Next Article in Journal
Vitamin D Inadequacy Affects Skeletal Muscle Index and Physical Performance in Lumbar Disc Degeneration
Next Article in Special Issue
First Characterization of the Transcriptome of Lung Fibroblasts of SSc Patients and Healthy Donors of African Ancestry
Previous Article in Journal
Advances in the Synthesis and Analysis of Biologically Active Phosphometabolites
Previous Article in Special Issue
The Molecular Mechanisms of Systemic Sclerosis-Associated Lung Fibrosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Immune Mechanisms of Pulmonary Fibrosis with Bleomycin

Department of Forensic Medicine, Wakayama Medical University, 811-1 Kimiidera, Wakayama 641-8509, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(4), 3149; https://doi.org/10.3390/ijms24043149
Submission received: 28 December 2022 / Revised: 27 January 2023 / Accepted: 1 February 2023 / Published: 5 February 2023

Abstract

:
Fibrosis and structural remodeling of the lung tissue can significantly impair lung function, often with fatal consequences. The etiology of pulmonary fibrosis (PF) is diverse and includes different triggers such as allergens, chemicals, radiation, and environmental particles. However, the cause of idiopathic PF (IPF), one of the most common forms of PF, remains unknown. Experimental models have been developed to study the mechanisms of PF, and the murine bleomycin (BLM) model has received the most attention. Epithelial injury, inflammation, epithelial–mesenchymal transition (EMT), myofibroblast activation, and repeated tissue injury are important initiators of fibrosis. In this review, we examined the common mechanisms of lung wound-healing responses after BLM-induced lung injury as well as the pathogenesis of the most common PF. A three-stage model of wound repair involving injury, inflammation, and repair is outlined. Dysregulation of one or more of these three phases has been reported in many cases of PF. We reviewed the literature investigating PF pathogenesis, and the role of cytokines, chemokines, growth factors, and matrix feeding in an animal model of BLM-induced PF.

1. Introduction

Idiopathic pulmonary fibrosis (IPF) is a type of idiopathic interstitial pneumonia with high mortality and life expectancy of 2–3 years after diagnosis [1]. The main symptoms of IPF include dyspnea, hypoxemia, prominent lung infiltrates on radiographs, and accumulation of fibroblasts in lung tissue. These symptoms result from the scarring of lung tissue caused by IPF [2].
Although the mechanisms of fibrosis in IPF are not yet fully understood, a general phenomenon in its pathogenesis is a genetic predisposition to subclinical recruitment injuries into the alveolar epithelium, which is followed by pulmonary failure of cyst re-epithelialization and repair [3]. Activated cells within the alveoli release large amounts of cytokines and growth factors that promote pulmonary fibroblast recruitment, proliferation, and differentiation into myofibroblasts, resulting in excessive collagen deposition and progression of the lung parenchyma, causing scars and irreversible loss of function [4,5]. The latest guidelines for treating IPF recommend the use of pirfenidone and nintedanib: two compounds with pleiotropic mechanisms of action [6,7,8]. Unfortunately, both agents have limited efficacy in preventing disease progression and improving the quality of life, and they are associated with tolerance issues [4,5,9,10]. Lung transplantation is the only treatment option for patients with IPF; however, owing to age and comorbidities, it is viable only for a few patients [11]. Therefore, there is an urgent need to develop new therapeutic agents to treat PF.
Although not a single animal model of PF emulates all the features of the human PF, some present several symptoms of IPF. Among the various animal models of PF (BLM, fluorescein isothiocyanate (FITC), silica, radiation, etc.), the BLM model is the most widely used and best-characterized mouse model [12,13]. The BLM-induced PF model is advantageous because it can be easily induced in a short period with high reproducibility. However, a significant drawback of the BLM model is the self-limiting nature of the fibrosis, which is in contrast with the typical progressive chronic fibrosis observed in human IPF [12]. Despite its limitations, the BLM model is appropriate to elucidate the PF pathophysiology and for preclinical research owing to its several advantages and drug discovery capacity.
BLM are a family of complex glycopeptides with antitumor activity; they were originally isolated from the actinomycete Streptomyces verticillus [14,15]. Clinically, it is primarily used as an antitumor antibiotic against various cancers, including lymphomas [16,17,18]. BLM toxicity occurs primarily in the lungs, skin, and mucosal organs, and PF is a well-known side effect of BLM [16,19]. This pronounced undesirable side effect of BLM has prompted the studies to understand the mechanisms of PF associated with IPF in animal models.
Fibrogenesis is considered a defective and persistent wound-healing/connective tissue repair in response to repeated alveolar microinjuries. A prominent feature of this fibrotic repair process is the excessive deposition of extracellular matrix (ECM) components, such as hyaluronic acid, fibronectin, and interstitial collagen, that irreversibly restructures the lung tissue, thickening alveolar and bronchial walls and compromising gas exchange [20,21]. In wound healing, fibroblasts are the key cells responsible for ECM synthesis and deposition, as they provide the initial scaffolding for tissue regeneration [20,21]. When abnormal wound healing and fibrosis occur, fibroblasts respond by overproliferating at the site of injury, acquiring a profibrotic phenotype that is resistant to apoptosis and differentiates into contractile myofibroblasts that perpetuate the fibrotic process [20,21]. Activated fibroblasts/myofibroblasts are highly responsive to growth factors and cytokines.

2. Animal Models

BLM-treated animals have been reported to have histological features, such as the loss of endobronchial buds, collagen walls, and alveolar space, similar to those seen in patients with IPF [22]. This observation suggests that BLM recapitulates the typical features of human disease, popularizing the use of this model (Table 1). In addition, BLM models are advantageous because they are easy to handle, widely available, reproducible, and meet important criteria for a good animal model [23]. Different fibrotic patterns occur according to the route of administration, and consistent dosages have been established across species to achieve a fibrotic response. Intratracheal administration, the standard route of administration, accentuates central bronchial fibrosis, whereas intravenous or intraperitoneal administration produces subpleural scarring similar to that in human disease [24]. BLM models have significantly contributed to our understanding of the pathophysiological roles of cytokines, growth factors, and signaling pathways involved in PF. For example, transforming growth factor (TGF)-β has been identified as a key factor in the pathogenesis of PF [25].
However, the BLM model has significant limitations when it comes to elucidating the progressive nature of human IPF, despite the similarities in histological changes and their cause. BLM resembles acute lung injury in some respects, as it causes an inflammatory response triggered by the overproduction of free radicals, induction of proinflammatory cytokines, and activation of macrophages and neutrophils. Moreover, the subsequent progression of fibrosis is partially reversible [100]. Therefore, one of the most important features of human IPF is its absence in animal models, which should be considered when using the BLM model.
Several research groups have attempted to model IPF using different species, fibrotic disorders, routes, and methods of administration to induce a fibrotic response in animal lungs. However, there is little consistency across laboratories regarding the optimal protocols for these preclinical models. The BLM model is well characterized, clinically relevant, and allows multiple routes of administration to induce fibrosis. The disadvantage is that the disease may be self-limiting in mice. In the BLM model, fibrosis has been shown to recede spontaneously beyond 28 days [13,23], contrary to the majority of patients, where pulmonary fibrosis does not resolve. Therefore, the use of this model is limited to assess the prophylactic efficacy of potential antifibrotic compounds. Another variable introduced when investigating the potential of a novel antifibrotic compound in an animal model of pulmonary fibrosis is the dosing regimen used for the drug. The administration of compounds that are tested prophylactically is initiated before or on the same day that fibrosis is first induced. On the other hand, when tested therapeutically, this is usually initiated after fibrosis has been established. The current state of preclinical testing can be improved by addressing issues such as the time course of treatment, animal size and characteristics, clinically relevant treatment endpoints, and reproducibility of treatment results. While it is acknowledged that the natural occurrence of pulmonary fibrosis in domestic animals can provide valuable information, rodents are the most essential model for studying the pathogenesis of diseases and to evaluate preclinical treatments. Many traditional and newly developed experimental models provide valuable insights into pathogenesis and help to identify novel therapeutic targets that could be evaluated and validated in clinical trials [23,101,102]. As the role of animal models is to replicate specific aspects of disease, they must be carefully selected, designed, and implemented to bridge the translational gap between benchside and bedside. Currently, the BLM model of pulmonary fibrosis is the cheapest, easiest, fastest, most reproducible, and the most used animal model of IPF. This allows overcoming the poor reproducibility of human disease (Table 2). So far, valuable insights into the pathogenesis, prognosis, and treatment of IPF have been obtained with this model.

3. Mechanisms of Wound Healing and Fibrosis

The wound-healing response is often characterized by three distinct phases: injury, inflammation, and repair (Figure 1). Although not all PF cases fit this simple paradigm, it is a useful model for elucidating the general and particular mechanisms of PF.

3.1. Phase I, Injury

Injuries from autoimmune and allergic reactions, environmental particulates, infections, and mechanical injuries often disrupt the normal tissue architecture and initiate healing responses. Inflammation after injury also causes cell damage and tissue destruction. Damaged epithelial and endothelial cells must be replaced to maintain barrier function and integrity, respectively, and to prevent blood loss. Acute injury of endothelial cells triggers the release of inflammatory mediators and the initiation of the antifibrinolytic clotting cascade, temporarily blocking injured vessels with platelets and fibrin-rich thrombi [103]. The platelet differentiation factor, X-box binding protein-1, has a higher level of expression in the lung homogenates, epithelial cells, and bronchoalveolar lavage fluid (BALF) from patients with IPF expressed than in those from patients with chronic obstructive pulmonary disease (COPD) patients and controls [104,105], suggesting that clot-forming responses are continuously activated. Thrombin, a serine protease required to convert fibrinogen to fibrin, is also readily detected in the lung and alveolar lumen of some patients with PF and is associated with the activation of the coagulation pathway [106,107,108]. In addition, thrombin directly activates fibroblasts and promotes their proliferation and differentiation into collagen-producing myofibroblasts [109,110]. Damage to the airway epithelium, particularly alveolar lung cells, triggers a similar anti-fibrinolytic cascade, leading to interstitial edema, areas of acute inflammation, and detachment of the epithelium from the basement membrane [108,111].

Matrix Metalloproteases (MMPs)

There is increasing evidence that the majority of the MMPs are involved in the pathophysiological mechanisms of IPF. MMP-1, MMP-2, MMP-3, MMP-7, MMP-8, MMP-9, MMP-10, MMP-11, MMP-12, MMP-13, MMP-19, MMP-28, and tissue inhibitors of metalloproteases (TIMPs) exhibit roles in the progression and suppression of IPF [112,113,114]. MMP activity is regulated at multiple levels, including gene transcription, extracellular activation of zymogens, and inactivation by specific TIMPs [115]. Accumulating evidence suggests that an imbalance between MMPs and TIMPs can alter ECM metabolism in diverse lung diseases, including IPF [116,117,118]. Two gelatinases, MMP-2 and MMP-9, are of particular interest because they can degrade type IV collagen and gelatin, which are major components of basement membranes. Cinetto et al. reported that the levels of MMP-2, MMP-9, TIMP-1, and TIMP-2 were increased in the BALF of BLM-treated mice and that glycogen synthase 3 (GSK-3) inhibition may modulate MMP-2, MMP-9, TIMP-1, and TIMP-2 activity in BALF and lung tissue, thus preventing BLM-induced lung injury [93].
Platelet recruitment, degranulation, and thrombus formation rapidly progress to a phase of vasodilation phase with increased permeability, allowing the direct recruitment of leukocytes to the site of extravasation and injury [119]. The basement membrane, which forms the ECM underlying the epithelium and endothelium of the parenchymal tissue, precludes direct access to the damaged tissue. To break down this physical barrier, MMPs cleave one or more ECM components, allowing the cellular access to and from the site of injury. Specifically, MMP-2 and MMP-9 cleave two key components of the basement membrane, namely, N-type collagen and gelatin [120,121,122]. Several studies have reported an upregulation of MMP-2 and MMP-9 in fibrotic diseases, highlighting its role in the processes of tissue destruction and regeneration [123,124,125,126].
Kim et al. reported that MMP-2 and MMP-9 activities in BALF and lung parenchyma initially increased rapidly at the beginning, peaked on day 4, and then continuously decreased [127]. This is in agreement with previous studies that showed an initial increase in MMP-2 and MMP-9 activity followed by a decrease [125,128]. Both MMPs are predominantly expressed in inflammatory cells, more specifically MMP-2 in alveolar macrophages and MMP-9 in neutrophils [127]. Previous studies have revealed that these MMPs co-localize with type IV collagen on the basement membrane, suggesting that the early stages of MMP-2 and MMP-9 secretion by macrophages and neutrophils may play an important role in degrading the basement membrane and promoting inflammatory cell migration [116].
MMP gene-targeted studies in mice have indicated that many MMPs regulate processes involved in the pathogenesis of IPF [113]. MMP-9 levels are increased in IPF BALF samples and they are localized in alveolar and interstitial macrophages, metaplastic airway epithelial cells, and neutrophils in IPF lungs [129,130,131]. Membrane-bound MMP-9 on tumor cells activates latent TGB-β1, but the treatment of wild-type (WT) and Mmp9−/− mice with BLM results in similar lung collagen levels [30,132]. However, MMP-9 activity to regulate the lung fibrosis response is cell-specific; MMP-9 did not affect the lung fibrosis response in transgenic mice overexpressing profibrotic interleukin (IL)-13 in airway club cells [133]. In addition, BLM-treated transgenic mice overexpressing human MMP-9 in macrophages have been reported to experience less severe PF, which is preceded by a significant decrease in the number of neutrophils and lymphocytes in BAL samples and a decrease in TIMP-1 levels in BALF samples [31]. The overexpression of MMP-9 in alveolar macrophages limits the severity of PF induced by BLM in mice by cleaving the insulin-like growth factor (IGF)-binding protein-3 (IGFBP-3), which is a carrier protein that exerts its function through IGFs. However, IGFBP-3 also has IGF-independent effects mediated by binding to TGF-β receptors, and IGFBP-3 is closely related to IPF pathogenesis [134].
Owing to the complexity of MMP expression and regulation, a great deal remains unknown about MMP activity in the PF response to injury. MMPs are expressed in the cells of different tissues and may have beneficial or detrimental effects on different organs. Presently, there are no MMP inhibitors in clinical use for the treating IPF; however, with further basic research, MMP inhibitors can potentially be novel therapeutic agents for IPF.

3.2. Phase II, Inflammation

IPF is generally recognized as an inflammatory disease. Damage to the lung tissue causes surrounding tissue inflammation. Inflammatory cells gather and immune cells participate in the tissue repair process. Several chemical factors secreted by these inflammatory cells activate fibroblasts, causing an excessive deposition of ECM components. IPF is also considered to be an epithelium-driven disease [135]. Normal lungs can repair microdamaged type II alveolar epithelial cells [136]. However, when lung microinjury persists, abnormal wound healing activates alveolar epithelial cells, resulting in excessive secretion of mediators, such as TGF-β, platelet-derived growth factor (PDGF), tumor necrosis factor (TNF)-α, angiotensin, and chemokines [137,138]. These mediators promote fibroblast proliferation, migration, and differentiation into myofibroblasts that are resistant to cell apoptosis and secrete ECM components, such as collagen [139,140].
Neutrophils, eosinophils, lymphocytes, and macrophages are observed at the sites of acute injury, while phagocytic cells clear cell debris and necrotic areas. The impact of specific inflammatory cells on downstream fibrosis, particularly IPF, remains unknown [141,142,143,144]. One hypothesis stems from the observation that anti-inflammatory agents are largely ineffective in treating patients with IPF and interstitial pneumonia [145,146,147]. However, the timing of inflammatory events may determine the role that inflammatory processes play. Early inflammation, which decreases in the later stages of the disease, may promote wound healing and contribute to fibrosis. As an example, early recruitment of eosinophils, neutrophils, lymphocytes, and macrophages that produce inflammatory cytokines and chemokines may contribute to local TGF-β and IL-13 upregulation [148,149,150,151]. However, after the initial injury and inflammatory response, the late replenishment of inflammatory cells assists phagocytosis, removes cell debris, and controls excessive cell proliferation, which collectively may contribute to normal healing. Thus, delayed inflammation may play an antifibrotic role and could be necessary for terminating the wound-healing response. A late inflammatory profile rich in phagocytic macrophages that help eliminate fibroblasts and regulatory T cells that secrete IL-10 may suppress local chemokine and TGF-β production and prevent excessive fibroblast activation [152,153]. However, it should be recognized that the mechanisms leading to PF are diverse and that numerous genetic, environmental, and immunological interactions regulate the overall process.

3.2.1. Cytokines

The inflammatory response has profound effects on tissue-resident and other inflammatory cells. Inflammatory cells further propagate inflammation by secreting cytokines, chemokines, and growth factors. Many cytokines participate in wound healing and fibrotic responses, and cytokines such as interleukin (IL)-1α, IL-1β, TNF-α, TGF-β, and PDGF are frequently observed in patients with IPF [154]. Of the many cytokines involved in PF, IL-4, IL-13, and TGF-β are of particular interest. These cytokines act through the mobilization, activation, and proliferation of fibroblasts, macrophages, and myofibroblasts, respectively, and they can exhibit significant profibrotic activity [155,156,157,158,159].
IL-1β is considered to play a central role in PF as it can stimulate collagen expression in vitro in a dose-dependent manner [160]. The administration of recombinant mouse IL-1β to wild-type (WT) mice induces severe tissue destruction, with increased inflammation and collagen deposition [36]. Several studies have revealed that uric acid and ATP released from BLM-injured lung cells are the major endogenous danger signals that increase mitochondrial reactive oxygen species (ROS) generation and activate the NALP3 inflammasome, which is characterized by procaspase-1 maturation, leading to IL-1β production and lung fibrosis [94,161,162,163,164,165]. The administration of the caspase-1 inhibitor z-YVAD-fmk or caspase-1 in mice suppresses BLM-induced IL-1β production, lung inflammation, and fibrosis [94]. In addition, IL-1β activates the IL-1R/MyD88 complex in tissue resident cells, particularly in lung epithelial cells, and activates NF-κB. This leads to inflammation, including through neutrophil and lymphocyte recruitment and fibroblast activation [36]. Fluorenidone (FD), a small pyridine drug, suppresses experimental lung inflammation, fibrosis, and protein expression of IL-1β in the BALF of mice [95]. Furthermore, FD inhibits BLM-induced lung inflammation and fibrosis in mice via the activation of the NALP3 inflammasome and inhibition of the IL-1β/IL-1R1/ MyD88/NF-κB pathway [96].
Inflammation associated with BLM lung injury increased the production of the anti-inflammatory cytokine IL-10 by lung macrophages [38], while BLM-treated IL-10−/− mice displayed increased lung inflammation, as reflected by increased lymphocytes and granulocytes in the BALF. However, there was no significant difference in the severity of lung fibrosis between IL-10−/− and WT mice [38], suggesting an immunomodulatory role of IL-10 in the inflammatory response but not in BLM-induced lung fibrosis.
IL-12 participates in the differentiation of naive T cells into Th1 cells [166]. Intratracheally treated Il12p40−/− mice with BLM exhibited reduced lung mononuclear cell infiltration but increased lung fibrosis compared with WT controls [39]. Furthermore, the expression of CXCL10, CCL5, and CCL11 was lower in Il12p40−/− mice than in the controls. Thus, IL-12 may promote lymphoid tissue response to BLM and suppress PF development.
The Th2 cytokine IL-13 was observed at a high concentration in the blood and BALF of patients with IPF and correlated with disease severity [167]. IL-13 was found to be directly linked to the pathogenesis of PF, and the blocking of IL-13 or deletion of it in the germ line reduced collagen deposition after BLM exposure [168,169,170]. Nie et al. used Akt1−/− mice to modulate BLM-induced lung fibrosis by upregulating the production of the profibrotic cytokine IL-13 in macrophages [97].
Previous studies have shown that IL-17 is associated with profibrotic effects, such as EMT and collagen production, through its interaction with TGF-β signaling [171,172]. The B-cell activating factor is increased in the BALF of patients with IPF; it promotes IL-17 release from Th17 cells and participates in lung fibrosis caused by BLM [98]. IL-27 suppresses PF by inhibiting IL-17 secretion and the JAK/STAT and TGF-β/Smad signaling pathways [41]. In addition, IL-17 production by γδT cells in response to epithelial cell injury is mediated by IL-23 in PF [42].
IL-24 is a member of the IL-20 subfamily of cytokines and is widely expressed in tissues, such as the skin, lungs, and reproductive organs [173]. Current research on IL-24 is primarily focused on cutaneous wound healing and tumor suppression [174,175,176,177]. In Il24−/− mice, Rao et al. observed protection from BLM-induced lung injury and fibrosis, attenuated TGF-β1 production, and reduced M2 macrophage infiltration [40].
IL-33 is a member of the IL-1 family and functions as an inflammatory alamine when released after stress or cell death [178,179]. Fanny et al. investigated the role of IL-33 in a model of BLM-induced inflammation and fibrosis using mice deficient in the IL-33 receptor (chain suppression of tumorigenicity 2, ST2). The deficiency of ST2 causes acutely exacerbated lung inflammation, and the influx of neutrophils was associated with the increased expression of the chemokine CXCL1 in the lungs. In contrast, M2 macrophages were reduced in the lungs of ST2−/− mice, and lung fibrosis was decreased [43]. These results suggest that acute neutrophilic pneumonia can cause PF associated with the production of M2 polarization in an IL-33-ST2-dependent manner [180].

3.2.2. Chemokines

Experiments in humans and rodents have revealed that CC chemokines play an important role in IPF pathogenesis by recruiting and activating mononuclear cells. CCL2 and CCL3 have been reported to promote macrophage mobilization to the lungs, while CCL17 and CCL22 participate in Th2 cell influx [154]. Furthermore, the CCL3–CCR5 axis has also been proposed to play an important role in regulating the mobilization of fibrocytes and monocytes to fibrotic sites in the lungs [44].
The CCL2-CCR2 axis is a key regulator of monocyte trafficking and plays an important role in inflammatory diseases, including PF [181,182,183]. CCL2 concentration is higher in the BALF and serum of patients with IPF [184]. A study on PF in mice reported that ECM deposition was attenuated in Ccr2−/− mice and was associated with less macrophage infiltration and macrophage-derived MMP-2 and MMP-9 production [45]. Proteinase-activated receptor-1 (PAR1), a major thrombin receptor, increased CCL2 release. PAR1−/− mice were protected from BLM-induced lung inflammation and fibrosis, and this protection was associated with a significant attenuation of CCL2 induction [99]. In contrast, CCR2+ cells may be a potential marker of inflammation during fibrosis [185,186]. Moreover, the mobilization of Ly6ChighCCR2+ inflammatory monocytes from the bone marrow to the lungs in fibrotic lungs is signaled by the elevated CCL2 levels [61]. Importantly, Ccr2−/− mice show remarkably reduced lung fibrosis in the BLM model [46,47].
BLM treatment enhances the expression of various chemokines in the lungs, including CCL3, in the lungs [49,183,187,188]. In addition, specific receptors for CCL3 and CC chemokine receptors CCR1 and CCR5 are expressed in fibrocytes [189,190]. The passive immunization of BLM-loaded mice with anti-CCL3 antibodies suppresses fibrotic changes [49], suggesting that CCL3 is a mediator of fibrosis induced by BLM. Compared with WT mice, collagen accumulation was suppressed in Ccl3−/− and Ccr5−/− mice but not in Ccr1−/− mice, suggesting that locally produced CCL3 interacts with CCR5 and participates in the mobilization of bone marrow-derived macrophages and fibrocytes, the main producers of TGF-β1 by BLM, and in the subsequent development of PF [44]. In humans, the 32 bp deletion allele in the Ccr5 gene is more common in Caucasians with a frequency of 0.092 [191,192,193]. As this deletion causes a frameshift and results in a premature stop codon, homozygotes do not express a functional CCR5 protein. Thus, susceptibility to BLM may be determined using CCR5 polymorphisms. Therefore, typing human CCR5 polymorphisms could help to individualize BLM treatment for malignancies.
CCL18 is constitutively expressed at high levels in the lungs and is selectively chemotactic to T cells [194,195,196,197]. The synergistic effects of CCL18 overexpression and BLM injury increased the inflammation in the lungs, especially T-cell infiltration, TNF-α, IFN-γ, MMP-2, and MMP-9 [50]. Despite this synergistic effects on inflammation, CCL18 overexpression attenuated BLM-induced collagen accumulation. CCL18-stimulated T lymphocyte infiltration is profibrotic in healthy lungs, whereas in the inflammatory profibrotic lung setting, it is partially protective against lung fibrosis [50].
Activation of the receptor CXCR2 receptor by CXCL1 and CXCL2 is considered to participate in angiogenesis and the recruitment of neutrophils to inflammatory sites [198,199,200,201,202]. Although neutrophils play an essential role in the host defense against pathogens [203], inflammation and angiogenesis by neutrophils may also contribute to the etiology of various chronic syndromes, including PF [204,205]. DF2162, a CXCR2 antagonist, inhibits IL-8-induced angiogenesis and endothelial cell activation [51].
CXCL6 is enhanced in the BALF of IPF patients. Neutralization with anti-CXCL6 antibody reduced pulmonary neutrophil infiltration, IL-1β, CXCL1, TIMP-1 expression, and collagen deposition in BLM-treated mice [52].
Reportedly, the CXCL12–CXCR4 axis participates in BLM-induced PF. Neutralizing CXCL12 suppresses fibrocyte mobilization and lung collagen deposition [53]. Bone marrow-derived lung CXCR4+ cells migrate in response to CXCL12 and differentiate into collagen-producing lung fibroblasts [56]. Similarly, CXCR4 antagonists alleviated BLM-induced PF [54,55].
CX3CL1, a member of the CX3C chemokine family, can bind to its specific receptor, CX3CR1 [206,207]. The lack of CX3CR1 results in reduced macrophage mobilization, fibroblast accumulation, and impaired skin wound healing [208]. The interaction between CX3CL1 and CX3CR1 may contribute to the pathogenesis of lung diseases such as asthma and emphysema [209,210]. Ishida et al. showed that the CX3CL1–CX3CR1 axis regulates fibrocytes and M2 macrophages that can exhibit profibrotic activity and is essential for the development of PF by BLM [57].

3.2.3. Macrophages

Macrophages play a central role in lung homeostasis and organizing the immune response after injury [58,211,212,213], making them a target for future therapeutic agents. Macrophage abundance in tissues increases after animals are exposed to various lung toxins, including BLM [214,215,216], suggesting that macrophages play a role in acute lung injury. Furthermore, macrophages that accumulate in the lungs early in the course of lung toxic injury are activated to an inflammatory M1 phenotype through the expression of inducible nitric oxide synthase (iNOS) and TNF-α [215,217,218,219]. In the BLM-induced PF model, inflammatory macrophages increased their concentration immediately, peaked on day 3, and gradually decreased until day 21. M1 macrophages are the most abundant in BALF, but after BLM exposure, M2 macrophages gradually increased their concentration, reaching maximum levels on day 14, and correlated with collagen deposition [58,59].
Direct evidence of the role of inflammatory/cytotoxic M1 macrophages in acute lung injury comes from the finding that tissue damage is directly correlated with macrophage functional status. In several experimental models using different approaches, such as pharmacological, genetic, or using macrophage-deficient mice, the suppression or depletion of macrophages improved or prevented lung damage. Blocking the cytotoxic/inflammatory activity of M1 macrophages with anti-inflammatory steroids reduces BLM-induced lung injury [62] and suppresses BLM-induced acute lung injury and iNOS expression in Ccr4−/− mice that cannot generate M1 macrophages [48]. CCR2 knockout mice deficient in M1 macrophage migration to injury sites are also protected from BLM-induced oxidative stress and tissue injury [47]. Furthermore, the activation of macrophages exacerbates the tissue damage induced by lung toxins, reinforcing the participation of inflammatory M1 macrophages in the pathogenesis of acute lung injury. The pretreatment of rodents with macrophage activators such as lipopolysaccharides (LPS) or Bacillus Calmette–Guerin enhances the acute lung injury induced by endotoxins, radiation, BLM, and ozone [215]. A similar increase in toxicity was observed in response to BLM in mice lacking surfactant protein D, which is a lung collectin that suppresses the inflammatory activity of lung macrophages under homeostasis [63].
The resolution of acute injury and restoration of normal lung structure and function are precisely regulated processes. There is evidence that M2 macrophages stimulate a counter-regulatory mechanism that inhibits the release of inflammatory mediators and activates tissue repair. The levels of anti-inflammatory/wound-repair M2 macrophages are higher in the lungs after exposure to BLM [215]. The emergence of M2 macrophages compared with that of M1 macrophages is delayed, which is consistent with the role of these cells in tissue repair [220]. M2 macrophages promote fibrosis by secreting multiple cytokines that enhance the fibroblast-to-myofibroblast transition and epithelial cell wound healing. Targeting M2 macrophages to suppress the fibrosis-promoting phenotype is a potential therapeutic approach for PF. As discussed previously (in the inflammation section), locally produced CX3CL1 can promote the development of BLM-induced PF, primarily by attracting CX3CR1+ M2 macrophages and fibrocytes to the lungs [57]. Li et al. reported that clevudine, approved for treating patients with hepatitis B virus (HBV), can inhibit the profibrotic phenotype (such as CD206 and arginase (Arg)) and promote the antifibrotic phenotype (such as CD86 and IL-10) of M2 macrophages by inhibiting the PI3K/Akt signaling pathway [64]. This effect further reduces myofibroblast activation. EMT by M2 decreases collagen deposition and the secretion of fibrosis-promoting cytokines and restores pulmonary function. Furthermore, RP-832c, which specifically targets CD206 receptors on M2 macrophages, significantly suppresses BLM-induced fibrosis in mouse lungs and downregulates TGF-β1 and α-SMA expression [65]. M2 macrophages produce fibroblast-activating factors and cytokines, such as TGF-β, IL-10, and Arg, which promote myofibroblast proliferation and collagen secretion and are involved in the progression of fibrosis [40,66,67]. Thus, the M2 macrophage marker CD206 may be a potential target for PF treatment. Furthermore, JAK/STAT3 signaling plays a central role in M2 macrophage polarization, suggesting that the JAK/STAT3 signaling pathway participates in tissue fibrosis [68,69].
EMT is a process by which epithelial cells lose polarity, intercellular adhesion, and transition to mesenchymal cells [221]. In patients with PF, the EMT of alveolar epithelial cells results in myofibroblasts [222]. TGF-β activates the Smad signaling pathway and plays an important role in alveolar epithelial cell EMT [223,224]. Zhu et al. reported that M2 macrophages are mobilized in the lungs and that the TGF-β–Smad2 signaling pathway is activated in BLM-induced PF in mice [60]. Furthermore, M2 macrophages secrete large amounts of TGF-β and can induce EMT in lung epithelial cells via the TGF-β–Smad2 signaling pathway in vitro [60].
Macrophage migration inhibitory factor (MIF) inhibition attenuated lung injury and ECM deposition in rats with BLM-induced pulmonary fibrosis [225]. The inhibition of MIF is associated with the reduction in macrophage infiltration in lungs. Furthermore, MIF deficiency suppressed myofibroblast accumulation and fibrosis via the TGF-β1/Smads signaling pathway. Thus, the specific inhibition of MIF is expected to contribute to the treatment of PF.

3.2.4. Dendritic Cells

Dendritic cells (DCs) are potent antigen-presenting cells that are abundant in human fibrotic interstitial lung disease [226,227,228]. However, the role of DCs in the pathogenesis of PF remains unknown. Mature DCs and activated memory T cells accumulate in the lungs of BLM-induced PF mice, while the pharmacological inactivation of DCs attenuates the features of BLM-induced lung injury, suggesting that DCs may maintain lung inflammation and fibrosis in the BLM model fibrosis in BLM models [70]. DCs are composed of heterogeneous populations including myeloid DCs (mDCs), tissue-resident DCs, and plasmacytoid DCs (pDCs). These DC subsets may play distinct roles in different stages of fibrosis. pDC depletion reduced the expression of genes and proteins involved in lung fibrosis, inflammation, and DC differentiation in BLM-treated mice [71].

3.2.5. T Cells

Emerging evidence suggests a regulatory role of Th1/Th2 imbalance in the inflammatory phase of PF [21,229]. Systemic T cell depletion by anti-CD3 antibody attenuates ECM accumulation in a mouse model of BLM-induced PF [72].
Among T lymphocytes, CD4+CD25+FoxP3+ regulatory T cells (Treg) play a pivotal role in maintaining immune system homeostasis, and Treg activity is significantly reduced in patients with IPF [230]; however, conflicting results have been reported regarding the role of Tregs in PF [229,231]. Kotsianidis et al. [230] reported a protective role for Tregs using antibody-based depletion in the fibrosis phase of BLM-induced mouse PF [232]. Furthermore, Kamio et al. demonstrated that splenocytes potently ameliorated PF in vivo using a BLM-induced mouse model of PF [73]. This effect was abolished by the antibody-mediated depletion of Tregs [73].

3.3. Phase III, Tissue Repair

Myofibroblast-derived collagen and α-SMA form fibrin scaffolds and an extracellular matrix [233,234]. These structures are collectively referred to as granulation tissue; primary fibroblasts and alveolar macrophages isolated from patients with IPF produce significantly more fibronectin and α-SMA than controls, indicating a state of increased fibroblast activation in patients with IPF [74,235]. Mice lacking the type III domain of fibronectin exhibit significantly less lung fibrosis after BLM treatment than WT mice, suggesting that fibronectin is required for PF development [74,76].
In addition to fibronectin, the ECM is composed of glycoproteins such as PDGF and glycosaminoglycans such as hyaluronic acid, proteoglycans, and elastin [236,237,238,239,240]. Fibroblasts activated by growth factors migrate along the ECM network to repair wounds. Within skin wounds, TGF-β also induces a contractile response and regulates collagen fiber orientation [241]. The differentiation of fibroblasts into myofibroblasts leads to the neoexpression of stress fibers and α-SMA, both of which confer myofibroblasts high contractile activity [242,243]. The extent of ECM bedding and the number of activated myofibroblasts determine the amount of collagen deposited [244]. Thus, the balance of MMPs and TIMPs as well as that of collagen and collagenase shifts from promoting collagen synthesis and increasing collagen deposition to control collagen concentration and prevent it from increasing [117,245,246]. For successful wound healing, this balance is achieved when fibroblasts undergo apoptosis, inflammation begins to end, granulation tissue recedes, and collagen-rich lesions remain. The removal of inflammatory cells, especially α-SMA+ myofibroblasts, is essential to end collagen deposition [247]. The signals that initiate fibroblast apoptosis in IPF are poorly understood, although several factors could participate, including cytokine imbalance, genetic causes, and constitutive antiapoptotic pathways similar to those of some cancer cells [118,244,248,249].

3.3.1. Growth Factors

TGF-β is probably the most well-studied growth factor in fibrosis and is regarded as a prototypic mediator of fibrosis [250]. Of the three isoforms, TGF-β1 has been reported to be primarily involved in PF [251]. It increases the transcription of downstream target genes such as procollagen I and III via the transmembrane receptor serine/threonine kinase and cytosolic Smad2/3 signaling pathways [252]. In particular, the loss of Smad3 ameliorates BLM-induced PF [25]. Secreted modular calcium-binding protein 2 (SMOC2), a factor expressed in almost all tissues, was found to inhibit lung fibrosis progression in SMOC2−/−. This was indicated by the decreased levels of TGF-β1, α-SMA, p-Smad2, and p-Smad3 in lung tissue samples [253]. Kurarinone, a prenylated flavonoid isolated from Sophora Flavescens, suppresses TGF-β-induced EMT in lung epithelial cells and BLM-induced PF by inhibiting the phosphorylation of Smad2/3 and AKT signaling [77]. Anemarrhenae rhizoma is a traditional Chinese herbal medicine. Shen et al. explored the therapeutic effects of the total extract of anemarrhenae rhizoma (TEAR) on BLM-induced pulmonary fibrosis in rat [254]. TEAR protected rats from fibrosis in a dose-dependent manner, and the antifibrotic activity of TEAR was associated with modulation of the TGF-β1/Smad signaling pathway. In addition, the activation of TGF-β1 production and signaling are considered the cornerstone in the EMT process [255]. Ticagrelor, an antiplatelet, inhibited TGF-β1 production and suppressed Smad3 activation and the signaling pathway in BLM-treated rats [256]. Moreover, ticagrelor suppressed EMT, which was indicated by the increased expression of E-cadherin and decreased expression of vimentin and α-SMA.
In addition to TGF-β, PDGF is another potent fibrotic growth factor that promotes lung fibrosis via fibroblast activation [257]. The expression of PDGF is increased in epithelial cells and macrophages in the lungs of patients with IPF [258]. Imatinib, a PDGF tyrosine kinase inhibitor, exhibits strong anti-fibrotic effects via the inhibition of mesenchymal cell proliferation in BLM-induced PF [78]. Nintedanib is also a multiple inhibitor of tyrosine kinase receptors that participate in the development of lung fibrosis, including PDGF receptors alpha and beta, VEGF receptors 1, 2, and 3, and FGF receptors 1, 2, and 3 [259]. Further, nintedanib can inhibit the development of lung fibrosis in the BLM mouse model [79].
The main downstream element of TGF-β signaling [260], connective growth factor (CTGF), is a profibrotic growth factor that stimulates fibroblast proliferation and increases ECM deposition [261]. CTGF has been detected in the lung tissue of patients with IPF in both type II alveolar cells and interstitial fibroblasts [262]. Wang et al. reported that the increased expression of α-SMA by recombinant CTGF was reduced in human embryonic lung fibroblasts (HELF) treated with anti-CTGF single chain variable fragment antibody (scFv). Anti-CTGF scFv significantly reduced the number of inflammatory leukocytes in BALF and the hydroxyproline content in the lung tissue of BLM-treated mice [80].
Insulin-like growth factor (IGF-1) is another mediator that may be participate in the pathogenesis of IPF; IGF-1 regulates cell migration and differentiation, and its levels are elevated in the lungs of patients with IPF [263]. Blockade of the IGF-1 pathway using a monoclonal antibody against the IGF-1 receptor (IGF-1R) reduces BLM-induced lung injury [81]. Among the IGF-1R-mediated signaling pathways, the PI3K–AKT–mTOR pathway is fundamental in cell growth and survival [264]. Metformin is one of the most common oral biguanide drugs used to lower blood glucose levels in patients with type 2 diabetes. Metformin suppresses collagen deposition, increases fibronectin and α-SMA levels, and increases IGF-1 and PI3K expression in BLM-treated mice [82]. Somatostatin, also known as somatotropin release inhibitory factor (SRIF), is a member of a family of cyclopeptides produced primarily by normal endocrine cells, gastrointestinal cells, immune cells, neurons, and certain tumors. Lung fibroblasts express somatostatin receptors [83,265]. Somatostatin analogs reduced leukocyte counts and IGF-1 levels in BALF and inhibited PF in mouse BLM [84].

3.3.2. Mesenchymal Stem Cells (MSCs)

Stem cells, particularly MSCs, have several advantages such as migrating into damaged lung tissue, modulation of endothelial and epithelial cell permeability by secreting various paracrine factors, suppressing inflammatory responses, promoting tissue repair, and inhibiting bacterial growth [266]. MSCs are pluripotent stem cells that can differentiate into a variety of cell types, including bone cells (osteoblasts), cartilage cells (chondrocytes), muscle cells (myocytes), and adipocytes, which give rise to bone marrow adipose tissue (fat cells). In addition, MSCs in the stromal and perivascular areas near fibroblasts can regulate myofibroblast activity [267]. These results indicate that MSCs are effective for the treatment of BLM-induced fibrotic lung lesions in animals. Stem cells modulate B cell activity by inhibiting B cell maturation and mobilization to the site of lung injury in IPF. The modulation of B cell activity leads to the suppression of the chronic inflammatory process with the subsequent formation of fibrotic lesions [85]. In a model of BLM-induced lung injury, MSCs reduced the expression of co-stimulatory proteins in DCs and monocyte-derived macrophages, reduced their ability to induce antigen-specific T cell immune responses, and promoted the production of immune cells such as Tregs and cytokines such as IL-10, which have immune regulatory functions [86].
Gingivally derived MSCs (GMSCs), which are an abundant source of MSCs and can be harvested at a low cost, were isolated and characterized for the first time in 2009 [268]. Intervention with GMSCs has been reported to reduce BLM-induced PF, inflammation, edema, and apoptosis [87]. The administration of BLM notably increased the gene expression of IL-1β, TNF-α, TGF-β, and MMP-9, and decreased IL-10 expression in lung tissue, but these effects were restored by GMSC intervention [87].
The involvement of Toll-like receptor (TLR) signaling and the therapeutic role of MSCs in MyD88−/− mice were analyzed using human placental-derived MSCs (hfPMSCs) [37]. Compared with WT mice, lung injury in MyD88−/− mice treated with BLM reduced inflammation and fibrosis, along with less TGF-β signaling and ECM production. Thus, hfPMSCs may attenuate BLM-induced lung inflammation and fibrosis, partly through the MyD88-mediated attenuation of inflammation in mice.

3.3.3. Fibrocytes

Traditionally, fibroblasts have been considered mesenchymal tissue-derived/endogenous cells. However, recent research has established the concept of circulating bone marrow-derived cells called fibrocytes that migrate into tissues and differentiate into fibroblasts and myofibroblasts [269]. Fibroblasts express myeloid markers, such as CD45 and CD34, chemokine receptor CXCR4, and collagen-I [199,269,270,271,272]. Fibrocytes produce ECM components (collagen I, collagen III, fibronectin, and vimentin), cross-linking enzymes (lysyl oxidase family), cytokines (TNF-α, IL-6, IL-8, and IL-10), chemokines (CCL2, CCL3, CCL4, and CXCL4), growth factors (VEGF, PDGF, GM-CSF, and others), and various MMPs [273,274,275]. Moeller et al. discovered that circulating fibrocytes are increased in patients with IPF and are a prognostic marker and predictor of early death [276]. CXCL12 participates in the mobilization of circulating fibrocytes at the site of lung injury [53]. The neuronal guidance protein slit guidance ligand 2 secreted by fibroblasts suppresses fibrocyte differentiation and BLM-induced lung fibrosis [88].

3.3.4. Fibroblasts and Myofibroblasts

IPF is characterized by fibroblast accumulation, collagen deposition, and ECM remodeling [277,278]. A large amount of myofibroblasts is found in lung tissues from patients with PF, and myofibroblasts are believed to be key effector cells responsible for matrix deposition and structural remodeling [279,280]. Myofibroblasts arise from circulating progenitors, resident fibroblasts, and from alveolar epithelial cells that have undergone EMT, which can be induced by TGF-β [280,281]. In addition, M2 macrophages secrete large amounts of TGF-β. Zhu et al. have demonstrated that M2 macrophages induce EMT through the TGF-β/Smad2 signaling pathway [60].
In humans with IPF and in mouse model of lung fibrosis, adenosine monophosphate (AMP)-activated protein kinase (AMPK) activity is lower in fibrotic regions associated with metabolically active and apoptosis-resistant myofibroblasts [282]. Metformin is a safe and widely used agent for non-insulin-dependent diabetes, and it has therapeutic potential to restore glucose and lipid metabolic homeostasis [283]. Further analysis has revealed a deficient AMPK activation in non-resolving, pathologic fibrotic processes. This supports the role of metformin in reversing the established fibrosis by facilitating deactivation and apoptosis of myofibroblasts [282]. In addition, nintedanib is a potent small molecule inhibitor of the receptor tyrosine kinases PDGF receptor, FGF, and VEGF receptors. Nintedanib has shown consistent antifibrotic activity by reducing fibroblast proliferation, migration and differentiation, and the secretion of ECM, in animal models of lung fibrosis [284]. Therefore, these data provide a strong rationale for the clinical efficacy of nintedanib in patients with IPF.
Previous studies have reported that the intermediate filament protein nestin plays an important role in tissue regeneration and wound healing in various organs [285,286]. Wang et al. found that nestin expression was primarily localized to lung myofibroblasts and increased in a mouse model of pulmonary fibrosis and in patients with IPF [287]. The deficiency of nestin significantly alleviated pulmonary fibrosis in a mouse model. Therefore, targeting nestin expression in lung myofibroblasts may alleviate lung fibrosis and is a promising therapeutic strategy for IPF. On the other hand, tripartite motif-containing 33 (TRIM33) is an E3 ubiquitin ligase known as a negative regulator of TGF-β1 signaling through Smad4 ubiquitination and turnover of TGF-β receptors [288,289]. TRIM33 was overexpressed in alveolar fibroblasts in IPF patients and lungs of BLM-treated mice [290]. In primary lung fibroblasts, Trim33 deficiency increased the expression of genes downstream of TGF-β1. Since TRIM33 is a negative regulator of fibrosis, the potential therapeutic effect of IPF is expected.
In a rat model of BLM-induced PF, baicalin, the biologically active form of the dried roots of Scutellaria baicalensis Georgi, suppressed the severity of PF by reducing the amount of hydroxyproline in lung tissue [291]. Furthermore, BLM promoted fibroblast viability in vivo in a dose-dependent manner, which was limited after treatment with baicalin. These findings suggest that baicalin exerts an inhibitory effect on PF and fibroblast proliferation by BLM.

3.3.5. Induced Pluripotent Stem (iPS) Cells

iPS cells are cells primed by transferring specific pluripotency genes from adult somatic cells, paving the way for new stem cell-based therapies [292]. Zhou et al. demonstrated that by suppressing the TGF-β1-Smad2/3 signaling pathway, EMT, and inflammatory responses, iPS cells have a BLM-induced therapeutic effect in PF mice [89].

3.3.6. Exosomes

Exosomes are extracellular vesicles (EVs), and they are an important point of discussion in the field of EV research. They are nanoscale vesicles secreted by cells under physiological or pathological conditions [293,294,295,296]. The function of exosomes as important mediators of intercellular signaling depends on their loading profiles. Loading profiles include miRNAs, mRNAs, DNA, and proteins that contribute to both the physiological functions and the pathophysiology of various diseases, including PF [293,297,298]. Exosomes mediate intracellular communication between macrophages and fibroblasts and are thought to play an important role in BLM-induced PF. Sun et al. observed that exosomes with macrophage-derived angiotensin II type 1 receptor (AT1R) activate the renin–angiotensin system (RAS) and the TGF-β-Smad2/3 pathway through a shift to the Ang II–AT1R axis, promoting collagen synthesis and mediating BLM-induced lung fibrosis [90]. In addition, a single intravenous administration of purified exosomes derived from human bone marrow mesenchymal stem cells (BM MSC) effectively prevented and restored the core features of BLM-induced PF, improved lung morphology, blunted collagen deposition, and restored lung structure [91].
Exosome-mediated intercellular communication is essential for activating fibrosis-promoting pathways in IPF [299,300]. Similarly, evidence suggests increased exosome synthesis in patients with IPF [92,301]. Martin-Medina et al. discovered more EVs in BALF and BLM-challenged mice with IPF compared to healthy controls. They also reported an increase in WNT5a, a protein associated with fibroblast proliferation and activation, in EVs obtained from primary human fibrotic lung fibroblasts (HFLF) and fibroblasts activated with TGF-β [92].

4. Conclusions and Future Directions

In this review, we discussed the use of BLM models of PF for describing disease states in vivo and identifying novel drug targets, and we also addressed their use in demonstrating the efficacy of potential compounds. However, BLM models may have limited use for the detailed evaluation and assessment of novel agents for clinical use. The degree of inflammation and ECM deposition contribute to the net deposition of collagen, which ultimately determines whether a fibrotic lesion develops. Therapeutic interventions that inhibit fibroblast activation, proliferation, and apoptosis require an understanding of all stages of wound repair. These three stages of wound repair often appear sequentially; however, in chronic or repetitive injury, these processes occur in parallel, placing a heavy burden on regulatory mechanisms. In humans with IPF and mouse BLM models of PF, inflammation is considered to play a role in disease onset and progression. However, the underlying mechanisms remain unclear. One abnormal response propagates into other mechanisms, and the well-controlled healing response gradually transforms into a fibrotic lesion. In PF, the onset and progression of the healing response becomes uncontrolled, and many delicate balances are broken. Despite significant advances in our understanding of these pathways, therapeutic interventions and new treatments for PF are lacking. Because PF has numerous causes, forms, and stages, the heterogeneity of the disease must be considered when evaluating the results from individual mouse models, and most important, when designing and implementing new therapeutic strategies. Lung wound repair is a highly dynamic process in which immunology, structural biology, and physiology participate. Cooperation between these systems is essential for successful repair.
Recently, single-cell RNA sequencing (scRNA-seq) has revealed the potential to overcome the large-scale changes and spatial heterogeneity in cell types and allowed the reliable identification of related cell populations and confirmation of the complex molecular procedures in IPF. For example, scRNA-seq data provided insight into the heterogeneity and plasticity of endothelial cells (ECs) in normal and fibrotic lungs and suggested that the potential cross-talk between ECs, macrophages, and stromal cells contributes to pathologic IPF [302]. Adipose-derived mesenchymal stem cells (ADSCs) have been reported to exert therapeutic effects in PF, but the underlying mechanisms remain unclear. Rahman et al. injected ADSCs intratracheally into BLM-treated mice and performed scRNA-seq with lung-derived cells. ADSC administration dramatically altered the transcriptome profile and composition of lung macrophages and reduced lung inflammation and fibrosis [303]. Moreover, Aran et al. have identified a profibrotic macrophage subpopulation that localizes to sites of fibrotic scar with activating effects on the mesenchyme [304]. Using scRNA-seq techniques with isolated mouse lung fibroblasts, it was observed that although fibroblast numbers did not increase after BLM treatment, different populations of activated fibroblasts could be identified in fibrotic mouse lungs [305]. In addition, to comprehensively classify lung fibroblast populations with a nonbiased approach, Xie et al. performed scRNA-seq on mesenchymal preparations of lungs from intact or BLM-treated mice. Single-cell transcriptome analysis classified and defined six types of mesenchymal cells in normal lungs and seven types in fibrotic lungs [306]. The collection of these single-cell transcriptomes and the clear classification of the various cell subsets will provide a new resource for understanding the various cellular landscapes and their role in fibrotic disease.
Although treatments for pulmonary fibrosis have been developed in recent years, the prognosis is still discouraging, and there is a need to develop new therapeutic agents. When a receptor activator of NF-κB ligand (RANKL) partial peptide, MHP1-AcN, was administrated in the BLM-induced lung fibrosis model, less collagen deposition was observed [307]. Mansour et al. investigated the antifibrotic potential of tadalafil for BLM-induced pulmonary fibrosis. Tadalafil inhibited TGF-β production and collagen deposition [308]. In addition, sodium houttuyfonate (SH) could attenuate BLM-induced lung injury by reducing the inflammation in fibrogenesis [309]. DRDE-30, an analog of amifostine, suppressed BLM-induced increases in BALF TGF-β and pulmonary hydroxyproline levels while decreasing α-SMA expression, suggesting the suppression of EMT [310]. Moreover, irbesartan, an angiotensin receptor blocker (ARB) with peroxisome proliferator-activated receptor (PPAR)γ activity, reduced lung hydroxyproline levels, BALF leukocyte counts, TGF-β1, and CCL2 levels [311]. These studies indicate that these treatment options could ameliorate pulmonary fibrosis caused by BLM, reinforcing their potential use as adjuvants to alleviate the side effects of BLM [312].

Author Contributions

Conceptualization, Y.I. and T.K.; methodology, Y.I.; investigation, Y.K. and N.M.; writing—original draft preparation, Y.I.; writing—review and editing, N.M. and T.K. All authors have read and agreed to the published version of the manuscript.

Funding

This study was financially supported in part by Grants-in-Aid for Scientific Research (B, 20H03957, Y.I.), (C, 21K06966, N.M.) and (B, 22H03366, T.K.) from Japan Society for the Promotion of Science (JSPS).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are deeply grateful to Mariko Kawaguchi for her excellent assistance in the preparation of this manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ley, B.; Collard, H.R.; King, T.E., Jr. Clinical Course and Prediction of Survival in Idiopathic Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2011, 183, 431–440. [Google Scholar] [CrossRef] [PubMed]
  2. Gross, T.J.; Hunninghake, G.W. Idiopathic Pulmonary Fibrosis. N. Engl. J. Med. 2001, 345, 517–525. [Google Scholar] [CrossRef] [PubMed]
  3. Spagnolo, P.; Cottin, V. Genetics of idiopathic pulmonary fibrosis: From mechanistic pathways to personalised medicine. J. Med. Genet. 2016, 54, 93–99. [Google Scholar] [CrossRef] [PubMed]
  4. Bellaye, P.-S.; Kolb, M. Why do patients get idiopathic pulmonary fibrosis? Current concepts in the pathogenesis of pulmonary fibrosis. BMC Med. 2015, 13, 176. [Google Scholar] [CrossRef] [PubMed]
  5. Collard, H.R.; Jones, K.D. Pathogenesis of Idiopathic Pulmonary Fibrosis. Annu. Rev. Pathol. Mech. Dis. 2014, 9, 157–179. [Google Scholar] [CrossRef]
  6. Raghu, G.; Rochwerg, B.; Zhang, Y.; Cuello-Garcia, C.; Azuma, A.; Behr, J.; Brozek, J.L.; Collard, H.R.; Cunningham, W.; Homma, S.; et al. An Official ATS/ERS/JRS/ALAT Clinical Practice Guideline: Treatment of Idiopathic Pulmonary Fibrosis. An Update of the 2011 Clinical Practice Guideline. Am. J. Respir. Crit. Care Med. 2015, 192, e3–e19. [Google Scholar] [CrossRef]
  7. Spagnolo, P.; Maher, T.; Richeldi, L. Idiopathic pulmonary fibrosis: Recent advances on pharmacological therapy. Pharmacol. Ther. 2015, 152, 18–27. [Google Scholar] [CrossRef]
  8. Spagnolo, P.; Tzouvelekis, A.; Bonella, F. The Management of Patients with Idiopathic Pulmonary Fibrosis. Front. Med. 2018, 5, 148. [Google Scholar] [CrossRef]
  9. Galli, J.A.; Pandya, A.; Vega-Olivo, M.; Dass, C.; Zhao, H.; Criner, G.J. Pirfenidone and nintedanib for pulmonary fibrosis in clinical practice: Tolerability and adverse drug reactions. Respirology 2017, 22, 1171–1178. [Google Scholar] [CrossRef]
  10. Bando, M.; Yamauchi, H.; Ogura, T.; Taniguchi, H.; Watanabe, K.; Azuma, A.; Homma, S.; Sugiyama, Y. The Japan Pirfenidone Clinical Study Group Clinical Experience of the Long-term Use of Pirfenidone for Idiopathic Pulmonary Fibrosis. Intern. Med. 2016, 55, 443–448. [Google Scholar] [CrossRef] [Green Version]
  11. George, P.M.; Patterson, C.M.; Reed, A.K.; Thillai, M. Lung transplantation for idiopathic pulmonary fibrosis. Lancet Respir. Med. 2019, 7, 271–282. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, T.; Santos, F.G.D.L.; Phan, S.H. The Bleomycin Model of Pulmonary Fibrosis. Methods Mol. Biol. 2017, 1627, 27–42. [Google Scholar] [CrossRef] [PubMed]
  13. Moore, B.B.; Hogaboam, C.M. Murine models of pulmonary fibrosis. Am. J. Physiol. Cell. Mol. Physiol. 2008, 294, L152–L160. [Google Scholar] [CrossRef] [PubMed]
  14. Umezawa, H.; Ishizuka, M.; Maeda, K.; Takeuchi, T. Studies on bleomycin. Cancer 1967, 20, 891–895. [Google Scholar] [CrossRef]
  15. Umezawa, H.; Maeda, K.; Takeuchi, T.; Okami, Y. New antibiotics, bleomycin A and B. J. Antibiot. 1966, 19, 200–209. [Google Scholar]
  16. Yagoda, A.; Mukherji, B.; Young, C.; Etcubanas, E.; Lamonte, C.; Smith, J.R.; Tan, C.T.C.; Krakoff, I.H. Bleomycin, an Antitumor Antibiotic. Ann. Intern. Med. 1972, 77, 861–870. [Google Scholar] [CrossRef]
  17. Blum, R.H.; Carter, S.K.; Agre, K. A clinical review of bleomycin—A new antineoplastic agent. Cancer 1973, 31, 903–914. [Google Scholar] [CrossRef]
  18. Umezawa, H. Chemistry and mechanism of action of bleomycin. Fed. Proc. 1974, 33, 2296–2302. [Google Scholar]
  19. Rudders, R.A.; Hensley, G.T. Bleomycin Pulmonary Toxicity. Chest 1973, 63, 626–628. [Google Scholar] [CrossRef]
  20. Fernandez, I.E.; Eickelberg, O. New cellular and molecular mechanisms of lung injury and fibrosis in idiopathic pulmonary fibrosis. Lancet 2012, 380, 680–688. [Google Scholar] [CrossRef]
  21. Wynn, T.A. Integrating mechanisms of pulmonary fibrosis. J. Exp. Med. 2011, 208, 1339–1350. [Google Scholar] [CrossRef] [PubMed]
  22. Usuki, J.; Fukuda, Y. Evolution of three patterns of intra-alveolar fibrosis produced by bleomycin in rats. Pathol. Int. 1995, 45, 552–564. [Google Scholar] [CrossRef] [PubMed]
  23. Moeller, A.; Ask, K.; Warburton, D.; Gauldie, J.; Kolb, M. The bleomycin animal model: A useful tool to investigate treatment options for idiopathic pulmonary fibrosis? Int. J. Biochem. Cell Biol. 2008, 40, 362–382. [Google Scholar] [CrossRef] [PubMed]
  24. Chua, F.; Gauldie, J.; Laurent, G.J. Pulmonary Fibrosis: Searching for model answers. Am. J. Respir. Cell Mol. Biol. 2005, 33, 9–13. [Google Scholar] [CrossRef]
  25. Zhao, J.; Shi, W.; Wang, Y.-L.; Chen, H.; Bringas, P., Jr.; Datto, M.B.; Frederick, J.P.; Wang, X.-F.; Warburton, D. Smad3 deficiency attenuates bleomycin-induced pulmonary fibrosis in mice. Am. J. Physiol.-Lung Cell. Mol. Physiol. 2002, 282, L585–L593. [Google Scholar] [CrossRef]
  26. Yamashita, C.M.; Dolgonos, L.; Zemans, R.L.; Young, S.K.; Robertson, J.; Briones, N.; Suzuki, T.; Campbell, M.N.; Gauldie, J.; Radisky, D.C.; et al. Matrix Metalloproteinase 3 Is a Mediator of Pulmonary Fibrosis. Am. J. Pathol. 2011, 179, 1733–1745. [Google Scholar] [CrossRef]
  27. Zuo, F.; Kaminski, N.; Eugui, E.; Allard, J.; Yakhini, Z.; Ben-Dor, A.; Lollini, L.; Morris, D.; Kim, Y.; DeLustro, B.; et al. Gene expression analysis reveals matrilysin as a key regulator of pulmonary fibrosis in mice and humans. Proc. Natl. Acad. Sci. USA 2002, 99, 6292–6297. [Google Scholar] [CrossRef]
  28. Craig, V.J.; Quintero, P.A.; Fyfe, S.E.; Patel, A.S.; Knolle, M.D.; Kobzik, L.; Owen, C.A. Profibrotic Activities for Matrix Metalloproteinase-8 during Bleomycin-Mediated Lung Injury. J. Immunol. 2013, 190, 4283–4296. [Google Scholar] [CrossRef]
  29. García-Prieto, E.; Gonzalez-Lopez, A.; Cabrera, S.; Astudillo, A.; Gutierrez-Fernandez, A.; Fanjul-Fernandez, M.; Batalla-Solís, E.; Suarez-Puente, X.; Fueyo-Silva, A.; López-Otín, C.; et al. Resistance to Bleomycin-Induced Lung Fibrosis in MMP-8 Deficient Mice Is Mediated by Interleukin-10. PLoS ONE 2010, 5, e13242. [Google Scholar] [CrossRef]
  30. Betsuyaku, T.; Fukuda, Y.; Parks, W.C.; Shipley, J.M.; Senior, R.M. Gelatinase B Is Required for Alveolar Bronchiolization after Intratracheal Bleomycin. Am. J. Pathol. 2000, 157, 525–535. [Google Scholar] [CrossRef]
  31. Cabrera, S.; Gaxiola, M.; Arreola, J.L.; Ramírez, R.; Jara, P.; D’Armiento, J.; Richards, T.; Selman, M.; Pardo, A. Overexpression of MMP9 in macrophages attenuates pulmonary fibrosis induced by bleomycin. Int. J. Biochem. Cell Biol. 2007, 39, 2324–2338. [Google Scholar] [CrossRef] [PubMed]
  32. Manoury, B.; Nenan, S.; Guenon, I.; Boichot, E.; Planquois, J.-M.; Bertrand, C.P.; Lagente, V. Macrophage metalloelastase (MMP-12) deficiency does not alter bleomycin-induced pulmonary fibrosis in mice. J. Inflamm. 2006, 3, 2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Nkyimbeng, T.; Ruppert, C.; Shiomi, T.; Dahal, B.; Lang, G.; Seeger, W.; Okada, Y.; D’Armiento, J.; Günther, A. Pivotal Role of Matrix Metalloproteinase 13 in Extracellular Matrix Turnover in Idiopathic Pulmonary Fibrosis. PLoS ONE 2013, 8, e73279. [Google Scholar] [CrossRef] [PubMed]
  34. Yu, G.; Kovkarova-Naumovski, E.; Jara, P.; Parwani, A.; Kass, D.; Ruiz, V.; Lopez-Otín, C.; Rosas, I.O.; Gibson, K.F.; Cabrera, S.; et al. Matrix Metalloproteinase-19 Is a Key Regulator of Lung Fibrosis in Mice and Humans. Am. J. Respir. Crit. Care Med. 2012, 186, 752–762. [Google Scholar] [CrossRef] [PubMed]
  35. Gharib, S.A.; Johnston, L.K.; Huizar, I.; Birkland, T.P.; Hanson, J.; Wang, Y.; Parks, W.C.; Manicone, A.M. MMP28 promotes macrophage polarization toward M2 cells and augments pulmonary fibrosis. J. Leukoc. Biol. 2013, 95, 9–18. [Google Scholar] [CrossRef] [PubMed]
  36. Gasse, P.; Mary, C.; Guenon, I.; Noulin, N.; Charron, S.; Schnyder-Candrian, S.; Schnyder, B.; Akira, S.; Quesniaux, V.F.; Lagente, V.; et al. IL-1R1/MyD88 signaling and the inflammasome are essential in pulmonary inflammation and fibrosis in mice. J. Clin. Investig. 2007, 117, 3786–3799. [Google Scholar] [CrossRef]
  37. Li, F.; Han, F.; Li, H.; Zhang, J.; Qiao, X.; Shi, J.; Yang, L.; Dong, J.; Luo, M.; Wei, J.; et al. Human placental mesenchymal stem cells of fetal origins-alleviated inflammation and fibrosis by attenuating MyD88 signaling in bleomycin-induced pulmonary fibrosis mice. Mol. Immunol. 2017, 90, 11–21. [Google Scholar] [CrossRef]
  38. Kradin, R.L.; Sakamoto, H.; Jain, F.; Zhao, L.-H.; Hymowitz, G.; Preffer, F. IL-10 inhibits inflammation but does not affect fibrosis in the pulmonary response to bleomycin. Exp. Mol. Pathol. 2004, 76, 205–211. [Google Scholar] [CrossRef]
  39. Sakamoto, H.; Zhao, L.-H.; Jain, F.; Kradin, R. IL-12p40−/− Mice Treated with Intratracheal Bleomycin Exhibit Decreased Pulmonary Inflammation and Increased Fibrosis. Exp. Mol. Pathol. 2002, 72, 1–9. [Google Scholar] [CrossRef]
  40. Rao, L.-Z.; Wang, Y.; Zhang, L.; Wu, G.; Zhang, L.; Wang, F.-X.; Chen, L.-M.; Sun, F.; Jia, S.; Zhang, S.; et al. IL-24 deficiency protects mice against bleomycin-induced pulmonary fibrosis by repressing IL-4-induced M2 program in macrophages. Cell Death Differ. 2020, 28, 1270–1283. [Google Scholar] [CrossRef]
  41. Dong, Z.; Lu, X.; Yang, Y.; Zhang, T.; Li, Y.; Chai, Y.; Lei, W.; Li, C.; Ai, L.; Tai, W. IL-27 alleviates the bleomycin-induced pulmonary fibrosis by regulating the Th17 cell differentiation. BMC Pulm. Med. 2015, 15, 13. [Google Scholar] [CrossRef] [PubMed]
  42. Gasse, P.; Riteau, N.; Vacher, R.; Michel, M.-L.; Fautrel, A.; Di Padova, F.; Fick, L.; Charron, S.; Lagente, V.; Eberl, G.; et al. IL-1 and IL-23 Mediate Early IL-17A Production in Pulmonary Inflammation Leading to Late Fibrosis. PLoS ONE 2011, 6, e23185. [Google Scholar] [CrossRef] [PubMed]
  43. Fanny, M.; Nascimento, M.; Baron, L.; Schricke, C.; Maillet, I.; Akbal, M.; Riteau, N.; Le Bert, M.; Quesniaux, V.; Ryffel, B.; et al. The IL-33 Receptor ST2 Regulates Pulmonary Inflammation and Fibrosis to Bleomycin. Front. Immunol. 2018, 9, 1476. [Google Scholar] [CrossRef] [PubMed]
  44. Ishida, Y.; Kimura, A.; Kondo, T.; Hayashi, T.; Ueno, M.; Takakura, N.; Matsushima, K.; Mukaida, N. Essential Roles of the CC Chemokine Ligand 3-CC Chemokine Receptor 5 Axis in Bleomycin-Induced Pulmonary Fibrosis through Regulation of Macrophage and Fibrocyte Infiltration. Am. J. Pathol. 2007, 170, 843–854. [Google Scholar] [CrossRef]
  45. Okuma, T.; Terasaki, Y.; Kaikita, K.; Kobayashi, H.; Kuziel, W.A.; Kawasuji, M.; Takeya, M. C-C chemokine receptor 2 (CCR2) deficiency improves bleomycin-induced pulmonary fibrosis by attenuation of both macrophage infiltration and production of macrophage-derived matrix metalloproteinases. J. Pathol. 2004, 204, 594–604. [Google Scholar] [CrossRef]
  46. Moore, B.B.; Paine, R.; Christensen, P.J.; Moore, T.A.; Sitterding, S.; Ngan, R.; Wilke, C.A.; Kuziel, W.A.; Toews, G.B. Protection from Pulmonary Fibrosis in the Absence of CCR2 Signaling. J. Immunol. 2001, 167, 4368–4377. [Google Scholar] [CrossRef]
  47. Gharaee-Kermani, M.; McCullumsmith, R.E.; Charo, I.F.; Kunkel, S.L.; Phan, S.H. CC-chemokine receptor 2 required for bleomycin-induced pulmonary fibrosis. Cytokine 2003, 24, 266–276. [Google Scholar] [CrossRef]
  48. Trujillo, G.; O’Connor, E.C.; Kunkel, S.L.; Hogaboam, C.M. A Novel Mechanism for CCR4 in the Regulation of Macrophage Activation in Bleomycin-Induced Pulmonary Fibrosis. Am. J. Pathol. 2008, 172, 1209–1221. [Google Scholar] [CrossRef]
  49. Smith, R.E.; Strieter, R.M.; Phan, S.H.; Lukacs, N.W.; Huffnagle, G.B.; Wilke, C.A.; Burdick, M.D.; Lincoln, P.; Evanoff, H.; Kunkel, S.L. Production and function of murine macrophage inflammatory protein-1 alpha in bleomycin-induced lung injury. J. Immunol. 1994, 153, 4704–4712. [Google Scholar] [CrossRef]
  50. Pochetuhen, K.; Luzina, I.G.; Lockatell, V.; Choi, J.; Todd, N.W.; Atamas, S.P. Complex Regulation of Pulmonary Inflammation and Fibrosis by CCL18. Am. J. Pathol. 2007, 171, 428–437. [Google Scholar] [CrossRef]
  51. Russo, R.C.; Guabiraba, R.; Garcia, C.C.; Barcelos, L.S.; Roffê, E.; Souza, A.L.S.; Amaral, F.A.; Cisalpino, D.; Cassali, G.D.; Doni, A.; et al. Role of the Chemokine Receptor CXCR2 in Bleomycin-Induced Pulmonary Inflammation and Fibrosis. Am. J. Respir. Cell Mol. Biol. 2009, 40, 410–421. [Google Scholar] [CrossRef] [PubMed]
  52. Besnard, A.-G.; Struyf, S.; Guabiraba, R.; Fauconnier, L.; Rouxel, N.; Proost, P.; Uyttenhove, C.; van Snick, J.; Couillin, I.; Ryffel, B. CXCL6 antibody neutralization prevents lung inflammation and fibrosis in mice in the bleomycin model. J. Leukoc. Biol. 2013, 94, 1317–1323. [Google Scholar] [CrossRef] [PubMed]
  53. Phillips, R.J.; Burdick, M.D.; Hong, K.; Lutz, M.A.; Murray, L.A.; Xue, Y.Y.; Belperio, J.A.; Keane, M.P.; Strieter, R.M. Circulating fibrocytes traffic to the lungs in response to CXCL12 and mediate fibrosis. J. Clin. Investig. 2004, 114, 438–446. [Google Scholar] [CrossRef] [PubMed]
  54. Xu, J.; Mora, A.; Shim, H.; Stecenko, A.; Brigham, K.L.; Rojas, M. Role of the SDF-1/CXCR4 Axis in the Pathogenesis of Lung Injury and Fibrosis. Am. J. Respir. Cell Mol. Biol. 2007, 37, 291–299. [Google Scholar] [CrossRef] [PubMed]
  55. Makino, H.; Aono, Y.; Azuma, M.; Kishi, M.; Yokota, Y.; Kinoshita, K.; Takezaki, A.; Kishi, J.; Kawano, H.; Ogawa, H.; et al. Antifibrotic effects of CXCR4 antagonist in bleomycin-induced pulmonary fibrosis in mice. J. Med. Investig. 2013, 60, 127–137. [Google Scholar] [CrossRef]
  56. Hashimoto, N.; Jin, H.; Liu, T.; Chensue, S.W.; Phan, S.H. Bone marrow–derived progenitor cells in pulmonary fibrosis. J. Clin. Investig. 2004, 113, 243–252. [Google Scholar] [CrossRef]
  57. Ishida, Y.; Kimura, A.; Nosaka, M.; Kuninaka, Y.; Hemmi, H.; Sasaki, I.; Kaisho, T.; Mukaida, N.; Kondo, T. Essential involvement of the CX3CL1-CX3CR1 axis in bleomycin-induced pulmonary fibrosis via regulation of fibrocyte and M2 macrophage migration. Sci. Rep. 2017, 7, 16833. [Google Scholar] [CrossRef]
  58. Hou, J.; Shi, J.; Chen, L.; Lv, Z.; Chen, X.; Cao, H.; Xiang, Z.; Han, X. M2 macrophages promote myofibroblast differentiation of LR-MSCs and are associated with pulmonary fibrogenesis. Cell Commun. Signal. 2018, 16, 89. [Google Scholar] [CrossRef]
  59. Ji, W.-J.; Ma, Y.-Q.; Zhou, X.; Zhang, Y.-D.; Lu, R.-Y.; Sun, H.-Y.; Guo, Z.-Z.; Zhang, Z.; Li, Y.-M.; Wei, L.-Q. Temporal and spatial characterization of mononuclear phagocytes in circulating, lung alveolar and interstitial compartments in a mouse model of bleomycin-induced pulmonary injury. J. Immunol. Methods 2014, 403, 7–16. [Google Scholar] [CrossRef]
  60. Zhu, L.; Fu, X.; Chen, X.; Han, X.; Dong, P. M2 macrophages induce EMT through the TGF-β/Smad2 signaling pathway. Cell Biol. Int. 2017, 41, 960–968. [Google Scholar] [CrossRef]
  61. Gibbons, M.A.; MacKinnon, A.C.; Ramachandran, P.; Dhaliwal, K.; Duffin, R.; Phythian-Adams, A.T.; van Rooijen, N.; Haslett, C.; Howie, S.E.; Simpson, A.J.; et al. Ly6Chi Monocytes Direct Alternatively Activated Profibrotic Macrophage Regulation of Lung Fibrosis. Am. J. Respir. Crit. Care Med. 2011, 184, 569–581. [Google Scholar] [CrossRef]
  62. Chen, F.; Gong, L.; Zhang, L.; Wang, H.; Qi, X.; Wu, X.; Xiao, Y.; Cai, Y.; Liu, L.; Li, X.; et al. Short courses of low dose dexamethasone delay bleomycin-induced lung fibrosis in rats. Eur. J. Pharmacol. 2006, 536, 287–295. [Google Scholar] [CrossRef]
  63. Casey, J.; Kaplan, J.; Atochina-Vasserman, E.N.; Gow, A.J.; Kadire, H.; Tomer, Y.; Fisher, J.H.; Hawgood, S.; Savani, R.C.; Beers, M.F. Alveolar Surfactant Protein D Content Modulates Bleomycin-induced Lung Injury. Am. J. Respir. Crit. Care Med. 2005, 172, 869–877. [Google Scholar] [CrossRef]
  64. Li, S.; Gao, S.; Jiang, Q.; Liang, Q.; Luan, J.; Zhang, R.; Zhang, F.; Ruan, H.; Li, X.; Li, X.; et al. Clevudine attenuates bleomycin-induced early pulmonary fibrosis via regulating M2 macrophage polarization. Int. Immunopharmacol. 2021, 101, 108271. [Google Scholar] [CrossRef]
  65. Gheberemedhin, A.; Salam, A.B.; Adu-Addai, B.; Noonan, S.; Stratton, R.; Ahmed, M.S.; Martin, G.; Huixian, L.; Andrews, C.; Balasubramanyam, K.; et al. A novel CD206 targeting peptide inhibits bleomycin induced pulmonary fibrosis in mice. bioRxiv 2020. [Google Scholar] [CrossRef]
  66. Zhao, P.; Cai, Z.; Tian, Y.; Li, J.; Li, K.; Li, M.; Bai, Y.; Li, J. Effective-compound combination inhibits the M2-like polarization of macrophages and attenuates the development of pulmonary fibrosis by increasing autophagy through mTOR signaling. Int. Immunopharmacol. 2021, 101, 108360. [Google Scholar] [CrossRef] [PubMed]
  67. Yao, Y.; Wang, Y.; Zhang, Z.; He, L.; Zhu, J.; Zhang, M.; He, X.; Cheng, Z.; Ao, Q.; Cao, Y.; et al. Chop Deficiency Protects Mice Against Bleomycin-induced Pulmonary Fibrosis by Attenuating M2 Macrophage Production. Mol. Ther. 2016, 24, 915–925. [Google Scholar] [CrossRef] [PubMed]
  68. Milara, J.; Hernandez, G.; Ballester, B.; Morell, A.; Roger, I.; Montero, P.; Escrivá, J.; Lloris, J.M.; Molina-Molina, M.; Morcillo, E.; et al. The JAK2 pathway is activated in idiopathic pulmonary fibrosis. Respir. Res. 2018, 19, 24. [Google Scholar] [CrossRef] [PubMed]
  69. Liu, B.; Jiang, Q.; Chen, R.; Gao, S.; Xia, Q.; Zhu, J.; Zhang, F.; Shao, C.; Liu, X.; Li, X.; et al. Tacrolimus ameliorates bleomycin-induced pulmonary fibrosis by inhibiting M2 macrophage polarization via JAK2/STAT3 signaling. Int. Immunopharmacol. 2022, 113, 109424. [Google Scholar] [CrossRef]
  70. Bantsimba-Malanda, C.; Marchal-Sommé, J.; Goven, D.; Freynet, O.; Michel, L.; Crestani, B.; Soler, P. A Role for Dendritic Cells in Bleomycin-induced Pulmonary Fibrosis in Mice? Am. J. Respir. Crit. Care Med. 2010, 182, 385–395. [Google Scholar] [CrossRef]
  71. Kafaja, S.; Valera, I.; Divekar, A.A.; Saggar, R.; Abtin, F.; Furst, D.E.; Khanna, D.; Singh, R.R. pDCs in lung and skin fibrosis in a bleomycin-induced model and patients with systemic sclerosis. J. Clin. Investig. 2018, 3, e98380. [Google Scholar] [CrossRef] [PubMed]
  72. Sharma, S.K.; MacLean, J.A.; Pinto, C.; Kradin, R.L. The effect of an anti-CD3 monoclonal antibody on bleomycin-induced lymphokine production and lung injury. Am. J. Respir. Crit. Care Med. 1996, 154, 193–200. [Google Scholar] [CrossRef] [PubMed]
  73. Kamio, K.; Azuma, A.; Matsuda, K.; Usuki, J.; Inomata, M.; Morinaga, A.; Kashiwada, T.; Nishijima, N.; Itakura, S.; Kokuho, N.; et al. Resolution of bleomycin-induced murine pulmonary fibrosis via a splenic lymphocyte subpopulation. Respir. Res. 2018, 19, 71. [Google Scholar] [CrossRef] [PubMed]
  74. Muro, A.F.; Moretti, F.A.; Moore, B.B.; Yan, M.; Atrasz, R.G.; Wilke, C.A.; Flaherty, K.R.; Martinez, F.J.; Tsui, J.L.; Sheppard, D.; et al. An Essential Role for Fibronectin Extra Type III Domain A in Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2008, 177, 638–645. [Google Scholar] [CrossRef] [PubMed]
  75. Hernnäs, J.; Nettelbladt, O.; Bjermer, L.; Särnstrand, B.; Malmström, A.; Hällgren, R. Alveolar accumulation of fibronectin and hyaluronan precedes bleomycin-induced pulmonary fibrosis in the rat. Eur. Respir. J. 1992, 5, 404–410. [Google Scholar] [CrossRef] [PubMed]
  76. Lazenby, A.J.; Crouch, E.C.; McDonald, J.A.; Kuhn, C. Remodeling of the Lung in Bleomycin-induced Pulmonary Fibrosis in the Rat: An Immunohistochemical Study of Laminin, Type IV Collagen, and Fibronectin. Am. Rev. Respir. Dis. 1990, 142, 206–214. [Google Scholar] [CrossRef]
  77. Park, S.-J.; Kim, T.-H.; Lee, K.; Kang, M.-A.; Jang, H.-J.; Ryu, H.-W.; Oh, S.-R.; Lee, H.-J. Kurarinone Attenuates BLM-Induced Pulmonary Fibrosis via Inhibiting TGF-β Signaling Pathways. Int. J. Mol. Sci. 2021, 22, 8388. [Google Scholar] [CrossRef]
  78. Aono, Y.; Nishioka, Y.; Inayama, M.; Ugai, M.; Kishi, J.; Uehara, H.; Izumi, K.; Sone, S. Imatinib as a Novel Antifibrotic Agent in Bleomycin-induced Pulmonary Fibrosis in Mice. Am. J. Respir. Crit. Care Med. 2005, 171, 1279–1285. [Google Scholar] [CrossRef]
  79. Chaudhary, N.I.; Roth, G.J.; Hilberg, F.; Muller-Quernheim, J.; Prasse, A.; Zissel, G.; Schnapp, A.; Park, J.E. Inhibition of PDGF, VEGF and FGF signalling attenuates fibrosis. Eur. Respir. J. 2007, 29, 976–985. [Google Scholar] [CrossRef]
  80. Wang, X.; Wu, G.; Gou, L.; Liu, Z.; Wang, X.; Fan, X.; Wu, L.; Liu, N. A novel single-chain-Fv antibody against connective tissue growth factor attenuates bleomycin-induced pulmonary fibrosis in mice. Respirology 2011, 16, 500–507. [Google Scholar] [CrossRef]
  81. Choi, J.-E.; Lee, S.-S.; Sunde, D.A.; Huizar, I.; Haugk, K.L.; Thannickal, V.J.; Vittal, R.; Plymate, S.R.; Schnapp, L.M. Insulin-like Growth Factor-I Receptor Blockade Improves Outcome in Mouse Model of Lung Injury. Am. J. Respir. Crit. Care Med. 2009, 179, 212–219. [Google Scholar] [CrossRef] [PubMed]
  82. Xiao, H.; Huang, X.; Wang, S.; Liu, Z.; Dong, R.; Song, D.; Dai, H. Metformin ameliorates bleomycin-induced pulmonary fibrosis in mice by suppressing IGF-1. Am. J. Transl. Res. 2020, 12, 940–949. [Google Scholar] [PubMed]
  83. Borie, R.; Fabre, A.; Prost, F.; Marchal-Somme, J.; Lebtahi, R.; Marchand-Adam, S.; Aubier, M.; Soler, P.; Crestani, B. Activation of somatostatin receptors attenuates pulmonary fibrosis. Thorax 2008, 63, 251–258. [Google Scholar] [CrossRef] [PubMed]
  84. Hosono, T.; Bando, M.; Mizushina, Y.; Sata, M.; Hagiwara, K.; Sugiyama, Y. Inhibitory effects of somatostatin analogue in bleomycin-induced pulmonary fibrosis. Exp. Lung Res. 2021, 47, 280–288. [Google Scholar] [CrossRef]
  85. Komura, K.; Yanaba, K.; Horikawa, M.; Ogawa, F.; Fujimoto, M.; Tedder, T.F.; Sato, S. CD19 regulates the development of bleomycin-induced pulmonary fibrosis in a mouse model. Arthritis Rheum. 2008, 58, 3574–3584. [Google Scholar] [CrossRef]
  86. Cargnoni, A.; Romele, P.; Signoroni, P.B.; Farigu, S.; Magatti, M.; Vertua, E.; Toschi, I.; Cesari, V.; Silini, A.R.; Stefani, F.R.; et al. Amniotic MSCs reduce pulmonary fibrosis by hampering lung B-cell recruitment, retention, and maturation. Stem Cells Transl. Med. 2020, 9, 1023–1035. [Google Scholar] [CrossRef]
  87. Wang, X.; Zhao, S.; Lai, J.; Guan, W.; Gao, Y. Anti-Inflammatory, Antioxidant, and Antifibrotic Effects of Gingival-Derived MSCs on Bleomycin-Induced Pulmonary Fibrosis in Mice. Int. J. Mol. Sci. 2021, 23, 99. [Google Scholar] [CrossRef]
  88. Pilling, D.; Zheng, Z.; Vakil, V.; Gomer, R.H. Fibroblasts secrete Slit2 to inhibit fibrocyte differentiation and fibrosis. Proc. Natl. Acad. Sci. USA 2014, 111, 18291–18296. [Google Scholar] [CrossRef]
  89. Zhou, Y.; He, Z.; Gao, Y.; Zheng, R.; Zhang, X.; Zhao, L.; Tan, M. Induced Pluripotent Stem Cells Inhibit Bleomycin-Induced Pulmonary Fibrosis in Mice through Suppressing TGF-β1/Smad-Mediated Epithelial to Mesenchymal Transition. Front. Pharmacol. 2016, 7, 430. [Google Scholar] [CrossRef]
  90. Sun, N.-N.; Zhang, Y.; Huang, W.-H.; Zheng, B.-J.; Jin, S.-Y.; Li, X.; Meng, Y. Macrophage exosomes transfer angiotensin II type 1 receptor to lung fibroblasts mediating bleomycin-induced pulmonary fibrosis. Chin. Med. J. 2021, 134, 2175–2185. [Google Scholar] [CrossRef]
  91. Mansouri, N.; Willis, G.R.; Fernandez-Gonzalez, A.; Reis, M.; Nassiri, S.; Mitsialis, S.A.; Kourembanas, S. Mesenchymal stromal cell exosomes prevent and revert experimental pulmonary fibrosis through modulation of monocyte phenotypes. J. Clin. Investig. 2019, 4, e128060. [Google Scholar] [CrossRef]
  92. Martin-Medina, A.; Lehmann, M.; Burgy, O.; Hermann, S.; Baarsma, H.; Wagner, D.E.; De Santis, M.M.; Ciolek, F.; Hofer, T.P.; Frankenberger, M.; et al. Increased Extracellular Vesicles Mediate WNT5A Signaling in Idiopathic Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2018, 198, 1527–1538. [Google Scholar] [CrossRef] [PubMed]
  93. Cinetto, F.; Ceccato, J.; Caputo, I.; Cangiano, D.; Montini, B.; Lunardi, F.; Piazza, M.; Agostini, C.; Calabrese, F.; Semenzato, G.; et al. GSK-3 Inhibition Modulates Metalloproteases in a Model of Lung Inflammation and Fibrosis. Front. Mol. Biosci. 2021, 8, 633054. [Google Scholar] [CrossRef] [PubMed]
  94. Gasse, P.; Riteau, N.; Charron, S.; Girre, S.; Fick, L.; Pétrilli, V.; Tschopp, J.; Lagente, V.; Quesniaux, V.F.J.; Ryffel, B.; et al. Uric Acid Is a Danger Signal Activating NALP3 Inflammasome in Lung Injury Inflammation and Fibrosis. Am. J. Respir. Crit. Care Med. 2009, 179, 903–913. [Google Scholar] [CrossRef] [PubMed]
  95. Meng, J.; Zou, Y.; Hu, C.; Zhu, Y.; Peng, Z.; Hu, G.; Wang, Z.; Tao, L. Fluorofenidone Attenuates Bleomycin-Induced Pulmonary Inflammation and Fibrosis in Mice Via Restoring Caveolin 1 Expression and Inhibiting Mitogen-Activated Protein Kinase Signaling Pathway. Shock 2012, 38, 567–573. [Google Scholar] [CrossRef] [PubMed]
  96. Song, C.; He, L.; Zhang, J.; Ma, H.; Yuan, X.; Hu, G.; Tao, L.; Meng, J. Fluorofenidone attenuates pulmonary inflammation and fibrosis via inhibiting the activation of NALP 3 inflammasome and IL-1β/IL-1R1/MyD88/NF-κB pathway. J. Cell. Mol. Med. 2016, 20, 2064–2077. [Google Scholar] [CrossRef]
  97. Nie, Y.; Hu, Y.; Yu, K.; Zhang, D.; Shi, Y.; Li, Y.; Sun, L.; Qian, F. Akt1 regulates pulmonary fibrosis via modulating IL-13 expression in macrophages. J. Endotoxin Res. 2019, 25, 451–461. [Google Scholar] [CrossRef]
  98. François, A.; Gombault, A.; Villeret, B.; Alsaleh, G.; Fanny, M.; Gasse, P.; Adam, S.M.; Crestani, B.; Sibilia, J.; Schneider, P.; et al. B cell activating factor is central to bleomycin- and IL-17-mediated experimental pulmonary fibrosis. J. Autoimmun. 2015, 56, 1–11. [Google Scholar] [CrossRef]
  99. Mercer, P.F.; Johns, R.H.; Scotton, C.J.; Krupiczojc, M.A.; Königshoff, M.; Howell, D.C.J.; McAnulty, R.J.; Das, A.; Thorley, A.J.; Tetley, T.D.; et al. Pulmonary Epithelium Is a Prominent Source of Proteinase-activated Receptor-1–inducible CCL2 in Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2009, 179, 414–425. [Google Scholar] [CrossRef]
  100. Izbicki, G.; Segel, M.; Christensen, T.; Conner, M.; Breuer, R. Time course of bleomycin-induced lung fibrosis. Int. J. Exp. Pathol. 2002, 83, 111–119. [Google Scholar] [CrossRef]
  101. Mouratis, M.A.; Aidinis, V. Modeling pulmonary fibrosis with bleomycin. Curr. Opin. Pulm. Med. 2011, 17, 355–361. [Google Scholar] [CrossRef] [PubMed]
  102. Moore, B.B.; Lawson, W.E.; Oury, T.D.; Sisson, T.H.; Raghavendran, K.; Hogaboam, C.M. Animal Models of Fibrotic Lung Disease. Am. J. Respir. Cell Mol. Biol. 2013, 49, 167–179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Chambers, R. Role of coagulation cascade proteases in lung repair and fibrosis. Eur. Respir. J. 2003, 22, 33s–35s. [Google Scholar] [CrossRef] [PubMed]
  104. Hamada, N.; Maeyama, T.; Kawaguchi, T.; Yoshimi, M.; Fukumoto, J.; Yamada, M.; Yamada, S.; Kuwano, K.; Nakanishi, Y. The Role of High Mobility Group Box1 in Pulmonary Fibrosis. Am. J. Respir. Cell Mol. Biol. 2008, 39, 440–447. [Google Scholar] [CrossRef]
  105. Korfei, M.; Ruppert, C.; Mahavadi, P.; Henneke, I.; Markart, P.; Koch, M.; Lang, G.; Fink, L.; Bohle, R.-M.; Seeger, W.; et al. Epithelial Endoplasmic Reticulum Stress and Apoptosis in Sporadic Idiopathic Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2008, 178, 838–846. [Google Scholar] [CrossRef]
  106. Dik, W.A.; Zimmermann, L.J.; Naber, B.A.; Janssen, D.J.; van Kaam, A.H.; Versnel, M.A. Thrombin contributes to bronchoalveolar lavage fluid mitogenicity in lung disease of the premature infant. Pediatr. Pulmonol. 2002, 35, 34–41. [Google Scholar] [CrossRef]
  107. Hernández-Rodríguez, N.; Cambrey, A.; Chambers, R.; Gray, A.; McAnulty, R.; Laurent, G.; Harrison, N.; Southcott, A.; Dubois, R.; Black, C.; et al. Role of thrombin in pulmonary fibrosis. Lancet 1995, 346, 1071–1073. [Google Scholar] [CrossRef]
  108. Gabazza, E.C.; Taguchi, O.; Tamaki, S.; Takeya, H.; Kobayashi, H.; Yasui, H.; Hataji, O.; Urano, H.; Zhou, H.; Suzuki, K.; et al. Thrombin in the Airways of Asthmatic Patients. Lung 1999, 177, 253–262. [Google Scholar] [CrossRef]
  109. Bogatkevich, G.S.; Tourkina, E.; Silver, R.M.; Ludwicka-Bradley, A. Thrombin Differentiates Normal Lung Fibroblasts to a Myofibroblast Phenotype via the Proteolytically Activated Receptor-1 and a Protein Kinase C-dependent Pathway. J. Biol. Chem. 2001, 276, 45184–45192. [Google Scholar] [CrossRef]
  110. Chambers, R.C.; Dabbagh, K.; McAnulty, R.J.; Gray, A.J.; Blanc-Brude, O.P.; Laurent, G.J. Thrombin stimulates fibroblast procollagen production via proteolytic activation of protease-activated receptor 1. Biochem. J. 1998, 333, 121–127. [Google Scholar] [CrossRef]
  111. Vaccaro, C.A.; Brody, J.S.; Snider, G.L. Alveolar wall basement membranes in bleomycin-induced pulmonary fibrosis. Am. Rev. Respir. Dis. 1985, 132, 905–912. [Google Scholar] [CrossRef] [PubMed]
  112. Mahalanobish, S.; Saha, S.; Dutta, S.; Sil, P.C. Matrix metalloproteinase: An upcoming therapeutic approach for idiopathic pulmonary fibrosis. Pharmacol. Res. 2020, 152, 104591. [Google Scholar] [CrossRef] [PubMed]
  113. Craig, V.J.; Zhang, L.; Hagood, J.S.; Owen, C.A. Matrix Metalloproteinases as Therapeutic Targets for Idiopathic Pulmonary Fibrosis. Am. J. Respir. Cell Mol. Biol. 2015, 53, 585–600. [Google Scholar] [CrossRef] [PubMed]
  114. Pardo, A.; Cabrera, S.; Maldonado, M.; Selman, M. Role of matrix metalloproteinases in the pathogenesis of idiopathic pulmonary fibrosis. Respir. Res. 2016, 17, 23. [Google Scholar] [CrossRef] [PubMed]
  115. Chakraborti, S.; Mandal, M.; Das, S.; Mandal, A.; Chakraborti, T. Regulation of matrix metalloproteinases: An overview. Mol. Cell. Biochem. 2003, 253, 269–285. [Google Scholar] [CrossRef] [PubMed]
  116. Hayashi, T.; Stetler-Stevenson, W.G.; Fleming, M.V.; Fishback, N.; Koss, M.N.; Liotta, L.A.; Ferrans, V.J.; Travis, W.D. Immunohistochemical study of metalloproteinases and their tissue inhibitors in the lungs of patients with diffuse alveolar damage and idiopathic pulmonary fibrosis. Am. J. Pathol. 1996, 149, 1241–1256. [Google Scholar] [PubMed]
  117. Selman, M.; Ruiz, V.; Cabrera, S.; Segura, L.; Ramírez, R.; Barrios, R.; Pardo, A. TIMP-1, -2, -3, and -4 in idiopathic pulmonary fibrosis. A prevailing nondegradative lung microenvironment? Am. J. Physiol. Cell. Mol. Physiol. 2000, 279, L562–L574. [Google Scholar] [CrossRef]
  118. Ramos, C.; Montaño, M.; García-Alvarez, J.; Ruiz, V.; Uhal, B.D.; Selman, M.; Pardo, A. Fibroblasts from Idiopathic Pulmonary Fibrosis and Normal Lungs Differ in Growth Rate, Apoptosis, and Tissue Inhibitor of Metalloproteinases Expression. Am. J. Respir. Cell Mol. Biol. 2001, 24, 591–598. [Google Scholar] [CrossRef]
  119. McKeown, S.; Richter, A.G.; O’Kane, C.; McAuley, D.F.; Thickett, D.R. MMP expression and abnormal lung permeability are important determinants of outcome in IPF. Eur. Respir. J. 2009, 33, 77–84. [Google Scholar] [CrossRef]
  120. Hoshino, M.; Nakamura, Y.; Sim, J.; Shimojo, J.; Isogai, S. Bronchial subepithelial fibrosis and expression of matrix metalloproteinase-9 in asthmatic airway inflammation. J. Allergy Clin. Immunol. 1998, 102, 783–788. [Google Scholar] [CrossRef]
  121. Murphy, G.; Docherty, A.J.P. The Matrix Metalloproteinases and Their Inhibitors. Am. J. Respir. Cell Mol. Biol. 1992, 7, 120–125. [Google Scholar] [CrossRef] [PubMed]
  122. Corbel, M.; Belleguic, C.; Boichot, E.; Lagente, V. Involvement of gelatinases (MMP-2 and MMP-9) in the development of airway inflammation and pulmonary fibrosis. Cell Biol. Toxicol. 2002, 18, 51–61. [Google Scholar] [CrossRef] [PubMed]
  123. Ruiz, V.; Ordóñez, R.M.; Berumen, J.; Ramírez, R.; Uhal, B.; Becerril, C.; Pardo, A.; Selman, M. Unbalanced collagenases/TIMP-1 expression and epithelial apoptosis in experimental lung fibrosis. Am. J. Physiol. Cell. Mol. Physiol. 2003, 285, L1026–L1036. [Google Scholar] [CrossRef] [PubMed]
  124. Gadek, J.E.; Kelman, J.A.; Fells, G.; Weinberger, S.E.; Horwitz, A.L.; Reynolds, H.Y.; Fulmer, J.D.; Crystal, R.G. Collagenase in the Lower Respiratory Tract of Patients with Idiopathic Pulmonary Fibrosis. N. Engl. J. Med. 1979, 301, 737–742. [Google Scholar] [CrossRef] [PubMed]
  125. Oggionni, T.; Morbini, P.; Inghilleri, S.; Palladini, G.; Tozzi, R.; Vitulo, P.; Fenoglio, C.; Perlini, S.; Pozzi, E. Time course of matrix metalloproteases and tissue inhibitors in bleomycin-induced pulmonary fibrosis. Eur. J. Histochem. 2007, 50, 317–326. [Google Scholar]
  126. O’Connor, C.M.; FitzGerald, M.X. Matrix metalloproteases and lung disease. Thorax 1994, 49, 602–609. [Google Scholar] [CrossRef]
  127. Kim, J.Y.; Choeng, H.C.; Ahn, C.; Cho, S.-H. Early and Late Changes of MMP-2 and MMP-9 in Bleomycin-Induced Pulmonary Fibrosis. Yonsei Med. J. 2009, 50, 68–77. [Google Scholar] [CrossRef]
  128. Yaguchi, T.; Fukuda, Y.; Ishizaki, M.; Yamanaka, N. Immunohistochemical and gelatin zymography studies for matrix metalloproteinases in bleomycin-induced pulmonary fibrosis. Pathol. Int. 1998, 48, 954–963. [Google Scholar] [CrossRef]
  129. Henry, M.; McMahon, K.; Mackarel, A.; Prikk, K.; Sorsa, T.; Maisi, P.; Sepper, R.; FitzGerald, M.; O’Connor, C. Matrix metalloproteinases and tissue inhibitor of metalloproteinase-1 in sarcoidosis and IPF. Eur. Respir. J. 2002, 20, 1220–1227. [Google Scholar] [CrossRef]
  130. Lemjabbar, H.; Gosset, P.; Lechapt-Zalcman, E.; Franco-Montoya, M.-L.; Wallaert, B.; Harf, A.; Lafuma, C. Overexpression of Alveolar Macrophage Gelatinase B (MMP-9) in Patients with Idiopathic Pulmonary Fibrosis. Am. J. Respir. Cell Mol. Biol. 1999, 20, 903–913. [Google Scholar] [CrossRef]
  131. Suga, M.; Iyonaga, K.; Okamoto, T.; Gushima, Y.; Miyakawa, H.; Akaike, T.; Ando, M. Characteristic Elevation of Matrix Metalloproteinase Activity in Idiopathic Interstitial Pneumonias. Am. J. Respir. Crit. Care Med. 2000, 162, 1949–1956. [Google Scholar] [CrossRef] [PubMed]
  132. Yu, W.-H.; Woessner, J.J.F.; McNeish, J.D.; Stamenkovic, I. CD44 anchors the assembly of matrilysin/MMP-7 with heparin-binding epidermal growth factor precursor and ErbB4 and regulates female reproductive organ remodeling. Genes Dev. 2002, 16, 307–323. [Google Scholar] [CrossRef] [PubMed]
  133. Lanone, S.; Zheng, T.; Zhu, Z.; Liu, W.; Lee, C.G.; Ma, B.; Chen, Q.; Homer, R.J.; Wang, J.; Rabach, L.A.; et al. Overlapping and enzyme-specific contributions of matrix metalloproteinases-9 and -12 in IL-13–induced inflammation and remodeling. J. Clin. Investig. 2002, 110, 463–474. [Google Scholar] [CrossRef] [PubMed]
  134. Pilewski, J.M.; Liu, L.; Henry, A.C.; Knauer, A.V.; Feghali-Bostwick, C.A. Insulin-Like Growth Factor Binding Proteins 3 and 5 Are Overexpressed in Idiopathic Pulmonary Fibrosis and Contribute to Extracellular Matrix Deposition. Am. J. Pathol. 2005, 166, 399–407. [Google Scholar] [CrossRef]
  135. Liu, Y.-M.; Nepali, K.; Liou, J.-P. Idiopathic Pulmonary Fibrosis: Current Status, Recent Progress, and Emerging Targets. J. Med. Chem. 2016, 60, 527–553. [Google Scholar] [CrossRef]
  136. Chambers, R.C.; Mercer, P.F. Mechanisms of Alveolar Epithelial Injury, Repair, and Fibrosis. Ann. Am. Thorac. Soc. 2015, 12, S16–S20. [Google Scholar] [CrossRef]
  137. Drakopanagiotakis, F.; Wujak, L.; Wygrecka, M.; Markart, P. Biomarkers in idiopathic pulmonary fibrosis. Matrix Biol. 2018, 68-69, 404–421. [Google Scholar] [CrossRef]
  138. Liu, H.; Wu, X.; Gan, C.; Wang, L.; Wang, G.; Yue, L.; Liu, Z.; Wei, W.; Su, X.; Zhang, Q.; et al. A novel multikinase inhibitor SKLB-YTH-60 ameliorates inflammation and fibrosis in bleomycin-induced lung fibrosis mouse models. Cell Prolif. 2021, 54, e13081. [Google Scholar] [CrossRef]
  139. Upagupta, C.; Shimbori, C.; Alsilmi, R.; Kolb, M. Matrix abnormalities in pulmonary fibrosis. Eur. Respir. Rev. 2018, 27, 180033. [Google Scholar] [CrossRef]
  140. Martinez, F.J.; Collard, H.R.; Pardo, A.; Raghu, G.; Richeldi, L.; Selman, M.; Swigris, J.J.; Taniguchi, H.; Wells, A.U. Idiopathic pulmonary fibrosis. Nat. Rev. Dis. Prim. 2017, 3, 17074. [Google Scholar] [CrossRef]
  141. Gauldie, J. Inflammatory Mechanisms Are a Minor Component of the Pathogenesis of Idiopathic Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2002, 165, 1205–1206. [Google Scholar] [CrossRef] [PubMed]
  142. Piguet, P.F. Inflammation in idiopathic pulmonary fibrosis. Am. J. Respir. Crit. Care Med. 2003, 167, 1037. [Google Scholar] [CrossRef] [PubMed]
  143. Strieter, R.M. Inflammatory Mechanisms Are Not a Minor Component of the Pathogenesis of Idiopathic Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2002, 165, 1206–1207. [Google Scholar] [CrossRef]
  144. Bringardner, B.D.; Baran, C.P.; Eubank, T.D.; Marsh, C.B. The Role of Inflammation in the Pathogenesis of Idiopathic Pulmonary Fibrosis. Antioxid. Redox Signal. 2008, 10, 287–302. [Google Scholar] [CrossRef]
  145. Davies, H.R.H.R.; Richeldi, L.; Walters, E.H. Immunomodulatory agents for idiopathic pulmonary fibrosis. Cochrane Database Syst. Rev. 2003, CD003134. [Google Scholar] [CrossRef]
  146. Richeldi, L.; Davies, H.R.H.R.; Spagnolo, P.; Luppi, F. Corticosteroids for idiopathic pulmonary fibrosis. Cochrane Database Syst. Rev. 2003, CD002880. [Google Scholar] [CrossRef] [PubMed]
  147. Selman, M.; King, T.E., Jr.; Pardo, A. Idiopathic Pulmonary Fibrosis: Prevailing and Evolving Hypotheses about Its Pathogenesis and Implications for Therapy. Ann. Intern. Med. 2001, 134, 136–151. [Google Scholar] [CrossRef]
  148. Elovic, A.E.; Ohyama, H.; Sauty, A.; McBride, J.; Tsuji, T.; Nagai, M.; Weller, P.F.; Wong, D.T. IL-4-dependent regulation of TGF-α and TGF-β1 expression in human eosinophils. J. Immunol. 1998, 160, 6121–6127. [Google Scholar] [CrossRef]
  149. Minshall, E.M.; Leung, D.Y.M.; Martin, R.J.; Song, Y.L.; Cameron, L.; Ernst, P.; Hamid, Q. Eosinophil-associated TGF- β1 mRNA Expression and Airways Fibrosis in Bronchial Asthma. Am. J. Respir. Cell Mol. Biol. 1997, 17, 326–333. [Google Scholar] [CrossRef]
  150. Ohno, I.; Nitta, Y.; Yamauchi, K.; Hoshi, H.; Honma, M.; Woolley, K.; O’Byrne, P.; Tamura, G.; Jordana, M.; Shirato, K. Transforming growth factor beta 1 (TGF β1) gene expression by eosinophils in asthmatic airway inflammation. Am. J. Respir. Cell Mol. Biol. 1996, 15, 404–409. [Google Scholar] [CrossRef]
  151. Wenzel, S.E.; Trudeau, J.B.; Barnes, S.; Zhou, X.; Cundall, M.; Westcott, J.Y.; McCord, K.; Chu, H.W. TGF-β and IL-13 Synergistically Increase Eotaxin-1 Production in Human Airway Fibroblasts. J. Immunol. 2002, 169, 4613–4619. [Google Scholar] [CrossRef] [PubMed]
  152. Nakagome, K.; Dohi, M.; Okunishi, K.; Tanaka, R.; Miyazaki, J.; Yamamoto, K. In vivo IL-10 gene delivery attenuates bleomycin induced pulmonary fibrosis by inhibiting the production and activation of TGF-β in the lung. Thorax 2006, 61, 886–894. [Google Scholar] [CrossRef] [PubMed]
  153. Moodley, Y.; Rigby, P.; Bundell, C.; Bunt, S.; Hayashi, H.; Misso, N.; McAnulty, R.; Laurent, G.; Scaffidi, A.; Thompson, P.; et al. Macrophage Recognition and Phagocytosis of Apoptotic Fibroblasts Is Critically Dependent on Fibroblast-Derived Thrombospondin 1 and CD36. Am. J. Pathol. 2003, 162, 771–779. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Agostini, C. Chemokine/Cytokine Cocktail in Idiopathic Pulmonary Fibrosis. Proc. Am. Thorac. Soc. 2006, 3, 357–363. [Google Scholar] [CrossRef]
  155. Fertin, C.; Nicolas, J.F.; Gillery, P.; Kalis, B.; Banchereau, J.; Maquart, F.X. Interleukin-4 stimulates collagen synthesis by normal and scleroderma fibroblasts in dermal equivalents. Cell. Mol. Biol. 1991, 37, 823–829. [Google Scholar]
  156. Sempowski, G.D.; Beckmann, M.P.; Derdak, S.; Phipps, R.P. Subsets of murine lung fibroblasts express membrane-bound and soluble IL-4 receptors. Role of IL-4 in enhancing fibroblast proliferation and collagen synthesis. J. Immunol. 1994, 152, 3606–3614. [Google Scholar] [CrossRef]
  157. Strutz, F.; Zeisberg, M.; Renziehausen, A.; Raschke, B.; Becker, V.; Van Kooten, C.; Müller, G.A. TGF-β1 induces proliferation in human renal fibroblasts via induction of basic fibroblast growth factor (FGF-2). Kidney Int. 2001, 59, 579–592. [Google Scholar] [CrossRef]
  158. Wynn, T.A. IL-13 Effector Functions. Annu. Rev. Immunol. 2003, 21, 425–456. [Google Scholar] [CrossRef]
  159. Wynn, T.A. Fibrotic disease and the TH1/TH2 paradigm. Nat. Rev. Immunol. 2004, 4, 583–594. [Google Scholar] [CrossRef]
  160. Postlethwaite, A.E.; Raghow, R.; Stricklin, G.P.; Poppleton, H.; Seyer, J.M.; Kang, A.H. Modulation of fibroblast functions by interleukin 1: Increased steady-state accumulation of type I procollagen messenger RNAs and stimulation of other functions but not chemotaxis by human recombinant interleukin 1 alpha and beta. J. Cell Biol. 1988, 106, 311–318. [Google Scholar] [CrossRef]
  161. Riteau, N.; Gasse, P.; Fauconnier, L.; Gombault, A.; Couegnat, M.; Fick, L.; Kanellopoulos, J.; Quesniaux, V.F.J.; Marchand-Adam, S.; Crestani, B.; et al. Extracellular ATP Is a Danger Signal Activating P2X7 Receptor in Lung Inflammation and Fibrosis. Am. J. Respir. Crit. Care Med. 2010, 182, 774–783. [Google Scholar] [CrossRef]
  162. Kirsten, D. Sarkoidose: Aktuelle Diagnostik und Therapie. DMW Dtsch. Med. Wochenschr. 2013, 138, 537–541. [Google Scholar] [CrossRef]
  163. Kim, S.-K.; Choe, J.-Y.; Park, K.-Y. Rebamipide Suppresses Monosodium Urate Crystal-Induced Interleukin-1β Production Through Regulation of Oxidative Stress and Caspase-1 in THP-1 Cells. Inflammation 2015, 39, 473–482. [Google Scholar] [CrossRef] [PubMed]
  164. Haddad, J.J. Antioxidant and prooxidant mechanisms in the regulation of redox(y)-sensitive transcription factors. Cell. Signal. 2002, 14, 879–897. [Google Scholar] [CrossRef] [PubMed]
  165. Guzik, T.J.; Korbut, R.; Adamek-Guzik, T. Nitric oxide and superoxide in inflammation and immune regulation. J. Physiol. Pharmacol. 2003, 54, 469–487. [Google Scholar] [PubMed]
  166. Hsieh, C.-S.; Macatonia, S.E.; Tripp, C.S.; Wolf, S.F.; O’Garra, A.; Murphy, K.M. Development of TH 1 CD4+ T Cells Through IL-12 Produced by Listeria-Induced Macrophages. Science 1993, 260, 547–549. [Google Scholar] [CrossRef]
  167. Park, S.-W.; Ahn, M.-H.; Jang, H.K.; Jang, A.S.; Kim, D.-J.; Koh, E.-S.; Park, J.-S.; Uh, S.-T.; Kim, Y.H.; Park, J.S.; et al. Interleukin-13 and Its Receptors in Idiopathic Interstitial Pneumonia: Clinical Implications for Lung Function. J. Korean Med. Sci. 2009, 24, 614–620. [Google Scholar] [CrossRef]
  168. Belperio, J.A.; Dy, M.; Burdick, M.D.; Xue, Y.Y.; Li, K.; Elias, J.A.; Keane, M.P. Interaction of IL-13 and C10 in the Pathogenesis of Bleomycin-Induced Pulmonary Fibrosis. Am. J. Respir. Cell Mol. Biol. 2002, 27, 419–427. [Google Scholar] [CrossRef]
  169. Jakubzick, C.; Choi, E.S.; Joshi, B.H.; Keane, M.P.; Kunkel, S.L.; Puri, R.K.; Hogaboam, C.M. Therapeutic Attenuation of Pulmonary Fibrosis Via Targeting of IL-4- and IL-13-Responsive Cells. J. Immunol. 2003, 171, 2684–2693. [Google Scholar] [CrossRef]
  170. Li, C.; Du, S.; Lu, Y.; Lu, X.; Liu, F.; Chen, Y.; Weng, D.; Chen, J. Blocking the 4-1BB Pathway Ameliorates Crystalline Silica-induced Lung Inflammation and Fibrosis in Mice. Theranostics 2016, 6, 2052–2067. [Google Scholar] [CrossRef]
  171. Mi, S.; Li, Z.; Yang, H.-Z.; Liu, H.; Wang, J.-P.; Ma, Y.-G.; Wang, X.-X.; Liu, H.-Z.; Sun, W.; Hu, Z.-W. Blocking IL-17A Promotes the Resolution of Pulmonary Inflammation and Fibrosis Via TGF-β1–Dependent and –Independent Mechanisms. J. Immunol. 2011, 187, 3003–3014. [Google Scholar] [CrossRef] [PubMed]
  172. Wilson, M.S.; Madala, S.K.; Ramalingam, T.R.; Gochuico, B.R.; Rosas, I.O.; Cheever, A.W.; Wynn, T.A. Bleomycin and IL-1β–mediated pulmonary fibrosis is IL-17A dependent. J. Exp. Med. 2010, 207, 535–552. [Google Scholar] [CrossRef] [PubMed]
  173. Wang, M.; Liang, P. Interleukin-24 and its receptors. Immunology 2005, 114, 166–170. [Google Scholar] [CrossRef] [PubMed]
  174. Menezes, M.E.; Bhatia, S.; Bhoopathi, P.; Das, S.K.; Emdad, L.; Dasgupta, S.; Dent, P.; Wang, X.-Y.; Sarkar, D.; Fisher, P.B. MDA-7/IL-24: Multifunctional Cancer Killing Cytokine. Anticancer. Genes 2014, 818, 127–153. [Google Scholar] [CrossRef] [Green Version]
  175. Poindexter, N.J.; Williams, R.R.; Powis, G.; Jen, E.; Caudle, A.S.; Chada, S.; Grimm, E.A. IL-24 is expressed during wound repair and inhibits TGFα-induced migration and proliferation of keratinocytes. Exp. Dermatol. 2009, 19, 714–722. [Google Scholar] [CrossRef]
  176. Shefler, I.; Pasmanik-Chor, M.; Kidron, D.; Mekori, Y.A.; Hershko, A.Y. T cell-derived microvesicles induce mast cell production of IL-24: Relevance to inflammatory skin diseases. J. Allergy Clin. Immunol. 2014, 133, 217–224.e3. [Google Scholar] [CrossRef]
  177. Kumari, S.; Bonnet, M.C.; Ulvmar, M.H.; Wolk, K.; Karagianni, N.; Witte, E.; Uthoff-Hachenberg, C.; Renauld, J.-C.; Kollias, G.; Toftgard, R.; et al. Tumor Necrosis Factor Receptor Signaling in Keratinocytes Triggers Interleukin-24-Dependent Psoriasis-like Skin Inflammation in Mice. Immunity 2013, 39, 899–911. [Google Scholar] [CrossRef]
  178. Küchler, A.M.; Pollheimer, J.; Balogh, J.; Sponheim, J.; Manley, L.; Sorensen, D.R.; De Angelis, P.M.; Scott, H.; Haraldsen, G. Nuclear Interleukin-33 Is Generally Expressed in Resting Endothelium but Rapidly Lost upon Angiogenic or Proinflammatory Activation. Am. J. Pathol. 2008, 173, 1229–1242. [Google Scholar] [CrossRef]
  179. Moussion, C.; Ortega, N.; Girard, J.-P. The IL-1-Like Cytokine IL-33 Is Constitutively Expressed in the Nucleus of Endothelial Cells and Epithelial Cells In Vivo: A Novel ‘Alarmin’? PLoS ONE 2008, 3, e3331. [Google Scholar] [CrossRef]
  180. Li, D.; Guabiraba, R.; Besnard, A.-G.; Komai-Koma, M.; Jabir, M.S.; Zhang, L.; Graham, G.J.; Kurowska-Stolarska, M.; Liew, F.Y.; McSharry, C.; et al. IL-33 promotes ST2-dependent lung fibrosis by the induction of alternatively activated macrophages and innate lymphoid cells in mice. J. Allergy Clin. Immunol. 2014, 134, 1422–1432.e11. [Google Scholar] [CrossRef]
  181. Zhang, K.; Gharaee-Kermani, M.; Jones, M.L.; Warren, J.S.; Phan, S. Lung monocyte chemoattractant protein-1 gene expression in bleomycin-induced pulmonary fibrosis. J. Immunol. 1994, 153, 4733–4741. [Google Scholar] [CrossRef] [PubMed]
  182. Iyonaga, K.; Takeya, M.; Saita, N.; Sakamoto, O.; Yoshimura, T.; Ando, M.; Takahashi, K. Monocyte chemoattractant protein-1 in idiopathic pulmonary fibrosis and other interstitial lung diseases. Hum. Pathol. 1994, 25, 455–463. [Google Scholar] [CrossRef] [PubMed]
  183. Smith, R.E.; Stricter, R.M.; Zhang, K.; Phan, S.; Standiford, T.J.; Lukacs, N.W.; Kunkel, S.L. A role for C-C chemokines in fibrotic lung disease. J. Leukoc. Biol. 1995, 57, 782–787. [Google Scholar] [CrossRef] [PubMed]
  184. Suga, M.; Iyonaga, K.; Ichiyasu, H.; Saita, N.; Yamasaki, H.; Ando, M. Clinical significance of MCP-1 levels in BALF and serum in patients with interstitial lung diseases. Eur. Respir. J. 1999, 14, 376–382. [Google Scholar] [CrossRef]
  185. Misharin, A.V.; Morales-Nebreda, L.; Reyfman, P.A.; Cuda, C.M.; Walter, J.M.; McQuattie-Pimentel, A.C.; Chen, C.-I.; Anekalla, K.R.; Joshi, N.; Williams, K.J.N.; et al. Monocyte-derived alveolar macrophages drive lung fibrosis and persist in the lung over the life span. J. Exp. Med. 2017, 214, 2387–2404. [Google Scholar] [CrossRef]
  186. Reyfman, P.A.; Walter, J.M.; Joshi, N.; Anekalla, K.R.; McQuattie-Pimentel, A.C.; Chiu, S.; Fernandez, R.; Akbarpour, M.; Chen, C.-I.; Ren, Z.; et al. Single-Cell Transcriptomic Analysis of Human Lung Provides Insights into the Pathobiology of Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2019, 199, 1517–1536. [Google Scholar] [CrossRef]
  187. Smith, R.E.; Strieter, R.M.; Phan, S.H.; Kunkel, S.L. C-C chemokines: Novel mediators of the profibrotic inflammatory response to bleomycin challenge. Am. J. Respir. Cell Mol. Biol. 1996, 15, 693–702. [Google Scholar] [CrossRef]
  188. Tokuda, A.; Itakura, M.; Onai, N.; Kimura, H.; Kuriyama, T.; Matsushima, K. Pivotal Role of CCR1-Positive Leukocytes in Bleomycin- Induced Lung Fibrosis in Mice. J. Immunol. 2000, 164, 2745–2751. [Google Scholar] [CrossRef]
  189. Abe, R.; Donnelly, S.C.; Peng, T.; Bucala, R.; Metz, C.N. Peripheral Blood Fibrocytes: Differentiation Pathway and Migration to Wound Sites. J. Immunol. 2001, 166, 7556–7562. [Google Scholar] [CrossRef]
  190. Moore, B.B.; Kolodsick, J.E.; Thannickal, V.J.; Cooke, K.; Moore, T.A.; Hogaboam, C.; Wilke, C.A.; Toews, G.B. CCR2-Mediated Recruitment of Fibrocytes to the Alveolar Space after Fibrotic Injury. Am. J. Pathol. 2005, 166, 675–684. [Google Scholar] [CrossRef]
  191. Bamshad, M.J.; Mummidi, S.; Gonzalez, E.; Ahuja, S.S.; Dunn, D.M.; Watkins, W.S.; Wooding, S.; Stone, A.C.; Jorde, L.B.; Weiss, R.B.; et al. A strong signature of balancing selection in the 5′ cis-regulatory region of CCR5. Proc. Natl. Acad. Sci. USA 2002, 99, 10539–10544. [Google Scholar] [CrossRef] [PubMed]
  192. Blanpain, C.; Lee, B.; Tackoen, M.; Puffer, B.; Boom, A.; Libert, F.; Sharron, M.; Wittamer, V.; Vassart, G.; Doms, R.W.; et al. Multiple nonfunctional alleles of CCR5 are frequent in various human populations. Blood 2000, 96, 1638–1645. [Google Scholar] [CrossRef] [PubMed]
  193. Mummidi, S.; Bamshad, M.; Ahuja, S.S.; Gonzalez, E.; Feuillet, P.M.; Begum, K.; Galvis, M.C.; Kostecki, V.; Valente, A.J.; Murthy, K.K.; et al. Evolution of Human and Non-human Primate CC Chemokine Receptor 5 Gene and mRNA. J. Biol. Chem. 2000, 275, 18946–18961. [Google Scholar] [CrossRef] [PubMed]
  194. Curci, M.; Little, K.; Sieber, W.; Kiesewetter, W. Peptic ulcer disease in childhood reexamined. J. Pediatr. Surg. 1976, 11, 329–335. [Google Scholar] [CrossRef]
  195. Hieshima, K.; Imai, T.; Baba, M.; Shoudai, K.; Ishizuka, K.; Nakagawa, T.; Tsuruta, J.; Takeya, M.; Sakaki, Y.; Takatsuki, K.; et al. A novel human CC chemokine PARC that is most homologous to macrophage-inflammatory protein-1 alpha/LD78 alpha and chemotactic for T lymphocytes, but not for monocytes. J. Immunol. 1997, 159, 1140–1149. [Google Scholar] [CrossRef]
  196. Guan, P.; Burghes, A.; Cunningham, A.; Lira, P.; Brissette, W.H.; Neote, K.; McColl, S.R. Genomic Organization and Biological Characterization of the Novel Human CC Chemokine DC-CK-1/PARC/MIP-4/SCYA18. Genomics 1999, 56, 296–302. [Google Scholar] [CrossRef]
  197. Luzina, I.G.; Papadimitriou, J.C.; Anderson, R.; Pochetuhen, K.; Atamas, S.P. Induction of prolonged infiltration of T lymphocytes and transient T lymphocyte-dependent collagen deposition in mouse lungs following adenoviral gene transfer of CCL18. Arthritis Rheum. 2006, 54, 2643–2655. [Google Scholar] [CrossRef]
  198. Belperio, J.A.; Keane, M.P.; Arenberg, D.A.; Addison, C.L.; Ehlert, J.E.; Burdick, M.D.; Strieter, R.M. CXC chemokines in angiogenesis. J. Leukoc. Biol. 2000, 68, 1–8. [Google Scholar] [CrossRef]
  199. Strieter, R.M.; Gomperts, B.N.; Keane, M.P. The role of CXC chemokines in pulmonary fibrosis. J. Clin. Investig. 2007, 117, 549–556. [Google Scholar] [CrossRef]
  200. Bertini, R.; Allegretti, M.; Bizzarri, C.; Moriconi, A.; Locati, M.; Zampella, G.; Cervellera, M.N.; Di Cioccio, V.; Cesta, M.C.; Galliera, E.; et al. Noncompetitive allosteric inhibitors of the inflammatory chemokine receptors CXCR1 and CXCR2: Prevention of reperfusion injury. Proc. Natl. Acad. Sci. USA 2004, 101, 11791–11796. [Google Scholar] [CrossRef]
  201. Souza, D.G.; Bertini, R.; Vieira, A.T.; Cunha, F.Q.; Poole, S.; Allegretti, M.; Colotta, F.; Teixeira, M.M. Repertaxin, a novel inhibitor of rat CXCR2 function, inhibits inflammatory responses that follow intestinal ischaemia and reperfusion injury. Br. J. Pharmacol. 2004, 143, 132–142. [Google Scholar] [CrossRef] [PubMed]
  202. Barsante, M.M.; Cunha, T.; Allegretti, M.; Cattani, F.; Policani, F.; Bizzarri, C.; Tafuri, W.L.; Poole, S.; Cunha, F.Q.; Bertini, R.; et al. Blockade of the chemokine receptor CXCR2 ameliorates adjuvant-induced arthritis in rats. Br. J. Pharmacol. 2008, 153, 992–1002. [Google Scholar] [CrossRef] [PubMed]
  203. Burg, N.D.; Pillinger, M. The Neutrophil: Function and Regulation in Innate and Humoral Immunity. Clin. Immunol. 2001, 99, 7–17. [Google Scholar] [CrossRef]
  204. Barnes, P.J. New molecular targets for the treatment of neutrophilic diseases. J. Allergy Clin. Immunol. 2007, 119, 1055–1062. [Google Scholar] [CrossRef] [PubMed]
  205. Keane, M.P.; Belperio, J.A.; Moore, T.A.; Moore, B.B.; Arenberg, D.A.; Smith, R.E.; Burdick, M.D.; Kunkel, S.L.; Strieter, R.M. Neutralization of the CXC chemokine, macro-phage inflammatory protein-2, attenuates bleomycin-induced pulmonary fibrosis. J. Immunol. 1999, 162, 5511–5518. [Google Scholar] [CrossRef]
  206. Julia, V. CX3CL1 in allergic diseases: Not just a chemotactic molecule. Allergy 2012, 67, 1106–1110. [Google Scholar] [CrossRef] [PubMed]
  207. White, G.E.; Greaves, D.R.; Umehara, H.; Bloom, E.T.; Okazaki, T.; Nagano, Y.; Yoshie, O.; Imai, T.; Saederup, N.; Chan, L.; et al. Fractalkine: A Survivor’s Guide. Arter. Thromb. Vasc. Biol. 2012, 32, 589–594. [Google Scholar] [CrossRef]
  208. Ishida, Y.; Gao, J.-L.; Murphy, P.M. Chemokine Receptor CX3CR1 Mediates Skin Wound Healing by Promoting Macrophage and Fibroblast Accumulation and Function. J. Immunol. 2008, 180, 569–579. [Google Scholar] [CrossRef]
  209. El-Shazly, A.; Berger, P.; Girodet, P.-O.; Ousova, O.; Fayon, M.; Vernejoux, J.-M.; Marthan, R.; Tunon-De-Lara, J.M. Fraktalkine Produced by Airway Smooth Muscle Cells Contributes to Mast Cell Recruitment in Asthma. J. Immunol. 2006, 176, 1860–1868. [Google Scholar] [CrossRef]
  210. Xiong, Z.; Leme, A.S.; Ray, P.; Shapiro, S.D.; Lee, J.S. CX3CR1+ Lung Mononuclear Phagocytes Spatially Confined to the Interstitium Produce TNF-α and IL-6 and Promote Cigarette Smoke-Induced Emphysema. J. Immunol. 2011, 186, 3206–3214. [Google Scholar] [CrossRef]
  211. Sari, E.; He, C.; Margaroli, C. Plasticity towards Rigidity: A Macrophage Conundrum in Pulmonary Fibrosis. Int. J. Mol. Sci. 2022, 23, 11443. [Google Scholar] [CrossRef] [PubMed]
  212. Eapen, M.S.; Hansbro, P.M.; McAlinden, K.; Kim, R.Y.; Ward, C.; Hackett, T.-L.; Walters, E.H.; Sohal, S.S. Abnormal M1/M2 macrophage phenotype profiles in the small airway wall and lumen in smokers and chronic obstructive pulmonary disease (COPD). Sci. Rep. 2017, 7, 13392. [Google Scholar] [CrossRef] [PubMed]
  213. Mora, A.L.; Torres-González, E.; Rojas, M.; Corredor, C.; Ritzenthaler, J.; Xu, J.; Roman, J.; Brigham, K.; Stecenko, A. Activation of Alveolar Macrophages via the Alternative Pathway in Herpesvirus-Induced Lung Fibrosis. Am. J. Respir. Cell Mol. Biol. 2006, 35, 466–473. [Google Scholar] [CrossRef] [PubMed]
  214. Hiraiwa, K.; van Eeden, S.F. Contribution of Lung Macrophages to the Inflammatory Responses Induced by Exposure to Air Pollutants. Mediat. Inflamm. 2013, 2013, 619523. [Google Scholar] [CrossRef] [Green Version]
  215. Laskin, D.L.; Sunil, V.R.; Gardner, C.R.; Laskin, J.D. Macrophages and Tissue Injury: Agents of Defense or Destruction? Annu. Rev. Pharmacol. Toxicol. 2011, 51, 267–288. [Google Scholar] [CrossRef]
  216. Weinberger, B.; Laskin, J.D.; Sunil, V.R.; Sinko, P.J.; Heck, D.E.; Laskin, D.L. Sulfur mustard-induced pulmonary injury: Therapeutic approaches to mitigating toxicity. Pulm. Pharmacol. Ther. 2011, 24, 92–99. [Google Scholar] [CrossRef]
  217. Aggarwal, N.R.; King, L.S.; D’Alessio, F.R. Diverse macrophage populations mediate acute lung inflammation and resolution. Am. J. Physiol. Cell. Mol. Physiol. 2014, 306, L709–L725. [Google Scholar] [CrossRef]
  218. Malaviya, R.; Laskin, J.D.; Laskin, D.L. Anti-TNFα therapy in inflammatory lung diseases. Pharmacol. Ther. 2017, 180, 90–98. [Google Scholar] [CrossRef]
  219. Sugiura, H.; Ichinose, M. Nitrative stress in inflammatory lung diseases. Nitric Oxide 2011, 25, 138–144. [Google Scholar] [CrossRef]
  220. Cheng, P.; Li, S.; Chen, H. Macrophages in Lung Injury, Repair, and Fibrosis. Cells 2021, 10, 436. [Google Scholar] [CrossRef]
  221. Du, B.; Shim, J.S. Targeting Epithelial–Mesenchymal Transition (EMT) to Overcome Drug Resistance in Cancer. Molecules 2016, 21, 965. [Google Scholar] [CrossRef] [PubMed]
  222. Willis, B.C. Epithelial Origin of Myofibroblasts during Fibrosis in the Lung. Proc. Am. Thorac. Soc. 2006, 3, 377–382. [Google Scholar] [CrossRef] [PubMed]
  223. Ji, Y.; Dou, Y.-N.; Zhao, Q.-W.; Zhang, J.-Z.; Yang, Y.; Wang, T.; Xia, Y.-F.; Dai, Y.; Wei, Z.-F. Paeoniflorin suppresses TGF-β mediated epithelial-mesenchymal transition in pulmonary fibrosis through a Smad-dependent pathway. Acta Pharmacol. Sin. 2016, 37, 794–804. [Google Scholar] [CrossRef] [PubMed]
  224. Zhang, J.; Tian, X.-J.; Xing, J. Signal Transduction Pathways of EMT Induced by TGF-β, SHH, and WNT and Their Crosstalks. J. Clin. Med. 2016, 5, 41. [Google Scholar] [CrossRef] [PubMed]
  225. Luo, Y.; Yi, H.; Huang, X.; Lin, G.; Kuang, Y.; Guo, Y.; Xie, C. Inhibition of macrophage migration inhibitory factor (MIF) as a therapeutic target in bleomycin-induced pulmonary fibrosis rats. Am. J. Physiol. Cell. Mol. Physiol. 2021, 321, L6–L16. [Google Scholar] [CrossRef] [PubMed]
  226. Vermaelen, K.; Pauwels, R. Pulmonary Dendritic Cells. Am. J. Respir. Crit. Care Med. 2005, 172, 530–551. [Google Scholar] [CrossRef]
  227. Perros, F.; Dorfmüller, P.; Souza, R.; Durand-Gasselin, I.; Mussot, S.; Mazmanian, M.; Herve, P.; Emilie, D.; Simonneau, G.; Humbert, M. Dendritic cell recruitment in lesions of human and experimental pulmonary hypertension. Eur. Respir. J. 2007, 29, 462–468. [Google Scholar] [CrossRef]
  228. Van Rijt, L.; Jung, S.; Kleinjan, A.; Vos, N.; Willart, M.; Duez, C.; Hoogsteden, H.C.; Lambrecht, B.N. In vivo depletion of lung CD11c+ dendritic cells during allergen challenge abrogates the characteristic features of asthma. J. Exp. Med. 2005, 201, 981–991. [Google Scholar] [CrossRef]
  229. Luzina, I.G.; Todd, N.W.; Iacono, A.T.; Atamas, S.P. Roles of T lymphocytes in pulmonary fibrosis. J. Leukoc. Biol. 2007, 83, 237–244. [Google Scholar] [CrossRef]
  230. Kotsianidis, I.; Nakou, E.; Bouchliou, I.; Tzouvelekis, A.; Spanoudakis, E.; Steiropoulos, P.; Sotiriou, I.; Aidinis, V.; Margaritis, D.; Tsatalas, C.; et al. Global Impairment of CD4+CD25+FOXP3+ Regulatory T Cells in Idiopathic Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2009, 179, 1121–1130. [Google Scholar] [CrossRef]
  231. Galati, D.; De Martino, M.; Trotta, A.; Rea, G.; Bruzzese, D.; Cicchitto, G.; Stanziola, A.A.; Napolitano, M.; Sanduzzi, A.; Bocchino, M. Peripheral depletion of NK cells and imbalance of the Treg/Th17 axis in idiopathic pulmonary fibrosis patients. Cytokine 2014, 66, 119–126. [Google Scholar] [CrossRef] [PubMed]
  232. Boveda-Ruiz, D.; D’Alessandro-Gabazza, C.N.; Toda, M.; Takagi, T.; Naito, M.; Matsushima, Y.; Matsumoto, T.; Kobayashi, T.; Gil-Bernabe, P.; Chelakkot-Govindalayathil, A.-L.; et al. Differential role of regulatory T cells in early and late stages of pulmonary fibrosis. Immunobiology 2012, 218, 245–254. [Google Scholar] [CrossRef] [PubMed]
  233. Driscoll, K.E.; Maurer, J.K.; Poynter, J.; Higgins, J.; Asquith, T.; Miller, N.S. Stimulation of rat alveolar macrophage fibronectin release in a cadmium chloride model of lung injury and fibrosis. Toxicol. Appl. Pharmacol. 1992, 116, 30–37. [Google Scholar] [CrossRef] [PubMed]
  234. Kadler, K.E.; Hill, A.; Canty-Laird, E.G. Collagen fibrillogenesis: Fibronectin, integrins, and minor collagens as organizers and nucleators. Curr. Opin. Cell Biol. 2008, 20, 495–501. [Google Scholar] [CrossRef] [PubMed]
  235. Lacronique, J.G.; Rennard, S.I.; Bitterman, P.B.; Ozaki, T.; Crystal, R.G. Alveolar macrophages in idiopathic pulmonary fibrosis have glucocorticoid receptors, but glucocorticoid therapy does not suppress alveolar macrophage release of fibronectin and alveolar macrophage derived growth factor. Am. Rev. Respir. Dis. 1984, 130, 450–456. [Google Scholar] [CrossRef]
  236. Iwano, M.; Plieth, D.; Danoff, T.M.; Xue, C.; Okada, H.; Neilson, E.G. Evidence that fibroblasts derive from epithelium during tissue fibrosis. J. Clin. Investig. 2002, 110, 341–350. [Google Scholar] [CrossRef]
  237. Venkatesan, N.; Roughley, P.J.; Ludwig, M.S. Proteoglycan expression in bleomycin lung fibroblasts: Role of transforming growth factor-β1 and interferon-γ. Am. J. Physiol. Cell. Mol. Physiol. 2002, 283, L806–L814. [Google Scholar] [CrossRef]
  238. Bensadoun, E.S.; Burke, A.K.; Hogg, J.C.; Roberts, C.R. Proteoglycan deposition in pulmonary fibrosis. Am. J. Respir. Crit. Care Med. 1996, 154, 1819–1828. [Google Scholar] [CrossRef]
  239. McGowan, S.E. Extracellular matrix and the regulation of lung development and repair. FASEB J. 1992, 6, 2895–2904. [Google Scholar] [CrossRef]
  240. Parra, E.R.; Kairalla, R.A.; de Carvalho, C.R.R.; Capelozzi, V.L. Abnormal deposition of collagen/elastic vascular fibres and prognostic significance in idiopathic interstitial pneumonias. Thorax 2007, 62, 428–437. [Google Scholar] [CrossRef]
  241. Farahani, R.M.; Kloth, L.C. The hypothesis of ‘biophysical matrix contraction’: Wound contraction revisited. Int. Wound J. 2008, 5, 477–482. [Google Scholar] [CrossRef] [PubMed]
  242. Mirastschijski, U.; Haaksma, C.J.; Tomasek, J.J.; Ågren, M.S. Matrix metalloproteinase inhibitor GM 6001 attenuates keratinocyte migration, contraction and myofibroblast formation in skin wounds. Exp. Cell Res. 2004, 299, 465–475. [Google Scholar] [CrossRef] [PubMed]
  243. Hinz, B. Masters and servants of the force: The role of matrix adhesions in myofibroblast force perception and transmission. Eur. J. Cell Biol. 2006, 85, 175–181. [Google Scholar] [CrossRef] [PubMed]
  244. Thannickal, V.J. Evolving Concepts of Apoptosis in Idiopathic Pulmonary Fibrosis. Proc. Am. Thorac. Soc. 2006, 3, 350–356. [Google Scholar] [CrossRef]
  245. García-Alvarez, J.; Ramirez, R.; Checa, M.; Nuttall, R.K.; Sampieri, C.L.; Edwards, D.R.; Selman, M.; Pardo, A. Tissue inhibitor of metalloproteinase-3 is up-regulated by transforming growth factor-β1 in vitro and expressed in fibroblastic foci in vivo in idiopathic pulmonary fibrosis. Exp. Lung Res. 2006, 32, 201–214. [Google Scholar] [CrossRef] [PubMed]
  246. Gill, S.E.; Parks, W.C. Metalloproteinases and their inhibitors: Regulators of wound healing. Int. J. Biochem. Cell Biol. 2008, 40, 1334–1347. [Google Scholar] [CrossRef]
  247. Fattman, C.L. Apoptosis in Pulmonary Fibrosis: Too Much or Not Enough? Antioxidants Redox Signal. 2008, 10, 379–386. [Google Scholar] [CrossRef]
  248. Moodley, Y.P.; Caterina, P.; Scaffidi, A.K.; Misso, N.L.; Papadimitriou, J.M.; McAnulty, R.J.; Laurent, G.J.; Thompson, P.J.; Knight, D.A. Comparison of the morphological and biochemical changes in normal human lung fibroblasts and fibroblasts derived from lungs of patients with idiopathic pulmonary fibrosis during FasL-induced apoptosis. J. Pathol. 2004, 202, 486–495. [Google Scholar] [CrossRef]
  249. Frankel, S.K.; Cosgrove, G.P.; Cha, S.-I.; Cool, C.D.; Wynes, M.W.; Edelman, B.L.; Brown, K.K.; Riches, D.W.H. TNF-α Sensitizes Normal and Fibrotic Human Lung Fibroblasts to Fas-Induced Apoptosis. Am. J. Respir. Cell Mol. Biol. 2006, 34, 293–304. [Google Scholar] [CrossRef]
  250. Fernandez, I.E.; Eickelberg, O. The Impact of TGF-β on Lung Fibrosis: From targeting to biomarkers. Proc. Am. Thorac. Soc. 2012, 9, 111–116. [Google Scholar] [CrossRef]
  251. Ask, K.; Bonniaud, P.; Maass, K.; Eickelberg, O.; Margetts, P.J.; Warburton, D.; Groffen, J.; Gauldie, J.; Kolb, M. Progressive pulmonary fibrosis is mediated by TGF-β isoform 1 but not TGF-β3. Int. J. Biochem. Cell Biol. 2008, 40, 484–495. [Google Scholar] [CrossRef]
  252. Roberts, A.B.; Russo, A.; Felici, A.; Flanders, K.C. Smad3: A Key Player in Pathogenetic Mechanisms Dependent on TGF-β. Ann. N. Y. Acad. Sci. 2003, 995, 1–10. [Google Scholar] [CrossRef] [PubMed]
  253. Luo, L.; Wang, C.-C.; Song, X.-P.; Wang, H.-M.; Zhou, H.; Sun, Y.; Wang, X.-K.; Hou, S.; Pei, F.-Y. Suppression of SMOC2 reduces bleomycin (BLM)-induced pulmonary fibrosis by inhibition of TGF-β1/SMADs pathway. Biomed. Pharmacother. 2018, 105, 841–847. [Google Scholar] [CrossRef] [PubMed]
  254. Shen, X.-B.; Ding, D.-L.; Yu, L.-Z.; Ni, J.-Z.; Liu, Y.; Wang, W.; Liu, L.-M.; Nian, S.-H. Total extract of Anemarrhenae Rhizoma attenuates bleomycin-induced pulmonary fibrosis in rats. Bioorg. Chem. 2021, 119, 105546. [Google Scholar] [CrossRef] [PubMed]
  255. Câmara, J.; Jarai, G. Epithelial-mesenchymal transition in primary human bronchial epithelial cells is Smad-dependent and enhanced by fibronectin and TNF-α. Fibrogenesis Tissue Repair 2010, 3, 2. [Google Scholar] [CrossRef]
  256. Wanas, H.; El Shereef, Z.; Rashed, L.; Aboulhoda, B.E. Ticagrelor Ameliorates Bleomycin-Induced Pulmonary Fibrosis in Rats by the Inhibition of TGF-β1/Smad3 and PI3K/AKT/mTOR Pathways. Curr. Mol. Pharmacol. 2021, 15, 227–238. [Google Scholar] [CrossRef]
  257. Bonner, J.C. Regulation of PDGF and its receptors in fibrotic diseases. Cytokine Growth Factor Rev. 2004, 15, 255–273. [Google Scholar] [CrossRef]
  258. Antoniades, H.N.; Bravo, M.A.; Avila, R.E.; Galanopoulos, T.; Neville-Golden, J.; Maxwell, M.; Selman, M. Platelet-derived growth factor in idiopathic pulmonary fibrosis. J. Clin. Investig. 1990, 86, 1055–1064. [Google Scholar] [CrossRef]
  259. Hilberg, F.; Roth, G.J.; Krssak, M.; Kautschitsch, S.; Sommergruber, W.; Tontsch-Grunt, U.; Garin-Chesa, P.; Bader, G.; Zoephel, A.; Quant, J.; et al. BIBF 1120: Triple Angiokinase Inhibitor with Sustained Receptor Blockade and Good Antitumor Efficacy. Cancer Res. 2008, 68, 4774–4782. [Google Scholar] [CrossRef]
  260. Abreu, J.G.; Ketpura, N.I.; Reversade, B.; De Robertis, E.M. Connective-tissue growth factor (CTGF) modulates cell signalling by BMP and TGF-β. Nature 2002, 4, 599–604. [Google Scholar] [CrossRef]
  261. Lasky, J.A.; Ortiz, L.A.; Tonthat, B.; Hoyle, G.W.; Corti, M.; Athas, G.; Lungarella, G.; Brody, A.; Friedman, M. Connective tissue growth factor mRNA expression is upregulated in bleomycin-induced lung fibrosis. Am. J. Physiol. Cell. Mol. Physiol. 1998, 275, L365–L371. [Google Scholar] [CrossRef] [PubMed]
  262. Pan, L.-H.; Yamauchi, K.; Uzuki, M.; Nakanishi, T.; Takigawa, M.; Inoue, H.; Sawai, T. Type II alveolar epithelial cells and interstitial fibroblasts express connective tissue growth factor in IPF. Eur. Respir. J. 2001, 17, 1220–1227. [Google Scholar] [CrossRef] [PubMed]
  263. Krein, P.M.; Winston, B.W. Roles for Insulin-Like Growth Factor I and Transforming Growth Factor-β in Fibrotic Lung Disease. Chest 2002, 122, 289S–293S. [Google Scholar] [CrossRef] [PubMed]
  264. Klement, R.J.; Fink, M.K. Dietary and pharmacological modification of the insulin/IGF-1 system: Exploiting the full repertoire against cancer. Oncogenesis 2016, 5, e193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Bokum, T.; Hofland, L.J.; Jong, D.; Bouma, J.; Melief, M.J.; Kwekkeboom, D.J.; Schonbrunn, A.; Mooy, C.M.; Laman, J.D.; Lamberts, J.; et al. Immunohistochemical localization of somatostatin receptor sst2A in sarcoid granulomas. Eur. J. Clin. Investig. 1999, 29, 630–636. [Google Scholar] [CrossRef] [PubMed]
  266. Zhao, Y.; Yan, Z.; Liu, Y.; Zhang, Y.; Shi, J.; Li, J.; Ji, F. Effectivity of mesenchymal stem cells for bleomycin-induced pulmonary fibrosis: A systematic review and implication for clinical application. Stem Cell Res. Ther. 2021, 12, 470. [Google Scholar] [CrossRef]
  267. Kletukhina, S.; Mutallapova, G.; Titova, A.; Gomzikova, M. Role of Mesenchymal Stem Cells and Extracellular Vesicles in Idiopathic Pulmonary Fibrosis. Int. J. Mol. Sci. 2022, 23, 11212. [Google Scholar] [CrossRef] [PubMed]
  268. Zhang, Q.; Shi, S.; Liu, Y.; Uyanne, J.; Shi, Y.; Shi, S.; Le, A.D. Mesenchymal Stem Cells Derived from Human Gingiva Are Capable of Immunomodulatory Functions and Ameliorate Inflammation-Related Tissue Destruction in Experimental Colitis. J. Immunol. 2009, 183, 7787–7798. [Google Scholar] [CrossRef]
  269. Strieter, R.M.; Keeley, E.C.; Hughes, M.A.; Burdick, M.D.; Mehrad, B. The role of circulating mesenchymal progenitor cells (fibrocytes) in the pathogenesis of pulmonary fibrosis. J. Leukoc. Biol. 2009, 86, 1111–1118. [Google Scholar] [CrossRef]
  270. Pilling, D.; Fan, T.; Huang, D.; Kaul, B.; Gomer, R. Identification of Markers that Distinguish Monocyte-Derived Fibrocytes from Monocytes, Macrophages, and Fibroblasts. PLoS ONE 2009, 4, e7475. [Google Scholar] [CrossRef]
  271. Gomperts, B.N.; Strieter, R.M. Fibrocytes in lung disease. J. Leukoc. Biol. 2007, 82, 449–456. [Google Scholar] [CrossRef] [PubMed]
  272. Moore, B.B.; Thannickal, V.J.; Toews, G.B. Bone Marrow-Derived Cells in the Pathogenesis of Lung Fibrosis. Curr. Respir. Med. Rev. 2005, 1, 69–76. [Google Scholar] [CrossRef] [PubMed]
  273. Quan, T.E.; Cowper, S.E.; Bucala, R. The role of circulating fibrocytes in fibrosis. Curr. Rheumatol. Rep. 2006, 8, 145–150. [Google Scholar] [CrossRef] [PubMed]
  274. Chesney, J.; Metz, C.; Stavitsky, A.B.; Bacher, M.; Bucala, R. Regulated production of type I collagen and inflammatory cytokines by peripheral blood fibrocytes. J. Immunol. 1998, 160, 419–425. [Google Scholar] [CrossRef]
  275. Kleaveland, K.R.; Moore, B.B.; Kim, K.K. Paracrine functions of fibrocytes to promote lung fibrosis. Expert Rev. Respir. Med. 2014, 8, 163–172. [Google Scholar] [CrossRef]
  276. Moeller, A.; Gilpin, S.E.; Ask, K.; Cox, G.; Cook, D.; Gauldie, J.; Margetts, P.J.; Farkas, L.; Dobranowski, J.; Boylan, C.; et al. Circulating Fibrocytes Are an Indicator of Poor Prognosis in Idiopathic Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2009, 179, 588–594. [Google Scholar] [CrossRef]
  277. Li, Y.; Jiang, D.; Liang, J.; Meltzer, E.B.; Gray, A.; Miura, R.; Wogensen, L.; Yamaguchi, Y.; Noble, P.W. Severe lung fibrosis requires an invasive fibroblast phenotype regulated by hyaluronan and CD44. J. Exp. Med. 2011, 208, 1459–1471. [Google Scholar] [CrossRef]
  278. Chen, X.; Shi, C.; Meng, X.; Zhang, K.; Li, X.; Wang, C.; Xiang, Z.; Hu, K.; Han, X. Inhibition of Wnt/β-catenin signaling suppresses bleomycin-induced pulmonary fibrosis by attenuating the expression of TGF-β1 and FGF-2. Exp. Mol. Pathol. 2016, 101, 22–30. [Google Scholar] [CrossRef]
  279. Holgate, S.T.; Peters-Golden, M.; Panettieri, R.A.; Henderson, W.R. Roles of cysteinyl leukotrienes in airway inflammation, smooth muscle function, and remodeling. J. Allergy Clin. Immunol. 2003, 111, S18–S36. [Google Scholar] [CrossRef]
  280. Wang, C.; Zhu, H.; Sun, Z.; Xiang, Z.; Ge, Y.; Ni, C.; Luo, Z.; Qian, W.; Han, X. Inhibition of Wnt/β-catenin signaling promotes epithelial differentiation of mesenchymal stem cells and repairs bleomycin-induced lung injury. Am. J. Physiol. Physiol. 2014, 307, C234–C244. [Google Scholar] [CrossRef]
  281. Chen, T.; Nie, H.; Gao, X.; Yang, J.; Pu, J.; Chen, Z.; Cui, X.; Wang, Y.; Wang, H.; Jia, G. Epithelial–mesenchymal transition involved in pulmonary fibrosis induced by multi-walled carbon nanotubes via TGF-β/Smad signaling pathway. Toxicol. Lett. 2014, 226, 150–162. [Google Scholar] [CrossRef] [PubMed]
  282. Rangarajan, S.; Bone, N.B.; Zmijewska, A.A.; Jiang, S.; Park, D.W.; Bernard, K.; Locy, M.L.; Ravi, S.; Deshane, J.; Mannon, R.B.; et al. Metformin reverses established lung fibrosis in a bleomycin model. Nat. Med. 2018, 24, 1121–1127. [Google Scholar] [CrossRef] [PubMed]
  283. Foretz, M.; Guigas, B.; Bertrand, L.; Pollak, M.; Viollet, B. Metformin: From Mechanisms of Action to Therapies. Cell Metab. 2014, 20, 953–966. [Google Scholar] [CrossRef] [PubMed]
  284. Wollin, L.; Wex, E.; Pautsch, A.; Schnapp, G.; Hostettler, K.E.; Stowasser, S.; Kolb, M. Mode of action of nintedanib in the treatment of idiopathic pulmonary fibrosis. Eur. Respir. J. 2015, 45, 1434–1445. [Google Scholar] [CrossRef] [Green Version]
  285. Méndez-Ferrer, S.; Michurina, T.V.; Ferraro, F.; Mazloom, A.R.; MacArthur, B.D.; Lira, S.A.; Scadden, D.T.; Ma’Ayan, A.; Enikolopov, G.N.; Frenette, P.S. Mesenchymal and haematopoietic stem cells form a unique bone marrow niche. Nature 2010, 466, 829–834. [Google Scholar] [CrossRef]
  286. Jiang, M.H.; Cai, B.; Tuo, Y.; Wang, J.; Zang, Z.J.; Tu, X.; Gao, Y.; Su, Z.; Li, W.; Li, G.; et al. Characterization of Nestin-positive stem Leydig cells as a potential source for the treatment of testicular Leydig cell dysfunction. Cell Res. 2014, 24, 1466–1485. [Google Scholar] [CrossRef]
  287. Wang, J.; Lai, X.; Yao, S.; Chen, H.; Cai, J.; Luo, Y.; Wang, Y.; Qiu, Y.; Huang, Y.; Wei, X.; et al. Nestin promotes pulmonary fibrosis via facilitating recycling of TGF-β receptor I. Eur. Respir. J. 2021, 59, 2003721. [Google Scholar] [CrossRef]
  288. Dupont, S.; Mamidi, A.; Cordenonsi, M.; Montagner, M.; Zacchigna, L.; Adorno, M.; Martello, G.; Stinchfield, M.J.; Soligo, S.; Morsut, L.; et al. FAM/USP9x, a Deubiquitinating Enzyme Essential for TGFβ Signaling, Controls Smad4 Monoubiquitination. Cell 2009, 136, 123–135. [Google Scholar] [CrossRef]
  289. Quéré, R.; Saint-Paul, L.; Carmignac, V.; Martin, R.Z.; Chrétien, M.-L.; Largeot, A.; Hammann, A.; de Barros, J.-P.P.; Bastie, J.-N.; Delva, L. Tif1γ regulates the TGF-β1 receptor and promotes physiological aging of hematopoietic stem cells. Proc. Natl. Acad. Sci. USA 2014, 111, 10592–10597. [Google Scholar] [CrossRef]
  290. Boutanquoi, P.-M.; Burgy, O.; Beltramo, G.; Bellaye, P.-S.; Dondaine, L.; Marcion, G.; Pommerolle, L.; Vadel, A.; Spanjaard, M.; Demidov, O.; et al. TRIM33 prevents pulmonary fibrosis by impairing TGF-β1 signalling. Eur. Respir. J. 2020, 55, 1901346. [Google Scholar] [CrossRef]
  291. Zhao, H.; Li, C.; Li, L.; Liu, J.; Gao, Y.; Mu, K.; Chen, D.; Lu, A.; Ren, Y.; Li, Z. Baicalin alleviates bleomycin-induced pulmonary fibrosis and fibroblast proliferation in rats via the PI3K/AKT signaling pathway. Mol. Med. Rep. 2020, 21, 2321–2334. [Google Scholar] [CrossRef] [PubMed]
  292. Takahashi, K.; Tanabe, K.; Ohnuki, M.; Narita, M.; Ichisaka, T.; Tomoda, K.; Yamanaka, S. Induction of Pluripotent Stem Cells from Adult Human Fibroblasts by Defined Factors. Cell 2007, 131, 861–872. [Google Scholar] [CrossRef] [PubMed]
  293. Cocozza, F.; Grisard, E.; Jaular, L.M.; Mathieu, M.; Théry, C. SnapShot: Extracellular Vesicles. Cell 2020, 182, 262.e1. [Google Scholar] [CrossRef] [PubMed]
  294. Negrete-García, M.C.; Ramos-Abundis, J.D.J.; Alvarado-Vasquez, N.; Montes-Martínez, E.; Montaño, M.; Ramos, C.; Sommer, B. Exosomal Micro-RNAs as Intercellular Communicators in Idiopathic Pulmonary Fibrosis. Int. J. Mol. Sci. 2022, 23, 11047. [Google Scholar] [CrossRef] [PubMed]
  295. Stahl, P.D.; Raposo, G. Extracellular Vesicles: Exosomes and Microvesicles, Integrators of Homeostasis. Physiology 2019, 34, 169–177. [Google Scholar] [CrossRef]
  296. Wen, C.; Seeger, R.C.; Fabbri, M.; Wang, L.; Wayne, A.S.; Jong, A.Y. Biological roles and potential applications of immune cell-derived extracellular vesicles. J. Extracell. Vesicles 2017, 6, 1400370. [Google Scholar] [CrossRef]
  297. Akbari, A.; Rezaie, J. Potential therapeutic application of mesenchymal stem cell-derived exosomes in SARS-CoV-2 pneumonia. Stem Cell Res. Ther. 2020, 11, 356. [Google Scholar] [CrossRef]
  298. Khalaj, K.; Figueira, R.L.; Antounians, L.; Lauriti, G.; Zani, A. Systematic review of extracellular vesicle-based treatments for lung injury: Are EVs a potential therapy for COVID-19? J. Extracell. Vesicles 2020, 9, 1795365. [Google Scholar] [CrossRef]
  299. Chen, L.; Yang, Y.; Yue, R.; Peng, X.; Yu, H.; Huang, X. Exosomes derived from hypoxia-induced alveolar epithelial cells stimulate interstitial pulmonary fibrosis through a HOTAIRM1-dependent mechanism. Lab. Investig. 2022, 102, 935–944. [Google Scholar] [CrossRef]
  300. Kadota, T.; Yoshioka, Y.; Fujita, Y.; Araya, J.; Minagawa, S.; Hara, H.; Miyamoto, A.; Suzuki, S.; Fujimori, S.; Kohno, T.; et al. Extracellular Vesicles from Fibroblasts Induce Epithelial-Cell Senescence in Pulmonary Fibrosis. Am. J. Respir. Cell Mol. Biol. 2020, 63, 623–636. [Google Scholar] [CrossRef]
  301. Chanda, D.; Otoupalova, E.; Hough, K.P.; Locy, M.L.; Bernard, K.; Deshane, J.S.; Sanderson, R.D.; Mobley, J.A.; Thannickal, V.J. Fibronectin on the Surface of Extracellular Vesicles Mediates Fibroblast Invasion. Am. J. Respir. Cell Mol. Biol. 2019, 60, 279–288. [Google Scholar] [CrossRef] [PubMed]
  302. Liu, X.; Qin, X.; Qin, H.; Jia, C.; Yuan, Y.; Sun, T.; Chen, B.; Chen, C.; Zhang, H. Characterization of the heterogeneity of endothelial cells in bleomycin-induced lung fibrosis using single-cell RNA sequencing. Angiogenesis 2021, 24, 809–821. [Google Scholar] [CrossRef] [PubMed]
  303. Rahman, M.; Wang, Z.-Y.; Li, J.-X.; Xu, H.-W.; Wang, R.; Wu, Q. Single-Cell RNA Sequencing Reveals the Interaction of Injected ADSCs with Lung-Originated Cells in Mouse Pulmonary Fibrosis. Stem Cells Int. 2022, 2022, 9483166. [Google Scholar] [CrossRef] [PubMed]
  304. Aran, D.; Looney, A.P.; Liu, L.; Wu, E.; Fong, V.; Hsu, A.; Chak, S.; Naikawadi, R.P.; Wolters, P.J.; Abate, A.R.; et al. Reference-based analysis of lung single-cell sequencing reveals a transitional profibrotic macrophage. Nat. Immunol. 2019, 20, 163–172. [Google Scholar] [CrossRef]
  305. Peyser, R.; MacDonnell, S.; Gao, Y.; Cheng, L.; Kim, Y.; Kaplan, T.; Ruan, Q.; Wei, Y.; Ni, M.; Adler, C.; et al. Defining the Activated Fibroblast Population in Lung Fibrosis Using Single-Cell Sequencing. Am. J. Respir. Cell Mol. Biol. 2019, 61, 74–85. [Google Scholar] [CrossRef]
  306. Xie, T.; Wang, Y.; Deng, N.; Huang, G.; Taghavifar, F.; Geng, Y.; Liu, N.; Kulur, V.; Yao, C.; Chen, P.; et al. Single-Cell Deconvolution of Fibroblast Heterogeneity in Mouse Pulmonary Fibrosis. Cell Rep. 2018, 22, 3625–3640. [Google Scholar] [CrossRef]
  307. Ju, N.; Hayashi, H.; Shimamura, M.; Baba, S.; Yoshida, S.; Morishita, R.; Rakugi, H.; Nakagami, H. Prevention of bleomycin-induced pulmonary fibrosis by a RANKL peptide in mice. Sci. Rep. 2022, 12, 12474. [Google Scholar] [CrossRef]
  308. Mansour, S.M.; El-Abhar, H.S.; Soubh, A.A. MiR-200a inversely correlates with Hedgehog and TGF-β canonical/non-canonical trajectories to orchestrate the anti-fibrotic effect of Tadalafil in a bleomycin-induced pulmonary fibrosis model. Inflammopharmacology 2020, 29, 167–182. [Google Scholar] [CrossRef]
  309. Shen, Y.-H.; Cheng, M.-H.; Liu, X.-Y.; Zhu, D.-W.; Gao, J. Sodium Houttuyfonate Inhibits Bleomycin Induced Pulmonary Fibrosis in Mice. Front. Pharmacol. 2021, 12, 204. [Google Scholar] [CrossRef]
  310. Arora, A.; Bhuria, V.; Hazari, P.P.; Pathak, U.; Mathur, S.; Roy, B.G.; Sandhir, R.; Soni, R.; Dwarakanath, B.S.; Bhatt, A.N. Amifostine Analog, DRDE-30, Attenuates Bleomycin-Induced Pulmonary Fibrosis in Mice. Front. Pharmacol. 2018, 9, 394. [Google Scholar] [CrossRef]
  311. Tanaka, J.; Tajima, S.; Asakawa, K.; Sakagami, T.; Moriyama, H.; Takada, T.; Suzuki, E.; Narita, I. Preventive effect of irbesartan on bleomycin-induced lung injury in mice. Respir. Investig. 2013, 51, 76–83. [Google Scholar] [CrossRef] [PubMed]
  312. Kuwano, K.; Hagimoto, N.; Kawasaki, M.; Yatomi, T.; Nakamura, N.; Nagata, S.; Suda, T.; Kunitake, R.; Maeyama, T.; Miyazaki, H.; et al. Essential roles of the Fas-Fas ligand pathway in the development of pulmonary fibrosis. J. Clin. Investig. 1999, 104, 13–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Mechanism of wound healing and pathophysiology of pulmonary fibrosis. The three-step model of injury and wound healing describes the distinct stages of a successful response to injury. After lung injury, epithelial cells release inflammatory mediators and initiate an antifibrinolytic coagulation cascade that triggers platelet activation and thrombus formation. Leukocyte infiltration then begins. Activated macrophages and neutrophils eliminate dead cells and invading organisms. Subsequently, bone marrow-derived fibroblasts and intrinsic fibroblasts proliferate and differentiate into myofibroblasts, which release extracellular matrix (ECM) components. Fibroblasts and myofibroblasts can also be derived from epithelial cells that have undergone epithelial–mesenchymal transition (EMT). In the final stage, activated myofibroblasts can promote wound repair, resulting in wound contraction and vascular recovery. However, if any stage of the tissue repair program becomes out of control or if lung-damaging stimuli persist, fibrosis often develops.
Figure 1. Mechanism of wound healing and pathophysiology of pulmonary fibrosis. The three-step model of injury and wound healing describes the distinct stages of a successful response to injury. After lung injury, epithelial cells release inflammatory mediators and initiate an antifibrinolytic coagulation cascade that triggers platelet activation and thrombus formation. Leukocyte infiltration then begins. Activated macrophages and neutrophils eliminate dead cells and invading organisms. Subsequently, bone marrow-derived fibroblasts and intrinsic fibroblasts proliferate and differentiate into myofibroblasts, which release extracellular matrix (ECM) components. Fibroblasts and myofibroblasts can also be derived from epithelial cells that have undergone epithelial–mesenchymal transition (EMT). In the final stage, activated myofibroblasts can promote wound repair, resulting in wound contraction and vascular recovery. However, if any stage of the tissue repair program becomes out of control or if lung-damaging stimuli persist, fibrosis often develops.
Ijms 24 03149 g001
Table 1. Studies of BLM-induced PF model focused on immune mechanism.
Table 1. Studies of BLM-induced PF model focused on immune mechanism.
ApproachPF vs. ControlsDescriptionReferences
Mmp3−/−MMP-3 induced EMT through activation of β-catenin signaling.[26]
Mmp7−/−Decreased pulmonary inflammation and fibrosis.[27]
Mmp8−/−Increased lung inflammation, but reduced fibrosis by enhancing lung levels of IP-10, MIP-1α, and IL-10.[28,29]
Mmp9−/−No changeMinimal alveolar bronchiolization.[30]
Transgenic mice overexpressing MMP-9 in alveolar macrophagesReduced neutrophils and lymphocytes in BAL. Antifibrotic effect via IGFBP3 degradation.[31]
Mmp12−/−No changeSimilar number of pulmonary macrophages and levels of MMP-2 and MMP-9 as in WT mice.[32]
Mmp13−/−Exaggerated pulmonary inflammation and fibrosis.[33]
Mmp19−/−Antifibrotic effect through decreased PTGS2.[34]
Mmp28−/−Deficits in M2 polarization.[35]
IL-1R1−/− and MyD88−/−Blockade of IL-1R1 by IL-1 receptor antagonist also reduced inflammation and fibrosis. BLM-induced PF required the inflammasome and IL-1R1/MyD88 signaling.[36]
MyD88−/−Reduced lung inflammation and fibrosis by reducing TGF-β signaling. Human placental MSCs of fetal origin (hfPMSCs) also attenuated fibrosis with reduced MyD88 and TGF-β signaling activation.[37]
IL-10−/−No changeIncreased inflammatory cells in BALF.[38]
IL-12p40−/−Less pulmonary mononuclear cell inflammation, whereas increased PF.[39]
IL-24−/−Attenuated M2 macrophage recruitment and TGF-β1 production.[40]
IL-27 treatmentAlleviate pulmonary fibrosis by regulating Th17 differentiation.[41]
IL-1R1−/−, IL-23p19−/−, and IL-17RA−/−IL-1β-IL-23-IL-17 axis is essential for pulmonary inflammation and fibrosis.[42]
IL-24−/−Attenuated TGF-β1 production and reduced IL-4-induced production of M2 macrophages.[40]
IL-33 receptor ST2−/−Augmented acute inflammation, but decreased lung fibrosis with reduced M2 macrophages.[43]
CCL3−/− and CCR5−/−Collagen accumulation was reduced in CCL3−/− and CCR5−/− but not CCR1−/− mice. The CCL3–CCR5 axis was involved in the accumulation of macrophages and fibrocytes.[44]
CCR2−/−Fewer macrophages in BALF and macrophage-derived MMP-2 and MMP-9 production. Reduced TNF-α and TGF-β1.[45,46,47]
CCR4−/−CCL17-dependent activation of CCR4+ macrophages played a central role in free radical-induced pulmonary injury.[48]
Anti-MIP-1α antibodyReduced pulmonary mononuclear phagocyte accumulation and fibrosis.[49]
CCL18 overexpressionExacerbated inflammation but attenuated collagen accumulation.[50]
CXCR2 antagonist, DF2162Attenuated collagen deposition with reduced TGF-β1, IL-13, and IL-8.[51]
anti-CXCL6 antibodyReduced pulmonary neutrophil influx and collagen deposition.[52]
anti-CXCL12 antibodyAttenuated recruitment of CD45+Col I+CXCR4+ fibrocytes and attenuated fibrosis.[53]
CXCR4 antagonist, TN14003 or AMD3100Inhibited the migration of human fibrocytes in response to CXCL12 in vitro and reduced the trafficking of fibrocytes into the murine lungs in vivo.[54,55]
Recombinant CXCL12 and CCL21N.A.Fibrocytes expressed CXCR4 and CCR7 and could chemotactically respond to both CXCL12 and CCL21.[56]
CX3CR1−/−Similar levels of inflammation but reduced fibrosis. CX3CR1+ macrophages displayed profibrotic M2 phenotypes. Fibrocytes expressed CX3CR1.[57]
Blocking pulmonary macrophage infiltrationSuppression of the Wnt/β-catenin signaling pathway could attenuate myofibroblast differentiation of lung resident mesenchymal stem cells induced by M2 macrophages and fibrosis.[58]
M2 macrophagesN.A.F4/80+CD11c+CD206+ M2 alveolar macrophages peaked on day 14 and were correlated with the magnitude of fibrosis.[59]
M2 macrophagesM2 macrophages induced EMT through the TGF-β/Smad2 signaling pathway.[60]
Depletion of lung macrophagesDepletion of Ly6Chi circulating monocytes also reduced PF.[61]
Dexamethasone treatmentReduced the inflammation, lung damage, and fibrosis.[62]
Surfactant protein D (SP-D)−/− or SP-D-overexpressionSP-D−/−, ↑; SP-D-overexpression, ↓SP-D had an anti-inflammatory role via modulation of oxidative–nitrative stress.[63]
Clevudine treatmentSuppressed profibrotic phenotype of M2 macrophages by inhibiting PI3K/Akt signaling pathway.[64]
Targeting CD206 receptor, RP-832cDecreased expression of inflammatory cytokines and TGF-β1 and subsequent fibrosis.[65]
Effective-compound combination (ECC)Reduced M2 polarization through the promotion of autophagy via mTOR signaling suppression.[66]
C/EBP homologous protein (Chop)−/−Attenuated M2 macrophage recruitment and TGF-β1 production.[67]
p-JAK2/p-STAT3 inhibitor, JSI-124JSI-124 could modulate fibrosis, autophagy, senescence, and anti-apoptosis.[68]
A macrolide antibiotic, tacrolimusInhibited JAK2/STAT3 signaling in macrophages and reduced the release of profibrotic factors secreted by M2 macrophages.[69]
DC inhibitor, VAG539Decreased inflammation and fibrosis.[70]
pDC depletion, anti-PDCA-1Reduces immune cell infiltration and reduces the expression of genes and proteins involved in chemotaxis, inflammation and fibrosis.[71]
Anti-CD3 antibodySuppressed IL-2 and IL-4 production by lung lymphocytes.[72]
Adoptive Treg transferSuppressed FGF9 and CCL2 production. IL-10−/− abolished the ameliorative effect of Tregs on PF.[73]
Extra type III domain A (EDA)−/−Decreased TGF-β expression and fibrosis. EDA-containing fibronectin played a critical role in fibrogenesis.[74]
FibronectinN.A.Fibronectin accumulates during the early inflammatory phase, parallelling hyaluronan accumulation and preceding the development of PF.[75,76]
Smad3−/−Suppressed type I procollagen gene expression and reduced collagen deposition.[25]
A prenylated flavonoid, kurarinoneSuppressed the TGF-β-induced EMT via suppressed phosphorylation of Smad2/3 and AKT.[77]
A tyrosine kinase inhibitor, imatinibSuppressed fibrosis by reducing the number of mesenchymal cells incorporating bromodeoxyuridine.[78]
A small molecule inhibitor targeting the receptor kinases of PDGF, bFGF and VEGF, BIBF 1000Reduced collagen deposition and profibrotic gene expression.[79]
Anti-CTGF scFvReduced the numbers of inflammatory leukocytes in BALF and the hydroxyproline content of lung tissue.[80]
Anti-IGF-I receptor antibody, A12Increased fibroblast apoptosis and subsequent resolution of PF. IGF-I increased lung fibroblast migration via the insulin receptor substrate-2/phosphatidylinositol 3-kinase/Akt axis.[81]
Antidiabetic agent, MetforminInhibited enhanced IGF-1 and PI3K expression.[82]
A somatostatin analogue, SOM230 (pasireotide)Reduced BAL inflammatory cell influx and expression of CTGF and TGF-β.[83]
Somatotropin-release inhibiting factors (SRIFs), somatostatinsReduced the number of neutrophils and lymphocytes, and IGF-1 level in BALF.[84]
CD19−/− and CD19 transgenic miceCD19−/−, ↓; CD19 Tg, ↑The levels of IL-6 and immunoglobulin in BALF correlated with CD19 expression levels and the severity of PF.[85]
Amniotic mesenchymal stromal cells (hAMSCs) injectionhAMSCs reduced pulmonary B-cell recruitment, retention, and maturation, and they counteract the formation and expansion of intrapulmonary lymphoid aggregates.[86]
Gingival-derived mesenchymal stem cells (GMSCs) administrationReduced neutrophil elastase, MMP-9, lysophosphatidic acid, lysophosphatidic acid receptor 1, and TGF-β release.[87]
Recombinant Slit2Fibroblast-derived Slit2 inhibits fibrocyte differentiation.[88]
iPS cellsRepressed the expression ratios of MMP-2/TIMP-2 and MMP-9/TIMP-1. Inhibited activation of TGF-β1-Smad2/3 signaling and EMT.[89]
An exosomes inhibitor, GW4869Macrophage-derived exosomes containing the angiotensin II type 1 receptor (AT1R) had profibrotic effects by upregulating the Ang II/AT1R axis in fibroblasts.[90]
MSC-derived exosomes (MEx)Transplantation of BM-derived monocytes that were preconditioned by MEx prevented collagen deposition, restored lung architecture.[91]
Extracellular vesicles (EVs)N.A.EVs was found in PF, carry fibrotic mediator such as WNT5A, and contribute to fibrogenesis.[92]
GSK-3 inhibitionDecreased MMP-2 and MMP-9, and enhanced ECM deposition.[93]
Uric acid releaseConstitutes an endogenous danger signal that activates the NALP3 inflammasome, leading to IL-1β production, and fibrosis.[94]
FluorofenidoneAntifibrotic effect include regulating caveolin 1 expression and blocking MAPK signaling pathways and/or the IL-1β/IL-1R1/MyD88/NF-κB pathway.[95,96]
Akt1−/−Decreased production of profibrotic cytokine, IL-13, by macrophages.[97]
Genetic ablation or neutralization of B-cell activating factor (BAFF)Reduced the production of TGF-β1 and collagen deposition.[98]
proteinase-activated receptor-1 (PAR1)−/−Reduced inflammation and fibrosis with attenuation in CCL2 induction.[99]
↓: alleviation of pulmonary fibrosis (PF) vs. controls, ↑: aggravation of PF vs. controls, N.A.: not applicable. Abbreviations: Mmp, matrix metalloproteinase; EMT, epithelial–mesenchymal transition; IL, interleukin; MIP-1α, macrophage inflammatory protein-1α; BAL, bronchoalveolar lavage; IGFBP3, insulin-like growth factor binding protein 3; WT, wild type; PTGS2, prostaglandin–endoperoxide synthase 2; MyD88, myeloid differentiation factor 88; IL-1R1, IL-1 receptor 1; TGF-β, transforming growth factor-β; MSC, mesenchymal stem cell; hfPMSC, human placental MSCs of fetal origin; BALF, BAL fluid; PF, pulmonary fibrosis; IL-17RA, IL-17 receptor; ST2, chain suppression of tumorigenicity 2; CCL, CC chemokine ligand; CCR, CC chemokine receptor; TNF, tumor necrosis factor; CXCR, CXC chemokine receptor; CXCL, CXC chemokine ligand; Col I, collagen type I; CX3CR, CX3C chemokine receptor; PI3K, phosphatidylinositol 3-kinase; mTOR, mammalian target of rapamycin; JAK, Janus kinase; p-STAT, phosphorylated-signal transducer and activator of transcription; DC, dendritic cell; FGF, fibroblast growth factor; Treg, regulatory T-cell; PDGF, platelet-derived growth factor; bFGF, basic fibroblast growth factor; VEGF, vascular endothelial growth factor; CTGF, connective tissue growth factor; scFv, single-chain variable fragment antibody; IGF-I, insulin-like growth factor I; iPS cell, induced pluripotent stem cell; BM, bone marrow; GSK-3, glycogen synthase kinase 3; ECM, extracellular matrix; NALP3, also called cryopyrin or NLRP3; MAPK, mitogen-activated protein kinase; NF-κB, nuclear factor-kappa B.
Table 2. Advantages and disadvantages of animal models using BLM.
Table 2. Advantages and disadvantages of animal models using BLM.
AdvantagesDisadvantages
  • The BLM model of PF is the cheapest, easiest, fastest, most reproducible, and most used animal model of IPF
  • In rodent models, pulmonary fibrosis may regress spontaneously after 28 days
  • The BLM model is well characterized, clinically relevant, and capable of inducing fibrosis by multiple routes of administration
  • Attempts have been made to model IPF using different species, routes, and administration methods, but there is no consistency across laboratories
  • Rodent models are the most useful models for elucidating disease pathogenesis and evaluating preclinical treatments
  • Administration of compounds tested prophylactically is initiated before fibrosis is induced; when tested therapeutically, it is initiated after fibrosis is established
  • The current state of preclinical trials can be improved by addressing issues such as time course of treatment, animal size and characteristics, clinically relevant treatment endpoints, and reproducibility of treatment results
  • Animal models must be carefully selected, designed, and implemented to bridge the translational gap between benchside and bedside
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ishida, Y.; Kuninaka, Y.; Mukaida, N.; Kondo, T. Immune Mechanisms of Pulmonary Fibrosis with Bleomycin. Int. J. Mol. Sci. 2023, 24, 3149. https://doi.org/10.3390/ijms24043149

AMA Style

Ishida Y, Kuninaka Y, Mukaida N, Kondo T. Immune Mechanisms of Pulmonary Fibrosis with Bleomycin. International Journal of Molecular Sciences. 2023; 24(4):3149. https://doi.org/10.3390/ijms24043149

Chicago/Turabian Style

Ishida, Yuko, Yumi Kuninaka, Naofumi Mukaida, and Toshikazu Kondo. 2023. "Immune Mechanisms of Pulmonary Fibrosis with Bleomycin" International Journal of Molecular Sciences 24, no. 4: 3149. https://doi.org/10.3390/ijms24043149

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop