Next Article in Journal
Spinal Muscular Atrophy: An Evolving Scenario through New Perspectives in Diagnosis and Advances in Therapies
Previous Article in Journal
Interaction of VvDELLA2 and VvCEB1 Mediates Expression of Expansion-Related Gene during GA-Induced Enlargement of Grape Fruit
Previous Article in Special Issue
Extracellular Heat Shock Protein 70 Increases the Glucocorticoid Receptor and Dual-Specificity Phosphatase 1 via Toll-like Receptor 4 and Attenuates Inflammation in Airway Epithelial Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Efferocytosis and Respiratory Disease

The Key Laboratory of Pathobiology, Ministry of Education, College of Basic Medical Sciences, Jilin University, Changchun 130021, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(19), 14871; https://doi.org/10.3390/ijms241914871
Submission received: 29 August 2023 / Revised: 26 September 2023 / Accepted: 29 September 2023 / Published: 3 October 2023
(This article belongs to the Special Issue Molecular Research and Therapy of Asthma)

Abstract

:
Cells are the smallest units that make up living organisms, which constantly undergo the processes of proliferation, differentiation, senescence and death. Dead cells need to be removed in time to maintain the homeostasis of the organism and keep it healthy. This process is called efferocytosis. If the process fails, this may cause different types of diseases. More and more evidence suggests that a faulty efferocytosis process is closely related to the pathological processes of respiratory diseases. In this review, we will first introduce the process and the related mechanisms of efferocytosis of the macrophage. Secondly, we will propose some methods that can regulate the function of efferocytosis at different stages of the process. Next, we will discuss the role of efferocytosis in different lung diseases and the related treatment approaches. Finally, we will summarize the drugs that have been applied in clinical practice that can act upon efferocytosis, in order to provide new ideas for the treatment of lung diseases.

1. Introduction

Under homeostasis, cells, as the smallest functional unit of the body, go through a regular process of dying and being replaced by new cells [1]. There are various forms of cell death, such as apoptosis, necrosis, pyroptosis and ferroptosis, with apoptosis predominating in the steady state [2,3]. Dead cells need to be promptly cleared to form new structures, maintain organ function and prevent secondary necrosis. Otherwise, the contents of dead cells are released into the surrounding environment, damaging surrounding cells and tissues, which may cause inflammation and autoimmune diseases [4].
The timely clearance of dying and dead cells by phagocytes is crucial for maintaining host homeostasis. In most tissues, professional and non-professional phagocytes exist to clear cell corpses, a process known as efferocytosis. Professional phagocytes are mostly tissue-resident cells, such as macrophages and dendritic cells, whereas non-professional phagocytes are mostly adjacent cells, such as epithelial cells and fibroblasts [5], while macrophages are the main cells involved in efferocytosis (Figure 1) [4]. Effective efferocytosis can prevent the release of cellular contents, prevent secondary necrosis and inhibit inflammation. It can also promote the release of anti-inflammatory mediators, such as IL-10, to facilitate the resolution of inflammation [6].

2. Mechanisms of Efferocytosis

The phagocytosis of cellular corpses is mainly divided into four steps: (1) the smell” stage, which recruits phagocytes through the “find-me” signal; (2) the “taste” stage, which identifies dead cells through the “eat-me” and “don’t eat-me” signals; (3) the phagocytic stage, which ingests cellular corpses; (4) the digestion and degradation of dead cells [7] (Figure 2).

2.1. The “Smell” Stage

Apoptotic cells are capable of releasing various “find-me” signals to recruit phagocytes to the dead site [8], such as adenosine triphosphate (ATP) and uridine triphosphate (UTP), which are recognized by macrophage purinergic receptor P2Y2 [9,10], lyso-phosphatidylcholine (LPC) and sphingosine 1-phosphate (S1P), which can, respectively, bind to macrophage surface G protein-coupled receptors G2A and S1P1-5 [11,12,13], and chemokine CX3CL1 [14], which is released from apoptotic lymphocytes in a caspase and Bcl-2-dependent manner and mediates the macrophage recognition of apoptotic cells by interacting with macrophage CX3CR1 receptors through an unknown mechanism [8] (Table 1). CX3CL1 can also enhance efferocytosis by stimulating the expression of milk-fat globulin-EGF-factor 8 (MFG-E8), a phagocytic ligand, in phagocytes [10]. These signals have a dual effect; firstly, they can attract phagocytes to approach apoptotic cells; secondly, they can enhance the expression of phagocytic receptors by regulating the cytoskeleton of phagocytes [5], and some find-me signals may improve the efferocytotic function of non-professional phagocytes and play a role in immune regulation [15,16,17]. In addition, intercellular adhesion molecule 3 (ICAM3 or CD50) and thrombospondin 1 (TSP-1) on dying cells can, respectively, act upon CD14 and CD36 on macrophages, to promote the macrophage recognition of dying cells [18].

2.2. The “Taste” Stage

After phagocytic cells are recruited to the site of dying cells, it is necessary to distinguish between apoptotic cells and healthy cells through the “eat-me” signals expressed on the surface of dying cells for precise phagocytosis in the next step. Different changes occur on the surface of apoptotic cells, such as exposure of phosphatidylserine (PS) to the outer membrane lobules, epitope changes in the intracellular adhesion molecule-1 (ICAM1), exposure of calreticulin and changes in cell surface charge and glycosylation conformation [25]. Correspondingly, the “eat-me” signals that can be expressed on the surface of apoptotic cells include phosphatidylserine (PS), modified intracellular adhesion molecules, calreticulin, oxidized low-density lipoprotein, cell-bound thrombospondin (TSP), antibodies and complement opsonins [26] (Table 1).
The most effective, evolutionarily conserved and well-studied “eat-me” signal is PS exposed to the outer leaves of the plasma membrane of apoptotic cells [5]. In healthy cells, PS is kept on the inner leaves of the plasma membrane by two specific enzymes called flippase and scramble, which help maintain the asymmetry of PS on the cell membrane. When apoptosis occurs, these enzymes are cleaved by caspase 3, leading to the cleavage and inactivation of flippase and the activation of scramble; then, PS flips outward to the cell surface and is recognized by phagocytes [5,27,28]. Apoptotic cells can also expose their surface calreticulin, as an “eat-me” signal. For example, in cancer cells, the translocation of calreticulin from the endoplasmic reticulum to the plasma membrane occurs during apoptosis induction accompanied by endoplasmic reticulum stress, and then the CD91 (low-density lipoprotein receptor protein) mediates calreticulin recognition and subsequent phagocytosis [23,24].
The recognition of eat-me signals by phagocytes requires phagocytic receptors on the surface of phagocytes. There are two ways in this process, one is direct binding, and the other is indirect binding through bridging molecules. Phagocytic receptors such as brain-specific angio-genesis inhibitor 1 (BAI1), T cell immunoglobulin mucin receptor TIM1-TIM4 (TIM family) and Stabilin2 (Stab2) can directly bind to PS, while tyrosine kinase Tyro3/Axl/Mer (TAM) family receptors, scavenger receptor SCARF1, integrin receptor family complexes with CD36 or transglutaminase 2 (TG2) are bound to PS through soluble bridging molecules, such as Gas6 and protein S [22].
After BAI1 directly binds to PS through thrombospondin repeat (TSR), it recruits and activates the ELMO1-DOCK180 complex to initiate intracellular signaling, induce actin cytoskeleton rearrangement and promote efferocytosis [19]. The TIM receptor family is a type I cell surface glycoprotein that binds to PS through the N-terminal IgV domain and mediates the phagocytosis of apoptotic cells [10]. TIM-3 can promote the cross-presentation of apoptotic cell-associated antigens [29], while TIM-4 mediated phagocytosis can trigger a unique non-classical autophagy form called LC3-associated phagocytosis (LAP) [10,30]. It is worth noting that, although TIM-4 can directly interact with PS, it needs to cooperate with integrins or other receptors to transduce cell signals, and cannot be carried out independently [5]. Stabilin2 is a type I surface receptor that mediates the direct interaction with PS and triggers efferocytosis via adaptor protein GULP and thymosin β4, which regulates actin polymerization [20,31], while Stab2 can also promote binding to apoptotic cells in cultured cells [32].
The TAM receptor family is involved in a variety of physiological processes and has distinct but closely related roles in efferocytosis; all three receptors can mediate efferocytosis. The TAM receptor indirectly recognizes PS on apoptotic cells by binding the PS-binding serum protein Gas6 and ProS1, and subsequently dimerizes, activates Rac and promotes phagocytosis in apoptotic cells [10,21]. The integrin αvβ3 receptor family can also indirectly bind to PS, the integrin αv/β3/CD36 receptor complex binding to PS via TSP-1, while the integrin αv/β3/TG2 receptor complex interacts with PS via MFG-E8 [22]. Serum protein C1q can act as a bridging molecular by binding to the SCARF1 scavenger receptor on phagocytes, but also to apoptotic cell membrane-associated receptors A2 or A5 (AnnexinA2/A5) [33].
The reason why phagocytes recognize the “eat-me” signals is to accurately phagocytize apoptotic cells; in addition, phagocytes can also combine with the “don’t eat-me” signals on healthy cells through anti-phagocytosis receptors, and then activate phosphatase, mainly SH1/2, to prevent normal cells from being phagocytosed. Interestingly, when both the “eat-me” signals and the “don’t eat-me” signals are present on the cell, the “don’t eat-me” signals are superior to the “eat-me” signals, so phagocytosis does not occur [34].
The “don’t eat-me” signals that have been discovered so far mainly include CD47, CD24 and CD31 [5] (Table 1). CD47, which is expressed in normal cells but is changed or decreased in terms of expression in apoptotic cells, can activate the signal inhibitory regulatory protein-α (SIRP α) on macrophages, phosphorylate their cytoplasmic domains and activate phosphatase SHP1/2, thereby inhibiting phagocytosis [35]. CD24 is highly expressed on tumor cells and it binds to sialic acid-binding Ig like lectin 10 (Siglec10) to inhibit the clearance of tumor cells [36]. And CD31 has homologous interactions between healthy cells and macrophages, so it can also inhibit phagocytosis [37].

2.3. The Phagocytic Stage

The uptake of dead cells is mediated by the Rho family of small GTP enzymes, including RhoA, Rac, Cdc42, etc., which transition between resting, inactive GDP binding and active GTP binding states under the action of specific guanine nucleotide exchange factors (GEFs) [38,39]. RhoA activation can inhibit the phagocytosis of apoptotic cells, while Rac and Cdc42 can enhance phagocytosis [39]. Once the phagocytic cell recognizes the dying cell, the Rho family relocates on the plasma membrane and triggers a series of signaling pathways; then, the cytoskeleton rearranges through the dynamic network of actin fibers below the plasma membrane, the plasma membrane invaginates, local extravasation occurs [4] and a phagocytic cup is formed on the side facing the apoptotic cell [40], ultimately forming a phagosome. There are two main pathways involved in this process. One is the recruitment of the adapter protein GULP by binding LRP1 (CD91) to calreticulin or Stab2 to PS, thereby activating Rac [20,41]; the other is that, after the Mer tyrosine kinase (MerTK) or integrin receptor family indirectly binds to PS, engulfment and cell motility protein 1 (ELMO1) and dedicator of cytokinesis protein 1 (DOCK180) bind, and are then recruited to the cell membrane by the adapter protein CrkII, forming the ELMO1-DOCK180-CrkII complex [19,42,43], which further activates the RHO family, especially Rac1 [40]. Both of the above pathways promote phagocytosis by activating Rac, indicating that Rac plays a crucial role in the phagocytic process. Moreover, Cdc42 can promote phagocytosis by stimulating the formation of filopodia [44].

2.4. Digestion Phase

The completion of phagocytosis is not the end, but just the beginning. The phagosomes formed by phagocytosis need to fuse with lysosomes, which is also a multi-step process. The phagosomes need to mature first, then target lysosomes and fuse with lysosomes. This process is mainly mediated by the Rab GTP enzyme family, which, when activated, has functions such as transportation, vesicular fusion, Rab-related signaling and activation of other downstream Rab family members [45,46]. After completion of phagocytosis, Rab5 is recruited and activates early endosome antigen 1 (EEA1) and class III phosphatidylinositol 3-kinase VPS34 in an unknown way, which mediates the fusion of phagosomes with early endosomes, forming early phagosomes [44,47]. EEA1 endows vesicles with targeting from the plasma membrane to early endosomes, and there are two Rab5 binding sites on it, but neither site alone can mediate its stable binding with Rab, so synergistic action with phosphatidylinositol-3-phosphate (PI3P) is required [47]; VPS34 is rapidly recruited after the formation of early phagosomes, catalyzing the formation of PI3P by phosphatidylinositol, and further catalyzing the maturation of phagosomes [44].
Early phagosomes continue to transform into late phagosomes, with the most significant marker being the loss of Rab5 but the acquisition of Rab7 [44]. Rab7 is crucial for phagocytic maturation, acidification and fusion with lysosomes. The most critical proteins in this process are the Rab5 effector Mon1a/b and the homologous fusion and protein sorting (HOPS) complex [44]. When Mon1 still binds to Rab5, it can recruit the Rab7 binding protein Ccz-1 to promote binding to GDP-Rab7. At the same time, Mon1 can inactivate Rab5, and the Mon1/Ccz-1 complex plays a role in the activation process of Rab7 [48,49]. After binding to Rab7, the HOPS complex can serve as a GDI displacement factor to remove Rab7 from GDP dissociation inhibitors (GDIs) in order to activate Rab7 [50]. Rab7-interacting lysosomal protein (RILP) and oxysterol-binding protein-related protein 1 (ORPL1) directly combine with molecular motor motilin/motilin to mediate effective lysosomal fusion, which also requires complete microtubules [4].
After the fusion of late phagosomes and lysosomes, the newly formed phagolysosomes rely on proteases, nucleases and lipases in the lysosome to digest the phagocytic cell corpses [4]. The digested products are transformed and reused by macrophages or excreted from the body; otherwise, their accumulation can trigger inflammation and cause diseases [51].

3. Regulation of Efferocytosis

3.1. Approaches to Target the Taste Stage

Due to the fact that the taste stage is the recognition of apoptotic cells by macrophages, that is, the “eat-me” signal sent by apoptotic cells interacts with phagocytic receptors on macrophages, the regulation of phagocytic receptors may be an important way to improve efferocytosis. For example, IL-4 may enhance efferocytosis by increasing the expression of the phagocytic receptors Stabilin1 and Stabilin2 [52]. In a wild-type mouse model of colitis induced by the dextran sulfate sodium (DSS), BAI-1 mRNA levels were reduced in intestinal epithelial cells and colon tissue, while apoptotic cells were reduced in the colon epithelium of BAI-1 transgenic mice, and the pathology of colitis was also alleviated [53]. In cancer, reducing the number of anti-phagocytosis receptors allows for cancer cells to be engulfed, which may be a potential target for the treatment [5].
Many inhibitors of efferocytosis can exert their effects by blocking the recognition of apoptotic cells by macrophages. High mobility group box protein 1 (HMGB1) can bind to various macrophage receptors, for example, binding to the soluble receptor for advanced glycation end products (RAGE) can block macrophage recognition of PS, binding to integrin αV can block the interaction with the bridging molecule MFG-E8 [6,54,55,56]. RAGE and Annexin A5 can also downregulate efferocytosis by blocking the recognition of apoptotic cells by masking PS on apoptotic cells [56,57].
IL-4 and IL-13 induce the production of potential PPAR γ-activating ligands 13-hydroxyoctadecadienoic acid (13-HODE) and 15-hydroxyeicosatetraenoic acid (15-HETE) through STAT6 or 15-lipoxygenase to increase the expression and activity of PPAR γ, so as to increase the “alternative activation” of phagocyte surface receptors and the secretion of the bridging molecule adiponectin, enhance efferocytosis and inhibit inflammation [6,58,59,60]. IL-4 and IL-13 can also enhance PPAR δ of macrophages, promote the expression and release of the bridging molecules and make macrophages acquire anti-inflammatory function [6]. However, both of these receptors obtain these functions by combining with other nuclear receptors to form heterodimers, so liver × receptor (LXR) and retinoid × receptor α (RXRα) may play a role in the enhancement of efferocytosis by increasing the expression of MerTK [5], but whether IL-4 and IL-13 can directly interact with LXR and RXRα is unknown [6]. Although all LXR ligands can effectively enhance efferocytosis, some LXR agonists can increase liver triglyceride in some mouse models; therefore, developing new powerful and effective LXR agonists without adverse side effects may be beneficial to clinical use [22,61,62].

3.2. Approaches to Target the Phagocytic Stage

The phagocytic stage is mainly influenced by affecting the balance of Rac-1/RhoA. The enhancement of this stage mainly improves the ability of macrophages to phagocytose apoptotic cells by increasing the activity of Rac1, most of which are modified lipids. A modified phosphatidylserine, lyso-phosphatidylserine (LPS), is produced by the oxidation of the fatty acyl groups of PS in apoptotic neutrophils through the NADPH oxidase-dependent pathway [63,64]. LPS interacts with G2A receptors on macrophages to stimulate the production of prostaglandin E2 (PGE2), thereby enhancing the activity of cAMP and PKA-dependent Rac1 [6]. However, it should be noted that the effect of PGE2 on efferocytosis is concentration-dependent, with an enhanced efferocytosis effect within a concentration range of less than or equal to 1 nM, while an inhibitory effect is observed at doses greater than 10 nM [65]. Similar to LPS, lysophosphatidic acid (LPA) inhibits efferocytosis by activating RhoA [66]. Another molecule that can affect the Rac-1/RhoA balance is lipoprotein A4, which can activate Rac and RhoA at the same time, but the final result is more inclined toward the activation of Rac, which depends on protein kinase A [22,67].
When the mitochondrial oxidative stress level is elevated, the production of reactive oxygen species (ROS) also increases relatively, which activates the NOD-like receptor thermal protein domain-associated protein 3 (NLRP3), an inflammasome, and its related signaling pathways, ultimately leading to the inhibition of Rac1 expression at the gene level [68]. Mitochondrial uncoupling protein 2 (Ucp2) can regulate the NLRP3 signaling pathway by reducing the formation of ROS, so as to enhance the efferocytosis effect [69]. The efferocytosis effect can also be improved by reducing the membrane potential of macrophage mitochondria [18].

3.3. Approaches to Target the Digestion Stage

After the fusion of phagosomes and lysosomes, the cell corpse is digested, and the products obtained from digestion can affect the subsequent phagocytosis.

3.3.1. Sugar

The deficiency of glycolytic enzyme 6-phosphofructose-2-kinase and fructose-2,6-bisphosphatase (PFKFB3) in macrophages can lead to a decrease in efferocytosis ability [70]. In addition, the reuse of glucose derived from apoptotic cells can serve as a part of the signaling pathway and/or provide energy to enhance subsequent efferocytosis [71]. That is to say, enhanced glycolysis in efferocytotic macrophages can promote actin aggregation and the sustained uptake of apoptotic cells.

3.3.2. Lipid

Lipid metabolism induces the activation of PPAR γ and δ, LXR α and β and RXR α and β in macrophages, to promote efferocytosis (as previously mentioned) [72]. LXR signaling can regulate the expression of transglutaminase 2 (Tgm2), a co-receptor of αvβ3- integrin, and promote the formation of phagocytes [73,74]. LXR is also involved in Del-1-dependent efferocytosis and macrophage reprogramming to a pro-resolving phenotype [26,75]. In addition, the oxysterols produced by lipid metabolism are active inhibitors of the activation of the inflammasomes NLRP3, which promote subsequent efferocytosis [68,76].

3.3.3. Amino Acid

Macrophages have been proved to utilize arginine and ornithine produced by the decomposition of apoptotic cells from the first round of phagocytosis to promote subsequent efferocytosis [5,77]. Yurdagul and colleagues found that, under the catalysis of inducible nitric oxide synthase (iNOS), arginine produces the precursor of polyamines, ornithine, in macrophages with a pro-catabolic phenotype. Ornithine is then converted into putrescine under the action of ornithine decarboxylase (ODC) [71,77,78]. Putrescine can enhance the mRNA stability of GTP exchange factor Dbl through the RNA binding protein HuR, which in turn upregulates and activates Rac1, leading to changes in the actin cytoskeleton and promoting the internalization of apoptotic cells [18,71,79,80].

3.3.4. Transporter

Several members of the solute carrier (SLC) membrane transporter family, including glucose transport protein type 1 (GLUT1; encoded by gene Slc2a1) that promotes lactate release and monocarboxylate transporter 1 (MCT1; encoded by Slc16a1), are upregulated during efferocytosis [18]. SLC12A2 and SLC12A4 play different roles in the chloride-sensing pathway. SLC12A2 mediates the influx of Cl, while SLC12A4 has the opposite effect. The influx of Cl plays a brake role in efferocytosis, so the inhibition or loss of SLC12A2 can lead to an enhanced efferocytosis [11,81]. By altering the activity or expression of selected solute carrier proteins (such as SLC2a1) while regulating the activity of cholesterol transporters ABCA1 and SLC16a1, efferocytosis can be promoted [5].

4. Efferocytosis and Respiratory Disease

In healthy individuals, apoptotic cells can rarely be detected in lung tissue [82], which may be due to the fact that normal airway macrophages can phagocytize and remove apoptotic cells in time. However, in acute or chronic lung inflammatory diseases, the rate of apoptosis is increased or the clearance of apoptotic cells is defective, which leads to the accumulation of apoptotic cells [83,84]. Experiments have proved that this pathological process is mainly due to the faulty efferocytosis of airway macrophages [85].

4.1. COPD

Chronic obstructive pulmonary disease (COPD) is a common chronic respiratory disease characterized by airflow restriction and airway inflammation. COPD is often caused by long-term exposure to toxic particles or gases, such as cigarette smoke, which is why the disease can be prevented and treated. The inflammatory characteristics of this disease have sparked a series of explorations into anti-inflammatory drugs, but most of them are limited to improving inflammation and alleviating respiratory symptoms and cannot stop the development of the disease. At present, there is a large amount of evidence indicating that there are defects in the efferocytotic function during the occurrence and development of COPD, leading to secondary necrosis and inflammation. The targets related to efferocytosis have also been extensively studied as new therapeutic directions for COPD.
LC3-related phagocytosis (LAP) is known to be essential for effective efferocytosis. Asare and others discovered the imbalance in the LAP pathway in COPD and cigarette smoke exposure for the first time. In the process of COPD, the expression of Rubicon, a key mediator in the lap pathway, decreases, leading to weakened efferocytosis and accumulation of apoptotic cells; subsequently, the release of inflammatory mediators exacerbates inflammation [86]. The downregulation of CD44, CD163, VSIG4 and MARCO in alveolar macrophages can also lead to their dysfunction, making it difficult for inflammation to subside [87,88]. The activation pathway of CR3 mediated by CD44 relies on divalent cations, especially Zn2+, but not Ca2+. This pathway is related to efferocytosis, so there is a deficiency in the efferocytosis of Zn-deficient alveolar macrophages [89,90].
The interaction between surfactant proteins A (SP-A) and SP-D and SIRP- α can inhibit efferocytosis, especially SP-D which is significantly reduced in COPD patients [82,91,92]. Cigarette smoke also damages the clearance of apoptotic cells through the oxidant-dependent activation of RhoA and the inhibition of Rac1, leading to actin polymerization defects commonly required for effective efferocytosis [93,94,95]. In addition, a reduction in PPAR-c, a decrease in methylation, an increase in ROCK or the colonization of microorganisms in the airway, especially Hemophilus influenzae, may lead to the defects in the efferocytotic function [96,97,98,99].
The damaged efferocytotic function is of great significance for the occurrence and development of COPD, so many studies have been conducted to achieve the goal of treating COPD by restoring efferocytotic function. Lovastatin can increase efferocytosis in various ways, such as inhibiting HMG-CoA reductase, altering the balance of membrane-bound RhoA and Rac-1, inhibiting oxidative stress and inhibiting matrix metalloproteinase-9 (MMP-9) [89]. Statins can also increase efferocytosis by increasing PPAR [66]. There is a positive correlation between available glutathione and phagocytosis, as Galactin-3 can induce macrophages to transform into a more “M2” phenotype through its interaction with CD98, increasing the phagocytic capacity and promoting the production of available glutathione [100,101]. Azithromycin and Rosiglitazone can also play a certain therapeutic role in COPD by increasing the phagocytosis of apoptotic cells by alveolar macrophages [98,101].

4.2. Asthma

Asthma is also a common heterogeneous chronic respiratory inflammatory disease, characterized by airway hyperresponsiveness and airway obstruction. Unlike COPD, the airway obstruction caused by asthma is reversible [102]. Efferocytosis can trigger a series of anti-inflammatory events, mediated by receptors such as MerTK, and executed by the production of interleukin-10 (IL-10) [103,104]. In asthma, there are fewer macrophages that produce IL-10, indicating that efferocytosis may be impaired [103,105]. Felton and others found that the enhanced efferocytosis of alveolar macrophages is helpful to solve allergic airway inflammation in mice, making the clearance of apoptotic cells a potential therapeutic strategy. This depends on MerTK, and enhancing Mer activity may be a specific mechanism of targeted efferocytosis [103,106].
In neutrophilic asthma, a decrease in the level of galactose agglutinin 3 (gal-3) results in the inability of apoptotic neutrophils to be completely cleared, leading to impaired efferocytosis and the persistence of inflammation. However, after the addition of exogenous gal-3, the ability of macrophages to clear apoptotic cells is restored. IL-1β, an inflammatory mediator in the airway of patients with neutrophilic asthma, is visibly increased, and its level is negatively correlated with gal-3, so adding exogenous gal-3 can inhibit the production of IL-1β and promote the resolution of inflammation [102]. Axl, an apoptotic cell recognition receptor only expressed in human airways or alveolar macrophages, was significantly reduced in asthma, and macrophages could not recognize apoptotic cells, so they could not be eliminated. The accumulated apoptotic cells suppressed the release of anti-inflammatory mediators such as IL-10 and TGF-β. And the shedding of Axl was also related to the damage of efferocytosis, while IFN could restore the expression of Axl on airway macrophages [107,108,109]. Research has shown that the expression of Neurofibroin-2 (Nrp2) is increased in neutrophil asthma mice and is a negative regulatory factor for neutrophil asthma. Therefore, it can regulate the efferocytosis of macrophages by promoting the degradation of endocytic apoptotic cells [109].

4.3. COVID-19

In recent years, the world has been plagued by the COVID-19 epidemic. Vaccination has greatly alleviated the impact of the epidemic in a short time, but the rapid mutation of the virus has brought great difficulties to the research and development of vaccines. Among the numerous variant strains, the Delta variant strain dominates, and people who have received vaccines can also become infected. Vaccines alone cannot control the epidemic; therefore, finding treatment methods that can reduce the severity of the disease seems more important [110]. The cause of this major public health and safety incident is severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), which mainly infects ciliated airway epithelial cells [111,112]. After infection, this virus damages the airway epithelium, alveolar epithelium, causing various clinical symptoms, mainly pneumonia, acute respiratory distress syndrome and multiple organ dysfunction [111,113,114,115]. SARS-CoV-2 infection can cause various lung cells to die in different ways, such as apoptosis, necrosis or pyroptosis [116]. At the same time, these apoptotic cells accumulate in large quantities due to impaired efferocytotic function, resulting in the production and release of pro-inflammatory cytokines. This may also be the main reason for various complications of COVID-19.
COVID-19 infection has been shown to hinder the normal function of alveolar macrophages. Recent studies have shown that the phagocytosis of SARS-CoV-2-AC reduces the transcription of PS receptors, including scavenger receptors CD36, SRA-1, integrin α V β5 (ITGB5) and T cell immunoglobulin mucin receptor 4 (TIM4), but the transcription of MerTK will not be decreased [111,117]. That is to say, the expression of receptors related to efferocytosis on dying cells infected with SARS-CoV-2 is inhibited, thereby impairing the sustained clearance of apoptotic cells by macrophages. It is speculated that SARS-CoV-2 infection can also damage the release of sphingosine-1-phosphate (S1P) involved in the process of efferocytosis, which can also slow down the process of eliminating apoptotic cells caused by SARS-CoV-2 infection [111,118].
In summary, targeting damaged efferocytosis may be a new therapeutic strategy against COVID-19. Gas6, a bridging molecule, can promote the combination of phosphatidylserine (PS) exposed on apoptotic cells with TAM receptors on macrophages, activating the receptors and triggering subsequent cellular signal transduction, promoting the phagocytosis of apoptotic cells, inhibiting innate immune responses in affected tissues, preventing inflammation and autoimmunity and protecting surrounding healthy tissues [10,119,120,121]. ADAM-17, a disintegrin and metalloprotein 17, can cleave phosphatidylserine receptors, and its deletion has been shown to increase efferocytosis in vivo and enhance the anti-inflammatory response [122,123,124]. The inhibition of ADAM-17 activity can reduce cytokine storm and the excessive aggregation of neutrophils [110,125]. Some immunomodulators used in COVID-19 may adjust the increased cytokine level by enhancing ADAM-17 activity [126]. Some compounds have been selected and have entered clinical trials for their ability to inhibit the function of ADAM-17, but they have been discontinued due to adverse events [110,127]. Therefore, drugs used to inhibit ADAM-17 need to be carefully selected.

4.4. Influenza

Influenza is caused by the influenza virus, which can be divided into three types: A, B and C. Among them, the influenza A virus (IAV) is the most serious, which can cause major public health and security incidents all over the world. IAV mainly replicates in respiratory epithelial cells. After the virus enters, it inhibits host cell protein synthesis and rapidly begins replication, causing clinical symptoms, ranging from upper respiratory tract infection to severe pneumonia [128]. At present, the prevention and treatment methods are very limited, and the main measure to be taken against it is vaccination. However, the antigenicity of the influenza A virus is prone to mutation and the effectiveness of influenza vaccines is not lasting, which further increases the difficulty of prevention and treatment. As a result, it has caused multiple global pandemics throughout history.
Pulmonary alveolar macrophages are key cells in controlling influenza, which can inhibit the replication of IAV and promote its clearance. They can also promote efferocytosis and the regression of lung inflammation. The increased efferocytosis of alveolar phagocytes can prevent fatal acute lung injury caused by lung infection, while the defect in efferocytosis of alveolar phagocytes can lead to chronic pneumonia [129,130]. Mannose-binding lectin (MBL) is an innate immune pattern recognition molecule existing in the lung that can bind to IAV. The downregulation of the tumor necrosis factor (TNF) release and reduced apoptosis mediated by MBL in human macrophages may represent its mechanism for promoting IAV clearance, reducing local tissue inflammation and oxidative damage and limiting collateral damage to the lungs through increased efferocytosis [130]. Angiotensin-(1-7) [Ang-(1-7)], a pro-apoptotic mediator, is able to increase the binding and phagocytosis of alveolar macrophages to apoptotic neutrophils in vitro, as was found by subjecting it to action in mice, and it can also enhance neutrophil apoptosis. The effects of Ang-(1-7) may be significantly beneficial in reducing lung injury and achieving better clinical outcomes after IAV infection [131]. The Bacillus Calmette–Guerin (BCG) vaccine is currently the only vaccine available for anti-tuberculosis, but its more important role is as an immunotherapeutic agent for treating other diseases, as it can provide non-specific protection for other diseases by increasing the recruitment and activation of macrophages [129]. Sanjay Mukherjee gave the BCG vaccine via nasal drops to mice, which showed a significant increase in the efferocytosis of alveolar phagocytes, and all of them survived after infection with the IAV, while all mice in the PBS-treated group died. After inserting apoptotic 5,6-carboxyfluorescein diacetate succinimidyl ester positive (CFSE+) mouse lung epithelial (MLE) cells into the alveoli of mice, the clearance efficiency of the BCG intervention group was significantly higher. BCG was also found to increase the expression of TIM4 on apoptotic cells, which is the first time that the role of TIM-4 in efferocytosis and the mechanism of BCG increasing efferocytosis were reported [129].

4.5. Other Lung Diseases

Acute respiratory distress syndrome (ARDS), the most severe form of acute lung injury (ALI), is characterized by alveolar injury, pulmonary edema and hypoxic respiratory failure caused by alveolar epithelial and endothelial cell death, alveolar neutrophil infiltration, etc. [132,133,134]. The pathogenesis of ALI/ARDS is not well understood, and the therapeutic approaches have limitations [133]. Fas/FasL of the tumor necrosis factor (TNF) family is known to promote apoptosis, and the inhibition of its associated signaling attenuates lung injury [133,135]. During the ARDS phase, neutrophils have an extended lifespan but eventually undergo apoptosis [136,137]. Therefore, the clearance of apoptotic cells by lung macrophages is crucial in attenuating lung injury. It has been found that the alveolar macrophage in ARDS patients does have an efferocytotic error, and IL-8 induces the onset of macrophage classical activation, which can inhibit efferocytosis; therefore, blocking IL-8 may make it possible to promote the clearance of apoptotic cells and reduce inflammation [138]. In addition, blocking high-mobility group box protein 1 (HMGB1) or promoting the activation of AMP-activated protein kinase (AMPK) can upregulate macrophage efferocytosis [139].
Cystic fibrosis (CF) is an autosomal recessive disorder capable of invading multiple organs, most commonly in the respiratory and digestive systems. The lungs are primarily characterized by impaired mucosal ciliary clearance and subsequent bacterial colonization [140], caused by mutations in the CF transmembrane regulatory protein (CFTR) in the lung epithelium that result in reduced Cl uptake and hydration of airway fluids [141], which renders mucosal cilia incapable of blocking foreign bodies, such as bacteria, and results in the failure of neutrophilic inflammation to abate and damage the respiratory system. Bacterial colonization in the lungs of CF patients and persistent neutrophilic inflammation suggest a defect in phagocytosis by lung macrophages, but the exact mechanism is not clear; the only understood mechanism is that phagocytosis of apoptotic cells by airway macrophages is impaired by the cleavage of the PS receptor by neutrophil elastase [142].
The pathogenesis of idiopathic pulmonary fibrosis (IPF) is still unknown, and its main features are alveolar structural disorders as well as diffuse pulmonary fibrosis, leading to the obstruction of gas exchange in the lungs, which ultimately leads to respiratory failure and death, and the disease is characterized by rapid progression, high morbidity and mortality and poor prognosis. IPF is often accompanied by chronic neutrophilic inflammation, and it has been experimentally demonstrated that the phagocytosis of apoptotic cells in bronchoalveolar lavage fluid (BALF) is reduced in patients with IPF [143,144]. The drugs currently used to treat IPF are mainly pirfenidone and nintedanib, and no clinical drug has been found to be able to play a therapeutic role by increasing macrophage efferocytosis, and inducing the expression of PPAR γ in airway macrophages may be a potential therapeutic pathway [82,145].
The phagocytic errors of macrophages have been found in all three of the above respiratory diseases where the pathogenesis is not yet known; thus, targeting efferocytosis to develop drugs has a promising future.

5. Current Drugs Targeting Efferocytosis

Defective efferocytosis has been identified in a number of different types of diseases, and therefore developing drugs targeting efferocytosis may be a promising therapeutic direction for the future. The drugs that have been shown to be effective against efferocytosis to date are described below.
The most prominent drugs capable of enhancing phagocytosis are glucocorticoids, which can exert this effect through a variety of mechanisms: (i) enhancement of downstream signaling through Mer and Pro S, (ii) alteration of the Rac/RhoA balance and (iii) modulation of PPAR γ-induced macrophage reprogramming [6]. Dexamethasone-treated macrophages have increased Rac expression, altered cytoskeletal structure and increased cell motility, all of which are required to carry out effective efferocytosis [24,146].
Statins are 3-hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) reductase inhibitors that can help enhance efferocytosis by inhibiting the isoprenylation of RhoA [22,147]. Macrolide antibiotics may alter macrophage programming and increase the expression of bridging molecules to enhance uptake of apoptotic cells [6]. β-Hydroxybutyric acid, a dietary supplement, can increase the expression level of SLC2a1 in the presence of impaired sustained efferocytosis in phagocytes and the activity of the lysosomal arginine transporter protein Pqlc2, which can drive sustained phagocytosis in vitro and in vivo [5,77,148].

6. Conclusions

In recent years, the role of efferocytosis in the occurrence and development of diseases and the restoration of body homeostasis has been increasingly recognized. The clearance of apoptotic cells by phagocytic cells is a multi-step and interconnected process, and errors at any stage of the entire process may lead to inflammatory or autoimmune diseases. It has been proven that there is phagocytosis error by macrophages in respiratory system diseases, whether it is in chronic airway diseases such as COPD and asthma, which have been extensively studied, or in lung diseases such as ARDS and CF, whose pathogenesis is still unknown.
This review focuses on the role of efferocytosis, introducing its mechanisms, related respiratory diseases and targeted treatment methods. We focus on promoting inflammation resolution in respiratory diseases by improving the damaged efferocytosis. At present, with the efforts of researchers, some drugs that can act at different stages of the efferocytosis process to solve inflammation and treat diseases have been discovered, and they have achieved good therapeutic effects. Nevertheless, our understanding of the role of efferocytosis is still not profound enough. Whether there are any other “eat-me” signals, “don’t eat-me” signals, phagocytic receptors, anti-phagocytic receptors and bridging molecules; whether there are other signaling pathways that can be triggered after the recognition of apoptotic cells by the phagocytes; whether there is a synergistic effect between these signaling pathways; and more questions which all warrant further exploration. We firmly believe that the more thorough research is conducted on the role of the efferocytosis in the future, the more likely it is to provide more targets for the treatment of respiratory system diseases, which can develop more effective clinical drugs and provide more innovative and effective treatment strategies for respiratory system diseases and even other system diseases.

Author Contributions

Literature review and writing, W.Z., Z.Z. and X.G.; proofreading and editing, X.Z., W.Z., J.Z., Y.A., H.Z. and Y.Y.; supervision, G.W. and F.W. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the fellowship of China Postdoctoral Science Foundation (NO. 2021M701399), the Norman Bethune program of Jilin University (NO. 2020B10) and the Medical Basic Research Innovation Center of Airway Disease in North China, Ministry of Education.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Stanger, B.Z. Cellular homeostasis and repair in the mammalian liver. Annu. Rev. Physiol. 2015, 77, 179–200. [Google Scholar] [CrossRef]
  2. Morioka, S.; Maueröder, C.; Ravichandran, K.S. Living on the Edge: Efferocytosis at the Interface of Homeostasis and Pathology. Immunity 2019, 50, 1149–1162. [Google Scholar] [CrossRef] [PubMed]
  3. Pasparakis, M.; Vandenabeele, P. Necroptosis and its role in inflammation. Nature 2015, 517, 311–320. [Google Scholar] [CrossRef] [PubMed]
  4. Boada-Romero, E.; Martinez, J.; Heckmann, B.L.; Green, D.R. The clearance of dead cells by efferocytosis. Nat. Rev. Mol. Cell. Biol. 2020, 21, 398–414. [Google Scholar] [CrossRef] [PubMed]
  5. Mehrotra, P.; Ravichandran, K.S. Drugging the efferocytosis process: Concepts and opportunities. Nat. Rev. Drug Discov. 2022, 21, 601–620. [Google Scholar] [CrossRef] [PubMed]
  6. Korns, D.; Frasch, S.C.; Fernandez-Boyanapalli, R.; Henson, P.M.; Bratton, D.L. Modulation of macrophage efferocytosis in inflammation. Front. Immunol. 2011, 2, 57. [Google Scholar] [CrossRef]
  7. Ravichandran, K.S. Find-me and eat-me signals in apoptotic cell clearance: Progress and conundrums. J. Exp. Med. 2010, 207, 1807–1817. [Google Scholar] [CrossRef]
  8. Ge, Y.; Huang, M.; Yao, Y.-M. Efferocytosis and Its Role in Inflammatory Disorders. Front. Cell. Dev. Biol. 2022, 10, 839248. [Google Scholar] [CrossRef]
  9. Elliott, M.R.; Chekeni, F.B.; Trampont, P.C.; Lazarowski, E.R.; Kadl, A.; Walk, S.F.; Park, D.; Woodson, R.I.; Ostankovich, M.; Sharma, P.; et al. Nucleotides released by apoptotic cells act as a find-me signal to promote phagocytic clearance. Nature 2009, 461, 282–286. [Google Scholar] [CrossRef]
  10. Elliott, M.R.; Koster, K.M.; Murphy, P.S. Efferocytosis Signaling in the Regulation of Macrophage Inflammatory Responses. J. Immunol. 2017, 198, 1387–1394. [Google Scholar] [CrossRef] [PubMed]
  11. Ranta, A.; Kumar, S. Recent advancements in role of TAM receptors on efferocytosis, viral infection, autoimmunity, and tissue repair. Int. Rev. Cell Mol. Biol. 2020, 357, 1–19. [Google Scholar] [CrossRef] [PubMed]
  12. Lauber, K.; Bohn, E.; Kröber, S.M.; Xiao, Y.-J.; Blumenthal, S.G.; Lindemann, R.K.; Marini, P.; Wiedig, C.; Zobywalski, A.; Baksh, S.; et al. Apoptotic cells induce migration of phagocytes via caspase-3-mediated release of a lipid attraction signal. Cell 2003, 113, 717–730. [Google Scholar] [CrossRef]
  13. Gude, D.R.; Alvarez, S.E.; Paugh, S.W.; Mitra, P.; Yu, J.; Griffiths, R.; Barbour, S.E.; Milstien, S.; Spiegel, S. Apoptosis induces expression of sphingosine kinase 1 to release sphingosine-1-phosphate as a “come-and-get-me” signal. FASEB J. 2008, 22, 2629–2638. [Google Scholar] [CrossRef] [PubMed]
  14. Truman, L.A.; Ford, C.A.; Pasikowska, M.; Pound, J.D.; Wilkinson, S.J.; Dumitriu, I.E.; Melville, L.; Melrose, L.A.; Ogden, C.A.; Nibbs, R.; et al. CX3CL1/fractalkine is released from apoptotic lymphocytes to stimulate macrophage chemotaxis. Blood 2008, 112, 5026–5036. [Google Scholar] [CrossRef] [PubMed]
  15. Yamaguchi, H.; Maruyama, T.; Urade, Y.; Nagata, S. Immunosuppression via adenosine receptor activation by adenosine monophosphate released from apoptotic cells. Elife 2014, 3, e02172. [Google Scholar] [CrossRef] [PubMed]
  16. Hochreiter-Hufford, A.; Ravichandran, K.S. Clearing the dead: Apoptotic cell sensing, recognition, engulfment, and digestion. Cold Spring Harb. Perspect. Biol. 2013, 5, a008748. [Google Scholar] [CrossRef]
  17. Poon, I.K.H.; Lucas, C.D.; Rossi, A.G.; Ravichandran, K.S. Apoptotic cell clearance: Basic biology and therapeutic potential. Nat. Rev. Immunol. 2014, 14, 166–180. [Google Scholar] [CrossRef]
  18. Kourtzelis, I.; Hajishengallis, G.; Chavakis, T. Phagocytosis of Apoptotic Cells in Resolution of Inflammation. Front. Immunol. 2020, 11, 553. [Google Scholar] [CrossRef] [PubMed]
  19. Park, D.; Tosello-Trampont, A.-C.; Elliott, M.R.; Lu, M.; Haney, L.B.; Ma, Z.; Klibanov, A.L.; Mandell, J.W.; Ravichandran, K.S. BAI1 is an engulfment receptor for apoptotic cells upstream of the ELMO/Dock180/Rac module. Nature 2007, 450, 430–434. [Google Scholar] [CrossRef]
  20. Park, S.-Y.; Kang, K.-B.; Thapa, N.; Kim, S.-Y.; Lee, S.-J.; Kim, I.-S. Requirement of adaptor protein GULP during stabilin-2-mediated cell corpse engulfment. J. Biol. Chem. 2008, 283, 10593–10600. [Google Scholar] [CrossRef]
  21. Lemke, G. Biology of the TAM receptors. Cold Spring Harb. Perspect. Biol. 2013, 5, a009076. [Google Scholar] [CrossRef]
  22. Szondy, Z.; Garabuczi, E.; Joós, G.; Tsay, G.J.; Sarang, Z. Impaired clearance of apoptotic cells in chronic inflammatory diseases: Therapeutic implications. Front. Immunol. 2014, 5, 354. [Google Scholar] [CrossRef] [PubMed]
  23. Obeid, M.; Tesniere, A.; Ghiringhelli, F.; Fimia, G.M.; Apetoh, L.; Perfettini, J.-L.; Castedo, M.; Mignot, G.; Panaretakis, T.; Casares, N.; et al. Calreticulin exposure dictates the immunogenicity of cancer cell death. Nat. Med. 2007, 13, 54–61. [Google Scholar] [CrossRef] [PubMed]
  24. Robb, C.T.; Regan, K.H.; Dorward, D.A.; Rossi, A.G. Key mechanisms governing resolution of lung inflammation. Semin. Immunopathol. 2016, 38, 425–448. [Google Scholar] [CrossRef] [PubMed]
  25. Gardai, S.J.; Bratton, D.L.; Ogden, C.A.; Henson, P.M. Recognition ligands on apoptotic cells: A perspective. J. Leukoc. Biol. 2006, 79, 896–903. [Google Scholar] [CrossRef] [PubMed]
  26. Arandjelovic, S.; Ravichandran, K.S. Phagocytosis of apoptotic cells in homeostasis. Nat. Immunol. 2015, 16, 907–917. [Google Scholar] [CrossRef] [PubMed]
  27. Arashiki, N.; Takakuwa, Y.; Mohandas, N.; Hale, J.; Yoshida, K.; Ogura, H.; Utsugisawa, T.; Ohga, S.; Miyano, S.; Ogawa, S.; et al. ATP11C is a major flippase in human erythrocytes and its defect causes congenital hemolytic anemia. Haematologica 2016, 101, 559–565. [Google Scholar] [CrossRef]
  28. Segawa, K.; Nagata, S. An Apoptotic ‘Eat Me’ Signal: Phosphatidylserine Exposure. Trends Cell Biol. 2015, 25, 639–650. [Google Scholar] [CrossRef]
  29. Nakayama, M.; Akiba, H.; Takeda, K.; Kojima, Y.; Hashiguchi, M.; Azuma, M.; Yagita, H.; Okumura, K. Tim-3 mediates phagocytosis of apoptotic cells and cross-presentation. Blood 2009, 113, 3821–3830. [Google Scholar] [CrossRef]
  30. Martinez, J.; Almendinger, J.; Oberst, A.; Ness, R.; Dillon, C.P.; Fitzgerald, P.; Hengartner, M.O.; Green, D.R. Microtubule-associated protein 1 light chain 3 alpha (LC3)-associated phagocytosis is required for the efficient clearance of dead cells. Proc. Natl. Acad. Sci. USA 2011, 108, 17396–17401. [Google Scholar] [CrossRef] [PubMed]
  31. Lee, S.-J.; So, I.-S.; Park, S.-Y.; Kim, I.-S. Thymosin beta4 is involved in stabilin-2-mediated apoptotic cell engulfment. FEBS Lett. 2008, 582, 2161–2166. [Google Scholar] [CrossRef]
  32. Park, S.Y.; Jung, M.Y.; Kim, H.J.; Lee, S.J.; Kim, S.Y.; Lee, B.H.; Kwon, T.H.; Park, R.W.; Kim, I.S. Rapid cell corpse clearance by stabilin-2, a membrane phosphatidylserine receptor. Cell. Death Differ. 2008, 15, 192–201. [Google Scholar] [CrossRef] [PubMed]
  33. Martin, M.; Leffler, J.; Blom, A.M. Annexin A2 and A5 serve as new ligands for C1q on apoptotic cells. J. Biol. Chem. 2012, 287, 33733–33744. [Google Scholar] [CrossRef] [PubMed]
  34. Gheibi Hayat, S.M.; Bianconi, V.; Pirro, M.; Sahebkar, A. Efferocytosis: Molecular mechanisms and pathophysiological perspectives. Immunol. Cell. Biol. 2019, 97, 124–133. [Google Scholar] [CrossRef] [PubMed]
  35. Tsai, R.K.; Discher, D.E. Inhibition of “self” engulfment through deactivation of myosin-II at the phagocytic synapse between human cells. J. Cell. Biol. 2008, 180, 989–1003. [Google Scholar] [CrossRef]
  36. Barkal, A.A.; Brewer, R.E.; Markovic, M.; Kowarsky, M.; Barkal, S.A.; Zaro, B.W.; Krishnan, V.; Hatakeyama, J.; Dorigo, O.; Barkal, L.J.; et al. CD24 signalling through macrophage Siglec-10 is a target for cancer immunotherapy. Nature 2019, 572, 392–396. [Google Scholar] [CrossRef] [PubMed]
  37. Brown, S.; Heinisch, I.; Ross, E.; Shaw, K.; Buckley, C.D.; Savill, J. Apoptosis disables CD31-mediated cell detachment from phagocytes promoting binding and engulfment. Nature 2002, 418, 200–203. [Google Scholar] [CrossRef] [PubMed]
  38. Martinez, J. Prix Fixe: Efferocytosis as a Four-Course Meal. In Current Topics in Microbiology and Immunology; Springer: Berlin/Heidelberg, Germany, 2017; Volume 403. [Google Scholar] [CrossRef]
  39. Nakaya, M.; Tanaka, M.; Okabe, Y.; Hanayama, R.; Nagata, S. Opposite effects of rho family GTPases on engulfment of apoptotic cells by macrophages. J. Biol. Chem. 2006, 281, 8836–8842. [Google Scholar] [CrossRef] [PubMed]
  40. Doran, A.C.; Yurdagul, A.; Tabas, I. Efferocytosis in health and disease. Nat. Rev. Immunol. 2020, 20, 254–267. [Google Scholar] [CrossRef]
  41. Su, H.P.; Nakada-Tsukui, K.; Tosello-Trampont, A.-C.; Li, Y.; Bu, G.; Henson, P.M.; Ravichandran, K.S. Interaction of CED-6/GULP, an adapter protein involved in engulfment of apoptotic cells with CED-1 and CD91/low density lipoprotein receptor-related protein (LRP). J. Biol. Chem. 2002, 277, 11772–11779. [Google Scholar] [CrossRef]
  42. Albert, M.L.; Kim, J.I.; Birge, R.B. alphavbeta5 integrin recruits the CrkII-Dock180-rac1 complex for phagocytosis of apoptotic cells. Nat. Cell. Biol. 2000, 2, 899–905. [Google Scholar] [CrossRef] [PubMed]
  43. Wu, Y.; Singh, S.; Georgescu, M.-M.; Birge, R.B. A role for Mer tyrosine kinase in alphavbeta5 integrin-mediated phagocytosis of apoptotic cells. J. Cell. Sci. 2005, 118, 539–553. [Google Scholar] [CrossRef] [PubMed]
  44. Flannagan, R.S.; Jaumouillé, V.; Grinstein, S. The cell biology of phagocytosis. Annu. Rev. Pathol. 2012, 7, 61–98. [Google Scholar] [CrossRef]
  45. Rink, J.; Ghigo, E.; Kalaidzidis, Y.; Zerial, M. Rab conversion as a mechanism of progression from early to late endosomes. Cell 2005, 122, 735–749. [Google Scholar] [CrossRef] [PubMed]
  46. Stenmark, H. Rab GTPases as coordinators of vesicle traffic. Nat. Rev. Mol. Cell. Biol. 2009, 10, 513–525. [Google Scholar] [CrossRef]
  47. Rubino, M.; Miaczynska, M.; Lippé, R.; Zerial, M. Selective membrane recruitment of EEA1 suggests a role in directional transport of clathrin-coated vesicles to early endosomes. J. Biol. Chem. 2000, 275, 3745–3748. [Google Scholar] [CrossRef] [PubMed]
  48. Kinchen, J.M.; Ravichandran, K.S. Identification of two evolutionarily conserved genes regulating processing of engulfed apoptotic cells. Nature 2010, 464, 778–782. [Google Scholar] [CrossRef] [PubMed]
  49. Poteryaev, D.; Datta, S.; Ackema, K.; Zerial, M.; Spang, A. Identification of the switch in early-to-late endosome transition. Cell 2010, 141, 497–508. [Google Scholar] [CrossRef] [PubMed]
  50. Fairn, G.D.; Grinstein, S. How nascent phagosomes mature to become phagolysosomes. Trends Immunol. 2012, 33, 397–405. [Google Scholar] [CrossRef] [PubMed]
  51. Saas, P.; Vetter, M.; Maraux, M.; Bonnefoy, F.; Perruche, S. Resolution therapy: Harnessing efferocytic macrophages to trigger the resolution of inflammation. Front. Immunol. 2022, 13, 1021413. [Google Scholar] [CrossRef]
  52. Park, S.-Y.; Jung, M.-Y.; Lee, S.-J.; Kang, K.-B.; Gratchev, A.; Riabov, V.; Kzhyshkowska, J.; Kim, I.-S. Stabilin-1 mediates phosphatidylserine-dependent clearance of cell corpses in alternatively activated macrophages. J. Cell. Sci. 2009, 122, 3365–3373. [Google Scholar] [CrossRef] [PubMed]
  53. Lee, C.S.; Penberthy, K.K.; Wheeler, K.M.; Juncadella, I.J.; Vandenabeele, P.; Lysiak, J.J.; Ravichandran, K.S. Boosting Apoptotic Cell Clearance by Colonic Epithelial Cells Attenuates Inflammation In Vivo. Immunity 2016, 44, 807–820. [Google Scholar] [CrossRef] [PubMed]
  54. Banerjee, S.; Friggeri, A.; Liu, G.; Abraham, E. The C-terminal acidic tail is responsible for the inhibitory effects of HMGB1 on efferocytosis. J. Leukoc. Biol. 2010, 88, 973–979. [Google Scholar] [CrossRef] [PubMed]
  55. Friggeri, A.; Yang, Y.; Banerjee, S.; Park, Y.-J.; Liu, G.; Abraham, E. HMGB1 inhibits macrophage activity in efferocytosis through binding to the alphavbeta3-integrin. Am. J. Physiol. Cell. Physiol. 2010, 299, C1267–C1276. [Google Scholar] [CrossRef] [PubMed]
  56. He, M.; Kubo, H.; Morimoto, K.; Fujino, N.; Suzuki, T.; Takahasi, T.; Yamada, M.; Yamaya, M.; Maekawa, T.; Yamamoto, Y.; et al. Receptor for advanced glycation end products binds to phosphatidylserine and assists in the clearance of apoptotic cells. EMBO Rep. 2011, 12, 358–364. [Google Scholar] [CrossRef] [PubMed]
  57. Kenis, H.; van Genderen, H.; Deckers, N.M.; Lux, P.A.G.; Hofstra, L.; Narula, J.; Reutelingsperger, C.P.M. Annexin A5 inhibits engulfment through internalization of PS-expressing cell membrane patches. Exp. Cell. Res. 2006, 312, 719–726. [Google Scholar] [CrossRef]
  58. Berry, A.; Balard, P.; Coste, A.; Olagnier, D.; Lagane, C.; Authier, H.; Benoit-Vical, F.; Lepert, J.-C.; Séguéla, J.-P.; Magnaval, J.-F.; et al. IL-13 induces expression of CD36 in human monocytes through PPARgamma activation. Eur. J. Immunol. 2007, 37, 1642–1652. [Google Scholar] [CrossRef]
  59. Szanto, A.; Balint, B.L.; Nagy, Z.S.; Barta, E.; Dezso, B.; Pap, A.; Szeles, L.; Poliska, S.; Oros, M.; Evans, R.M.; et al. STAT6 transcription factor is a facilitator of the nuclear receptor PPARγ-regulated gene expression in macrophages and dendritic cells. Immunity 2010, 33, 699–712. [Google Scholar] [CrossRef]
  60. Welch, J.S.; Ricote, M.; Akiyama, T.E.; Gonzalez, F.J.; Glass, C.K. PPARgamma and PPARdelta negatively regulate specific subsets of lipopolysaccharide and IFN-gamma target genes in macrophages. Proc. Natl. Acad. Sci. USA 2003, 100, 6712–6717. [Google Scholar] [CrossRef]
  61. Im, S.-S.; Osborne, T.F. Liver x receptors in atherosclerosis and inflammation. Circ. Res. 2011, 108, 996–1001. [Google Scholar] [CrossRef] [PubMed]
  62. Loren, J.; Huang, Z.; Laffitte, B.A.; Molteni, V. Liver X receptor modulators: A review of recently patented compounds (2009–2012). Expert. Opin. Ther. Pat. 2013, 23, 1317–1335. [Google Scholar] [CrossRef] [PubMed]
  63. Frasch, S.C.; Berry, K.Z.; Fernandez-Boyanapalli, R.; Jin, H.-S.; Leslie, C.; Henson, P.M.; Murphy, R.C.; Bratton, D.L. NADPH oxidase-dependent generation of lysophosphatidylserine enhances clearance of activated and dying neutrophils via G2A. J. Biol. Chem. 2008, 283, 33736–33749. [Google Scholar] [CrossRef]
  64. Greenberg, M.E.; Sun, M.; Zhang, R.; Febbraio, M.; Silverstein, R.; Hazen, S.L. Oxidized phosphatidylserine-CD36 interactions play an essential role in macrophage-dependent phagocytosis of apoptotic cells. J. Exp. Med. 2006, 203, 2613–2625. [Google Scholar] [CrossRef]
  65. Frasch, S.C.; Fernandez-Boyanapalli, R.F.; Berry, K.Z.; Leslie, C.C.; Bonventre, J.V.; Murphy, R.C.; Henson, P.M.; Bratton, D.L. Signaling via macrophage G2A enhances efferocytosis of dying neutrophils by augmentation of Rac activity. J. Biol. Chem. 2011, 286, 12108–12122. [Google Scholar] [CrossRef] [PubMed]
  66. Morimoto, K.; Janssen, W.J.; Fessler, M.B.; McPhillips, K.A.; Borges, V.M.; Bowler, R.P.; Xiao, Y.-Q.; Kench, J.A.; Henson, P.M.; Vandivier, R.W. Lovastatin enhances clearance of apoptotic cells (efferocytosis) with implications for chronic obstructive pulmonary disease. J. Immunol. 2006, 176, 7657–7665. [Google Scholar] [CrossRef] [PubMed]
  67. Godson, C.; Mitchell, S.; Harvey, K.; Petasis, N.A.; Hogg, N.; Brady, H.R. Cutting edge: Lipoxins rapidly stimulate nonphlogistic phagocytosis of apoptotic neutrophils by monocyte-derived macrophages. J. Immunol. 2000, 164, 1663–1667. [Google Scholar] [CrossRef] [PubMed]
  68. Viaud, M.; Ivanov, S.; Vujic, N.; Duta-Mare, M.; Aira, L.-E.; Barouillet, T.; Garcia, E.; Orange, F.; Dugail, I.; Hainault, I.; et al. Lysosomal Cholesterol Hydrolysis Couples Efferocytosis to Anti-Inflammatory Oxysterol Production. Circ. Res. 2018, 122, 1369–1384. [Google Scholar] [CrossRef] [PubMed]
  69. Park, D.; Han, C.Z.; Elliott, M.R.; Kinchen, J.M.; Trampont, P.C.; Das, S.; Collins, S.; Lysiak, J.J.; Hoehn, K.L.; Ravichandran, K.S. Continued clearance of apoptotic cells critically depends on the phagocyte Ucp2 protein. Nature 2011, 477, 220–224. [Google Scholar] [CrossRef]
  70. Jiang, H.; Shi, H.; Sun, M.; Wang, Y.; Meng, Q.; Guo, P.; Cao, Y.; Chen, J.; Gao, X.; Li, E.; et al. PFKFB3-Driven Macrophage Glycolytic Metabolism Is a Crucial Component of Innate Antiviral Defense. J. Immunol. 2016, 197, 2880–2890. [Google Scholar] [CrossRef]
  71. Mao, Y. Apoptotic cell-derived metabolites in efferocytosis-mediated resolution of inflammation. Cytokine Growth Factor Rev. 2021, 62, 42–53. [Google Scholar] [CrossRef]
  72. A-Gonzalez, N.; Hidalgo, A. Nuclear Receptors and Clearance of Apoptotic Cells: Stimulating the Macrophage’s Appetite. Front. Immunol. 2014, 5, 211. [Google Scholar] [CrossRef] [PubMed]
  73. Rébé, C.; Raveneau, M.; Chevriaux, A.; Lakomy, D.; Sberna, A.-L.; Costa, A.; Bessède, G.; Athias, A.; Steinmetz, E.; Lobaccaro, J.M.A.; et al. Induction of transglutaminase 2 by a liver X receptor/retinoic acid receptor alpha pathway increases the clearance of apoptotic cells by human macrophages. Circ. Res. 2009, 105, 393–401. [Google Scholar] [CrossRef] [PubMed]
  74. Tóth, B.; Garabuczi, E.; Sarang, Z.; Vereb, G.; Vámosi, G.; Aeschlimann, D.; Blaskó, B.; Bécsi, B.; Erdõdi, F.; Lacy-Hulbert, A.; et al. Transglutaminase 2 is needed for the formation of an efficient phagocyte portal in macrophages engulfing apoptotic cells. J. Immunol. 2009, 182, 2084–2092. [Google Scholar] [CrossRef] [PubMed]
  75. Kourtzelis, I.; Li, X.; Mitroulis, I.; Grosser, D.; Kajikawa, T.; Wang, B.; Grzybek, M.; von Renesse, J.; Czogalla, A.; Troullinaki, M.; et al. DEL-1 promotes macrophage efferocytosis and clearance of inflammation. Nat. Immunol. 2019, 20, 40–49. [Google Scholar] [CrossRef]
  76. Xian, X.; Ding, Y.; Dieckmann, M.; Zhou, L.; Plattner, F.; Liu, M.; Parks, J.S.; Hammer, R.E.; Boucher, P.; Tsai, S.; et al. LRP1 integrates murine macrophage cholesterol homeostasis and inflammatory responses in atherosclerosis. Elife 2017, 6, e29292. [Google Scholar] [CrossRef] [PubMed]
  77. Yurdagul, A.; Subramanian, M.; Wang, X.; Crown, S.B.; Ilkayeva, O.R.; Darville, L.; Kolluru, G.K.; Rymond, C.C.; Gerlach, B.D.; Zheng, Z.; et al. Macrophage Metabolism of Apoptotic Cell-Derived Arginine Promotes Continual Efferocytosis and Resolution of Injury. Cell. Metab. 2020, 31, 518–533.e10. [Google Scholar] [CrossRef]
  78. Rath, M.; Müller, I.; Kropf, P.; Closs, E.I.; Munder, M. Metabolism via Arginase or Nitric Oxide Synthase: Two Competing Arginine Pathways in Macrophages. Front. Immunol. 2014, 5, 532. [Google Scholar] [CrossRef]
  79. Ray, R.M.; Bavaria, M.N.; Bhattacharya, S.; Johnson, L.R. Activation of Dbl restores migration in polyamine-depleted intestinal epithelial cells via Rho-GTPases. Am. J. Physiol. Gastrointest Liver Physiol. 2011, 300, G988–G997. [Google Scholar] [CrossRef] [PubMed]
  80. Ray, R.M.; McCormack, S.A.; Covington, C.; Viar, M.J.; Zheng, Y.; Johnson, L.R. The requirement for polyamines for intestinal epithelial cell migration is mediated through Rac1. J. Biol. Chem. 2003, 278, 13039–13046. [Google Scholar] [CrossRef] [PubMed]
  81. Perry, J.S.A.; Morioka, S.; Medina, C.B.; Iker Etchegaray, J.; Barron, B.; Raymond, M.H.; Lucas, C.D.; Onengut-Gumuscu, S.; Delpire, E.; Ravichandran, K.S. Interpreting an apoptotic corpse as anti-inflammatory involves a chloride sensing pathway. Nat. Cell. Biol. 2019, 21, 1532–1543. [Google Scholar] [CrossRef]
  82. Grabiec, A.M.; Hussell, T. The role of airway macrophages in apoptotic cell clearance following acute and chronic lung inflammation. Semin. Immunopathol. 2016, 38, 409–423. [Google Scholar] [CrossRef] [PubMed]
  83. Krysko, O.; Vandenabeele, P.; Krysko, D.V.; Bachert, C. Impairment of phagocytosis of apoptotic cells and its role in chronic airway diseases. Apoptosis 2010, 15, 1137–1146. [Google Scholar] [CrossRef] [PubMed]
  84. Schmidt, E.P.; Tuder, R.M. Role of Apoptosis in Amplifying Inflammatory Responses in Lung Diseases. J. Cell. Death 2010, 3, 41–53. [Google Scholar] [CrossRef] [PubMed]
  85. McCubbrey, A.L.; Curtis, J.L. Efferocytosis and lung disease. Chest 2013, 143, 1750–1757. [Google Scholar] [CrossRef]
  86. Asare, P.F.; Tran, H.B.; Hurtado, P.R.; Perkins, G.B.; Nguyen, P.; Jersmann, H.; Roscioli, E.; Hodge, S. Inhibition of LC3-associated phagocytosis in COPD and in response to cigarette smoke. Ther. Adv. Respir. Dis. 2021, 15, 17534666211039769. [Google Scholar] [CrossRef]
  87. Liu, Y.; Liu, H.; Li, C.; Ma, C.; Ge, W. Proteome Profiling of Lung Tissues in Chronic Obstructive Pulmonary Disease (COPD): Platelet and Macrophage Dysfunction Contribute to the Pathogenesis of COPD. Int. J. Chron. Obstruct Pulmon. Dis. 2020, 15, 973–980. [Google Scholar] [CrossRef]
  88. Noguera, A.; Gomez, C.; Faner, R.; Cosio, B.; González-Périz, A.; Clària, J.; Carvajal, A.; Agustí, A. An investigation of the resolution of inflammation (catabasis) in COPD. Respir. Res. 2012, 13, 101. [Google Scholar] [CrossRef] [PubMed]
  89. Hamon, R.; Homan, C.C.; Tran, H.B.; Mukaro, V.R.; Lester, S.E.; Roscioli, E.; Bosco, M.D.; Murgia, C.M.; Ackland, M.L.; Jersmann, H.P.; et al. Zinc and zinc transporters in macrophages and their roles in efferocytosis in COPD. PLoS ONE 2014, 9, e110056. [Google Scholar] [CrossRef] [PubMed]
  90. Vachon, E.; Martin, R.; Kwok, V.; Cherepanov, V.; Chow, C.-W.; Doerschuk, C.M.; Plumb, J.; Grinstein, S.; Downey, G.P. CD44-mediated phagocytosis induces inside-out activation of complement receptor-3 in murine macrophages. Blood 2007, 110, 4492–4502. [Google Scholar] [CrossRef]
  91. Janssen, W.J.; McPhillips, K.A.; Dickinson, M.G.; Linderman, D.J.; Morimoto, K.; Xiao, Y.Q.; Oldham, K.M.; Vandivier, R.W.; Henson, P.M.; Gardai, S.J. Surfactant proteins A and D suppress alveolar macrophage phagocytosis via interaction with SIRP alpha. Am. J. Respir. Crit. Care Med. 2008, 178, 158–167. [Google Scholar] [CrossRef]
  92. Sims, M.W.; Tal-Singer, R.M.; Kierstein, S.; Musani, A.I.; Beers, M.F.; Panettieri, R.A.; Haczku, A. Chronic obstructive pulmonary disease and inhaled steroids alter surfactant protein D (SP-D) levels: A cross-sectional study. Respir. Res. 2008, 9, 13. [Google Scholar] [CrossRef] [PubMed]
  93. Minematsu, N.; Blumental-Perry, A.; Shapiro, S.D. Cigarette smoke inhibits engulfment of apoptotic cells by macrophages through inhibition of actin rearrangement. Am. J. Respir. Cell. Mol. Biol. 2011, 44, 474–482. [Google Scholar] [CrossRef] [PubMed]
  94. Richens, T.R.; Linderman, D.J.; Horstmann, S.A.; Lambert, C.; Xiao, Y.-Q.; Keith, R.L.; Boé, D.M.; Morimoto, K.; Bowler, R.P.; Day, B.J.; et al. Cigarette smoke impairs clearance of apoptotic cells through oxidant-dependent activation of RhoA. Am. J. Respir. Crit. Care Med. 2009, 179, 1011–1021. [Google Scholar] [CrossRef]
  95. Vlahos, R.; Bozinovski, S. Role of alveolar macrophages in chronic obstructive pulmonary disease. Front. Immunol. 2014, 5, 435. [Google Scholar] [CrossRef]
  96. Barnawi, J.; Jersmann, H.; Haberberger, R.; Hodge, S.; Meech, R. Reduced DNA methylation of sphingosine-1 phosphate receptor 5 in alveolar macrophages in COPD: A potential link to failed efferocytosis. Respirology 2017, 22, 315–321. [Google Scholar] [CrossRef] [PubMed]
  97. Bewley, M.A.; Belchamber, K.B.R.; Chana, K.K.; Budd, R.C.; Donaldson, G.; Wedzicha, J.A.; Brightling, C.E.; Kilty, I.; Donnelly, L.E.; Barnes, P.J.; et al. Differential Effects of p38, MAPK, PI3K or Rho Kinase Inhibitors on Bacterial Phagocytosis and Efferocytosis by Macrophages in COPD. PLoS ONE 2016, 11, e0163139. [Google Scholar] [CrossRef] [PubMed]
  98. Hodge, S.; Jersmann, H.; Reynolds, P.N. The Effect of Colonization with Potentially Pathogenic Microorganisms on Efferocytosis in Chronic Obstructive Pulmonary Disease. Am. J. Respir. Crit. Care Med. 2016, 194, 912–915. [Google Scholar] [CrossRef] [PubMed]
  99. Lea, S.; Plumb, J.; Metcalfe, H.; Spicer, D.; Woodman, P.; Fox, J.C.; Singh, D. The effect of peroxisome proliferator-activated receptor-γ ligands on in vitro and in vivo models of COPD. Eur. Respir. J. 2014, 43, 409–420. [Google Scholar] [CrossRef]
  100. Mukaro, V.R.; Bylund, J.; Hodge, G.; Holmes, M.; Jersmann, H.; Reynolds, P.N.; Hodge, S. Lectins offer new perspectives in the development of macrophage-targeted therapies for COPD/emphysema. PLoS ONE 2013, 8, e56147. [Google Scholar] [CrossRef]
  101. Mukaro, V.R.S.; Hodge, S. Airway clearance of apoptotic cells in COPD. Curr. Drug Targets 2011, 12, 460–468. [Google Scholar] [CrossRef] [PubMed]
  102. Gao, P.; Gibson, P.G.; Baines, K.J.; Yang, I.A.; Upham, J.W.; Reynolds, P.N.; Hodge, S.; James, A.L.; Jenkins, C.; Peters, M.J.; et al. Anti-inflammatory deficiencies in neutrophilic asthma: Reduced galectin-3 and IL-1RA/IL-1β. Respir. Res. 2015, 16, 5. [Google Scholar] [CrossRef] [PubMed]
  103. van der Veen, T.A.; de Groot, L.E.S.; Melgert, B.N. The different faces of the macrophage in asthma. Curr. Opin. Pulm. Med. 2020, 26, 62–68. [Google Scholar] [CrossRef] [PubMed]
  104. Zizzo, G.; Hilliard, B.A.; Monestier, M.; Cohen, P.L. Efficient clearance of early apoptotic cells by human macrophages requires M2c polarization and MerTK induction. J. Immunol. 2012, 189, 3508–3520. [Google Scholar] [CrossRef]
  105. Draijer, C.; Boorsma, C.E.; Robbe, P.; Timens, W.; Hylkema, M.N.; Ten Hacken, N.H.; van den Berge, M.; Postma, D.S.; Melgert, B.N. Human asthma is characterized by more IRF5+ M1 and CD206+ M2 macrophages and less IL-10+ M2-like macrophages around airways compared with healthy airways. J. Allergy Clin. Immunol. 2017, 140, 280–283.e3. [Google Scholar] [CrossRef] [PubMed]
  106. Felton, J.M.; Lucas, C.D.; Dorward, D.A.; Duffin, R.; Kipari, T.; Vermeren, S.; Robb, C.T.; MacLeod, K.G.; Serrels, B.; Schwarze, J.; et al. Mer-mediated eosinophil efferocytosis regulates resolution of allergic airway inflammation. J. Allergy Clin. Immunol. 2018, 142, 1884–1893.e6. [Google Scholar] [CrossRef] [PubMed]
  107. Grabiec, A.M.; Denny, N.; Doherty, J.A.; Happonen, K.E.; Hankinson, J.; Connolly, E.; Fife, M.E.; Fujimori, T.; Fujino, N.; Goenka, A.; et al. Diminished airway macrophage expression of the Axl receptor tyrosine kinase is associated with defective efferocytosis in asthma. J. Allergy Clin. Immunol. 2017, 140, 1144–1146.e4. [Google Scholar] [CrossRef]
  108. Huang, Q.; Weng, D.; Yao, S.; Shen, H.; Gao, S.; Zhang, Y.; Huang, W.; Wang, Y.; Wang, H.; Xu, W. Progranulin deficiency suppresses allergic asthma and enhances efferocytosis via PPAR-γ/MFG-E8 regulation in macrophages. Immun. Inflamm. Dis. 2023, 11, e779. [Google Scholar] [CrossRef] [PubMed]
  109. Immormino, R.M.; Jania, C.M.; Tilley, S.L.; Moran, T.P. Neuropilin-2 regulates airway inflammation in a neutrophilic asthma model. Immun. Inflamm. Dis. 2022, 10, e575. [Google Scholar] [CrossRef] [PubMed]
  110. Erol, A. Importance of Efferocytosis in COVID-19 Mortality. Infect. Drug Resist. 2022, 15, 995–1007. [Google Scholar] [CrossRef] [PubMed]
  111. Dutta, S.; Mukherjee, A.; Nongthomba, U. Before the “cytokine storm”: Boosting efferocytosis as an effective strategy against SARS-CoV-2 infection and associated complications. Cytokine Growth Factor Rev. 2022, 63, 108–118. [Google Scholar] [CrossRef]
  112. 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]
  113. Astuti, I.; Ysrafil. Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2): An overview of viral structure and host response. Diabetes Metab. Syndr. 2020, 14, 407–412. [Google Scholar] [CrossRef] [PubMed]
  114. Shereen, M.A.; Khan, S.; Kazmi, A.; Bashir, N.; Siddique, R. COVID-19 infection: Origin, transmission, and characteristics of human coronaviruses. J. Adv. Res. 2020, 24, 91–98. [Google Scholar] [CrossRef] [PubMed]
  115. Soy, M.; Keser, G.; Atagündüz, P.; Tabak, F.; Atagündüz, I.; Kayhan, S. Cytokine storm in COVID-19: Pathogenesis and overview of anti-inflammatory agents used in treatment. Clin. Rheumatol. 2020, 39, 2085–2094. [Google Scholar] [CrossRef]
  116. Habib, H.M.; Ibrahim, S.; Zaim, A.; Ibrahim, W.H. The role of iron in the pathogenesis of COVID-19 and possible treatment with lactoferrin and other iron chelators. Biomed. Pharmacother. 2021, 136, 111228. [Google Scholar] [CrossRef]
  117. Salina, A.C.G.; Dos-Santos, D.; Rodrigues, T.S.; Fortes-Rocha, M.; Freitas-Filho, E.G.; Alzamora-Terrel, D.L.; Castro, I.M.S.; Fraga da Silva, T.F.C.; de Lima, M.H.F.; Nascimento, D.C.; et al. Efferocytosis of SARS-CoV-2-infected dying cells impairs macrophage anti-inflammatory functions and clearance of apoptotic cells. Elife 2022, 11, e74443. [Google Scholar] [CrossRef]
  118. Meacci, E.; Garcia-Gil, M.; Pierucci, F. SARS-CoV-2 Infection: A Role for S1P/S1P Receptor Signaling in the Nervous System? Int. J. Mol. Sci. 2020, 21, 6773. [Google Scholar] [CrossRef] [PubMed]
  119. Nepal, S.; Tiruppathi, C.; Tsukasaki, Y.; Farahany, J.; Mittal, M.; Rehman, J.; Prockop, D.J.; Malik, A.B. STAT6 induces expression of Gas6 in macrophages to clear apoptotic neutrophils and resolve inflammation. Proc. Natl. Acad. Sci. USA 2019, 116, 16513–16518. [Google Scholar] [CrossRef]
  120. Sim, M.M.S.; Wood, J.P. Dysregulation of Protein S in COVID-19. Best Pract. Res. Clin. Haematol. 2022, 35, 101376. [Google Scholar] [CrossRef]
  121. Ubil, E.; Caskey, L.; Holtzhausen, A.; Hunter, D.; Story, C.; Earp, H.S. Tumor-secreted Pros1 inhibits macrophage M1 polarization to reduce antitumor immune response. J. Clin. Investig. 2018, 128, 2356–2369. [Google Scholar] [CrossRef]
  122. Driscoll, W.S.; Vaisar, T.; Tang, J.; Wilson, C.L.; Raines, E.W. Macrophage ADAM17 deficiency augments CD36-dependent apoptotic cell uptake and the linked anti-inflammatory phenotype. Circ. Res. 2013, 113, 52–61. [Google Scholar] [CrossRef] [PubMed]
  123. Moller-Tank, S.; Maury, W. Phosphatidylserine receptors: Enhancers of enveloped virus entry and infection. Virology 2014, 468–470, 565–580. [Google Scholar] [CrossRef] [PubMed]
  124. Tajbakhsh, A.; Gheibi Hayat, S.M.; Butler, A.E.; Sahebkar, A. Effect of soluble cleavage products of important receptors/ligands on efferocytosis: Their role in inflammatory, autoimmune and cardiovascular disease. Ageing Res. Rev. 2019, 50, 43–57. [Google Scholar] [CrossRef]
  125. Lartey, N.L.; Valle-Reyes, S.; Vargas-Robles, H.; Jiménez-Camacho, K.E.; Guerrero-Fonseca, I.M.; Castellanos-Martínez, R.; Montoya-García, A.; García-Cordero, J.; Cedillo-Barrón, L.; Nava, P.; et al. ADAM17/MMP inhibition prevents neutrophilia and lung injury in a mouse model of COVID-19. J. Leukoc. Biol. 2022, 111, 1147–1158. [Google Scholar] [CrossRef] [PubMed]
  126. Burrage, D.R.; Koushesh, S.; Sofat, N. Immunomodulatory Drugs in the Management of SARS-CoV-2. Front. Immunol. 2020, 11, 1844. [Google Scholar] [CrossRef] [PubMed]
  127. Calligaris, M.; Cuffaro, D.; Bonelli, S.; Spanò, D.P.; Rossello, A.; Nuti, E.; Scilabra, S.D. Strategies to Target ADAM17 in Disease: From its Discovery to the iRhom Revolution. Molecules 2021, 26, 944. [Google Scholar] [CrossRef] [PubMed]
  128. Oltean, T.; Ibanez, L.I.; Divert, T.; Ysenbaert, T.; Van Eeckhoutte, H.; Goossens, V.; Schotsaert, M.; Bracke, K.; Schepens, B.; Maelfait, J.; et al. Reduced protection of RIPK3-deficient mice against influenza by matrix protein 2 ectodomain targeted active and passive vaccination strategies. Cell. Death Dis. 2022, 13, 280. [Google Scholar] [CrossRef]
  129. Mukherjee, S.; Subramaniam, R.; Chen, H.; Smith, A.; Keshava, S.; Shams, H. Boosting efferocytosis in alveolar space using BCG vaccine to protect host against influenza pneumonia. PLoS ONE 2017, 12, e0180143. [Google Scholar] [CrossRef] [PubMed]
  130. Nelson, B.; Zhou, X.; White, M.; Hartshorn, K.; Takahashi, K.; Kinane, T.B.; Anandaiah, A.; Koziel, H. Recombinant human mannose-binding lectin dampens human alveolar macrophage inflammatory responses to influenza A virus in vitro. J. Leukoc. Biol. 2014, 95, 715–722. [Google Scholar] [CrossRef] [PubMed]
  131. Melo, E.M.; Del Sarto, J.; Vago, J.P.; Tavares, L.P.; Rago, F.; Gonçalves, A.P.F.; Machado, M.G.; Aranda-Pardos, I.; Valiate, B.V.S.; Cassali, G.D.; et al. Relevance of angiotensin-(1-7) and its receptor Mas in pneumonia caused by influenza virus and post-influenza pneumococcal infection. Pharmacol. Res. 2021, 163, 105292. [Google Scholar] [CrossRef] [PubMed]
  132. Burnham, E.L.; Janssen, W.J.; Riches, D.W.H.; Moss, M.; Downey, G.P. The fibroproliferative response in acute respiratory distress syndrome: Mechanisms and clinical significance. Eur. Respir. J. 2014, 43, 276–285. [Google Scholar] [CrossRef] [PubMed]
  133. Noone, P.M.; Reddy, S.P. Recent advances in dead cell clearance during acute lung injury and repair. Fac. Rev. 2021, 10, 33. [Google Scholar] [CrossRef]
  134. Thille, A.W.; Esteban, A.; Fernández-Segoviano, P.; Rodriguez, J.-M.; Aramburu, J.-A.; Peñuelas, O.; Cortés-Puch, I.; Cardinal-Fernández, P.; Lorente, J.A.; Frutos-Vivar, F. Comparison of the Berlin definition for acute respiratory distress syndrome with autopsy. Am. J. Respir. Crit. Care Med. 2013, 187, 761–767. [Google Scholar] [CrossRef]
  135. Matsuda, N.; Yamamoto, S.; Takano, K.-I.; Kageyama, S.-I.; Kurobe, Y.; Yoshihara, Y.; Takano, Y.; Hattori, Y. Silencing of fas-associated death domain protects mice from septic lung inflammation and apoptosis. Am. J. Respir. Crit. Care Med. 2009, 179, 806–815. [Google Scholar] [CrossRef] [PubMed]
  136. Abraham, E. What role does neutrophil apoptosis play in acute respiratory distress syndrome? Crit. Care Med. 2000, 28, 253–254. [Google Scholar] [CrossRef]
  137. Hussain, N.; Wu, F.; Zhu, L.; Thrall, R.S.; Kresch, M.J. Neutrophil apoptosis during the development and resolution of oleic acid-induced acute lung injury in the rat. Am. J. Respir. Cell. Mol. Biol. 1998, 19, 867–874. [Google Scholar] [CrossRef] [PubMed]
  138. Mahida, R.Y.; Scott, A.; Parekh, D.; Lugg, S.T.; Hardy, R.S.; Lavery, G.G.; Matthay, M.A.; Naidu, B.; Perkins, G.D.; Thickett, D.R. Acute respiratory distress syndrome is associated with impaired alveolar macrophage efferocytosis. Eur. Respir. J. 2021, 58, 2100829. [Google Scholar] [CrossRef] [PubMed]
  139. Grégoire, M.; Uhel, F.; Lesouhaitier, M.; Gacouin, A.; Guirriec, M.; Mourcin, F.; Dumontet, E.; Chalin, A.; Samson, M.; Berthelot, L.-L.; et al. Impaired efferocytosis and neutrophil extracellular trap clearance by macrophages in ARDS. Eur. Respir. J. 2018, 52, 1702590. [Google Scholar] [CrossRef]
  140. Donnelly, L.E.; Barnes, P.J. Defective phagocytosis in airways disease. Chest 2012, 141, 1055–1062. [Google Scholar] [CrossRef]
  141. Chmiel, J.F.; Berger, M.; Konstan, M.W. The role of inflammation in the pathophysiology of CF lung disease. Clin. Rev. Allergy Immunol. 2002, 23, 5–27. [Google Scholar] [CrossRef]
  142. Vandivier, R.W.; Fadok, V.A.; Hoffmann, P.R.; Bratton, D.L.; Penvari, C.; Brown, K.K.; Brain, J.D.; Accurso, F.J.; Henson, P.M. Elastase-mediated phosphatidylserine receptor cleavage impairs apoptotic cell clearance in cystic fibrosis and bronchiectasis. J. Clin. Investig. 2002, 109, 661–670. [Google Scholar] [CrossRef]
  143. Morimoto, K.; Janssen, W.J.; Terada, M. Defective efferocytosis by alveolar macrophages in IPF patients. Respir. Med. 2012, 106, 1800–1803. [Google Scholar] [CrossRef] [PubMed]
  144. Ryu, J.H.; Moua, T.; Daniels, C.E.; Hartman, T.E.; Yi, E.S.; Utz, J.P.; Limper, A.H. Idiopathic pulmonary fibrosis: Evolving concepts. Mayo Clin. Proc. 2014, 89, 1130–1142. [Google Scholar] [CrossRef] [PubMed]
  145. Finnerty, J.P.; Ponnuswamy, A.; Dutta, P.; Abdelaziz, A.; Kamil, H. Efficacy of antifibrotic drugs, nintedanib and pirfenidone, in treatment of progressive pulmonary fibrosis in both idiopathic pulmonary fibrosis (IPF) and non-IPF: A systematic review and meta-analysis. BMC Pulm. Med. 2021, 21, 411. [Google Scholar] [CrossRef] [PubMed]
  146. Giles, K.M.; Ross, K.; Rossi, A.G.; Hotchin, N.A.; Haslett, C.; Dransfield, I. Glucocorticoid augmentation of macrophage capacity for phagocytosis of apoptotic cells is associated with reduced p130Cas expression, loss of paxillin/pyk2 phosphorylation, and high levels of active Rac. J. Immunol. 2001, 167, 976–986. [Google Scholar] [CrossRef] [PubMed]
  147. Tosello-Trampont, A.-C.; Nakada-Tsukui, K.; Ravichandran, K.S. Engulfment of apoptotic cells is negatively regulated by Rho-mediated signaling. J. Biol. Chem. 2003, 278, 49911–49919. [Google Scholar] [CrossRef] [PubMed]
  148. Tanegashima, K.; Sato-Miyata, Y.; Funakoshi, M.; Nishito, Y.; Aigaki, T.; Hara, T. Epigenetic regulation of the glucose transporter gene Slc2a1 by β-hydroxybutyrate underlies preferential glucose supply to the brain of fasted mice. Genes Cells 2017, 22, 71–83. [Google Scholar] [CrossRef]
Figure 1. The professional phagocytes and non-professional phagocytes in the human body.
Figure 1. The professional phagocytes and non-professional phagocytes in the human body.
Ijms 24 14871 g001
Figure 2. The mechanisms of efferocytosis. Take macrophages as an example. (A) The smell stage: apoptotic cells release “find-me” signals to recruit macrophages to the site of death. (B) The taste stage: Macrophages recognize apoptotic cells via “eat-me” signals, which are mainly phosphatidylserine (PS). This process occurs in two ways: one is direct binding of macrophage surface receptors to PS, such as BAI-1, Tim4 and Stabilin2, and the other is indirect binding to PS in the presence of the bridging molecules. Calreticulin is also an eat-me signal that can directly bind to LRP1/CD91 on the surface of macrophages. (C) The phagocytic stage: After phagocytes recognize apoptotic cells, Rac on the plasma membrane is activated and the cytoskeleton is rearranged under the plasma membrane, eventually forming phagosomes. There are two main pathways in this process. One is the recruitment of the articulin GULP after binding of LRP1 (CD91) to calreticulin or Stab2 to PS, which activates Rac. The other is the recruitment of the articulin CrkII to the cell membrane after indirect binding of MerTK to PS. Then, the articulin CrkII binds to the ELMO1-DOCK180 complex that is formed and activated by the binding of BAI-1 to PS. Finally, the CrkII-DOCK180-ELMO complex activates the RHO family, in particular, Rac. (D) The digestion stage: Phagosomes fuse with early endosomes in the presence of activated Rab5, VPS34 and PI3P to form early phagosomes. The transformation of early phagosomes into late phagosomes is marked by the absence of Rab5 as well as the acquisition of Rab7. Late phagosomes in turn fuse with lysosomes to form phagolysosomes under the synergistic action of activated Rab7, RILP and ORP1. The newly formed phagolysosome relies on proteases and other enzymes within it to digest the phagocytosed cell corpses, and the digested products are reused by macrophages or excreted from the body.
Figure 2. The mechanisms of efferocytosis. Take macrophages as an example. (A) The smell stage: apoptotic cells release “find-me” signals to recruit macrophages to the site of death. (B) The taste stage: Macrophages recognize apoptotic cells via “eat-me” signals, which are mainly phosphatidylserine (PS). This process occurs in two ways: one is direct binding of macrophage surface receptors to PS, such as BAI-1, Tim4 and Stabilin2, and the other is indirect binding to PS in the presence of the bridging molecules. Calreticulin is also an eat-me signal that can directly bind to LRP1/CD91 on the surface of macrophages. (C) The phagocytic stage: After phagocytes recognize apoptotic cells, Rac on the plasma membrane is activated and the cytoskeleton is rearranged under the plasma membrane, eventually forming phagosomes. There are two main pathways in this process. One is the recruitment of the articulin GULP after binding of LRP1 (CD91) to calreticulin or Stab2 to PS, which activates Rac. The other is the recruitment of the articulin CrkII to the cell membrane after indirect binding of MerTK to PS. Then, the articulin CrkII binds to the ELMO1-DOCK180 complex that is formed and activated by the binding of BAI-1 to PS. Finally, the CrkII-DOCK180-ELMO complex activates the RHO family, in particular, Rac. (D) The digestion stage: Phagosomes fuse with early endosomes in the presence of activated Rab5, VPS34 and PI3P to form early phagosomes. The transformation of early phagosomes into late phagosomes is marked by the absence of Rab5 as well as the acquisition of Rab7. Late phagosomes in turn fuse with lysosomes to form phagolysosomes under the synergistic action of activated Rab7, RILP and ORP1. The newly formed phagolysosome relies on proteases and other enzymes within it to digest the phagocytosed cell corpses, and the digested products are reused by macrophages or excreted from the body.
Ijms 24 14871 g002
Table 1. Signals involved in the efferocytosis process.
Table 1. Signals involved in the efferocytosis process.
SignalReceptorFunction
Find-meATP/UTPP2Y2 [9,10]Bind to receptors on phagocytes and recruit phagocytes to apoptotic sites
LPCG2A [11,12,13]
S1PS1P1-5 [11,12,13]
CX3CL1CX3CR1 [14]
ICAM-3CD14 [18]
TSP-1CD36 [18]
Eat-mePSBAI-1 [19]Directly integrate with PS, raise and activate ELMO1-DOCK180 complex
TIM family [10]Directly integrate with PS, TIM-4 triggers LC-3-associated phagocytosis
Stabilin1/2 [20]Directly integrate with PS, activate Rac via GULP
TAM family(Tyr3, Axl, MerTK) [10,21]Indirectly bind to PS, then dimerization occurs or recruit adapter protein CrkII, form ELMO1-DOCK180-CrkII complex, finally activate Rac
αv/β3/CD36 [22]Indirectly bind to PS via TSP-1
αv/β3/TG2 [22]Indirectly bind to PS via MFG-E8
Calreticulin LRP-1/CD91 [20,23,24]Recruit adaptor protein GULP, activate Rac
Don’t eat-meCD47SIRP α [5]Activate SIRPα and make it phosphorylated, activate phosphatase to inhibit phagocytosis
CD24Siglec 10 [5]Highly express on tumor cells, inhibit phagocytosis
CD31[5]Inhibit phagocytosis due to the homologous interactions between healthy cells and macrophages
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

Zheng, W.; Zhou, Z.; Guo, X.; Zuo, X.; Zhang, J.; An, Y.; Zheng, H.; Yue, Y.; Wang, G.; Wang, F. Efferocytosis and Respiratory Disease. Int. J. Mol. Sci. 2023, 24, 14871. https://doi.org/10.3390/ijms241914871

AMA Style

Zheng W, Zhou Z, Guo X, Zuo X, Zhang J, An Y, Zheng H, Yue Y, Wang G, Wang F. Efferocytosis and Respiratory Disease. International Journal of Molecular Sciences. 2023; 24(19):14871. https://doi.org/10.3390/ijms241914871

Chicago/Turabian Style

Zheng, Wenxue, Zhengjie Zhou, Xiaoping Guo, Xu Zuo, Jiaqi Zhang, Yiming An, Haoyu Zheng, Yuan Yue, Guoqiang Wang, and Fang Wang. 2023. "Efferocytosis and Respiratory Disease" International Journal of Molecular Sciences 24, no. 19: 14871. https://doi.org/10.3390/ijms241914871

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