Patient-specific iPSCs carrying an SFTPC mutation reveal the intrinsic alveolar epithelial dysfunction at the inception of interstitial lung disease

Konstantinos-Dionysios Alysandratos, Scott J Russo, Anton Petcherski, Evan P Taddeo, Rebeca Acín-Pérez, Carlos Villacorta-Martin, J C Jean, Surafel Mulugeta, Luis R Rodriguez, Benjamin C Blum, Ryan M Hekman, Olivia T Hix, Kasey Minakin, Marall Vedaie, Seunghyi Kook, Andrew M Tilston-Lunel, Xaralabos Varelas, Jennifer A Wambach, F Sessions Cole, Aaron Hamvas, Lisa R Young, Marc Liesa, Andrew Emili, Susan H Guttentag, Orian S Shirihai, Michael F Beers, Darrell N Kotton, Konstantinos-Dionysios Alysandratos, Scott J Russo, Anton Petcherski, Evan P Taddeo, Rebeca Acín-Pérez, Carlos Villacorta-Martin, J C Jean, Surafel Mulugeta, Luis R Rodriguez, Benjamin C Blum, Ryan M Hekman, Olivia T Hix, Kasey Minakin, Marall Vedaie, Seunghyi Kook, Andrew M Tilston-Lunel, Xaralabos Varelas, Jennifer A Wambach, F Sessions Cole, Aaron Hamvas, Lisa R Young, Marc Liesa, Andrew Emili, Susan H Guttentag, Orian S Shirihai, Michael F Beers, Darrell N Kotton

Abstract

Alveolar epithelial type 2 cell (AEC2) dysfunction is implicated in the pathogenesis of adult and pediatric interstitial lung disease (ILD), including idiopathic pulmonary fibrosis (IPF); however, identification of disease-initiating mechanisms has been impeded by inability to access primary AEC2s early on. Here, we present a human in vitro model permitting investigation of epithelial-intrinsic events culminating in AEC2 dysfunction, using patient-specific induced pluripotent stem cells (iPSCs) carrying an AEC2-exclusive disease-associated variant (SFTPCI73T). Comparing syngeneic mutant versus gene-corrected iPSCs after differentiation into AEC2s (iAEC2s), we find that mutant iAEC2s accumulate large amounts of misprocessed and mistrafficked pro-SFTPC protein, similar to in vivo changes, resulting in diminished AEC2 progenitor capacity, perturbed proteostasis, altered bioenergetic programs, time-dependent metabolic reprogramming, and nuclear factor κB (NF-κB) pathway activation. Treatment of SFTPCI73T-expressing iAEC2s with hydroxychloroquine, a medication used in pediatric ILD, aggravates the observed perturbations. Thus, iAEC2s provide a patient-specific preclinical platform for modeling the epithelial-intrinsic dysfunction at ILD inception.

Keywords: NF-κB; autophagy; bioenergetics; iPSC-derived alveolar epithelial type 2 cells; idiopathic pulmonary fibrosis; induced pluripotent stem cells; interstitial lung disease; metabolic reprogramming; proteostasis; surfactant protein C.

Conflict of interest statement

Declaration of interests The authors declare no competing interests.

Copyright © 2021 The Author(s). Published by Elsevier Inc. All rights reserved.

Figures

Figure 1.. Generation of patient-specific iPSC lines…
Figure 1.. Generation of patient-specific iPSC lines and their differentiation to alveolar epithelium
(A) Schematic showing the generation of patient-specific iPSCs from dermal fibroblasts or peripheral blood mononuclear cells (PBMCs) from three individuals carrying the most frequent SFTPC pathogenic variant (SFTPCI73T/WT). (B) Chest CT of the SPC2 donor reveals diffuse ground glass opacities and traction bronchiectasis. (C) H&E staining of the SPC2 donor lung explant shows end-stage lung disease with interstitial fibrosis, chronic inflammation, and alveolar remodeling with AEC2 hyperplasia and degenerating macrophages within the residual alveoli. Scale bar, 200 μm. (D) Transcription activator-like effector nucleases (TALEN) targeting strategy and edited SFTPC loci post Cre-mediated antibiotic cassette excision. (E) Schematic of directed differentiation protocol from iPSCs to day 30+ monolayered epithelial iAEC2 spheres (alveolospheres). (F) Schematic of CHIR modulation to achieve iAEC2 maturation and representative flow cytometry dot plots (mean ± SD; n = 6 biological replicates of independent differentiations). (G) Dot plots depicting the normalized expression level of AEC2-marker genes in day 114 and day 131 SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s compared to day 0 iPSCs and week 21 human fetal distal lung (HFL) controls, by bulk RNA sequencing (boxplots represent mean ± SD; n = 3 experimental replicates of independent wells of a differentiation).
Figure 2.. Derivation of parallel self-renewing corrected…
Figure 2.. Derivation of parallel self-renewing corrected (SFTPCtdT/WT) and mutant (SFTPCI73T/tdT) iAEC2s
(A) Representative flow cytometry dot plots of day 82 (p2), day 115 (p4), day 149 (p6), day 190 (p8), day 208 (p9), day 243 (p11), and day 285 (p13) SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s. MFI, mean fluorescence intensity. (B) Representative live-cell imaging of SFTPCtdT/WT and SFTPCI73T/tdT alveolospheres (bright-field/tdTomato overlay; day 149). Scale bars, 500 μm. (C) Graph showing yield per input tdT/WT or I73T/tdT SFTPCtdTomato+ cell sorted on day 51 and passaged without further sorting. **p < 0.01, ***p < 0.001, ****p < 0.0001 by unpaired, two-tailed Student’s t test. (D) Bar graph shows retention of the AEC2 cell fate in SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s maintained in culture for 302 days, measured by flow cytometry as the frequency of cells expressing the SFTPCtdTomato reporter. (E) Histograms show higher proliferation rates in SFTPCtdT/WT compared to SFTPCI73T/tdT iAEC2s, measured by flow cytometry as the frequency of SFTPCtdTomato+ cells that incorporate EdU. (F) Representative confocal immunofluorescence microscopy of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s stained for activated caspase 3 (green) and DNA (Hoechst, blue) shows absence of significant apoptosis. Scale bars, 50 μm. (A and C–E) Mean ± SD is shown; n = 3 experimental replicates of independent wells of a differentiation.
Figure 3.. SFTPC I73T/tdT iAEC2s demonstrate distinct…
Figure 3.. SFTPCI73T/tdT iAEC2s demonstrate distinct cellular morphology and misprocess and mistraffick pro-SFTPC similarly to in vivo SFTPCI73T-expressing AEC2s
(A) Representative H&E staining of formalin fixed and paraffin embedded sections (i, scale bars, 50 μm) and toluidine blue staining of plastic sections (ii) of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s. Arrowheads indicate lamellar body-like inclusions; eclipse indicates intraluminal inclusions. (B) Representative confocal immunofluorescence microscopy of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s for EPCAM (green) and DNA (Hoechst, blue). Scale bars, 10 μm. (C) Representative staining of E18.5 CFlp-SFTPCI73T/I73T and CFlp-SFTPCWT/WT embryos for HA-tagged SFTPC. Scale bars, 70 μm. (D) Representative confocal immunofluorescence microscopy of distal sections of SPC2 donor and healthy donor lung explants stained for EPCAM (red), pro-SFTPC (green), and DNA (Hoechst, blue). Scale bars, 10 μm. (E) Representative TEM images of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s depicting lamellar bodies. Scale bars, 1 μm. (F) Schematic representing the cellular compartments in which pro-SFTPC processing into mature SFTPC occurs. ER, endoplasmic reticulum; MVB, multi-vesicular body; LB, lamellar body. (G and H) Representative western blots of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2 lysates at the indicated time points were compared to freshly isolated primary human AEC2s lysates for pro-SFTPC (NPRO-SFTPC) (G, left panel), mature SFTPC (G, right panel), and mature SFTPB (H) with β actin as a loading control. (I) Representative confocal immunofluorescence microscopy of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s for pro-SFTPC (red; with zoom), EPCAM (green; with zoom), SFTPB (red), and DNA (Hoechst, blue). Scale bars, 10 μm. (J) Representative confocal immunofluorescence microscopy of distal sections of SPC2 donor and healthy donor lung explants for pro-SFTPC (green), EPCAM (red), SFTPB (green), and DNA (Hoechst, blue). Scale bars, 10 μm.
Figure 4.. Transcriptomic and proteomic/phosphoproteomic analyses identify…
Figure 4.. Transcriptomic and proteomic/phosphoproteomic analyses identify candidate disease-associated pathways
(A) Clustering of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2 scRNA-seq transcriptomes at two time points (days (d)30 and 113 of differentiation) using Uniform Manifold Approximation Projection (UMAP). (B) Average expression levels and frequencies (purple dots) for select genes profiled by scRNA-seq in SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s. Comparison is made to a publicly available adult primary distal lung dataset (Habermann et al., 2020), and genes are selected to indicate AEC2, AEC1, airway, endothelial, epithelial, leukocyte, or proliferation programs. (C) Heatmap of top 50 upregulated and top 50 downregulated genes comparing d113 SFTPCtdT/WT versus d113 SFTPCI73T/tdT iAEC2s by scRNA-seq (ranked by average log fold-change, FDR <0.05; row-normalized expression Z scores). A subset of differentially expressed genes is highlighted with large font. (D) Violin plots showing normalized expression for indicated genes or cell cycle phase in d30 and d113 SFTPCtdT/WT versus SFTPCI73T/tdT iAEC2s by scRNA-seq. (E) Heatmap of the 50 upregulated and top 50 downregulated genes in d114 SFTPCtdT/WT versus d114 SFTPCI73T/tdT iAEC2s by bulk RNA seq (ranked by FDR, FDR <0.05; row-normalized expression Z scores). AEC2-marker genes are highlighted with large font. (F) Volcano plots indicating differential protein expression in d113 SFTPCtdT/WT versus d113 SFTPCI73T/tdT iAEC2s by mass spectrometry. (G) Representative confocal immunofluorescence microscopy of distal sections of SPC2 donor and healthy donor lung explants stained for pro-SFTPC (green) and DNA (Hoechst, blue) shows an altered cellular localization pattern and a higher pro-SFTPC fluorescence intensity in the patient’s AEC2s, quantified by relative fluorescence units (RFU). Scale bars, 10 μm (mean ± SD; n = 35 individual SFTPC-positive cells treated as experimental replicates). (H) Top 10 upregulated pathways in d113 SFTPCI73T/tdT versus d113 SFTPCtdT/WT iAEC2s based on global proteomic analysis. (I) Integrated gene set enrichment analyses based on scRNA-seq transcriptomic, proteomic, and phosphoproteomic analyses in d113 SFTPCI73T/tdT versus d113 SFTPCtdT/WT iAEC2s with an FDR <0.1 in at least 2 out of 3 datasets. NES, normalized enrichment score; gene ratio, ratio of enriched genes for a given pathway to the total number of genes in the pathway. The last column represents the number of datasets (out of 3) with an FDR <0.05. **p

Figure 5.. Autophagy perturbations in SFTPC I73T…

Figure 5.. Autophagy perturbations in SFTPC I73T iAEC2s

(A) Schematic illustrating the autophagy pathway. Key proteins…

Figure 5.. Autophagy perturbations in SFTPCI73T iAEC2s
(A) Schematic illustrating the autophagy pathway. Key proteins involved in the pathway (p62/SQSTM1 and LC3) and the mechanisms of action of bafilomycin A1 and torin are depicted. (B) Representative western blot of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2 lysates for p62 and LC3 with GAPDH and β actin as loading controls with densitometric quantification (mean ± SD; n = 11 for LC3 and n = 7 for p62 independent experiments). (C) Western blots of cell lysates for LC3 and β actin from autophagic flux studies using SFTPCtdT/WT or SFTPCI73T/tdT iAEC2s treated with either torin (5 μM) for 4, 8, or 18 h or bafilomycin A1 (BafA1) (50 or 100 nM) for 18 h or vehicle (DMSO) for 4 or 18 h (blots shown are representative of n = 2 independent experiments). (D) A repeat autophagic flux study, as in (C) but using BafA1 (50 nM) for 1, 2, 4, and 18 h or vehicle (DMSO) for 18 h (representative of n = 2 independent experiments). (E) SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s transduced with a lentiviral LC3:GFP fusion protein and exposed to BafA1 (50 nM) to quantify autophagosomes and autophagolysosomes. Scale bar, 10 μm (representative confocal fluorescence microscopy images of n = 3 experimental replicates of independent wells of a differentiation). (F) Representative TEM image of SFTPCI73T/tdT iAEC2s shows a double-membrane autophagosome (inset, arrowheads). Scale bar, 1 μm. (G) Real-time qPCR showing fold change in gene expression compared to WT in primary mouse SftpcI73T AEC2s (mean ± SEM; n = 6 for SftpcI73T and n = 5 for WT mice). (H) Representative western blots of primary mouse WT and SftpcI73T AEC2 lysates for pro-SFTPC (NPRO), LC3, and p62 with β actin as a loading control (see also Figure 3G) (n = 3 WT and SftpcI73T mice). (I) Representative live-cell confocal fluorescence microscopy of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s stained with LysoTracker green. Scale bars, 10 μm. (J) Quantification of acidic organelles expressed as percentage of covered cell area (mean ± SD; n = 11–14 independent alveolospheres treated as experimental replicates). (K) Quantification of lysosomal pH based on Lysosensor yellow/blue-dextran ratio (mean ± SD; n = 2 independent experiments with 20–30 cells analyzed per experiment; see also Figure S4). *p

Figure 6.. SFTPC I73T -expressing AEC2s demonstrate…

Figure 6.. SFTPC I73T -expressing AEC2s demonstrate metabolic reprogramming

(A) Representative live-cell confocal fluorescence microscopy of SFTPC…

Figure 6.. SFTPCI73T-expressing AEC2s demonstrate metabolic reprogramming
(A) Representative live-cell confocal fluorescence microscopy of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s stained with MitoTracker green. Scale bars, 10 μm. (B–D) Quantitative analyses of morphometric data from fluorescence images (mean ± SD; n = 24 SFTPCtdT/WT and n = 18 SFTPCI73T/tdT independent alveolospheres treated as experimental replicates). (B) Increased mitochondrial mass in SFTPCI73T/tdT iAEC2s, quantified as the percentage of cell area covered by mitochondria. (C) Increased mitochondrial size in SFTPCI73T/tdT iAEC2s. (D) SFTPCI73T/tdT iAEC2 mitochondria are more fragmented, assessed by the aspect ratio. (E) Representative western blot of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2 lysates for TOM20 with β actin as loading control with densitometric quantification (mean ± SD; n = 4 [days 128, 176, 223, 223] SFTPCtdT/WT and n = 3 [days 128, 223, 223] SFTPCI73T/tdT independent experiments). (F) Early time point OCR was measured under basal conditions followed by addition of oligomycin, FCCP, and antimycin A (Ant A), as indicated. (G) SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s demonstrate no significant differences in proton leak, maximal respiration, and spare capacity measured as percent change over basal respiration (mean ± SEM; n = 2 independent experiments). (H) Early time point ECAR under basal conditions and following addition of oligomycin. (I) Late time point OCR. (J) SFTPCI73T/tdT iAEC2s demonstrate significantly lower maximal respiration and spare respiratory capacity (mean ± SEM; n = 3 independent experiments) measured as percent change over basal respiration. (K) Late time point reveals a significantly higher ECAR in SFTPCI73T/tdT iAEC2s (mean ± SD; n = 3 independent experiments). (L) Representative live-cell confocal fluorescence microscopy of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s stained with BODIPY 493/503 dye with quantitative analyses. Scale bars, 10 μm (mean ± SD; n = 10 SFTPCtdT/WT and n = 7 SFTPCI73T/tdT independent alveolospheres treated as experimental replicates). (M) OCR measured under basal conditions and following addition of oligomycin, FCCP, and Ant A, as indicated (mean ± SD; n = 7 SftpcI73T and n = 5 WT mice). (N) ECAR measured under basal conditions in WT and SftpcI73T mouse AEC2s (mean ± SEM; n = 5 WT mice and n = 7 SftpcI73T). OCR, oxygen consumption rate; ECAR, extracellular acidification rate. (G and J) Depict super plots: small shapes represent replicate values within each experiment and large shapes represent the average of each independent experiment, tdT/WT and I73T/tdT are color-matched between experiments. *p

Figure 7.. SFTPC I73T/tdT iAEC2s display activation…

Figure 7.. SFTPC I73T/tdT iAEC2s display activation of the NF-κB pathway and are more susceptible…

Figure 7.. SFTPCI73T/tdT iAEC2s display activation of the NF-κB pathway and are more susceptible to hydroxychloroquine treatment
(A) Unsupervised hierarchical clustered heatmap of differentially expressed proteins (FDR in vivo inactivation of the viral LTR promoter; Ψ, Psi lentiviral packaging sequence; GFP, green fluorescence protein. (C) Representative live-cell imaging of lentivirally transduced SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s (bright-field/tdTomato/GFP overlay). Scale bars, 500 μm. (D) Sort gates used to purify transduced (GFP+) SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s. (E) Bioluminescence quantification shows increased luciferase activity in SFTPCI73T/tdT versus SFTPCtdT/WT iAEC2s. *p < 0.05, **p < 0.01, ***p < 0.001 by unpaired, two-tailed Student’s t test for all panels. (F) Luminex analysis of supernatants collected from SFTPCI73T/tdT and SFTPCtdT/WT iAEC2s for the NF-κB target proteins GM-CSF, CXCL5, and MMP-1. *p < 0.05, **p < 0.01, ***p < 0.001 by unpaired, two-tailed Student’s t test for all panels. (G) SFTPCtdT/WT or SFTPCI73T/tdT iAEC2s were treated with hydroxychloroquine (HCQ; 10 μM) or vehicle (ddH2O) for 18 h or 7 days. Lysates were subjected to western blotting for LC3 with β actin as loading control. Densitometric quantification is shown. See also Figure S5. *p < 0.05, **p < 0.01, ***p < 0.001 by unpaired, two-tailed Student’s t test for all panels. (H) Graph showing yield in cell number per input SFTPCtdT/WT or SFTPCI73T/tdT iAEC2 with or without treatment with HCQ (10 μM). ****p < 0.0001 across all groups by one-way ANOVA; p < 0.001 for d10, p < 0.01 for d20, p < 0.0001 for d31, and p < 0.05 for d44 for corrected versus mutant iAEC2s by unpaired, two-tailed Student’s t test. Mean ± SD is shown; n = 3 (E and F) or n = 2 (G) experimental replicates of independent wells of a differentiation or n = 3 (H) biological replicates of independent differentiations.
All figures (7)
Similar articles
Cited by
References
    1. Acin-Perez R, Benador IY, Petcherski A, Veliova M, Benavides GA, Lagarrigue S, Caudal A, Vergnes L, Murphy AN, Karamanlidis G, et al. (2020). A novel approach to measure mitochondrial respiration in frozen biological samples. EMBO J 39, e104073. - PMC - PubMed
    1. Araya J, Kojima J, Takasaka N, Ito S, Fujii S, Hara H, Yanagisawa H, Kobayashi K, Tsurushige C, Kawaishi M, et al. (2013). Insufficient autophagy in idiopathic pulmonary fibrosis. Am. J. Physiol. Lung Cell. Mol. Physiol 304, L56–L69. - PubMed
    1. Assali EA, Shlomo D, Zeng J, Taddeo EP, Trudeau KM, Erion KA, Colby AH, Grinstaff MW, Liesa M, Las G, and Shirihai OS (2019). Nanoparticle-mediated lysosomal reacidification restores mitochondrial turnover and function in β cells under lipotoxicity. FASEB J 33, 4154–4165. - PubMed
    1. Atochina-Vasserman EN, Bates SR, Zhang P, Abramova H, Zhang Z, Gonzales L, Tao J-Q, Gochuico BR, Gahl W, Guo C-J, et al. (2011). Early alveolar epithelial dysfunction promotes lung inflammation in a mouse model of Hermansky-Pudlak syndrome. Am. J. Respir. Crit. Care Med 184, 449–458. - PMC - PubMed
    1. Barkauskas CE, and Noble PW (2014). Cellular mechanisms of tissue fibrosis. 7. New insights into the cellular mechanisms of pulmonary fibrosis. Am. J. Physiol. Cell Physiol 306, C987–C996. - PMC - PubMed
Show all 86 references
Publication types
MeSH terms
[x]
Cite
Copy Download .nbib .nbib
Format: AMA APA MLA NLM

NCBI Literature Resources

MeSH PMC Bookshelf Disclaimer

The PubMed wordmark and PubMed logo are registered trademarks of the U.S. Department of Health and Human Services (HHS). Unauthorized use of these marks is strictly prohibited.

Follow NCBI
Figure 5.. Autophagy perturbations in SFTPC I73T…
Figure 5.. Autophagy perturbations in SFTPCI73T iAEC2s
(A) Schematic illustrating the autophagy pathway. Key proteins involved in the pathway (p62/SQSTM1 and LC3) and the mechanisms of action of bafilomycin A1 and torin are depicted. (B) Representative western blot of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2 lysates for p62 and LC3 with GAPDH and β actin as loading controls with densitometric quantification (mean ± SD; n = 11 for LC3 and n = 7 for p62 independent experiments). (C) Western blots of cell lysates for LC3 and β actin from autophagic flux studies using SFTPCtdT/WT or SFTPCI73T/tdT iAEC2s treated with either torin (5 μM) for 4, 8, or 18 h or bafilomycin A1 (BafA1) (50 or 100 nM) for 18 h or vehicle (DMSO) for 4 or 18 h (blots shown are representative of n = 2 independent experiments). (D) A repeat autophagic flux study, as in (C) but using BafA1 (50 nM) for 1, 2, 4, and 18 h or vehicle (DMSO) for 18 h (representative of n = 2 independent experiments). (E) SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s transduced with a lentiviral LC3:GFP fusion protein and exposed to BafA1 (50 nM) to quantify autophagosomes and autophagolysosomes. Scale bar, 10 μm (representative confocal fluorescence microscopy images of n = 3 experimental replicates of independent wells of a differentiation). (F) Representative TEM image of SFTPCI73T/tdT iAEC2s shows a double-membrane autophagosome (inset, arrowheads). Scale bar, 1 μm. (G) Real-time qPCR showing fold change in gene expression compared to WT in primary mouse SftpcI73T AEC2s (mean ± SEM; n = 6 for SftpcI73T and n = 5 for WT mice). (H) Representative western blots of primary mouse WT and SftpcI73T AEC2 lysates for pro-SFTPC (NPRO), LC3, and p62 with β actin as a loading control (see also Figure 3G) (n = 3 WT and SftpcI73T mice). (I) Representative live-cell confocal fluorescence microscopy of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s stained with LysoTracker green. Scale bars, 10 μm. (J) Quantification of acidic organelles expressed as percentage of covered cell area (mean ± SD; n = 11–14 independent alveolospheres treated as experimental replicates). (K) Quantification of lysosomal pH based on Lysosensor yellow/blue-dextran ratio (mean ± SD; n = 2 independent experiments with 20–30 cells analyzed per experiment; see also Figure S4). *p

Figure 6.. SFTPC I73T -expressing AEC2s demonstrate…

Figure 6.. SFTPC I73T -expressing AEC2s demonstrate metabolic reprogramming

(A) Representative live-cell confocal fluorescence microscopy of SFTPC…

Figure 6.. SFTPCI73T-expressing AEC2s demonstrate metabolic reprogramming
(A) Representative live-cell confocal fluorescence microscopy of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s stained with MitoTracker green. Scale bars, 10 μm. (B–D) Quantitative analyses of morphometric data from fluorescence images (mean ± SD; n = 24 SFTPCtdT/WT and n = 18 SFTPCI73T/tdT independent alveolospheres treated as experimental replicates). (B) Increased mitochondrial mass in SFTPCI73T/tdT iAEC2s, quantified as the percentage of cell area covered by mitochondria. (C) Increased mitochondrial size in SFTPCI73T/tdT iAEC2s. (D) SFTPCI73T/tdT iAEC2 mitochondria are more fragmented, assessed by the aspect ratio. (E) Representative western blot of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2 lysates for TOM20 with β actin as loading control with densitometric quantification (mean ± SD; n = 4 [days 128, 176, 223, 223] SFTPCtdT/WT and n = 3 [days 128, 223, 223] SFTPCI73T/tdT independent experiments). (F) Early time point OCR was measured under basal conditions followed by addition of oligomycin, FCCP, and antimycin A (Ant A), as indicated. (G) SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s demonstrate no significant differences in proton leak, maximal respiration, and spare capacity measured as percent change over basal respiration (mean ± SEM; n = 2 independent experiments). (H) Early time point ECAR under basal conditions and following addition of oligomycin. (I) Late time point OCR. (J) SFTPCI73T/tdT iAEC2s demonstrate significantly lower maximal respiration and spare respiratory capacity (mean ± SEM; n = 3 independent experiments) measured as percent change over basal respiration. (K) Late time point reveals a significantly higher ECAR in SFTPCI73T/tdT iAEC2s (mean ± SD; n = 3 independent experiments). (L) Representative live-cell confocal fluorescence microscopy of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s stained with BODIPY 493/503 dye with quantitative analyses. Scale bars, 10 μm (mean ± SD; n = 10 SFTPCtdT/WT and n = 7 SFTPCI73T/tdT independent alveolospheres treated as experimental replicates). (M) OCR measured under basal conditions and following addition of oligomycin, FCCP, and Ant A, as indicated (mean ± SD; n = 7 SftpcI73T and n = 5 WT mice). (N) ECAR measured under basal conditions in WT and SftpcI73T mouse AEC2s (mean ± SEM; n = 5 WT mice and n = 7 SftpcI73T). OCR, oxygen consumption rate; ECAR, extracellular acidification rate. (G and J) Depict super plots: small shapes represent replicate values within each experiment and large shapes represent the average of each independent experiment, tdT/WT and I73T/tdT are color-matched between experiments. *p

Figure 7.. SFTPC I73T/tdT iAEC2s display activation…

Figure 7.. SFTPC I73T/tdT iAEC2s display activation of the NF-κB pathway and are more susceptible…

Figure 7.. SFTPCI73T/tdT iAEC2s display activation of the NF-κB pathway and are more susceptible to hydroxychloroquine treatment
(A) Unsupervised hierarchical clustered heatmap of differentially expressed proteins (FDR in vivo inactivation of the viral LTR promoter; Ψ, Psi lentiviral packaging sequence; GFP, green fluorescence protein. (C) Representative live-cell imaging of lentivirally transduced SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s (bright-field/tdTomato/GFP overlay). Scale bars, 500 μm. (D) Sort gates used to purify transduced (GFP+) SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s. (E) Bioluminescence quantification shows increased luciferase activity in SFTPCI73T/tdT versus SFTPCtdT/WT iAEC2s. *p < 0.05, **p < 0.01, ***p < 0.001 by unpaired, two-tailed Student’s t test for all panels. (F) Luminex analysis of supernatants collected from SFTPCI73T/tdT and SFTPCtdT/WT iAEC2s for the NF-κB target proteins GM-CSF, CXCL5, and MMP-1. *p < 0.05, **p < 0.01, ***p < 0.001 by unpaired, two-tailed Student’s t test for all panels. (G) SFTPCtdT/WT or SFTPCI73T/tdT iAEC2s were treated with hydroxychloroquine (HCQ; 10 μM) or vehicle (ddH2O) for 18 h or 7 days. Lysates were subjected to western blotting for LC3 with β actin as loading control. Densitometric quantification is shown. See also Figure S5. *p < 0.05, **p < 0.01, ***p < 0.001 by unpaired, two-tailed Student’s t test for all panels. (H) Graph showing yield in cell number per input SFTPCtdT/WT or SFTPCI73T/tdT iAEC2 with or without treatment with HCQ (10 μM). ****p < 0.0001 across all groups by one-way ANOVA; p < 0.001 for d10, p < 0.01 for d20, p < 0.0001 for d31, and p < 0.05 for d44 for corrected versus mutant iAEC2s by unpaired, two-tailed Student’s t test. Mean ± SD is shown; n = 3 (E and F) or n = 2 (G) experimental replicates of independent wells of a differentiation or n = 3 (H) biological replicates of independent differentiations.
All figures (7)
Similar articles
Cited by
References
    1. Acin-Perez R, Benador IY, Petcherski A, Veliova M, Benavides GA, Lagarrigue S, Caudal A, Vergnes L, Murphy AN, Karamanlidis G, et al. (2020). A novel approach to measure mitochondrial respiration in frozen biological samples. EMBO J 39, e104073. - PMC - PubMed
    1. Araya J, Kojima J, Takasaka N, Ito S, Fujii S, Hara H, Yanagisawa H, Kobayashi K, Tsurushige C, Kawaishi M, et al. (2013). Insufficient autophagy in idiopathic pulmonary fibrosis. Am. J. Physiol. Lung Cell. Mol. Physiol 304, L56–L69. - PubMed
    1. Assali EA, Shlomo D, Zeng J, Taddeo EP, Trudeau KM, Erion KA, Colby AH, Grinstaff MW, Liesa M, Las G, and Shirihai OS (2019). Nanoparticle-mediated lysosomal reacidification restores mitochondrial turnover and function in β cells under lipotoxicity. FASEB J 33, 4154–4165. - PubMed
    1. Atochina-Vasserman EN, Bates SR, Zhang P, Abramova H, Zhang Z, Gonzales L, Tao J-Q, Gochuico BR, Gahl W, Guo C-J, et al. (2011). Early alveolar epithelial dysfunction promotes lung inflammation in a mouse model of Hermansky-Pudlak syndrome. Am. J. Respir. Crit. Care Med 184, 449–458. - PMC - PubMed
    1. Barkauskas CE, and Noble PW (2014). Cellular mechanisms of tissue fibrosis. 7. New insights into the cellular mechanisms of pulmonary fibrosis. Am. J. Physiol. Cell Physiol 306, C987–C996. - PMC - PubMed
Show all 86 references
Publication types
MeSH terms
[x]
Cite
Copy Download .nbib .nbib
Format: AMA APA MLA NLM
Figure 6.. SFTPC I73T -expressing AEC2s demonstrate…
Figure 6.. SFTPCI73T-expressing AEC2s demonstrate metabolic reprogramming
(A) Representative live-cell confocal fluorescence microscopy of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s stained with MitoTracker green. Scale bars, 10 μm. (B–D) Quantitative analyses of morphometric data from fluorescence images (mean ± SD; n = 24 SFTPCtdT/WT and n = 18 SFTPCI73T/tdT independent alveolospheres treated as experimental replicates). (B) Increased mitochondrial mass in SFTPCI73T/tdT iAEC2s, quantified as the percentage of cell area covered by mitochondria. (C) Increased mitochondrial size in SFTPCI73T/tdT iAEC2s. (D) SFTPCI73T/tdT iAEC2 mitochondria are more fragmented, assessed by the aspect ratio. (E) Representative western blot of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2 lysates for TOM20 with β actin as loading control with densitometric quantification (mean ± SD; n = 4 [days 128, 176, 223, 223] SFTPCtdT/WT and n = 3 [days 128, 223, 223] SFTPCI73T/tdT independent experiments). (F) Early time point OCR was measured under basal conditions followed by addition of oligomycin, FCCP, and antimycin A (Ant A), as indicated. (G) SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s demonstrate no significant differences in proton leak, maximal respiration, and spare capacity measured as percent change over basal respiration (mean ± SEM; n = 2 independent experiments). (H) Early time point ECAR under basal conditions and following addition of oligomycin. (I) Late time point OCR. (J) SFTPCI73T/tdT iAEC2s demonstrate significantly lower maximal respiration and spare respiratory capacity (mean ± SEM; n = 3 independent experiments) measured as percent change over basal respiration. (K) Late time point reveals a significantly higher ECAR in SFTPCI73T/tdT iAEC2s (mean ± SD; n = 3 independent experiments). (L) Representative live-cell confocal fluorescence microscopy of SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s stained with BODIPY 493/503 dye with quantitative analyses. Scale bars, 10 μm (mean ± SD; n = 10 SFTPCtdT/WT and n = 7 SFTPCI73T/tdT independent alveolospheres treated as experimental replicates). (M) OCR measured under basal conditions and following addition of oligomycin, FCCP, and Ant A, as indicated (mean ± SD; n = 7 SftpcI73T and n = 5 WT mice). (N) ECAR measured under basal conditions in WT and SftpcI73T mouse AEC2s (mean ± SEM; n = 5 WT mice and n = 7 SftpcI73T). OCR, oxygen consumption rate; ECAR, extracellular acidification rate. (G and J) Depict super plots: small shapes represent replicate values within each experiment and large shapes represent the average of each independent experiment, tdT/WT and I73T/tdT are color-matched between experiments. *p

Figure 7.. SFTPC I73T/tdT iAEC2s display activation…

Figure 7.. SFTPC I73T/tdT iAEC2s display activation of the NF-κB pathway and are more susceptible…

Figure 7.. SFTPCI73T/tdT iAEC2s display activation of the NF-κB pathway and are more susceptible to hydroxychloroquine treatment
(A) Unsupervised hierarchical clustered heatmap of differentially expressed proteins (FDR in vivo inactivation of the viral LTR promoter; Ψ, Psi lentiviral packaging sequence; GFP, green fluorescence protein. (C) Representative live-cell imaging of lentivirally transduced SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s (bright-field/tdTomato/GFP overlay). Scale bars, 500 μm. (D) Sort gates used to purify transduced (GFP+) SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s. (E) Bioluminescence quantification shows increased luciferase activity in SFTPCI73T/tdT versus SFTPCtdT/WT iAEC2s. *p < 0.05, **p < 0.01, ***p < 0.001 by unpaired, two-tailed Student’s t test for all panels. (F) Luminex analysis of supernatants collected from SFTPCI73T/tdT and SFTPCtdT/WT iAEC2s for the NF-κB target proteins GM-CSF, CXCL5, and MMP-1. *p < 0.05, **p < 0.01, ***p < 0.001 by unpaired, two-tailed Student’s t test for all panels. (G) SFTPCtdT/WT or SFTPCI73T/tdT iAEC2s were treated with hydroxychloroquine (HCQ; 10 μM) or vehicle (ddH2O) for 18 h or 7 days. Lysates were subjected to western blotting for LC3 with β actin as loading control. Densitometric quantification is shown. See also Figure S5. *p < 0.05, **p < 0.01, ***p < 0.001 by unpaired, two-tailed Student’s t test for all panels. (H) Graph showing yield in cell number per input SFTPCtdT/WT or SFTPCI73T/tdT iAEC2 with or without treatment with HCQ (10 μM). ****p < 0.0001 across all groups by one-way ANOVA; p < 0.001 for d10, p < 0.01 for d20, p < 0.0001 for d31, and p < 0.05 for d44 for corrected versus mutant iAEC2s by unpaired, two-tailed Student’s t test. Mean ± SD is shown; n = 3 (E and F) or n = 2 (G) experimental replicates of independent wells of a differentiation or n = 3 (H) biological replicates of independent differentiations.
All figures (7)
Figure 7.. SFTPC I73T/tdT iAEC2s display activation…
Figure 7.. SFTPCI73T/tdT iAEC2s display activation of the NF-κB pathway and are more susceptible to hydroxychloroquine treatment
(A) Unsupervised hierarchical clustered heatmap of differentially expressed proteins (FDR in vivo inactivation of the viral LTR promoter; Ψ, Psi lentiviral packaging sequence; GFP, green fluorescence protein. (C) Representative live-cell imaging of lentivirally transduced SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s (bright-field/tdTomato/GFP overlay). Scale bars, 500 μm. (D) Sort gates used to purify transduced (GFP+) SFTPCtdT/WT and SFTPCI73T/tdT iAEC2s. (E) Bioluminescence quantification shows increased luciferase activity in SFTPCI73T/tdT versus SFTPCtdT/WT iAEC2s. *p < 0.05, **p < 0.01, ***p < 0.001 by unpaired, two-tailed Student’s t test for all panels. (F) Luminex analysis of supernatants collected from SFTPCI73T/tdT and SFTPCtdT/WT iAEC2s for the NF-κB target proteins GM-CSF, CXCL5, and MMP-1. *p < 0.05, **p < 0.01, ***p < 0.001 by unpaired, two-tailed Student’s t test for all panels. (G) SFTPCtdT/WT or SFTPCI73T/tdT iAEC2s were treated with hydroxychloroquine (HCQ; 10 μM) or vehicle (ddH2O) for 18 h or 7 days. Lysates were subjected to western blotting for LC3 with β actin as loading control. Densitometric quantification is shown. See also Figure S5. *p < 0.05, **p < 0.01, ***p < 0.001 by unpaired, two-tailed Student’s t test for all panels. (H) Graph showing yield in cell number per input SFTPCtdT/WT or SFTPCI73T/tdT iAEC2 with or without treatment with HCQ (10 μM). ****p < 0.0001 across all groups by one-way ANOVA; p < 0.001 for d10, p < 0.01 for d20, p < 0.0001 for d31, and p < 0.05 for d44 for corrected versus mutant iAEC2s by unpaired, two-tailed Student’s t test. Mean ± SD is shown; n = 3 (E and F) or n = 2 (G) experimental replicates of independent wells of a differentiation or n = 3 (H) biological replicates of independent differentiations.

References

    1. Acin-Perez R, Benador IY, Petcherski A, Veliova M, Benavides GA, Lagarrigue S, Caudal A, Vergnes L, Murphy AN, Karamanlidis G, et al. (2020). A novel approach to measure mitochondrial respiration in frozen biological samples. EMBO J 39, e104073.
    1. Araya J, Kojima J, Takasaka N, Ito S, Fujii S, Hara H, Yanagisawa H, Kobayashi K, Tsurushige C, Kawaishi M, et al. (2013). Insufficient autophagy in idiopathic pulmonary fibrosis. Am. J. Physiol. Lung Cell. Mol. Physiol 304, L56–L69.
    1. Assali EA, Shlomo D, Zeng J, Taddeo EP, Trudeau KM, Erion KA, Colby AH, Grinstaff MW, Liesa M, Las G, and Shirihai OS (2019). Nanoparticle-mediated lysosomal reacidification restores mitochondrial turnover and function in β cells under lipotoxicity. FASEB J 33, 4154–4165.
    1. Atochina-Vasserman EN, Bates SR, Zhang P, Abramova H, Zhang Z, Gonzales L, Tao J-Q, Gochuico BR, Gahl W, Guo C-J, et al. (2011). Early alveolar epithelial dysfunction promotes lung inflammation in a mouse model of Hermansky-Pudlak syndrome. Am. J. Respir. Crit. Care Med 184, 449–458.
    1. Barkauskas CE, and Noble PW (2014). Cellular mechanisms of tissue fibrosis. 7. New insights into the cellular mechanisms of pulmonary fibrosis. Am. J. Physiol. Cell Physiol 306, C987–C996.
    1. Barkauskas CE, Cronce MJ, Rackley CR, Bowie EJ, Keene DR, Stripp BR, Randell SH, Noble PW, and Hogan BLM (2013). Type 2 alveolar cells are stem cells in adult lung. J. Clin. Invest 123, 3025–3036.
    1. Beers MF, and Lomax C (1995). Synthesis and processing of hydrophobic surfactant protein C by isolated rat type II cells. Am. J. Physiol 269, L744–L753.
    1. Beers MF, and Mulugeta S (2005). Surfactant protein C biosynthesis and its emerging role in conformational lung disease. Annu. Rev. Physiol 67, 663–696.
    1. Beers MF, Bates SR, and Fisher AB (1992). Differential extraction for the rapid purification of bovine surfactant protein B. Am. J. Physiol 262, L773–L778.
    1. Beers MF, Kim CY, Dodia C, and Fisher AB (1994). Localization, synthesis, and processing of surfactant protein SP-C in rat lung analyzed by epitope-specific antipeptide antibodies. J. Biol. Chem 269, 20318–20328.
    1. Beers MF, Hawkins A, Maguire JA, Kotorashvili A, Zhao M, Newitt JL, Ding W, Russo S, Guttentag S, Gonzales L, and Mulugeta S (2011). A nonaggregating surfactant protein C mutant is misdirected to early endosomes and disrupts phospholipid recycling. Traffic 12, 1196–1210.
    1. Bock C, Kiskinis E, Verstappen G, Gu H, Boulting G, Smith ZD, Ziller M, Croft GF, Amoroso MW, Oakley DH, et al. (2011). Reference Maps of human ES and iPS cell variation enable high-throughput characterization of pluripotent cell lines. Cell 144, 439–452.
    1. Borok Z, Danto SI, Lubman RL, Cao Y, Williams MC, and Crandall ED (1998). Modulation of t1alpha expression with alveolar epithelial cell phenotype in vitro. Am. J. Physiol 275, L155–L164.
    1. Boulting GL, Kiskinis E, Croft GF, Amoroso MW, Oakley DH, Wainger BJ, Williams DJ, Kahler DJ, Yamaki M, Davidow L, et al. (2011). A functionally characterized test set of human induced pluripotent stem cells. Nat. Biotechnol 29, 279–286.
    1. Brasch F, Ten Brinke A, Johnen G, Ochs M, Kapp N, Müller KM, Beers MF, Fehrenbach H, Richter J, Batenburg JJ, and Bühling F (2002). Involvement of cathepsin H in the processing of the hydrophobic surfactant-associated protein C in type II pneumocytes. Am. J. Respir. Cell Mol. Biol 26, 659–670.
    1. Brasch F, Griese M, Tredano M, Johnen G, Ochs M, Rieger C, Mulugeta S, Müller KM, Bahuau M, and Beers MF (2004). Interstitial lung disease in a baby with a de novo mutation in the SFTPC gene. Eur. Respir. J 24, 30–39.
    1. Braun S, Ferner M, Kronfeld K, and Griese M (2015). Hydroxychloroquine in children with interstitial (diffuse parenchymal) lung diseases. Pediatr. Pulmonol 50, 410–419.
    1. Bueno M, Lai Y-C, Romero Y, Brands J, St Croix CM, Kamga C, Corey C, Herazo-Maya JD, Sembrat J, Lee JS, et al. (2015). PINK1 deficiency impairs mitochondrial homeostasis and promotes lung fibrosis. J. Clin. Invest 125, 521–538.
    1. Cottin V, Reix P, Khouatra C, Thivolet-Béjui F, Feldmann D, and Cordier J-F (2011). Combined pulmonary fibrosis and emphysema syndrome associated with familial SFTPC mutation. Thorax 66, 918–919.
    1. Crossno PF, Polosukhin VV, Blackwell TS, Johnson JE, Markin C, Moore PE, Worrell JA, Stahlman MT, Phillips JA 3rd, Loyd JE, et al. (2010). Identification of early interstitial lung disease in an individual with genetic variations in ABCA3 and SFTPC. Chest 137, 969–973.
    1. Deutsch EW, Bandeira N, Sharma V, Perez-Riverol Y, Carver JJ, Kundu DJ, García-Seisdedos D, Jarnuczak AF, Hewapathirana S, Pullman BS, et al. (2020). The ProteomeXchange consortium in 2020: enabling ‘big data’ approaches in proteomics. Nucleic Acids Res 48 (D1), D1145–D1152.
    1. Dobin A, Davis CA, Schlesinger F, Drenkow J, Zaleski C, Jha S, Batut P, Chaisson M, and Gingeras TR (2013). STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21.
    1. Foster CD, Varghese LS, Skalina RB, Gonzales LW, and Guttentag SH (2007). In vitro transdifferentiation of human fetal type II cells toward a type I-like cell. Pediatr. Res 61, 404–409.
    1. Galetskiy D, Woischnik M, Ripper J, Griese M, and Przybylski M (2008). Aberrant processing forms of lung surfactant proteins SP-B and SP-C revealed by high-resolution mass spectrometry. Eur. J. Mass Spectrom (Chi-chester) 14, 379–390.
    1. Garcia CK (2018). Insights from human genetic studies of lung and organ fibrosis. J. Clin. Invest 128, 36–44.
    1. Garcia O, Hiatt MJ, Lundin A, Lee J, Reddy R, Navarro S, Kikuchi A, and Driscoll B (2016). Targeted Type 2 Alveolar Cell Depletion. A Dynamic Functional Model for Lung Injury Repair. Am. J. Respir. Cell Mol. Biol 54, 319–330.
    1. Gu Z, Eils R, and Schlesner M (2016). Complex heatmaps reveal patterns and correlations in multidimensional genomic data. Bioinformatics 32, 2847–2849.
    1. Habermann AC, Gutierrez AJ, Bui LT, Yahn SL, Winters NI, Calvi CL, Peter L, Chung M-I, Taylor CJ, Jetter C, et al. (2020). Single-cell RNA sequencing reveals profibrotic roles of distinct epithelial and mesenchymal lineages in pulmonary fibrosis. Sci. Adv 6, eaba1972.
    1. Hawkins A, Guttentag SH, Deterding R, Funkhouser WK, Goralski JL, Chatterjee S, Mulugeta S, and Beers MF (2015). 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.
    1. Hawkins F, Kramer P, Jacob A, Driver I, Thomas DC, McCauley KB, Skvir N, Crane AM, Kurmann AA, Hollenberg AN, et al. (2017). Prospective isolation of NKX2-1-expressing human lung progenitors derived from pluripotent stem cells. J. Clin. Invest 127, 2277–2294.
    1. Hill C, Li J, Liu D, Conforti F, Brereton CJ, Yao L, Zhou Y, Alzetani A, Chee SJ, Marshall BG, et al. (2019). Autophagy inhibition-mediated epithelial-mesenchymal transition augments local myofibroblast differentiation in pulmonary fibrosis. Cell Death Dis 10, 591.
    1. Hurley K, Ding J, Villacorta-Martin C, Herriges MJ, Jacob A, Vedaie M, Alysandratos K-D, Sun YL, Lin C, Werder RB, et al. (2020). Reconstructed Single-Cell Fate Trajectories Define Lineage Plasticity Windows during Differentiation of Human PSC-Derived Distal Lung Progenitors. Cell Stem Cell 26, 593–608.e8.
    1. Jacob A, Morley M, Hawkins F, McCauley KB, Jean JC, Heins H, Na C-L, Weaver TE, Vedaie M, Hurley K, et al. (2017). Differentiation of Human Pluripotent Stem Cells into Functional Lung Alveolar Epithelial Cells. Cell Stem Cell 21, 472–488.e10.
    1. Jacob A, Vedaie M, Roberts DA, Thomas DC, Villacorta-Martin C, Alysandratos K-D, Hawkins F, and Kotton DN (2019). Derivation of self-renewing lung alveolar epithelial type II cells from human pluripotent stem cells. Nat. Protoc 14, 3303–3332.
    1. Jiang P, Gil de Rubio R, Hrycaj SM, Gurczynski SJ, Riemondy KA, Moore BB, Omary MB, Ridge KM, and Zemans RL (2020). Ineffectual Type 2-to-Type 1 Alveolar Epithelial Cell Differentiation in Idiopathic Pulmonary Fibrosis: Persistence of the KRT8hi Transitional State. Am. J. Respir. Crit. Care Med 201, 1443–1447.
    1. Katzen J, Wagner BD, Venosa A, Kopp M, Tomer Y, Russo SJ, Headen AC, Basil MC, Stark JM, Mulugeta S, et al. (2019). An SFTPC BRICHOS mutant links epithelial ER stress and spontaneous lung fibrosis. JCI Insight 4, e126125.
    1. Katzenstein AL (1985). Pathogenesis of “fibrosis” in interstitial pneumonia: an electron microscopic study. Hum. Pathol 16, 1015–1024.
    1. Kim K, Doi A, Wen B, Ng K, Zhao R, Cahan P, Kim J, Aryee MJ, Ji H, Ehrlich LIR, et al. (2010). Epigenetic memory in induced pluripotent stem cells. Nature 467, 285–290.
    1. Klay D, Hoffman TW, Harmsze AM, Grutters JC, and van Moorsel CHM (2018). Systematic review of drug effects in humans and models with surfactant-processing disease. Eur. Respir. Rev 27, 170135.
    1. Klionsky DJ, Abdelmohsen K, Abe A, Abedin MJ, Abeliovich H, Acevedo Arozena A, Adachi H, Adams CM, Adams PD, Adeli K, et al. (2016). Guidelines for the use and interpretation of assays for monitoring autophagy (3rd edition). Autophagy 12, 1–222.
    1. Korogi Y, Gotoh S, Ikeo S, Yamamoto Y, Sone N, Tamai K, Konishi S, Nagasaki T, Matsumoto H, Ito I, et al. (2019). In Vitro Disease Modeling of Hermansky-Pudlak Syndrome Type 2 Using Human Induced Pluripotent Stem Cell-Derived Alveolar Organoids. Stem Cell Reports 12, 431–440.
    1. Korotkevich G, Sukhov V, and Sergushichev A (2019). Fast gene set enrichment analysis. bioRxiv 10.1101/060012.
    1. Kröner C, Reu S, Teusch V, Schams A, Grimmelt A-C, Barker M, Brand J, Gappa M, Kitz R, Kramer BW, et al. (2015). Genotype alone does not predict the clinical course of SFTPC deficiency in paediatric patients. Eur. Respir. J 46, 197–206.
    1. Kropski JA, Blackwell TS, and Loyd JE (2015). The genetic basis of idiopathic pulmonary fibrosis. Eur. Respir. J 45, 1717–1727.
    1. Law CW, Alhamdoosh M, Su S, Dong X, Tian L, Smyth GK, and Ritchie ME (2016). RNA-seq analysis is easy as 1-2-3 with limma, Glimma and edgeR. F1000Res 5, 1408.
    1. Lederer DJ, and Martinez FJ (2018). Idiopathic Pulmonary Fibrosis. N. Engl. J. Med 378, 1811–1823.
    1. Maguire JA, Mulugeta S, and Beers MF (2012). 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.
    1. Massaro GD, Gail DB, and Massaro D (1975). Lung oxygen consumption and mitochondria of alveolar epithelial and endothelial cells. J. Appl. Physiol 38, 588–592.
    1. Mauthe M, Orhon I, Rocchi C, Zhou X, Luhr M, Hijlkema K-J, Coppes RP, Engedal N, Mari M, and Reggiori F (2018). Chloroquine inhibits autophagic flux by decreasing autophagosome-lysosome fusion. Autophagy 14, 1435–1455.
    1. Misharin AV, Morales-Nebreda L, Reyfman PA, Cuda CM, Walter JM, McQuattie-Pimentel AC, Chen C-I, Anekalla KR, Joshi N, Williams KJN, et al. (2017). Monocyte-derived alveolar macrophages drive lung fibrosis and persist in the lung over the life span. J. Exp. Med 214, 2387–2404.
    1. Mulugeta S, Nureki S, and Beers MF (2015). Lost after translation: insights from pulmonary surfactant for understanding the role of alveolar epithelial dysfunction and cellular quality control in fibrotic lung disease. Am. J. Physiol. Lung Cell. Mol. Physiol 309, L507–L525.
    1. Ni H-M, Bockus A, Wozniak AL, Jones K, Weinman S, Yin X-M, and Ding W-X (2011). Dissecting the dynamic turnover of GFP-LC3 in the autolysosome. Autophagy 7, 188–204.
    1. Nogee LM, Dunbar AE 3rd, Wert SE, Askin F, Hamvas A, and Whitsett JA (2001). A mutation in the surfactant protein C gene associated with familial interstitial lung disease. N. Engl. J. Med 344, 573–579.
    1. Nureki S-I, Tomer Y, Venosa A, Katzen J, Russo SJ, Jamil S, Barrett M, Nguyen V, Kopp M, Mulugeta S, and Beers MF (2018). Expression of mutant Sftpc in murine alveolar epithelia drives spontaneous lung fibrosis. J. Clin. Invest 128, 4008–4024.
    1. Ono S, Tanaka T, Ishida M, Kinoshita A, Fukuoka J, Takaki M, Sakamoto N, Ishimatsu Y, Kohno S, Hayashi T, et al. (2011). Surfactant protein C G100S mutation causes familial pulmonary fibrosis in Japanese kindred. Eur. Respir. J 38, 861–869.
    1. Patel AS, Lin L, Geyer A, Haspel JA, An CH, Cao J, Rosas IO, and Morse D (2012). Autophagy in idiopathic pulmonary fibrosis. PLoS ONE 7, e41394.
    1. Perez-Riverol Y, Csordas A, Bai J, Bernal-Llinares M, Hewapathirana S, Kundu DJ, Inuganti A, Griss J, Mayer G, Eisenacher M, et al. (2019). The PRIDE database and related tools and resources in 2019: improving support for quantification data. Nucleic Acids Res 47 (D1), D442–D450.
    1. Raghu G, Chen S-Y, Yeh W-S, Maroni B, Li Q, Lee Y-C, and Collard HR (2014). Idiopathic pulmonary fibrosis in US Medicare beneficiaries aged 65 years and older: incidence, prevalence, and survival, 2001–11. Lancet Respir. Med 2, 566–572.
    1. Raghu G, Chen S-Y, Hou Q, Yeh W-S, and Collard HR (2016). Incidence and prevalence of idiopathic pulmonary fibrosis in US adults 18–64 years old. Eur. Respir. J 48, 179–186.
    1. Reimand J, Isserlin R, Voisin V, Kucera M, Tannus-Lopes C, Rostamianfar A, Wadi L, Meyer M, Wong J, Xu C, et al. (2019). Pathway enrichment analysis and visualization of omics data using g:Profiler, GSEA, Cytoscape and EnrichmentMap. Nat. Protoc 14, 482–517.
    1. Riemondy KA, Jansing NL, Jiang P, Redente EF, Gillen AE, Fu R, Miller AJ, Spence JR, Gerber AN, Hesselberth JR, and Zemans RL (2019). Single cell RNA sequencing identifies TGFβ as a key regenerative cue following LPS-induced lung injury. JCI Insight 5, e123637.
    1. Ritchie ME, Phipson B, Wu D, Hu Y, Law CW, Shi W, and Smyth GK (2015). limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res 43, e47.
    1. Schrezenmeier E, and Dörner T (2020). Mechanisms of action of hydroxychloroquine and chloroquine: implications for rheumatology. Nat. Rev. Rheumatol 16, 155–166.
    1. Selman M, and Pardo A (2014). 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.
    1. Serra M, Alysandratos K-D, Hawkins F, McCauley KB, Jacob A, Choi J, Caballero IS, Vedaie M, Kurmann AA, Ikonomou L, et al. (2017). Pluripotent stem cell differentiation reveals distinct developmental pathways regulating lung- versus thyroid-lineage specification. Development 144, 3879–3893.
    1. Sisson TH, Mendez M, Choi K, Subbotina N, Courey A, Cunningham A, Dave A, Engelhardt JF, Liu X, White ES, et al. (2010). Targeted injury of type II alveolar epithelial cells induces pulmonary fibrosis. Am. J. Respir. Crit. Care Med 181, 254–263.
    1. Somers A, Jean J-C, Sommer CA, Omari A, Ford CC, Mills JA, Ying L, Sommer AG, Jean JM, Smith BW, et al. (2010). Generation of transgene-free lung disease-specific human induced pluripotent stem cells using a single excisable lentiviral stem cell cassette. Stem Cells 28, 1728–1740.
    1. Stewart GA, Ridsdale R, Martin EP, Na C-L, Xu Y, Mandapaka K, and Weaver TE (2012). 4-Phenylbutyric acid treatment rescues trafficking and processing of a mutant surfactant protein-C. Am. J. Respir. Cell Mol. Biol 47, 324–331.
    1. Strikoudis A, Cieślak A, Loffredo L, Chen Y-W, Patel N, Saqi A, Lederer DJ, and Snoeck H-W (2019). Modeling of Fibrotic Lung Disease Using 3D Organoids Derived from Human Pluripotent Stem Cells. Cell Rep 27, 3709–3723.e5.
    1. Subramanian A, Tamayo P, Mootha VK, Mukherjee S, Ebert BL, Gillette MA, Paulovich A, Pomeroy SL, Golub TR, Lander ES, and Mesirov JP (2005). Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc. Natl. Acad. Sci. USA 102, 15545–15550.
    1. Szymaniak AD, Mahoney JE, Cardoso WV, and Varelas X (2015). Crumbs3-Mediated Polarity Directs Airway Epithelial Cell Fate through the Hippo Pathway Effector Yap. Dev. Cell 34, 283–296.
    1. Taddeo EP, Alsabeeh N, Baghdasarian S, Wikstrom JD, Ritou E, Sereda S, Erion K, Li J, Stiles L, Abdulla M, et al. (2020). Mitochondrial Proton Leak Regulated by Cyclophilin D Elevates Insulin Secretion in Islets at Nonstimulatory Glucose Levels. Diabetes 69, 131–145.
    1. Thomas AQ, Lane K, Phillips J 3rd, Prince M, Markin C, Speer M, Schwartz DA, Gaddipati R, Marney A, Johnson J, et al. (2002). 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.
    1. Travis WD, Costabel U, Hansell DM, King TE Jr., Lynch DA, Nicholson AG, Ryerson CJ, Ryu JH, Selman M, Wells AU, et al.; ATS/ERS Committee on Idiopathic Interstitial Pneumonias (2013). An official American Thoracic Society/European Respiratory Society statement: Update of the international multidisciplinary classification of the idiopathic interstitial pneumonias. Am. J. Respir. Crit. Care Med 188, 733–748.
    1. Twig G, Elorza A, Molina AJA, Mohamed H, Wikstrom JD, Walzer G, Stiles L, Haigh SE, Katz S, Las G, et al. (2008). Fission and selective fusion govern mitochondrial segregation and elimination by autophagy. EMBO J 27, 433–446.
    1. van Moorsel CHM, van Oosterhout MFM, Barlo NP, de Jong PA, van der Vis JJ, Ruven HJT, van Es HW, van den Bosch JMM, and Grutters JC (2010). 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.
    1. Venosa A, Katzen J, Tomer Y, Kopp M, Jamil S, Russo SJ, Mulugeta S, and Beers MF (2019). Epithelial Expression of an Interstitial Lung Disease-Associated Mutation in Surfactant Protein-C Modulates Recruitment and Activation of Key Myeloid Cell Populations in Mice. J. Immunol 202, 2760–2771.
    1. Wade KC, Guttentag SH, Gonzales LW, Maschhoff KL, Gonzales J, Kolla V, Singhal S, and Ballard PL (2006). Gene induction during differentiation of human pulmonary type II cells in vitro. Am. J. Respir. Cell Mol. Biol 34, 727–737.
    1. Wilson AA, Kwok LW, Porter EL, Payne JG, McElroy GS, Ohle SJ, Greenhill SR, Blahna MT, Yamamoto K, Jean JC, et al. (2013). Lentiviral delivery of RNAi for in vivo lineage-specific modulation of gene expression in mouse lung macrophages. Mol. Ther 21, 825–833.
    1. Winters NI, Burman A, Kropski JA, and Blackwell TS (2019). Epithelial Injury and Dysfunction in the Pathogenesis of Idiopathic PulmonaryFibrosis. Am. J. Med. Sci 357, 374–378.
    1. Wolfe DM, Lee J-H, Kumar A, Lee S, Orenstein SJ, and Nixon RA (2013). Autophagy failure in Alzheimer’s disease and the role of defective lysosomal acidification. Eur. J. Neurosci 37, 1949–1961.
    1. Wu D, and Smyth GK (2012). Camera: a competitive gene set test accounting for inter-gene correlation. Nucleic Acids Res 40, e133.
    1. Yamamoto A, Tagawa Y, Yoshimori T, Moriyama Y, Masaki R, and Tashiro Y (1998). Bafilomycin A1 prevents maturation of autophagic vacuoles by inhibiting fusion between autophagosomes and lysosomes in rat hepatoma cell line, H-4-II-E cells. Cell Struct. Funct 23, 33–42.
    1. Yao C, Guan X, Carraro G, Parimon T, Liu X, Huang G, Mulay A, Soukiasian HJ, David G, Weigt SS, et al. (2020). Senescence of Alveolar Type 2 Cells Drives Progressive Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med 203, 707–717.
    1. Yin BWT, Kiyamova R, Chua R, Caballero OL, Gout I, Gryshkova V, Bhaskaran N, Souchelnytskyi S, Hellman U, Filonenko V, et al. (2008). Monoclonal antibody MX35 detects the membrane transporter NaPi2b (SLC34A2) in human carcinomas. Cancer Immun 8, 3.
    1. Yu G, Tzouvelekis A, Wang R, Herazo-Maya JD, Ibarra GH, Srivastava A, de Castro JPW, DeIuliis G, Ahangari F, Woolard T, et al. (2018). Thyroid hormone inhibits lung fibrosis in mice by improving epithelial mitochondrial function. Nat. Med 24, 39–49.

Source: PubMed

3
Tilaa