ER retention
Intracellular aggregation
Not secreted
(+) ER stress, cleared by ERAD (+)
TGFβ1 elaboration
99, 100, 175
SFTPC
Group A1
    ΔExon4 (BRICHOS)
 L188Q (BRICHOS)
 G100S (BRICHOS)
NSIP (Children)
IPF/UIP (Adult)
Absence of mature SP-C (humans)
Arrested lung development (mice)
ER stress (humans; mice)
↑Sensitivity to bleomycin (mice)
Epithelial cytotoxicity
ER retention→ aggresomes
Intracellular aggregates
ERAD requires Erdj 4/5
MG132 blocks degradation
4-PBA improves aggregates
(+) ER stress
(+) Apoptosis
(+) Incomplete or absent proSP-C processing
(+) IL-8/TGFβ1 expression
(+) Polyubiquitinated isoforms
21, 39, 97, 98, 100, 111, 112, 116, 117, 120, 153, 159, 160, 173, 193
Group A2
    L110R (BRICHOS)
 P115L (BRICHOS)
 A116D (BRICHOS)
Unspecified ILD
Unspecified ILD
Unspecified chILD
Phenotype not reported EEA-1 (+); Syntaxin2
(−) Intracellular aggregation
↓ PC secretion
(+) Aberrant processing, ↓ cell viability
↑ HSP response
(+) Congo red aggregates
160, 193
Group B1
    E66K (Linker)
 I73T (Linker)
NSIP/PAP (Child)
IPF/UIP (Adult)
↑ Phospholipid; ↑SP-A, PAS positive staining
Biopsy: PM and EE localization
Misprocessed SP-C (BAL)
Misprocessed SP-B (BAL)
Plasma membrane→EE→LE/MVB (+) Aberrantly processed protein
(+) Late autophagy block
↓ Mitophagy
↑ Mysfunctional mitochondria
1, 19, 24, 26, 49, 116, 118, 128, 152
Group B2
    Δ91–93 (Non-BRICHOS) NSIP/PAP ↓ BAL SP-B
↑ BAL SP-A ↓ Surfactant surface tension
(+) Intracellular aggregates
(+) Congo red staining
Plasma membrane
EEA1 (+) compartments
Not reported 55, 181
Group C
    P30L (NH2-terminal) Unspecified ILD Phenotype not reported (+) ER retention ↑ Bip expression
(+) Polyubiquitinated isoforms
13, 116, 160
ABCA3
Group I (Trafficking Defective)
    L101P (1st luminal loop)
 R280C (1st cytosolic loop)
 L982P (3rd luminal loop)
 G1221S (11th TM domain)
 L1553P (COOH-terminal)
 Q1591P (COOH-terminal)
Surfactant deficiency*
RDS*
chILD†
Phenotype not reported
Phenotype not reported
Phenotype not reported
(+) ER retention
Non-LRO cytosolic vesicles
(+) ER stress 30, 31, 103, 147, 172, 177
Group II (Functionally Defective)
    R43L (1st luminal loop)
 D253H (1st luminal loop)
 E292V (1st cytosolic loop)
 N568D (ABC1)
 E690K (ABC1)
 T1114M (8thTM domain)
 T1173R (1st luminal loop)
 L1580P (COOH-terminal)
Surfactant deficiency* RDS* chILD (CPI) Reduced SP-B and SP-C (−) ER retention
Lysosomes or LROs (normal)
Impaired lipid transport
Impaired ATP hydrolysis
Impaired ATP binding
Abnormal LBs
↑ IL8 secretion
20, 25, 103, 104, 147, 148, 177
*

Seen with homozygous or compound heterozygous ABCA3 expression;

found with heterozugous ABCA3 expression.

S. Mulugeta, unpublished observations. BAL, bronchoalveolar lavage; chILD, interstitial lung disease of childhood; CPI, chronic pneumonitis of infancy; CRD, carbohydrate recognition domain; EE, early endosomes; EEA1, early endosomal antigen-1; ER, endoplasmic reticulum; ERAD, ER-associated protein degradation; HSP, heat shock response; ILD, interstitial lung disease; IPF, idiopathic pulmonary fibrosis; LBs, ; LE/MVB, ; LRO, lysosome-related organelle; NSIP, nonspecific interstitial pneumonitis; PAP, pulmonary alveolar proteinosis; PAS, periodic acid Schiff; 4-PBA, 4-phenylbutyric acid; PC, phosphatidyl choline; PM, plasma membrane; RDS, respiratory distress syndrome of newborn; SP, surfactant protein; TM, ; UIP, usual interstitial pneumonitis.

Clues from SFTPC BRICHOS Mutations: Aggresomes, ER Stress, Cytokines, EMT, and Apoptosis

The Group A1 mutations prone to misfolding tend to cluster around the conserved cysteine residues (cys121/cys189) in the SP-C BRICHOS domain. When the prototypical SP-CΔexon4 and SP-CL188Q BRICHOS mutants are expressed in a variety of in vitro model systems, they are initially retained in the ER, fail to undergo appropriate proteolytic processing, do not end up in LB, and have a high propensity to form intracellular aggregates. Alanine mutagenesis of either or both cysteines generates laboratory-mutant isoforms that indicate these aggregates are actually aggresomes whose formation can be blocked with microtubule inhibitors, exacerbated by proteasome inhibitors, and rescued by the chemical chaperone 4-phenylbutyric acid (4-PBA) (69, 153). Aggresomes are also observed with clinical BRICHOS mutants (173). Additional mutations in the more proximal BRICHOS domain termed Group A2 (A116D; L110R) are less well characterized but are also excluded from LRO/LB, partially localize in LAMP3/EEA1 (LE/MVB) compartments, induce a UPR, and form Congo red aggregates in vitro (160, 193). The sole Group C mutant reported to date (P30L) is also ER retained and polyubiquitinated (presumably for ERAD) (13, 160).

Using SP-CΔexon4 and SP-CL188Q mutants as substrates, misfolded BRICHOS conformers are also capable of activating one of more of the three UPR sensors, IRE1/XBP1, ATF6, and PERK/eIF2α, in both cell lines and primary rodent AT2 cells (97, 98, 111, 112). The ultimate failure of the UPR to correct mutant protein load and alleviate ER stress appears to induce a series of downstream events that then contribute to AT2 cell dysfunction and promote aberrant injury/repair responses.

Sounding the alarm: inflammation.

Persistent UPR activation by these mutant isoforms appears to induce an ER-mediated “alarm” response. Transient expression of SP-C BRICHOS domain mutants promoted IL-8 release from cultured epithelial cells. The release of IL-8 was mechanistically linked to the activation of JNK signaling and was associated with activation of nuclear factor-kappa light-chain-enhancer of activated B cells (NF-κB) (97). Stably transfected cell lines expressing SP-CΔexon4 appear to adapt to chronic ER stress via an NF-κB-dependent pathway and suggest this may be a prosurvival response, but one capable both of elaborating additional inflammatory mediators and of being exploited by second hits such as viral infections (22).

Apoptosis.

Given the role of AT2 cells as progenitors (16), their depletion poses a threat to the overall capacity of the lung for repair/regeneration. ER stress induced by mutant proSP-C can induce intrinsic apoptotic death via induction of numerous pathways including caspase 4 activation, mitochondrial cytochrome c release, c-Jun NH2-terminal kinase (JNK) signaling, and caspase 3 activation (97, 98, 111, 112). Interestingly, apoptosis induced by pharmacological proteasome inhibition or overexpression of an SP-C BRICHOS mutant can be significantly inhibited by angiotensin receptor blockade (saralasin) or the angiotensin II counterregulatory peptide Ang1-7, showing that ER stress-induced intrinsic apoptosis in human AECs can involve autocrine pathways such as the ANGII/ANG1-7 system (163).

Changes in cell phenotype.

Although somewhat controversial as to their contribution to total fibrotic burden, a variety of studies have suggested that epithelial cells in the lungs and other organs can undergo phenotypic changes with the adoption of mesenchymal phenotypes. Termed epithelial-mesenchymal transition (EMT), this phenomenon probably represents an adaptive response to cellular stress. In the alveolar epithelium and in cell lines, EMT driven by ER stress induced either pharmacologically or by using proSP-C mutants has been shown. In primary rodent AT2 cells, Zhong et al. (195) have shown that treatment with tunicamycin or thapsigargin induces α-smooth muscle actin expression and decreased E-cadherin that was accompanied by morphological changes consistent with a shift to a mesenchymal phenotype. Overexpression of SP-CΔexon4 in this model largely recapitulated these effects. Tanjore et al. (157) reported similar findings following induction of ER stress by overexpression of the SP-CL188Q mutant and identified SMAD2/3 signaling as a mediator of these events. Importantly, both studies demonstrated a role for SRC family kinases and suggest that therapies based on mitigating ER stress may have pluripotent benefits.

Beyond SP-C: Aberrant Cellular Responses to Other Surfactant Component Substrates

The cellular responses observed above are not idiosyncratic with proSP-C BRICHOS mutants, given that studies utilizing mutant SFTPA2 and ABCA3 substrates have both corroborated the pathways discussed above and extended the findings with several new observations.

SFTPA (SP-A).

In 2009, Wang and colleagues (175) using genomewide linkage scanning, reported on two kindreds with pulmonary fibrosis and lung adenocarcinomas in which rare heterozygous missense mutations were identified in SFTPA2 alleles. Resulting from the substitution of valine for glycine at codon 231 (G231V) or serine for phenylalanine at codon 198 (F198S) in the CRD of the SP-A protein, when expressed in lung cell lines or primary murine AT2 cells, the resulting mutant SFTPA2 protein products were proximally retained, failed to be secreted, and activated the UPR (99). Subsequently the same group found elevated amounts of TGF-β1, a pleotropic profibrotic cytokine linked to both myofibroblast activation and EMT, in the bronchoalveolar lavage (BAL) of patients expressing the G231V mutation. Primary mouse AT2 cells expressing G231V in vitro also elaborated significant amounts of TGF-β as well as two TGF-β binding proteins (LTBP-1 and LTBP-4) (100).

ABCA3.

Although not restricted to any particular domain, functional characterization of a subgroup of ABCA3 mutations (including L101P, L982P, L1553P, Q1591P, G1221S) by transient or stable expression results in their total or partial retention in the ER (30, 103). Designated as “Type I” (Table 1), this mutant subgroup can enhance expression of Bip, increase XBP1 splicing, and induce apoptosis in vitro (177). Recent work with both the Type I clinical ABCA3 mutants as well as laboratory-generated glycosylation-deficient constructs suggest that the misfolded isoforms are substrates for ERAD (14, 177). Taken in total these data are highly supportive of a common model of ER stress and failed proteostasis from misfolded surfactant components.

A Role for Induced AT2 ER Stress and Apoptosis in IPF Pathogenesis

Mechanistic relationships between ER stress-induced epithelial dysfunction and lung fibrosis identified from the in vitro studies described above have been further strengthened using mouse models. Bridges et al. (21) reported that transgenic mice constitutively expressing SP-CΔexon4 exhibit a gene dose-dependent arrest of lung morphogenesis, induction of Bip expression, and aberrant proSP-C processing. More recently, conditional expression of SP-CL188Q in AT2 cells of adult mice at induced levels of ∼20% of the endogenous gene expression produced ER stress (90). Although lacking evidence of gross fibrosis at baseline, these mice exhibited a greater sensitivity to low-dose bleomycin with marked fibrotic changes and enhanced AT2 cell apoptosis. The direct role of ER stress in this process was further supported by exposure of wild-type mice to tunicamycin treatment to independently induce ER stress, which again enhanced the fibrotic response to bleomycin (90).

ER Stress Signatures Are Not Exclusive to Cells, Mice, or Even SFTPC IPF Families

The molecular signatures for ER stress and UPR pathways identified by the in vitro systems and mouse models have been subsequently detected in the lungs of IPF patients. Lawson and colleagues (91) demonstrated evidence of UPR activation markers (including BiP, EDEM, and XBP1) in FPF patients expressing SP-CL188Q mutations as well as in samples both from non-SFTPC FPF patients and in sporadic cases of IPF. Korfei et al. (83) extended these findings by reporting molecular signatures of multiple UPR pathways and apoptosis in AT2 cells including ATF6, ATF4, XBP1, CHOP, Bax, and caspase 3 exclusively in patients with sporadic IPF with pathological features of UIP. Thus chronic ER stress in the alveolar epithelium likely represents a broad mechanism for the pathogenesis of ILD (156).

AT2 Quality Control Responses

An important concept, especially in the context of quality control issues discussed above, is that the cellular pathology and molecular signatures of ER stress induced by BRICHOS proSP-C expression must invariably result from of an imbalance in or failure of the cellular quality control network to accommodate the offending protein conformers and dysfunctional organelles. Very little is currently known about the actual quality control repertoire of AT2 cells. Using microarray analyses, Dong et al. (39) identified genes specifically induced in response to expression of SP-CΔexon4 and SP-CL188Q and showed that two BiP cochaperones, ER-localized DnaJ homologues ERdj4 and ERdj5, selectively bind to mutant but not wild-type proSP-C and were required for successful retrotranslocation and proteasomal degradation of BRICHOS mutant SP-C substrates. In a preliminary report, we have shown that although primary human AT2 cells exhibit a significant degree of both proteasomal activity and autophagy, these cells appear to rely primarily on ERAD for control either of misfolded proSP-C conformers or more generally of a variety of PolyQ huntingin substrates (57). Similarly, Maitra and colleagues (99) have shown that primary mouse AT2 cells in vitro rely almost exclusively on MG132-sensitive proteasomal activity to clear misfolded SP-A2 mutants. Although not appearing to have a primary role in basal proteostasis, the function of autophagy either as a backup to ERAD and/or in the quality control of AT2 organelles such as mitochondria, LB, or nonprotein substrates remains an important area for further investigation.

Beyond ER Stress and Apoptosis: Dominant Negative or Loss-of-Function Effects

Patients heterozygous for the SP-CΔexon4 mutation do not express mature SP-C (117). Mechanistically, because proSP-C sorting involves homomeric association of proSP-C monomers mediated by the mature SP-C domain, through heterotypic interactions, SP-C BRICHOS mutants can act as dominant negatives to direct wild-type proSP-C away from normal routing and into aggregates (174). Because SP-B and SP-C are felt to be functionally redundant with respect to overall surfactant biophysical activity, the question of whether a lack of mature SP-C in surfactant would contribute to IPF pathogenesis is unanswered. As expected, SP-C null mice are viable at birth and grow normally without a lung phenotype (45) but exhibit enhanced sensitivity to bleomycin fibrosis (93) and respiratory syncytial virus infection (47). However, when the original strain of SP-C null mice are backcrossed onto a different genetic background (129/Sv), the absence of SP-C is associated with spontaneous inflammation and lung remodeling (46), suggesting that SP-C deficiency is a potential disease modifying factor.

In a related fashion, when expressed in primary rodent AT2 cells, the disease-causing human SPA2 mutants are not secreted (99). Furthermore, in epithelial cell lines simultaneously transfected with SP-A2 isoforms and wild-type SP-A1, coimmunoprecipitation resulted in recovery of wild-type SP-A2 from the medium but not SP-A2 mutants (G231V or F198S) (175). Although total SP-A levels do not appear to be altered by heterozygous expression of SP-A2 mutants in vivo, given that the SP-A trimer and 18-mer are from combinations of SP-A1 and SP-A2, the resultant consequences of an SP-A1 restricted SP-A on innate lung immunity and/or IPF pathogenesis remains to be clarified.

In contrast to posttranslationally acquired SP-C or SP-A deficiency, functionally impaired ABCA3 mutations (designated Type II mutations) have been identified, with homozygous expression producing severe loss of function phenotypes (30, 103, 104) (Table 1). Affected patients as well as ABCA3 null mice demonstrate complete absence of protein expression, deformed LBs, and die from respiratory distress shortly after birth (31, 43, 147, 172), indicating a critical role in production of pulmonary surfactant. However, heterozygous expression of ABCA3 mutations, with a resultant partial loss-of-function caused either by ER retention (Type I), improper lipid-pump function (Type II), or both (Type I/II compound heterozygotes) have each been reported and appear to act as genetic modifiers of lung diseases associated with SFTPC mutations in children (24). In one adult FPF kindred all carrying the same SFTPC mutation, a functional genetic variant in ABCA3 (E292V) also found in the previous pediatric study appeared to also affect disease penetrance, suggesting interplay between the two genotypes (37), not surprising given their overlapping biology within the AT2 cell.

SFTPC Trafficking Mutations: a New Phenotype of Acquired Quality Control Disruption

Beginning in 2004, a series of missense mutations clustering within a 40-amino-acid span encoded on exon 3 of human SFTPC was reported, heralding an emerging diversity of molecular and clinical phenotypes induced by SFTPC expression (19, 55, 152). The first of these detailed a de novo heterozygous missense mutation (g.1286T>C) resulting in substitution of threonine for isoleucine (I73T) in an infant with NSIP (19). Interestingly, the identical SP-CI73T mutation (both inherited and sporadic) has reappeared in several other published reports of parenchymal lung disease in children and adults and is now felt to be the most common SFTPC mutation (26, 116). Collectively the Group B mutations (E66K; I73T; Δ91–93) (55, 152) all reside within the linker domain of proSP-C (Fig. 1). Although presentation and course can vary (1, 128), overlapping features that emerge include interstitial lung remodeling with intra-alveolar accumulation of surfactant (lipids, SP-A, SP-B), the presence of aberrantly processed proSP-C, and abnormal AT2 cell cytoplasmic inclusions on EM, suggesting abnormalities in surfactant homeostasis (19, 44, 58, 152). Importantly since mature SP-C can be detected in BAL from these patients, the non-BRICHOS trafficking mutants do not appear to act as dominant negatives (19, 162).

The cellular phenotype induced by Group B SFTPC mutants is also distinct from that of Group A1 (Table 1). Functionally, Group B SFTPC mutants have been defined by in vitro expression studies (mainly using SP-CI73T), demonstrating a unique and markedly aberrant intracellular trafficking pattern for proSP-C first directly to the plasma membrane with subsequent internalization and accumulation within early endosomes and LE/MVB (15, 153) (Fig. 3). This recapitulated the expression pattern seen on a biopsy of a patient with SFTPCE66K showing proSP-C staining on the plasma membrane and in early endosomes of AT2 cells (19). Recently, two separate preliminary studies using knockin or transgenic expression strategies have each demonstrated that the distribution of mutant SP-CI73T in AT2 cells in vivo is nearly identical with the pattern noted in cell lines and patients (118, 176).

The link between the mistrafficking behavior and cellular dysfunction, which does not involve ER stress (111), has recently been defined in vitro with cell lines expressing hSP-CI73T. As illustrated in Fig. 3, SP-CI73T induces a novel quality control phenotype punctuated by a striking disruption of macroautophagy characterized by the accumulation of dysmorphic, autophagic vacuoles containing organellar and proteinaceous debris (58). The in vitro findings from hSP-CI73T cell lines phenocopied ultrastructural changes seen in AT2 cells on EM of a lung biopsy from a SFTPC I73T patient (58). Biochemically, hSP-CI73T cells exhibited increased expression of Atg8/LC3, SQST/p62, and RAB7, consistent with a late block in autophagic vacuole maturation that was confirmed by autophagy flux studies. The observed disrupted autophagy resulted in a diminished functional capacity of hSP-CI73T cells to degrade model protein aggregate substrates and impairment in mitophagy, producing an accumulation of dysfunctional mitochondria. The disruption of autophagy-dependent proteostasis and mitophagy induced by expression of SP-CI73T is reminiscent of cellular quality control phenotypes observed in organellar storage disorders such as Niemann-Pick disease (140, 146), which has been associated with a premature development of fibrotic lung disease.

In translational studies of IPF from two separate groups, lung tissue from IPF patients demonstrates evidence of decreased autophagy as defined by increases in LC3, p62, and polyubiquitinated protein expression and decreased numbers of autophagosomes (3, 126). These findings, when combined with the aforementioned work on aging (23), emphasize the potential important role of impaired autophagy, mitophagy, and mitochondrial dysfunction in lung epithelia to IPF.

Impact of Environmental Factors on Lung Quality Control

Substantial clinical variability in the age of onset and disease severity in both surfactant mutation-associated (SP-C, ABCA3, SP-A2) ILD and sporadic IPF suggests a possible role for environmental factors and/or for modifier genes. In addition to the aging lung and its effects on quality control discussed above, given the direct exposure of the lung epithelium to the “outside world,” environmental insults may participate in a two-hit model of AT2 dysfunction to promote fibrosis. In support of this, IPF is found more frequently in cigarette smokers (11, 142), and cigarette smoking is one of the strongest associated risk factors for the development of diffuse parenchymal lung disease (DPLD) (151). Cigarette smoke exposure induces ER stress and proteasomal dysfunction in alveolar epithelial cell lines in vitro (149, 168) and signatures of UPR activation have been detected in the smoke-exposed human lung in vivo (77).

Similarly, emerging evidence also implicates human herpesviruses (including herpesvirus-8; Epstein-Barr virus; cytomegalovirus) in IPF, and one or more of these viruses have been detected in the lungs of up to 97% of tested patients with IPF (40, 86, 91, 155, 188). In a mouse model of aging, infection with the murine herpesvirus lead to upregulation of multiple components of the UPR/ER stress pathway and accompanied by the development of lung fibrosis (161). In human IPF, cytomegalovirus antigens and the UPR transcription factor XBP-1 were colocalized in alveolar epithelial cells from an individual with UIP (91). More recently, sampling of asymptomatic relatives of patients with familial IPF showed evidence of herpesvirus infection and ER stress (87).

In a recent clinical study, IPF exacerbations were significantly associated with antecedent 6-wk increases in air pollution exposures of ozone and nitrogen dioxide, suggesting that air pollution may contribute to the development of this clinically meaningful event (64, 91). Although the mechanism was not assessed, given the oxidative-nitrative stress applied by these agents, it seems likely that additional proteotoxic and mitochondrial stress to the quality control network in the distal lung from the enhanced generation of “challenged substrates” by these exposures may be contributing factors.

CONCLUSIONS, UNANSWERED QUESTIONS, AND OPPORTUNITIES

Over the past 15 years much progress has been made in understanding the pathogenesis of IPF. There is now substantial evidence to show that the AT2 epithelium is altered in the fibroblastic foci of IPF lungs and that the AT2 cell is a pivotal regulator of lung injury, inflammation, and repair. Although SFTPC, SFTPA, and ABCA3 mutations are rare (2–20% in most ILD cohorts) (92, 167), understanding both the spectrum of clinical phenotypes and the molecular pathogenesis of the AT2 cell dysfunction caused by these mutations adds value to understanding the broader mechanisms that underlie the pathogenesis of sporadic IPF.

The data summarized here generated by using surfactant protein mutations as a platform has revealed the role for both ER stress and alterations in cellular quality control pathways in epithelial cells in IPF pathogenesis (Fig. 3). The fact that characteristic cellular phenotypes are induced by distinct classes of surfactant component mutations and that these converge at the epithelial cell should spur further focus on efforts to understand AT2 biology in health and disease. Coupled with the emergence of confounding factors in IPF pathogenesis such as age, environmental exposures, and infection, the findings summarized in this review strongly support a two-hit (or multiple-hit) model whereby the repetitively injured AT2 cell population serves as a driver of disordered repair and injury. Although not discussed here, the well-known associations of telomerase mutations with familial IPF (4), which can also be thought of as a “quality control disorder” (for telomeres), also converge on this epithelial-centric model of IPF and highlight the importance for aging-associated processes and senescence that were not previously linked to IPF. In this context, the role of the recently described Muc5B promoter polymorphism in IPF remains to be clarified (127, 143).

Given the above model, it should come as no surprise that the therapies to date have done little to stabilize or reverse the fibrotic process (18). Analogous to cancer, it is likely that future therapeutic options for IPF should involve directed therapies aimed at engaging multiple targets including the epithelial cell dysfunction, fibroblast activation, and matrix deposition. This approach is probably not amenable to single drug therapy. Furthermore, experience with the surfactant protein component mutations reveals that the final common histopathological pathway seen as UIP or chILD likely reflects distinct combinations of several different upstream molecular events. Identifying the specific pathogenic pathway signatures in patients will be critical to identify personalized therapies based on pathogenesis not histology.

Finally, given the protracted period for drug development and the overlap of many of the pathways discussed above (ER stress, misfolding, aggregation, mistrafficking, autophagy, apoptosis, and MAP kinase signaling) with those seen in other chronic degenerative diseases (Alzheimer's disease, Huntington's disease, cystic fibrosis, and α1-antitrypsin deficiency), it may be worth considering a cross-fertilization of IPF research/treatment strategies with these disease areas by adopting experimental paradigms and reagents, coupled with readily available compound libraries, and approved drugs that already target some of these signatures in an attempt to open new therapeutic avenues for treatment of this devastating disease.

GRANTS

This research was supported by VA Merit Award BX001176 (M. F. Beers); NIH HL 119436 (M. F. Beers), NIH P30 ES013508 (M. F. Beers), and American Heart Association Research Grant 13GRNT17070104 (M. F. Beers).

M. F. Beers is an Albert M. Rose Established Investigator of the Pulmonary Fibrosis Foundation.

DISCLOSURES

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

AUTHOR CONTRIBUTIONS

S.M., S.-I.N., and M.F.B. conception and design of research; S.M. and S.-I.N. performed experiments; S.M. prepared figures; S.M. and M.F.B. drafted manuscript; S.M., S.-I.N., and M.F.B. edited and revised manuscript; S.M., S.-I.N., and M.F.B. approved final version of manuscript.

ACKNOWLEDGMENTS

The authors acknowledge the help and support of many mentors, collaborators, technical staff, and trainees over the years including Aron B. Fisher, M.D.; Philip L. Ballard, M.D., Ph.D.; Susan H. Guttentag, M.D.; Sandra R. Bates, Ph.D.; Frank Brasch, M.D.; Paul N. Stevens, M.D.; Larry Nogee, M.D.; Aaron Hamvas, M.D.; Michael Koval, Ph.D.; Sheldon I. Feinstein, Ph.D.; Linda Gonzalez, Ph.D.; Albert Kabore, Ph.D.; Chi Y. Kim, M.D.; Catherine A. Lomax, Scott J. Russo, Amy L. Johnson, V.M.D.; and Wen-Jing Wang, M.D.

We apologize to the many additional colleagues in advance whose work, for the limits of space, could not be cited.

REFERENCES

  • 1.Abou TR, Jaubert F, Emond S, Le Bourgeois M, Epaud R, Karila C, Feldmann D, Scheinmann P, de Blic J. Familial interstitial disease with I73T mutation: a mid- and long-term study. Pediatr Pulmonol 44: 167–175, 2009. [DOI] [PubMed] [Google Scholar]
  • 2.Alder JK, Barkauskas CE, Limjunyawong N, Stanley SE, Kembou F, Tuder RM, Hogan BLM, Mitzner W, Armanios M. Telomere dysfunction causes alveolar stem cell failure. Proc Natl Acad Sci USA 112: 5099–5104, 2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Araya J, Kojima J, Takasaka N, Ito S, Fujii S, Hara H, Yanagisawa H, Kobayashi K, Tsurushige C, Kawaishi M, Kamiya N, Hirano J, Odaka M, Morikawa T, Nishimura SL, Kawabata Y, Hano H, Nakayama K, Kuwano K. Insufficient autophagy in idiopathic pulmonary fibrosis. Am J Physiol Lung Cell Mol Physiol 304: L56–L69, 2013. [DOI] [PubMed] [Google Scholar]
  • 4.Armanios MY, Chen JJL, Cogan JD, Alder JK, Ingersoll RG, Markin C, Lawson WE, Xie M, Vulto I, Phillips JA, Lansdorp PM, Greider CW, Loyd JE. Telomerase mutations in families with idiopathic pulmonary fibrosis. N Engl J Med 356: 1317–1326, 2007. [DOI] [PubMed] [Google Scholar]
  • 5.Atochina-Vasserman EN, Bates SR, Zhang P, Abramova H, Zhang Z, Gonzales L, Tao JQ, Gochuico BR, Gahl W, Guo CJ, Gow AJ, Beers MF, Guttentag S. Early alveolar epithelial dysfunction promotes lung inflammation in a mouse model of Hermansky-Pudlak syndrome. Am J Respir Crit Care Med 184: 449–458, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Balch WE, Morimoto RI, Dillin A, Kelly JW, Balch WE, Morimoto RI, Dillin A, Kelly JW. Adapting proteostasis for disease intervention. Science 319: 916–919, 2008. [DOI] [PubMed] [Google Scholar]
  • 7.Balch WE, Sznajder JI, Budinger S, Finley D, Laposky AD, Cuervo AM, Benjamin IJ, Barreiro E, Morimoto RI, Postow L, Weissman AM, Gail D, Banks-Schlegel S, Croxton T, Gan W. Malfolded protein structure and proteostasis in lung diseases. Am J Respir Crit Care Med 189: 96–103, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Ban N, Matsumura Y, Sakai H, Takanezawa Y, Sasaki M, Arai H, Inagaki N. ABCA3 as a lipid transporter in pulmonary surfactant biogenesis. J Biol Chem 282: 9628–9634, 2007. [DOI] [PubMed] [Google Scholar]
  • 9.Baritussio AG, Magoon MW, Goerke J, Clements JA. Precursor-product relationship between rabbit type II cell lamellar bodies and alveolar surface-active material. Surfactant turnover time. Biochim Biophys Acta 666: 382–393, 1981. [DOI] [PubMed] [Google Scholar]
  • 10.Barkauskas CE, Cronce MJ, Rackley CR, Bowie EJ, Keene DR, Stripp BR, Randell SH, Noble PW, Hogan BL. Type 2 alveolar cells are stem cells in adult lung. J Clin Invest 123: 3025–3036, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Baumgartner KB, Samet JM, Stidley CA, Colby TV, Waldron JA. Cigarette smoking: a risk factor for idiopathic pulmonary fibrosis. Am J Respir Crit Care Med 155: 242–248, 1997. [DOI] [PubMed] [Google Scholar]
  • 12.Beers MF, Hawkins A, Shuman H, Zhao M, Newitt JL, Maguire JA, Ding W, Mulugeta S. A novel conserved targeting motif found in ABCA transporters mediates trafficking to early post-Golgi compartments. J Lipid Res 52: 1471–1482, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Beers MF, Mulugeta S. Surfactant protein C biosynthesis and its emerging role in conformational lung disease. Annu Rev Physiol 67: 663–696, 2005. [DOI] [PubMed] [Google Scholar]
  • 14.Beers MF, Zhao M, Tomer Y, Russo SJ, Zhang P, Gonzales LW, Guttentag SH, Mulugeta S. Disruption of N-linked glycosylation promotes proteasomal degradation of the human ATP-binding cassette transporter ABCA3. Am J Physiol Lung Cell Mol Physiol 305: L970–L980, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Beers MF, Hawkins A, Maguire JA, Kotorashvili A, Zhao M, Newitt JL, Ding W, Russo S, Guttentag S, Gonzales L, Mulugeta S. A nonaggregating surfactant protein C mutant is misdirected to early endosomes and disrupts phospholipid recycling. Traffic 12: 1196–1210, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Beers MF, Morrisey EE. The three R's of lung health and disease: repair, remodeling, and regeneration. J Clin Invest 121: 2065–2073, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Bence NF, Sampat RM, Kopito RR. Impairment of the ubiquitin-proteasome system by protein aggregation. Science 292: 1552–1555, 2001. [DOI] [PubMed] [Google Scholar]
  • 18.Blackwell TS, Tager AM, Borok Z, Moore BB, Schwartz DA, Anstrom KJ, Bar-Joseph Z, Bitterman P, Blackburn MR, Bradford W, Brown KK, Chapman HA, Collard HR, Cosgrove GP, Deterding R, Doyle R, Flaherty KR, Garcia CK, Hagood JS, Henke CA, Herzog E, Hogaboam CM, Horowitz JC, King TE Jr, Loyd JE, Lawson WE, Marsh CB, Noble PW, Noth I, Sheppard D, Olsson J, Ortiz LA, O'Riordan TG, Oury TD, Raghu G, Roman J, Sime PJ, Sisson TH, Tschumperlin D, Violette SM, Weaver TE, Wells RG, White ES, Kaminski N, Martinez FJ, Wynn TA, Thannickal VJ, Eu JP. Future directions in idiopathic pulmonary fibrosis research. An NHLBI workshop report. Am J Respir Crit Care Med 189: 214–222, 2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Brasch F, Griese M, Tredano M, Johnen G, Ochs M, Rieger C, Mulugeta S, Muller KM, Bahuau M, Beers MF. Interstitial lung disease in a baby with a de novo mutation in the SFTPC gene. Eur Respir J 24: 30–39, 2004. [DOI] [PubMed] [Google Scholar]
  • 20.Brasch F, Schimanski S, Muhlfeld C, Barlage S, Langmann T, Aslanidis C, Boettcher A, Dada A, Schroten H, Mildenberger E, Prueter E, Ballmann M, Ochs M, Johnen G, Griese M, Schmitz G. Alteration of the pulmonary surfactant system in full-term infants with hereditary ABCA3 deficiency. Am J Respir Crit Care Med 174: 571–580, 2006. [DOI] [PubMed] [Google Scholar]
  • 21.Bridges JP, Wert SE, Nogee LM, Weaver TE. Expression of a human SP-C mutation associated with interstitial lung disease disrupts lung development in transgenic mice. J Biol Chem 278: 52739–52746, 2003. [DOI] [PubMed] [Google Scholar]
  • 22.Bridges JP, Xu Y, Na CL, Wong HR, Weaver TE. Adaptation and increased susceptibility to infection associated with constitutive expression of misfolded SP-C. J Cell Biol 172: 395–407, 2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Bueno M, Lai YC, Romero Y, Brands J, Croix CMS, Kamga C, Corey C, Herazo-Maya JD, Sembrat J, Lee JS, Duncan SR, Rojas M, Shiva S, Chu CT, Mora AL. PINK1 deficiency impairs mitochondrial homeostasis and promotes lung fibrosis. J Clin Invest 125: 521–538, 2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Bullard JE, Nogee LM. Heterozygosity for ABCA3 mutations modifies the severity of lung disease associated with a surfactant protein C gene (SFTPC) mutation. Pediatr Res 62: 176–179, 2007. [DOI] [PubMed] [Google Scholar]
  • 25.Bullard JE, Wert SE, Whitsett JA, Dean M, Nogee LM. ABCA3 mutations associated with pediatric interstitial lung disease. Am J Respir Crit Care Med 172: 1026–1031, 2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Cameron HS, Somaschini M, Carrera P, Hamvas A, Whitsett JA, Wert SE, Deutsch G, Nogee LM. A common mutation in the surfactant protein C gene associated with lung disease. J Pediatr 146: 370–375, 2005. [DOI] [PubMed] [Google Scholar]
  • 27.Caniard A, Adnot S, Eickelberg O, Meiners S. Proteasome function in aging lungs (Abstract). Am J Respir Crit Care Med 189: A5399, 2014. [Google Scholar]
  • 28.Cecarini V, Bonfili L, Cuccioloni M, Mozzicafreddo M, Rossi G, Buizza L, Uberti D, Angeletti M, Eleuteri AM. Crosstalk between the ubiquitin-proteasome system and autophagy in a human cellular model of Alzheimer's disease. Biochim Biophys Acta 1822: 1741–1751, 2012. [DOI] [PubMed] [Google Scholar]
  • 29.Chen F, Yang DS, Petanceska S, Yang A, Tandon A, Yu G, Rozmahel R, Ghiso J, Nishimura M, Zhang DM, Kawarai T, Levesque G, Mills J, Levesque L, Song YQ, Rogaeva E, Westaway D, Mount H, Gandy S, George-Hyslop P, Fraser PE. Carboxyl-terminal fragments of Alzheimer beta-amyloid precursor protein accumulate in restricted and unpredicted intracellular compartments in presenilin 1-deficient cells. J Biol Chem 275: 36794–36802, 2000. [DOI] [PubMed] [Google Scholar]
  • 30.Cheong N, Madesh M, Gonzales LW, Zhao M, Yu K, Ballard PL, Shuman H. Functional and trafficking defects in ATP binding cassette A3 mutants associated with respiratory distress syndrome. J Biol Chem 281: 9791–9800, 2006. [DOI] [PubMed] [Google Scholar]
  • 31.Cheong N, Zhang H, Madesh M, Zhao M, Yu K, Dodia C, Fisher AB, Savani RC, Shuman H. ABCA3 is critical for lamellar body biogenesis in vivo. J Biol Chem 282: 23811–23817, 2007. [DOI] [PubMed] [Google Scholar]
  • 32.Choi AMK, Ryter SW, Levine B. Mechanisms of disease autophagy in human health and disease. N Engl J Med 368: 651–662, 2013. [DOI] [PubMed] [Google Scholar]
  • 33.Conkright JJ, Bridges JP, Na CL, Voorhout WF, Trapnell B, Glasser SW, Weaver TE. Secretion of surfactant protein C, an integral membrane protein, requires the N-terminal propeptide. J Biol Chem 276: 14658–14664, 2001. [DOI] [PubMed] [Google Scholar]
  • 34.Conkright JJ, Apsley KS, Martin EP, Ridsdale R, Rice WR, Na CL, Yang B, Weaver TE. Nedd4-2-mediated ubiquitination facilitates processing of surfactant protein-C. Am J Respir Cell Mol Biol 42: 181–189, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Connors TD, Van Raay TJ, Petry LR, Klinger KW, Landes GM, Burn TC. The cloning of a human ABC gene (ABC3) mapping to chromosome 16p13.3. Genomics 39: 231–234, 1997. [DOI] [PubMed] [Google Scholar]
  • 36.Crapo JD, Young SL, Fram EK, Pinkerton KE, Barry BE, Crapo RO. Morphometric characteristics of cells in the alveolar region of mammalian lungs. Am Rev Respir Dis 128: S42–S46, 1983. [DOI] [PubMed] [Google Scholar]
  • 37.Crossno PF, Polosukhin VV, Blackwell TS, Johnson JE, Markin C, Moore PE, Worrell JA, Stahlman MT, Phillips JA, Loyd JE, Cogan JD, Lawson WE. Identification of early interstitial lung disease in an individual with genetic variations in ABCA3 and SFTPC. Chest 137: 969–973, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Deterding RR. Infants and young children with children's interstitial lung disease. Pediatr Allergy Immunol Pulmonol 23: 25–31, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Dong M, Bridges JP, Apsley K, Xu Y, Weaver TE. ERdj4 and ERdj5 are required for endoplasmic reticulum-associated protein degradation of misfolded surfactant protein C. Mol Biol Cell 19: 2620–2630, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Egan JJ, Stewart JP, Hasleton PS, Arrand JR, Carroll KB, Woodcock AA. Epstein-Barr virus replication within pulmonary epithelial cells in cryptogenic fibrosing alveolitis. Thorax 50: 1234–1239, 1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Engelbrecht S, Kaltenborn E, Griese M, Kern S. The surfactant lipid transporter ABCA3 is N-terminally cleaved inside LAMP3-positive vesicles. FEBS Lett 584: 4306–4312, 2010. [DOI] [PubMed] [Google Scholar]
  • 42.Fisher AB, Dodia C, Ruckert P, Tao JQ, Bates SR. Pathway to lamellar bodies for surfactant protein A. Am J Physiol Lung Cell Mol Physiol 299: L51–L58, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Fitzgerald ML, Xavier R, Haley KJ, Welti R, Goss JL, Brown CE, Zhuang DZ, Bell SA, Lu N, McKee M, Seed B, Freeman MW. ABCA3 inactivation in mice causes respiratory failure, loss of pulmonary surfactant, and depletion of lung phosphatidylglycerol. J Lipid Res 48: 621–632, 2007. [DOI] [PubMed] [Google Scholar]
  • 44.Galetskiy D, Woischnik M, Ripper J, Griese M, Przybylski M. Aberrant processing forms of lung surfactant proteins SP-B and SP-C revealed by high-resolution mass spectrometry. Eur J Mass Spectrom (Chichester, Eng) 14: 379–390, 2008. [DOI] [PubMed] [Google Scholar]
  • 45.Glasser SW, Burhans MS, Korfhagen TR, Na CL, Sly PD, Ross GF, Ikegami W, Whitsett JA. Altered stability of pulmonary surfactant in SP-C-deficient mice. Proc Natl Acad Sci USA 98: 6366–6371, 2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Glasser SW, Detmer EA, Ikegami M, Na CL, Stahlman MT, Whitsett JA. Pneumonitis and emphysema in SP-C gene targeted mice. J Biol Chem 278: 14291–14298, 2003. [DOI] [PubMed] [Google Scholar]
  • 47.Glasser SW, Witt TL, Senft AP, Baatz JE, Folger D, Maxfield MD, Akinbi HT, Newton DA, Prows DR, Korfhagen TR. Surfactant protein C-deficient mice are susceptible to respiratory syncytial virus infection. Am J Physiol Lung Cell Mol Physiol 297: L64–L72, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Groll M, Ditzel L, Lowe J, Stock D, Bochtler M, Bartunik HD, Huber R. Structure of 20S proteasome from yeast at 2.4A resolution. Nature 386: 463–471, 1997. [DOI] [PubMed] [Google Scholar]
  • 49.Guillot L, Epaud R, Thouvenin G, Jonard L, Mohsni A, Couderc R, Counil F, de Blic J, Taam RA, Le Bourgeois M, Reix P, Flamein F, Clement A, Feldmann D. New surfactant protein C gene mutations associated with diffuse lung disease. J Med Genet 46: 490–494, 2009. [DOI] [PubMed] [Google Scholar]
  • 50.Guo CJ, Atochina-Vasserman EN, Abramova E, Foley JP, Zaman A, Crouch E, Beers MF, Savani RC, Gow AJ. S-nitrosylation of surfactant protein-D controls inflammatory function. PLoS Biol 6: e266, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Gustafsson M, Thyberg J, Naslund J, Eliasson E, Johansson J. Amyloid fibril formation by pulmonary surfactant protein C. FEBS Lett 464: 138–142, 1999. [DOI] [PubMed] [Google Scholar]
  • 52.Hacki J, Egger L, Monney L, Conus S, Rosse T, Fellay I, Borner C. Apoptotic crosstalk between the endoplasmic reticulum and mitochondria controlled by Bcl-2. Oncogene 19: 2286–2295, 2000. [DOI] [PubMed] [Google Scholar]
  • 53.Haddad IY, Zhu S, Ischiropoulos H, Matalon S. Nitration of surfactant protein A results in decreased ability to aggregate lipids. Am J Physiol Lung Cell Mol Physiol 270: L281–L288, 1996. [DOI] [PubMed] [Google Scholar]
  • 54.Hahn DR, Na CL, Weaver TE. Reserve autophagic capacity in alveolar epithelia provides a replicative niche for influenza A virus. Am J Respir Cell Mol Biol 51: 400–412, 2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Hamvas A, Nogee LM, White FV, Schuler P, Hackett BP, Huddleston CB, Mendeloff EN, Hsu FF, Wert SE, Gonzales LW, Beers MF, Ballard PL. Progressive lung disease and surfactant dysfunction with a deletion of surfactant protein C gene. Am J Respir Cell Mol Biol 2004: 771–776, 2004. [DOI] [PubMed] [Google Scholar]
  • 56.Haschek WM, Witschi H. Pulmonary fibrosis—a possible mechanism. Toxicol Appl Pharmacol 51: 475–487, 1979. [DOI] [PubMed] [Google Scholar]
  • 57.Hawkins A, Gonzales L, Guttentag S, Mulugeta S, Beers M. Protein quality control in alveolar type 2 cells: the proteasome is essential for control of aggregation-prone SP-C mutations. FASEB J 29 Suppl: 1015–5., 2015. [Google Scholar]
  • 58.Hawkins A, Guttentag SH, Deterding R, Funkhouser WK, Goralski JL, Chatterjee S, Mulugeta S, Beers MF. A non-BRICHOS SFTPC mutant (SP-CI73T) linked to interstitial lung disease promotes a late block in macroautophagy disrupting cellular proteostasis and mitophagy. Am J Physiol Lung Cell Mol Physiol 308: L33–L47, 2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Hedlund J, Johansson J, Persson B. BRICHOS — a superfamily of multidomain proteins with diverse functions. BMC Res Notes 2: 180, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Hetz C, Russelakis-Carneiro M, Maundrell K, Castilla J, Soto C. Caspase-12 and endoplasmic reticulum stress mediate neurotoxicity of pathological prion protein. EMBO J 22: 5435–5445, 2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Hitomi J, Katayama T, Eguchi Y, Kudo T, Taniguchi M, Koyama Y, Manabe T, Yamagishi S, Bando Y, Imaizumi K, Tsujimoto Y, Tohyama M. Involvement of caspase-4 in endoplasmic reticulum stress-induced apoptosis and A beta-induced cell death. J Cell Biol 165: 347–356, 2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Hotamisligil GS. Role of endoplasmic reticulum stress and c-Jun NH2-terminal kinase pathways in inflammation and origin of obesity and diabetes. Diabetes 54: S73–S78, 2005. [DOI] [PubMed] [Google Scholar]
  • 62a.Idiopathic Pulmonary Fibrosis Clinical Research Network, Raghu G, Anstrom KJ, King TE Jr, Lasky JA, Martinez FJ. Prednisone, azathioprine, and N-acetylcysteine for pulmonary fibrosis. N Engl J Med 366: 1968–1977, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Iwata A, Riley BE, Johnston JA, Kopito RR. HDAC6 and microtubules are required for autophagic degradation of aggregated Huntingtin. J Biol Chem 280: 40282–40292, 2005. [DOI] [PubMed] [Google Scholar]
  • 64.Johannson KA, Vittinghoff E, Lee K, Balmes JR, Ji W, Kaplan GG, Kim DS, Collard HR. Acute exacerbation of idiopathic pulmonary fibrosis associated with air pollution exposure. Eur Respir J 43: 1124–1131, 2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Johansson H, Nordling K, Weaver TE, Johansson J, Johansson H, Nordling K, Weaver TE, Johansson J. The Brichos domain-containing C-terminal part of pro-surfactant protein C binds to an unfolded poly-val transmembrane segment. J Biol Chem 281: 21032–21039, 2006. [DOI] [PubMed] [Google Scholar]
  • 66.Johnson AL, Braidotti P, Pietra GG, Russo SJ, Kabore A, Wang WJ, Beers MF. Post-translational processing of surfactant protein-C proprotein- Targeting motifs in the NH2-terminal flanking domain are cleaved in late compartments. Am J Respir Cell Mol Biol 24: 253–263, 2001. [DOI] [PubMed] [Google Scholar]
  • 67.Johnson MA, Kwan S, Snell NJC, Nunn AJ, Darbyshire JH, Turnerwarwick M. Randomized controlled trial comparing prednisolone alone with cyclophosphamide and low-dose prednisolone in combination in cryptogenic fibrosing alveolitis. Thorax 44: 280–288, 1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Johnston JA, Ward CL, Kopito RR. Aggresomes: a cellular response to misfolded proteins. J Cell Biol 143: 1883–1898, 1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Kabore AF, Wang WJ, Russo SJ, Beers MF. Biosynthesis of surfactant protein C: characterization of aggresome formation by EGFP chimeras containing propeptide mutants lacking conserved cysteine residues. J Cell Sci 114: 293–302, 2001. [DOI] [PubMed] [Google Scholar]
  • 70.Kaganovich D, Kopito R, Frydman J. Misfolded proteins partition between two distinct quality control compartments. Nature 454: 1088–1095, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Kaminski WE, Piehler A, Wenzel JJ. ABC A-subfamily transporters: structure, function and disease. Biochim Biophys Acta 1762: 510–524, 2006. [DOI] [PubMed] [Google Scholar]
  • 72.Kammoun H, Chabanon H, Hainault I, Luquet S, Magnan C, Koike T, Ferré P, Foufelle F. GRP78 expression inhibits insulin and ER stress-induced SREBP-1c activation and reduces hepatic steatosis in mice. J Clin Invest 119: 1201–1215, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Kaufman RJ. Orchestrating the unfolded protein response in health and disease. J Clin Invest 110: 1389–1398, 2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Kawaguchi Y, Kovacs JJ, McLaurin A, Vance JM, Ito A, Yao TP. The deacetylase HDAC6 regulates aggresome formation and cell viability in response to misfolded protein stress. Cell 115: 727–738, 2003. [DOI] [PubMed] [Google Scholar]
  • 75.Keller A, Eistetter HR, Voss T, Schafer KP. The pulmonary surfactant protein C (SP-C) precursor is a type II transmembrane protein. Biochem J 277: 493–499, 1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Kelly JW. The environmental dependency of protein folding best explains prion and amyloid diseases. Proc Natl Acad Sci USA 95: 930–932, 1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Kelsen SG, Duan X, Ji R, Perez O, Liu C, Merali S. Cigarette smoke induces an unfolded protein response in the human lung. Am J Respir Cell Mol Biol 38: 541–550, 2008. [DOI] [PubMed] [Google Scholar]
  • 78.Khubchandani KR, Snyder JM. Surfactant protein A (SP-A): the alveolus and beyond. FASEB J 15: 59–69, 2001. [DOI] [PubMed] [Google Scholar]
  • 79.Kim I, Lemasters JJ. Mitophagy selectively degrades individual damaged mitochondria after photoirradiation. Antioxid Redox Signal 14: 1919–1928, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Kim I, Xu W, Reed JC. Cell death and endoplasmic reticulum stress: disease relevance and therapeutic opportunities. Nat Rev Drug Discov 7: 1013–1030, 2008. [DOI] [PubMed] [Google Scholar]
  • 81.King TE, Pardo A, Selman M. Idiopathic pulmonary fibrosis. Lancet 378: 1949–1961, 2011. [DOI] [PubMed] [Google Scholar]
  • 82.Klionsky DJ, Emr SD. Autophagy as a regulated pathway of cellular degradation. Science 290: 1717–1721, 2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Korfei M, Ruppert C, Mahavadi P, Henneke I, Markart P, Koch M, Lang G, Fink L, Bohle RM, Seeger W, Weaver TE, Guenther A. Epithelial endoplasmic reticulum stress and apoptosis in sporadic idiopathic pulmonary fibrosis. Am J Respir Crit Care Med 178: 838–846, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Kotorashvili A, Russo SJ, Mulugeta S, Guttentag S, Beers MF. Anterograde transport of surfactant protein C proprotein to distal processing compartments requires PPDY-mediated association with Nedd4 ubiquitin ligases. J Biol Chem 284: 16667–16678, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Kraft C, Peter M, Hofmann K. Selective autophagy: ubiquitin-mediated recognition and beyond. Nat Cell Biol 12: 836–841, 2010. [DOI] [PubMed] [Google Scholar]
  • 86.Kropski JA, Lawson WE, Blackwell TS. Right place, right time: the evolving role of herpesvirus infection as a “second hit” in idiopathic pulmonary fibrosis. Am J Physiol Lung Cell Mol Physiol 302: L441–L444, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Kropski JA, Pritchett JM, Zoz DF, Crossno PF, Markin C, Garnett ET, Degryse AL, Mitchell DB, Polosukhin VV, Rickman OB, Choi L, Cheng DS, McConaha ME, Jones BR, Gleaves LA, McMahon FB, Worrell JA, Solus JF, Ware LB, Lee JW, Massion PP, Zaynagetdinov R, White ES, Kurtis JD, Johnson JE, Groshong SD, Lancaster LH, Young LR, Steele MP, Phillips JA, Cogan JD, Loyd JE, Lawson WE, Blackwell TS. Extensive phenotyping of individuals at risk for familial interstitial pneumonia reveals clues to the pathogenesis of interstitial lung disease. Am J Respir Crit Care Med 191: 417–426, 2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Kurland G, Deterding RR, Hagood JS, Young LR, Brody AS, Castile RG, Dell S, Fan LL, Hamvas A, Hilman BC, Langston C, Nogee LM, Redding GJ. An official American Thoracic Society Clinical Practice Guideline: classification, evaluation, and management of childhood interstitial lung disease in infancy. Am J Respir Crit Care Med 188: 376–394, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Lamb CA, Yoshimori T, Tooze SA. The autophagosome: origins unknown, biogenesis complex. Nat Rev Mol Cell Biol 14: 759–774, 2013. [DOI] [PubMed] [Google Scholar]
  • 90.Lawson WE, Cheng DS, Degryse AL, Tanjore H, Polosukhin VV, Xu XCC, Newcomb DC, Jones BR, Roldan J, Lane KB, Morrisey EE, Beers MF, Yull FE, Blackwell TS. Endoplasmic reticulum stress enhances fibrotic remodeling in the lungs. Proc Natl Acad Sci USA 108: 10562–10567, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Lawson WE, Crossno PF, Polosukhin VV, Roldan J, Cheng DS, Lane KB, Blackwell TR, Xu C, Markin C, Ware LB, Miller GG, Loyd JE, Blackwell TS. Endoplasmic reticulum stress in alveolar epithelial cells is prominent in IPF: association with altered surfactant protein processing and herpesvirus infection. Am J Physiol Lung Cell Mol Physiol 294: L1119–L1126, 2008. [DOI] [PubMed] [Google Scholar]
  • 92.Lawson WE, Grant SW, Ambrosini V, Womble KE, Dawson EP, Lane KB, Markin C, Renzoni E, Lympany P, Thomas AQ, Roldan J, Scott TA, Blackwell TS, Phillips JA 3rd, Loyd JE, du Bois RM. Genetic mutations in surfactant protein C are a rare cause of sporadic cases of IPF. Thorax 59: 977–80, 2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Lawson WE, Polosukhin VV, Stathopoulos GT, Zoia O, Han W, Lane KB, Li B, Donnelly EF, Holburn GE, Lewis KG, Collins RD, Hull WM, Glasser SW, Whitsett JA, Blackwell TS, Lawson WE, Polosukhin VV, Stathopoulos GT, Zoia O, Han W, Lane KB, Li B, Donnelly EF, Holburn GE, Lewis KG, Collins RD, Hull WM, Glasser SW, Whitsett JA, Blackwell TS. Increased and prolonged pulmonary fibrosis in surfactant protein C-deficient mice following intratracheal bleomycin. Am J Pathol 167: 1267–1277, 2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Levine B, Kroemer G, Levine B, Kroemer G. Autophagy in the pathogenesis of disease. Cell 132: 27–42, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Low P. The role of ubiquitin-proteasome system in ageing. Gen Comp Endocrinol 172: 39–43, 2011. [DOI] [PubMed] [Google Scholar]
  • 96.Madeo F, Zimmermann A, Maiuri MC, Kroemer G. Essential role for autophagy in life span extension. J Clin Invest 125: 85–93, 2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Maguire JA, Mulugeta S, Beers MF. Endoplasmic reticulum stress induced by surfactant protein C BRICHOS mutants promotes proinflammatory signaling by epithelial cells. Am J Respir Cell Mol Biol 44: 404–414, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Maguire JA, Mulugeta S, Beers MF. Multiple ways to die: delineation of the unfolded protein response and apoptosis induced by surfactant protein C BRICHOS mutants. Int J Biochem Cell Biol 44: 101–112, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Maitra M, Wang YY, Gerard RD, Mendelson CR, Garcia CK. Surfactant protein A2 mutations associated with pulmonary fibrosis lead to protein instability and endoplasmic reticulum stress. J Biol Chem 285: 22103–22113, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Maitra M, Cano CA, Garcia CK. Mutant surfactant A2 proteins associated with familial pulmonary fibrosis and lung cancer induce TGF-β1 secretion. Proc Natl Acad Sci USA 109: 21064–21069, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 101.Matalon S, O'Brodovich H. Sodium channels in alveolar epithelial cells: molecular characterization, biophysical properties, and physiological significance. Annu Rev Physiol 61: 627–661, 1999. [DOI] [PubMed] [Google Scholar]
  • 102.Matalon S, Shrestha K, Kirk M, Waldheuser S, McDonald B, Smith K, Gao Z, Belaaouaj A, Crouch EC. Modification of surfactant protein D by reactive oxygen-nitrogen intermediates is accompanied by loss of aggregating activity, in vitro and in vivo. FASEB J 23: 1415–1430, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Matsumura Y, Ban N, Ueda K, Inagaki N. Characterization and classification of ATP-binding cassette transporter ABCA3 mutants in fatal surfactant deficiency. J Biol Chem 281: 34503–34514, 2006. [DOI] [PubMed] [Google Scholar]
  • 104.Matsumura Y, Ban N, Inagaki N. Aberrant catalytic cycle and impaired lipid transport into intracellular vesicles in ABCA3 mutants associated with nonfatal pediatric interstitial lung disease. Am J Physiol Lung Cell Mol Physiol 295: L698–L707, 2008. [DOI] [PubMed] [Google Scholar]
  • 105.Matus S, Nassif M, Glimcher LH, Hetz C. XBP-1 deficiency in the nervous system reveals a homeostatic switch to activate autophagy. Autophagy 5: 1226–1228, 2009. [DOI] [PubMed] [Google Scholar]
  • 106.Meiners S, Keller IE, Semren N, Caniard A. Regulation of the proteasome: evaluating the lung proteasome as a new therapeutic target. Antioxid Redox Signal 21: 2364–2382, 2014. [DOI] [PubMed] [Google Scholar]
  • 107.Meusser B, Hirsch C, Jarosch E, Sommer T. ERAD: the long road to destruction. Nat Cell Biol 7: 766–772, 2005. [DOI] [PubMed] [Google Scholar]
  • 108.Morimoto RI. Proteotoxic stress and inducible chaperone networks in neurodegenerative disease and aging. Genes Dev 22: 1427–1438, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Mulugeta S, Beers MF. Processing of surfactant protein C requires a type II transmembrane topology directed by juxtamembrane positively charged residues. J Biol Chem 278: 47979–47986, 2003. [DOI] [PubMed] [Google Scholar]
  • 110.Mulugeta S, Gray JM, Notarfrancesco KL, Gonzales LW, Koval M, Feinstein SI, Ballard PL, Fisher AB, Shuman H. Identification of LBM180, a lamellar body limiting membrane protein of alveolar type II cells, as the ABC transporter protein ABCA3. J Biol Chem 277: 22147–22155, 2002. [DOI] [PubMed] [Google Scholar]
  • 111.Mulugeta S, Maguire JA, Newitt JL, Russo SJ, Kotorashvili A, Beers MF. Misfolded BRICHOS SP-C mutant proteins induce apoptosis via caspase-4- and cytochrome c-related mechanisms. Am J Physiol Lung Cell Mol Physiol 293: L720–L729, 2007. [DOI] [PubMed] [Google Scholar]
  • 112.Mulugeta S, Nguyen V, Russo SJ, Muniswamy M, Beers MF. A surfactant protein C precursor protein BRICHOS domain mutation causes endoplasmic reticulum stress, proteasome dysfunction, and caspase 3 activation. Am J Respir Cell Mol Biol 32: 521–530, 2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Nakahira K, Choi AMK. Autophagy: a potential therapeutic target in lung diseases. Am J Physiol Lung Cell Mol Physiol 305: L93–L107, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Nakatani Y, Nakamura N, Sano J, Inayama Y, Kawano N, Yamanaka S, Miyagi Y, Nagashima Y, Ohbayashi C, Mizushima M, Manabe T, Kuroda M, Yokoi T, Matsubara O. Interstitial pneumonia in Hermansky-Pudlak syndrome: significance of florid foamy swelling/degeneration (giant lamellar body degeneration) of type-2 pneumocytes. Virchows Arch 437: 304–313, 2000. [DOI] [PubMed] [Google Scholar]
  • 115.Nishitoh H, Matsuzawa A, Tobiume K, Saegusa K, Takeda K, Inoue K, Hori S, Kakizuka A, Ichijo H. ASK1 is essential for endoplasmic reticulum stress-induced neuronal cell death triggered by expanded polyglutamine repeats. Genes Dev 16: 1345–1355, 2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Nogee LM, Dunbar AE III, Wert S, Askin F, Hamvas A, Whitsett JA. Mutations in the surfactant protein C gene associated with interstitial lung disease. Chest 121: 20S–21S, 2002. [DOI] [PubMed] [Google Scholar]
  • 117.Nogee LM, Dunbar AE, Wert SE, Askin F, Hamvas A, Whitsett JA. A mutation in the surfactant protein C gene associated with familial interstitial lung disease. N Engl J Med 344: 573–579, 2001. [DOI] [PubMed] [Google Scholar]
  • 118.Nureki SI, Mulugeta S, Tomer Y, Gonzales LW, Russo SJ, Hawkins A, Beers MF. Expression of the mutant surfactant ProteinC I73T allele in vivo results in abnormal ProSP-C trafficking and processing in association with diffuse parenchymal lung disease in mice (Abstract). Am J Respir Crit Care Med 191: A5414, 2015. [Google Scholar]
  • 119.Olzmann JA, Kopito RR, Christianson JC. The mammalian endoplasmic reticulum-associated degradation system. Cold Spring Harb Perspect Biol 5: a013185, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Ono S, Tanaka T, Ishida M, Kinoshita A, Fukuoka J, Takaki M, Sakamoto N, Ishimatsu Y, Kohno S, Hayashi T, Senba M, Yasunami M, Kubo Y, Yoshida LM, Kubo H, Ariyoshi K, Yoshiura K, Morimoto K. Surfactant protein C G100S mutation causes familial pulmonary fibrosis in Japanese kindred. Eur Respir J 38: 861–869, 2011. [DOI] [PubMed] [Google Scholar]
  • 121.Oosterlaken-Dijksterhuis MA, van Eijk M, van Buel BLM, van Golde LMG, Haagsman HP. Surfactant protein composition of lamellar bodies isolated from rat lung. Biochem J 274: 115–119, 1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Osanai K, Mason RJ, Voelker DR. Trafficking of newly synthesized surfactant protein A in isolated rat alveolar type II cells. Am J Respir Cell Mol Biol 19: 929–935, 1998. [DOI] [PubMed] [Google Scholar]
  • 123.Oyadomari S, Araki E, Mori M. Endoplasmic reticulum stress-mediated apoptosis in pancreatic beta-cells. Apoptosis 7: 335–345, 2002. [DOI] [PubMed] [Google Scholar]
  • 124.Pandey UB, Nie Z, Batlevi Y, McCray BA, Ritson GP, Nedelsky NB, Schwartz SL, DiProspero NA, Knight MA, Schuldiner O, Padmanabhan R, Hild M, Berry DL, Garza D, Hubbert CC, Yao TP, Baehrecke EH, Taylor JP. HDAC6 rescues neurodegeneration and provides an essential link between autophagy and the UPS. Nature 447: 859–863, 2007. [DOI] [PubMed] [Google Scholar]
  • 125.Pankiv S, Clausen TH, Lamark T, Brech A, Bruun JA, Outzen H, Overvatn A, Bjorkoy G, Johansen T. p62/SQSTM1 binds directly to Atg8/LC3 to facilitate degradation of ubiquitinated protein aggregates by autophagy. J Biol Chem 282: 24131–24145, 2007. [DOI] [PubMed] [Google Scholar]
  • 126.Patel AS, Lin L, Geyer A, Haspel JA, An CH, Cao J, Rosas IO, Morse D. Autophagy in idiopathic pulmonary fibrosis. PLoS One 7: e41394, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Peljto AL, Zhang YZ, Fingerlin TE, Ma SF, Garcia JGN, Richards TJ, Silveira LJ, Lindell KO, Steele MP, Loyd JE, Gibson KF, Seibold MA, Brown KK, Talbert JL, Markin C, Kossen K, Seiwert SD, Murphy E, Noth I, Schwarz MI, Kaminski N, Schwartz DA. Association between the MUC5B promoter polymorphism and survival in patients with idiopathic pulmonary fibrosis. JAMA 309: 2232–2239, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Percopo S, Cameron HS, Nogee LM, Pettinato G, Montella S, Santamaria F. Variable phenotype associated with SP-C gene mutations: fatal case with the I73T mutation. Eur Respir J 24: 1072–1073, 2004. [DOI] [PubMed] [Google Scholar]
  • 129.Perez-Gil J. Structure of pulmonary surfactant membranes and films: the role of proteins and lipid-protein interactions. Biochim Biophys Acta 1778: 1676–1695, 2008. [DOI] [PubMed] [Google Scholar]
  • 130.Pickrell AM, Youle RJ. The roles of PINK1, parkin, and mitochondrial fidelity in Parkinson's disease. Neuron 85: 257–273, 2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Possmayer F. A proposed nomenclature for pulmonary surfactant-associated proteins. Am Rev Respir Dis 138: 990–998, 1988. [DOI] [PubMed] [Google Scholar]
  • 132.Raghu G, Brown KK, Bradford WZ, Starko K, Noble PW, Schwartz DA, King TE, Adlakha A, Tarczynski S, Ainslie G, Kalam R, Bai T, Truchan H, Baughman R, Wingst D, Bhorade S, Norwick L, Brown KK, Kervitsky D, Calhoun W, DiNella L, Chan C, Jamieson L, Chan K, Turpen T, Chapman J, Slattery S, Chen L, Turner J, Clark M, Sanders R, Crain M, Pate D, Davis G, Lynn M, Dhar A, Hrytsytk M, Drent M, Horr ET, du Bois R, Goh N, Egan J, Anthony N, Enelow R, Haram T, Ettinger N, Merli S, Frost A, Holy R, Glassberg M, Brown A, Golden J, DesMarais M, Haim YD, Rzeszutko B, Hassoun P, Finlay G, Lawler L, Hollingsworth H, Goncalves P, Jackson R, Meadows T, Kallay M, Celebie A, King-Biggs M, Beyle A, Lancaster L, Sanderson R, Lasky J, Ditta S, Lorch D, Schoh T, Lynch J, Alousha A, Mageto Y, Meachom M, Martinez F, Dahlgren D, Mckee C, Hamilton T, Meyer K, Wilson S, Millar A, Hann A, Mohsenifar Z, Balfe D, Morell F, Reyes L, Nathan S, Theodoslod S, Noble PW, Shrauger K, Noth I, Strek M, Au J, Olivier K, Wells R, Padilla M, Behnegar A, Polito A, Bloom C, Raghu G, Snydsman A, Rosen G, Jacobs S, Ryu JH, Carlson K, Sahn SA, Oser R, Schaumberg T, Chesnutt M, Leonard S, Schluger N, Jellen P, Steele M, Willis C, Strieter R, Christedoulou V, Valentine V, Lanning T, Wencel M, Sturgeon D, Xaubert A, Zisman D, Douglas L, Lynch DA, Godwin JD, Webb WR, Fleming T, Reis A, Riley D. A placebo-controlled trial of interferon gamma-1b in patients with idiopathic pulmonary fibrosis. N Engl J Med 350: 125–133, 2004. [DOI] [PubMed] [Google Scholar]
  • 133.Raghu G, Depaso WJ, Cain K, Hammar SP, Wetzel CE, Dreis DF, Hutchinson J, Pardee NE, Winterbauer RH. Azathioprine combined with prednisone in the treatment of idiopathic pulmonary fibrosis: a prospective double-blind, randomized, placebo-controlled clinical trial. Am Rev Respir Dis 144: 291–296, 1991. [DOI] [PubMed] [Google Scholar]
  • 134.Rooney SA. Regulation of surfactant secretion. In: Lung Surfactant: Cellular and Molecular Mechanisms, edited by Rooney SA. Austin, TX: Landes, 1998, p. 139–163. [Google Scholar]
  • 135.Ross CA, Poirier MA. Protein aggregation and neurodegenerative disease. Nat Med 10 Suppl: S10–S17, 2004. [DOI] [PubMed] [Google Scholar]
  • 136.Rubinsztein D, Mariño G, Kroemer G. Autophagy and aging. Cell 146: 682–695, 2011. [DOI] [PubMed] [Google Scholar]
  • 137.Sanchez I, Mahlke C, Yuan JY. Pivotal role of oligomerization in expanded polyglutamine neurodegenerative disorders. Nature 421: 373–379, 2003. [DOI] [PubMed] [Google Scholar]
  • 138.Sanchez I, Xu CJ, Juo P, Kakizaka A, Blenis J, Yuan JY. Caspase-8 is required for cell death induced by expanded polyglutamine repeats. Neuron 22: 623–633, 1999. [DOI] [PubMed] [Google Scholar]
  • 139.Sanchez-Pulido L, Devos D, Valencia A. BRICHOS: a conserved domain in proteins associated with dementia, respiratory distress and cancer. Trends Biochem Sci 27: 329–332, 2002. [DOI] [PubMed] [Google Scholar]
  • 140.Sarkar S, Carroll B, Buganim Y, Maetzel D, Ng AHM, Cassady JP, Cohen MA, Chakraborty S, Wang HY, Spooner E, Ploegh H, Gsponer J, Korolchuk VI, Jaenisch R. Impaired autophagy in the lipid-storage disorder Niemann-Pick type C1 disease. Cell Rep 5: 1302–1315, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Schwartz AL, Ciechanover A. Targeting proteins for destruction by the ubiquitin system: implications for human pathobiology. Annu Rev Pharmacol Toxicol 49: 73–96, 2009. [DOI] [PubMed] [Google Scholar]
  • 142.Schwartz DA, Van Fossen DS, Davis CS, Helmers RA, Dayton CS, Burmeister LF, Hunninghake GW. Determinants of progression in idiopathic pulmonary fibrosis. Am J Respir Crit Care Med 149: 444–449, 1994. [DOI] [PubMed] [Google Scholar]
  • 143.Seibold MA, Wise AL, Speer MC, Steele MP, Brown KK, Loyd JE, Fingerlin TE, Zhang WM, Gudmundsson G, Groshong SD, Evans CM, Garantziotis S, Adler KB, Dickey BF, du Bois RM, Yang IV, Herron A, Kervitsky D, Talbert JL, Markin C, Park J, Crews AL, Slifer SH, Auerbach S, Roy MG, Lin J, Hennessy CE, Schwarz MI, Schwartz DA. A common MUC5B promoter polymorphism and pulmonary fibrosis. N Engl J Med 364: 1503–1512, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Selman M, King TE, Pardo A. Idiopathic pulmonary fibrosis: prevailing and evolving hypotheses about its pathogenesis and implications for therapy. Ann Intern Med 134: 136–151, 2001. [DOI] [PubMed] [Google Scholar]
  • 145.Selman M, Pardo A. Revealing the pathogenic and aging-related mechanisms of the enigmatic idiopathic pulmonary fibrosis. An integral model. Am J Respir Crit Care Med 189: 1161–1172, 2014. [DOI] [PubMed] [Google Scholar]
  • 146.Settembre C, Fraldi A, Jahreiss L, Spampanato C, Venturi C, Medina D, de Pablo R, Tacchetti C, Rubinsztein DC, Ballabio A. A block of autophagy in lysosomal storage disorders. Hum Mol Genet 17: 119–129, 2008. [DOI] [PubMed] [Google Scholar]
  • 147.Shulenin S, Nogee LM, Annilo T, Wert SE, Whitsett JA, Dean M. ABCA3 gene mutations in newborns with fatal surfactant deficiency. N Engl J Med 350: 1296–1303, 2004. [DOI] [PubMed] [Google Scholar]
  • 148.Somaschini M, Nogee LM, Sassi I, Danhaive O, Presi S, Boldrini R, Montrasio C, Ferrari M, Wert SE, Carrera P. Unexplained neonatal respiratory distress due to congenital surfactant deficiency. J Pediatr 150: 649–653, 2007. [DOI] [PubMed] [Google Scholar]
  • 149.Somborac-Bačura A, van der Toorn M, Franciosi L, Slebos DJ, Žanić-Grubišić T, Bischoff R, van Oosterhout AJM. Cigarette smoke induces endoplasmic reticulum stress response and proteasomal dysfunction in human alveolar epithelial cells. Exp Physiol 98: 316–325, 2013. [DOI] [PubMed] [Google Scholar]
  • 150.Stanley SE, Chen JJL, Podlevsky JD, Alder JK, Hanse NN, Mathias RA, Qi XD, Rafaels NM, Wise RA, Silverman EK, Barnes KC, Armanios M. Telomerase mutations in smokers with severe emphysema. J Clin Invest 125: 563–570, 2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Steele MP, Speer MC, Loyd JE, Brown KK, Herron A, Slifer SH, Burch LH, Wahidi MM, Phillips JA III, Sporn TA, McAdams HP, Schwarz MI, Schwartz DA. Clinical and pathologic features of familial interstitial pneumonia. Am J Respir Crit Care Med 172: 1146–1152, 2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Stevens PA, Pettenazzo A, Brasch F, Mulugeta S, Baritussio A, Ochs M, Morrison L, Russo SJ, Beers MF. Nonspecific interstitial pneumonia, alveolar proteinosis, and abnormal proprotein trafficking resulting from a spontaneous mutation in the surfactant protein C gene. Pediatr Res 57: 89–98, 2005. [DOI] [PubMed] [Google Scholar]
  • 153.Stewart GA, Ridsdale R, Martin EP, Na CL, Xu Y, Mandapaka K, Weaver TE. 4-Phenylbutyric acid treatment rescues trafficking and processing of a mutant surfactant protein-C. Am J Respir Cell Mol Biol 47: 324–331, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Szyperski T, Vandenbussche G, Curstedt T, Ruysschaert JM, Wuthrich K, Johansson J. Pulmonary surfactant-associated polypeptide C in a mixed organic solvent transforms from a monomeric alpha-helical state into insoluble beta-sheet aggregates. Protein Sci 7: 2533–2540, 1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Tang YW, Johnson JE, Browning PJ, Cruz-Gervis RA, Davis A, Graham BS, Brigham KL, Oates JA Jr, Loyd JE, Stecenko AA. Herpesvirus DNA is consistently detected in lungs of patients with idiopathic pulmonary fibrosis. J Clin Microbiol 41: 2633–2640, 2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Tanjore H, Blackwell TS, Lawson WE. Emerging evidence for endoplasmic reticulum stress in the pathogenesis of idiopathic pulmonary fibrosis. Am J Physiol Lung Cell Mol Physiol 302: L721–L729, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Tanjore H, Cheng DS, Degryse AL, Zoz DF, Abdolrasulnia R, Lawson WE, Blackwell TS. Alveolar epithelial cells undergo epithelial-to-mesenchymal transition in response to endoplasmic reticulum stress. J Biol Chem 286: 30972–30980, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Thomas AQ, Lane K, Phillips J III, Prince M, Markin C, Speer M, Schwartz DA, Gaddipati R, Marney A, Johnson J, Roberts R, Haines J, Stahlman M, Loyd JE. Heterozygosity for a surfactant protein C gene mutation associated with usual interstitial pneumonitis and cellular nonspecific interstitial pneumonitis in one kindred. Am J Respir Crit Care Med 165: 1322–1328, 2002. [DOI] [PubMed] [Google Scholar]
  • 160.Thurm T, Kaltenborn E, Kern S, Griese M, Zarbock R. SFTPC mutations cause SP-C degradation and aggregate formation without increasing ER stress. Eur J Clin Invest 43: 791–800, 2013. [DOI] [PubMed] [Google Scholar]
  • 161.Torres-Gonzalez E, Bueno M, Tanaka A, Krug LT, Cheng DS, Polosukhin VV, Sorescu D, Lawson WE, Blackwell TS, Rojas M, Mora AL. Role of endoplasmic reticulum stress in age-related susceptibility to lung fibrosis. Am J Respir Cell Mol Biol 46: 748–756, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Tredano M, Griese M, Brasch F, Schumacher S, de Blic J, Marque S, Houdayer C, Elion J, Couderc R, Bahuau M. Mutation of SFTPC in infantile pulmonary alveolar proteinosis with or without fibrosing lung disease. Am J Med Genet A 126A: 18–26, 2004. [DOI] [PubMed] [Google Scholar]
  • 163.Uhal BD, Nguyen H, Dang M, Gopallawa I, Jiang J, Dang V, Ono S, Morimoto K. Abrogation of ER stress-induced apoptosis of alveolar epithelial cells by angiotensin 1–7. Am J Physiol Lung Cell Mol Physiol 305: L33–L41, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Uhal BD. The role of apoptosis in pulmonary fibrosis. Eur Respir Rev 17: 138–144, 2008. [Google Scholar]
  • 165.Uhal BD, Nguyen H. The Witschi hypothesis revisited after 35 years: genetic proof from SP-C BRICHOS domain mutations. Am J Physiol Lung Cell Mol Physiol 305: L906–L911, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Urano F, Wang X, Bertolotti A, Zhang Y, Chung P, Harding HP, Ron D. Coupling of stress in the ER to activation of JNK protein kinases by transmembrane protein kinase IRE1. Science 287: 664–666, 2000. [DOI] [PubMed] [Google Scholar]
  • 167.van Moorsel CHM, van Oosterhout MFM, Barlo NP, de Jong PA, van der Vis JJ, Ruven HJT, van Es HW, van den Bosch JMM, Grutters JC. Surfactant protein C mutations are the basis of a significant portion of adult familial pulmonary fibrosis in a dutch cohort. Am J Respir Crit Care Med 182: 1419–1425, 2010. [DOI] [PubMed] [Google Scholar]
  • 168.van Rijt SH, Keller IE, John G, Kohse K, Yildirim AO, Eickelberg O, Meiners S. Acute cigarette smoke exposure impairs proteasome function in the lung. Am J Physiol Lung Cell Mol Physiol 303: L814–L823, 2012. [DOI] [PubMed] [Google Scholar]
  • 169.Vannuvel K, Renard P, Raes M, Arnould T. Functional and morphological impact of ER stress on mitochondria. J Cell Physiol 228: 1802–1818, 2013. [DOI] [PubMed] [Google Scholar]
  • 170.Vembar SS, Brodsky JL. One step at a time: endoplasmic reticulum-associated degradation. Nat Rev Mol Cell Biol 9: 944–957, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Vidal RL, Figueroa A, Court F, Thielen P, Molina C, Wirth C, Caballero B, Kiffin R, Segura-Aguilar J, Cuervo AM, Glimcher LH, Hetz C. Targeting the UPR transcription factor XBP1 protects against Huntington's disease through the regulation of FoxO1 and autophagy. Hum Mol Genet 21: 2245–2262, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Wambach JA, Casey AM, Fishman MP, Wegner DJ, Wert SE, Cole FS, Hamvas A, Nogee LM. Genotype-phenotype correlations for infants and children with ABCA3 deficiency. Am J Respir Crit Care Med 189: 1538–1543, 2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Wang WJ, Mulugeta S, Russo SJ, Beers MF. Deletion of exon 4 from human surfactant protein C results in aggresome formation and generation of a dominant negative. J Cell Sci 116: 683–692, 2003. [DOI] [PubMed] [Google Scholar]
  • 174.Wang WJ, Russo SJ, Mulugeta S, Beers MF. Biosynthesis of surfactant protein C (SP-C). Sorting of SP-C proprotein involves homomeric association via a signal anchor domain. J Biol Chem 277: 19929–19937, 2002. [DOI] [PubMed] [Google Scholar]
  • 175.Wang Y, Kuan PJ, Xing C, Cronkhite JT, Torres F, Rosenblatt RL, DiMaio JM, Kinch LN, Grishin NV, Garcia CK. Genetic defects in surfactant protein A2 are associated with pulmonary fibrosis and lung cancer. Am J Hum Genet 84: 52–59, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Weaver TE, Ridsdale R, Stewart GA, Martin E, Na CL, Basha S, Wert SE. Mice expressing a disease-associated sftpc allele are predisposed to alveolar injury (Abstract). Am J Respir Crit Care Med 187: A5156, 2013. [Google Scholar]
  • 177.Weichert N, Kaltenborn E, Hector A, Woischnik M, Schams A, Holzinger A, Kern S, Griese M. Some ABCA3 mutations elevate ER stress and initiate apoptosis of lung epithelial cells. Respir Res 12: 4, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Wert SE, Whitsett JA, Nogee LM. Genetic disorders of surfactant dysfunction. Pediatr Dev Pathol 12: 253–274, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Wey S, Luo BQ, Lee AS. Acute inducible ablation of GRP78 reveals its role in hematopoietic stem cell survival, lymphogenesis and regulation of stress signaling. PLoS One 7: e39047, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Whitsett JA, Weaver TE. Hydrophobic surfactant proteins in lung function and disease. N Engl J Med 347: 2141–2148, 2002. [DOI] [PubMed] [Google Scholar]
  • 181.Willander H, Askarieh G, Landreh M, Westermark P, Nordling K, Keranen H, Hermansson E, Hamvas A, Nogee LM, Bergman T, Saenz A, Casals C, Aqvist J, Jornvall H, Berglund H, Presto J, Knight SD, Johansson J. High-resolution structure of a BRICHOS domain and its implications for anti-amyloid chaperone activity on lung surfactant protein C. Proc Natl Acad Sci USA 109: 2325–2329, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Willander H, Hermansson E, Johansson J, Presto J. BRICHOS domain associated with lung fibrosis, dementia and cancer—a chaperone that prevents amyloid fibril formation? FEBS J 278: 3893–3904, 2011. [DOI] [PubMed] [Google Scholar]
  • 183.Wright JR. Immunoregulatory functions of surfactant proteins. Nat Rev Immunol 5: 58–68, 2005. [DOI] [PubMed] [Google Scholar]
  • 184.Wynn TA. Integrating mechanisms of pulmonary fibrosis. J Exp Med 208: 1339–1350, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Xu C, Bailly-Maitre B, Reed JC. Endoplasmic reticulum stress: cell life and death decisions. J Clin Invest 115: 2656–2664, 2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Yamano G, Funahashi H, Kawanami O, Zhao LX, Ban N, Uchida Y, Morohoshi T, Ogawa J, Shioda S, Inagaki N. ABCA3 is a lamellar body membrane protein in human lung alveolar type II cells. FEBS Lett 508: 221–225, 2001. [DOI] [PubMed] [Google Scholar]
  • 187.Yang JB, Velikoff M, Canalis E, Horowitz JC, Kim KK. Activated alveolar epithelial cells initiate fibrosis through autocrine and paracrine secretion of connective tissue growth factor. Am J Physiol Lung Cell Mol Physiol 306: L786–L796, 2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Yonemaru M, Kasuga I, Kusumoto H, Kunisawa A, Kiyokawa H, Kuwabara S, Ichinose Y, Toyama K. Elevation of antibodies to cytomegalovirus and other herpes viruses in pulmonary fibrosis. Eur Respir J 10: 2040–2045, 1997. [DOI] [PubMed] [Google Scholar]
  • 189.Youle RJ, Narendra DP. Mechanisms of mitophagy. Nat Rev Mol Cell Biol 12: 9–14, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Youle RJ, van der Bliek AM. Mitochondrial fission, fusion, and stress. Science 337: 1062–1065, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Young LR, Gulleman PM, Bridges JP, Weaver TE, Deutsch GH, Blackwell TS, McCormack FX. The alveolar epithelium determines susceptibility to lung fibrosis in Hermansky-Pudlak syndrome. Am J Respir Crit Care Med 186: 1014–1024, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Young LR, Pasula R, Gulleman PM, Deutsch GH, McCormack FX. Susceptibility of Hermansky-Pudlak mice to bleomycin-induced type II cell apoptosis and fibrosis. Am J Respir Cell Mol Biol 37: 67–74, 2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Zarbock R, Woischnik M, Sparr C, Thurm T, Kern S, Kaltenborn E, Hector A, Hartl D, Liebisch G, Schmitz G, Griese M. The surfactant protein C mutation A116D alters cellular processing, stress tolerance, surfactant lipid composition, and immune cell activation. BMC Pulm Med 12: 15, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Zhang K, Kaufman RJ. From endoplasmic-reticulum stress to the inflammatory response. Nature 454: 455–462, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Zhong Q, Zhou B, Ann DK, Minoo P, Liu Y, Banfalvi A, Krishnaveni MS, Dubourd M, Demaio L, Willis BC, Kim K-J, duBois RM, Crandall ED, Beers MF, Borok Z. Role of endoplasmic reticulum (ER) stress in EMT of alveolar epithelial cells (AEC): effects of misfolded surfactant protein C. Am J Respir Cell Mol Biol 45: 498–509, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Zhu S, Kachel DL, Martin WJ, Matalon S. Nitrated SP-A does not enhance adherence of Pneumocystis carinii to alveolar macrophages. Am J Physiol Lung Cell Mol Physiol 275: L1031–L1039, 1998. [DOI] [PubMed] [Google Scholar]

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

RESOURCES