Next Article in Journal
Systemic Administration of Pegylated Arginase-1 Attenuates the Progression of Diabetic Retinopathy
Next Article in Special Issue
Morbidity and Mortality of Neutropenic Patients in Visceral Surgery: A Narrative Review
Previous Article in Journal
The ABA-LANCL1/2 Hormone-Receptors System Protects H9c2 Cardiomyocytes from Hypoxia-Induced Mitochondrial Injury via an AMPK- and NO-Mediated Mechanism
Previous Article in Special Issue
hMRP8-ATTAC Mice: A New Model for Conditional and Reversible Neutrophil Ablation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Neutrophil Activated by the Famous and Potent PMA (Phorbol Myristate Acetate)

by
Hylane Luiz Damascena
,
Wendy Ann Assis Silveira
,
Mariana S. Castro
and
Wagner Fontes
*
Laboratory of Protein Chemistry and Biochemistry, Department of Cell Biology, Institute of Biological Sciences, University of Brasília, Federal District, Brasilia 70910-900, Brazil
*
Author to whom correspondence should be addressed.
Cells 2022, 11(18), 2889; https://doi.org/10.3390/cells11182889
Submission received: 24 July 2022 / Revised: 28 August 2022 / Accepted: 29 August 2022 / Published: 16 September 2022
(This article belongs to the Collection Emerging Roles of Neutrophil in Inflammation and Immunity)

Abstract

:
This review will briefly outline the major signaling pathways in PMA-activated neutrophils. PMA is widely used to understand neutrophil pathways and formation of NETs. PMA activates PKC; however, we highlight some isoforms that contribute to specific functions. PKC α, β and δ contribute to ROS production while PKC βII and PKC ζ are involved in cytoskeleton remodeling. Actin polymerization is important for the chemotaxis of neutrophils and its remodeling is connected to ROS balance. We suggest that, although ROS and production of NETs are usually observed together in PMA-activated neutrophils, there might be a regulatory mechanism balancing both. Interestingly, we suggest that serine proteases might determine the PAD4 action. PAD4 could be responsible for the activation of the NF-κB pathway that leads to IL-1β release, triggering the cleavage of gasdermin D by serine proteases such as elastase, leading to pore formation contributing to release of NETs. On the other hand, when serine proteases are inhibited, NETs are formed by citrullination through the PAD4 pathway. This review puts together results from the last 31 years of research on the effects of PMA on the neutrophil and proposes new insights on their interpretation.

1. Introduction

Neutrophils are the first cells to migrate to an inflammation site. They fight against pathogens using phagocytosis, reactive oxygen species (ROS), enzymes and DNA release in neutrophil extracellular traps (NETs) [1,2,3].
It is through pattern recognition receptors (PRRs) that neutrophils recognize molecular structures such as pathogen-associated molecular patterns (PAMPs) or damage-associated molecular patterns (DAMPs) and start the activation process [3]. Neutrophil activation and recruitment are regulated by pro-inflammatory cytokines, endothelial cells, platelet activation factor (PAF), and microbial products, among other substances [4,5,6], but that is not the case for PMA.
Phorbol 12-myristate 13-acetate (PMA) is a phorbol compound obtained from Croton tiglium, a plant of the Euphorbiaceae family. It was first observed to be a tumor promoter in mouse skin, then other studies showed that the carcinogenic effect was correlated to the proinflammatory activity and to protein kinase C (PKC) activation, producing a diversity of biological effects. Nowadays, PMA is widely used in experiments to promote tumor development and leukocyte activation [7,8]. Although it is an artificial stimulus, it is widely known and commonly used to study the neutrophil, because PMA remarkably induces pathways similar to other stimuli, with the participation of PKC, myeloperoxidase (MPO) and neutrophil elastase (NE) [7,8]. In addition, PMA is known to induce neutrophil degranulation, metabolic changes, ROS production and NET formation, in nanomolar concentrations [9,10,11,12]. PMA is widely used to mimic PKC regulation by the same mechanism as diacylglycerol (DAG), a lipidic second messenger generated by the enzyme phospholipase C (PLC) when catalyzing the degradation of phosphatidylinositol 4, 5 bisphosphate (PIP2). DAG is responsible for the activation of PKC [9].
It is known that PMA activates PKC, however neutrophils express different PKC isoforms, contributing to different functions. Here, we highlight that while α, β and δ isoforms are involved in the translocation and phosphorylation of the NADPH oxidase components [10], and contribute to ROS production [11,12,13], PKC βII and PKC ζ participate in adhesion and migration [13,14,15]. In addition, ROS acts on neutrophil migration at the same time, impairing it through the oxidation of actin [16,17]. PKC δ activates the phosphoinositide-3-kinases (PI3K) pathway through myristoylated alanine-rich C-kinase substrate (MARCKS) phosphorylation. PtdIns (3,4,5) P3 (PIP3), a product of the PI3K pathway, activates Akt that contributes to p47phox phosphorylation, however its contribution to the major death mechanism activated by PMA, a process named NETosis, is unclear. At the same time, PIP3 can activate PKC ζ, which activates protein arginine deiminase 4 (PAD4) promoting the release of NETs. However, there is NETs formation without histone citrullination. This review highlights relations among these aspects and brings together for the first time the two existing ways that PAD4 can contribute to the formation of NETs.
Neutrophil pathways activated by PMA have been well studied, but there are aspects that are not clear. As we did not find a previous comprehensive review about the effects of PMA on neutrophils, and due to the increased number of publications since the 1990s (Figure 1), we also put together an overview of the functional features and their respective molecular pathways that were rather scattered in the literature over the past 31 years, not excluding the most relevant publications before this period. This review will provide a focus on the available knowledge of PMA-activated neutrophil responses, as well as highlight the open questions and disagreement in the literature.

2. Neutrophils Are Activated by PMA

PMA crosses the plasma membrane without the aid of a membrane receptor and directly activates PKC by translocation of the cytosolic isoforms to the cell membrane [18,19].

2.1. Protein Kinase C

Protein kinase C (PKC) is a member of the serine/threonine kinases protein family. This family includes protein kinase A (PKA) and protein kinase B (PKB). The PKC isoforms are classified according to their regulatory domain that is variable among the three PKC subgroups. The conventional PKCs (cPKCs) are made up of four isoforms (α, βI, βII, γ) that are activated by DAG/PMA and are calcium-ion dependent. The novel PKCs (nPKCs) are made up of four isoforms (δ, θ, ξ, η), also activated by DAG/PMA, however they are calcium-independent. The atypical PKCs (aPKCs) are made up of two isoforms (ζ, ι/λ) which are DAG/PMA and calcium independent. They can also be activated by phosphatidylinositol (3,4,5)-trisphosphate (PIP3) or ceramide. In addition, all three groups need the phospholipid, phosphatidylserine, for activation [9].
Neutrophils express five PKC isoforms (α, βI, βII, δ, ζ) which have been suggested to participate in and regulate different functions such as adhesion, migration, degranulation, phagocytosis, apoptosis and ROS production by activation of the NADPH oxidase enzyme complex [20,21].

2.2. Protein Kinase C Isoform Locations and Translocation

Protein kinase C isoforms in unstimulated neutrophils are located in the cytosol and translocate to the plasma membrane upon activation by PMA [22], except PKC β isoforms which are partially located in the membrane [21]. Unlike nPKC δ, which has already fully translocated to the plasma membrane within two minutes [21], cPKC α [21,23], βI and βII are still initiating translocation upon activation [15,21,23]. Although the atypical PKC (aPKC) ζ has been cited for being independent of PMA, it exhibited a rapid translocation, within five minutes, to membrane fragments in PMA-activated cells [15]. PKC ζ was first detected in 1994 by Dang et al. and its activation by PMA is not direct, but happens downstream via PIP3 [9] (Figure 2). It is also implicated to have a role in neutrophil effector functions, such as adhesion [15,24]. However, there is a disagreement about PKC ζ acting as an effector of ROS production. Laudanna et al. (1998) and Dang et al. (2001), using an inhibitor based on myristoylated peptides from the pseudosubstrate region of PKCζ did not show effect on ROS production [13,25], while Kato et al. (2005) showed decrease of ROS consequent to the inhibition of PKC ζ [26].
The translocation of PKC δ leads to the phosphorylation of myristoylated alanine rich protein kinase C substrate (MARCKS) [21,27]. MARCKS translocates from the cytosol to the plasma membrane, which alters the mobility or affinity of the CD18 (β2 integrin), polymerizing actin, and therefore leading to cell degranulation, adhesion and migration [28].

2.3. PMA Triggers Degranulation

Cytoplasm of neutrophils contains granules and vesicles that have hundreds of proteins. The primary or azurophilic granules have about 69% of the total neutrophil myeloperoxidase, besides being rich in α-defensin, myeloblastin, elastase, azurocidin and cathepsin G [29]. The secondary or specific granules have about 54% of the gelatin-associated lipocalin (NGAL), 52% of the lactoferrin, and contain other proteins such as pentraxin 3 and haptoglobulins [29,30]. Both release antimicrobial proteins, while the tertiary or gelatinase granules are rich in gelatinase, a matrix metalloproteinase that degrades the extracellular matrix. In addition, after fusion with the cell membrane during degranulation, granular membrane proteins expose receptors for adhesion and migration [29,31,32]. The secretory vesicles, which contain proteins such as the Toll-like receptor (TLR) 2, 4 and 8, as well as β and α-actin, add proteins to the cell membrane to recognize PAMPs and for migration [29,33].
Initially, neutrophils activated by PMA release the content of secretory vesicles, tertiary, secondary and primary granules already in the first minutes, lasting for hours [34,35]. PKC δ gene deficient neutrophils showed markedly reduced lactoferrin degranulation following stimulation with PMA [36] and neutrophils preincubated with a MARCKS synthetic peptide showed increased degranulation [37]. The facts that adhesion receptors get exposed after degranulation and that MARCKS is involved in both processes reinforces the correlation between degranulation and adhesion, even though degranulation happens independent of an adhesion surface [38]. Downstream to the same pathway, Karlsson et al., 2000 showed that PI3K inhibition does not change the degranulation of any type of granules or secretory vesicles [34]. On the other hand, Chen et al., 2010, showed that Akt2 gene deletion decreased the primary granule degranulation, suggesting a role for Akt in degranulation, independent of PI3K [39]. Furthermore, PMA was ineffective at eliciting primary granule content release in Rac2 knockout (KO) mice, however it still induced the release of secondary granules [40]. Interestingly, the suppression of both pathways, Akt2 and Rac2, affect the release of primary granules and extracellular ROS formation [39,40], while PI3K inhibition does not change the degranulation of any type of granules; it does however, inhibit intracellular ROS [34]. After NADPH oxidase assembly, the specific granules have the necessary components for ROS production, with the exception of myeloperoxidase to convert hydrogen peroxide (H2O2) to hypochlorous acid (HOCl) [41,42,43]. As degranulation of primary granules, as well as extracellular ROS production are decreased in Akt2 or Rac2 deletion [39,40], this suggests that the NADPH oxidase at the cell membrane and at secondary granules might be impaired, and that the primary granules might contribute to the amount of extracellular ROS. In 2016, Björnsdottir et al. proposed a fusion of specific granules and azurophil granules, producing intracellular ROS in the absence of phagosome formation [44]; however, it seems that specific granules do not contribute as much to intracellular ROS as opposed to extracellular ROS, perhaps due to the incorporation of their NADPH oxidase to the cell membrane after degranulation [45,46]. Based on these publications, we suggest that there are two different pathways to intracellular ROS influenced by PI3K, while extracellular ROS is Akt- and Rac2-dependent with the possible participation of the primary granules.

2.4. Crosstalk between Adhesion/Migration and NADPH Oxidase

PMA activates an increase of the main adhesion molecules of neutrophils: LFA-1 (CD11a/CD18) [47], MAC-1 (CD11b/CD18) [48,49], ICAM-1 (CD54), complement receptor 1 (CR1,CD35) at the plasma membrane [50], and promotes L-selectin shedding, increasing soluble L-selectin [49,51]. Moreover, PMA triggers an increase of E-selectin on endothelial cells [52], while decreasing P-selectin glycoprotein ligand-1 (PSGL-1) on the surface of neutrophils, reducing their adhesion to P-selectin [53] on the surface of endothelial cells. These mechanisms of regulation are involved in the control of neutrophils’ rolling velocity on the endothelium, favoring their diapedesis. Therefore, PMA promotes actin polymerization [54], supporting the rearrangement of the cytoskeleton, which is important for the assembly of the NADPH oxidase complex [55]. Accordingly, ROS production also allows neutrophil adhesion. Shen et al. (1999) showed that a decrease in ROS reduces approximately 60% of neutrophil adhesion [56]. Another antiadhesive molecule, CD43 (leukosialin), binds to E-selectin, limiting neutrophil interaction with the endothelium [57] and is decreased on the surface of PMA-activated neutrophils [58]. The serine protease inhibitor (3,4-DCI) prevents the increase in ROS production, as well as it inhibits the decrease in CD43 activated by PMA [59]. The decrease in CD43 occurs because it undergoes juxtamembranous cleavage by a serine protease, cathepsin G [60], and is relevant to cell adhesion, which is limited by CD43.
The rearrangement of the cytoskeleton enables the activation of structural and regulatory proteins, such as actin, moesin and cofilin, that interact with the cytosolic components p47phox, p40phox and p67phox of the NADPH oxidase enzyme complex. This interaction is thought to enable the translocation of these cytosolic components to the plasma membrane, therefore allowing the binding of these components to the membrane bound subunits, gp91phox and gp22phox, which together form the flavocytochrome b558 subunit of this complex [61,62]. Interestingly, cytoskeletal components have been indicated in the regulation and organization of the NADPH oxidase complex, as well as the complex itself having a regulatory role in cytoskeletal conformation [63]. Actin polymerization contributes to the activation of the NADPH oxidase complex [64], allowing translocation of the components [62,65]. Nevertheless, NADPH oxidase-generated ROS reduces actin polymerization through actin glutathionylation, impairing neutrophil chemotaxis [17]. On the other hand, total inhibition of ROS leads to a more frequent formation of multiple pseudopodia and a loss of direction during migration [66]. This suggests that ROS contributes to the adhesion and migration of neutrophils by orchestrating the shedding of adhesion molecules and actin polymerization, although the ideal concentration of ROS remains unclear.
Actin polymerization is the consequence of a remarkable association between PKC isoforms [67] and the cytoskeleton. Although Downey et al. (1992) showed that actin assembly was not inhibited by conventional PKC inhibitors [68], Niggli et al. (1999) showed that the inhibitor of actin polymerization (latrunculin A) reduced cytoskeletal PKC βII by 43%, suggesting a physical binding between PKC βII and F-actin [14]. In addition, the fact that PKC ζ inhibited PMA-activated adhesion and not ROS production suggests its possible effector function to trigger pathways leading to adhesion and not ROS [13,15,26], reinforcing the idea that the regulation of cytoskeleton remodeling is established by these isoforms of PKC.

2.5. NADPH Oxidase Subunit Phosphorylation and Translocation

Apart from the suggested role of the cytoskeleton in the translocation of NADPH oxidase cytosolic subunits, other events are necessary to trigger the full activation of this enzyme complex, such as phosphorylation of the oxidase subunits, as well as the activation of the Rho GTPase Rac2, a non-specific participant in this process [19]. Phosphorylation of the NADPH oxidase components by PKC isoforms is essential for ROS production in the PMA-activated neutrophil [10,69,70,71,72,73]. PKC isoforms α, β and δ have been shown to either activate or inhibit p47phox depending on the phosphorylation site [74,75], and hyperphosphorylation of p47phox leads to the deactivation of the NADPH oxidase complex [76]. Previously, in 2002, Fontayne, et al. described how PKC isoforms α, βII, δ and ζ played roles in the translocation of p47phox and its subsequent binding to the gp22phox membrane subunit in a cell free system [10]. Additionally, they demonstrated that each PKC isoform could individually phosphorylate p47phox, therefore inducing activation of the NADPH oxidase complex. It was noted, however, that the ζ isoform phosphorylated fewer and different serine sites in comparison with the other isoforms (α, βII and δ), which resulted in a higher level of NADPH oxidase activity. They further suggested that the PKC isoforms α, βII and δ could be associated with attenuated oxidase activity, therefore indicating specific roles for these PKC isoforms [10]. In summary, different PKC isoforms are related to specific phosphorylation patterns of the p47phox, leading to the modulation of ROS production in the first minutes after PMA activation [11,77,78].
Translocation patterns of PKC isoforms α and βII coincide with the translocation pattern of p47phox, adding evidence to their participation in NADPH oxidase assembly. This was not the case with PKC δ. Even though it has a different translocation profile to p47phox, PKC δ translocates faster than PKC α and βII [23]. However, specific inhibition of PKC δ decreases 96% of ROS production, and the immunodepletion of PKC δ decreases approximately 50% of intracellular ROS [12]. Recently, PKC δ tyrosine 155 phosphorylation was demonstrated not to be required for ROS production in PMA-activated neutrophils [79]; however, PKC δ also presents serine/threonine phosphorylation sites [9]. Moreover, PKC δ phosphorylates NAD kinase (NADK) which is required for the synthesis of NADPH, the substrate of the oxidase complex [80]. Another study in 2000, by Dekker et al. demonstrated that PKC β is responsible for at least 50% of superoxide production. Interestingly, it was demonstrated that an increased concentration (1 µM) of the inhibitor 379196, a specific inhibitor of PKC β, decreased ROS production [11]. Besides that, PKC α inhibition abrogated ROS [13], which seemingly confirms the role of these isoforms in NADPH oxidase activation and assembly; however, it is not clear which PKC isoform is most important to ROS production. Moreover, as the sum of percentages assigned to each isoform clearly surpass 100%, it is likely that there is a crosstalk among the PKC isoform mechanisms.
Apart from the phosphorylation of the NADPH oxidase complex, PKC isoforms have been shown to phosphorylate the Raf/MEK/ERK/p38/Rac-2 pathways [81,82,83,84] that also phosphorylate NADPH oxidase complex components (Figure 2). The MAPK pathway is composed of three kinds of kinases, MAPK kinase kinase (e.g., Raf), MAPK kinase (e.g., MEK), and MAPK (e.g., ERK1/2, p38, JNK), that are sequentially activated by phosphorylation [85]. Three MAPK families have been defined in mammalian cells, classical MAPK or extracellular signal-regulated kinases 1/2 (ERK1/2), C-Jun N-terminal kinase/stress activated protein kinase (JNK/SAPK) and p38 kinase [86]. The reduction in abundance of ERK/MAPK leads to NADPH oxidase inhibition [87,88,89] because MEK 1/2 phosphorylates the p67phox subunit of NADPH oxidase complex [90], and p38 MAPK phosphorylates the p47phox component [91]. Moreover, ERK2 [92] and PKC δ phosphorylate MARCKS [21], which translocates to the plasma membrane after phosphorylation [93] and activates PI3K pathway through the release of PIP2 molecules that recruit multiple active PI3K to the membrane surface [94]. Active PI3K at the membrane surface, exposed to PIP2, will phosphorylate it to PIP3. Furthermore, PI3K inhibition did not affect the MAPK pathway activated by PMA [95]. Therefore, we propose that MAPK pathway occurs concomitantly to the PI3K pathway.

3. PI3K Participates in ROS Production

Phosphoinositide-3-kinases (PI3K) are a family of lipid kinases that are grouped into classes I, II and III. Class I PI3K is the most widely studied and is divided into IA (PI3Kα, PI3Kβ, PI3Kδ) and IB (PI3Kγ) [96,97]. PKC β isozymes phosphorylate the PI3Kγ isotype [98], which phosphorylates PtdIns (4,5) P2 (or PIP2) to PtdIns (3,4,5) P3 (or PIP3). This leads to the translocation of various proteins through the pleckstrin homology (PH) domain [99]. Proteins such as serine/threonine kinases Akt/PKB, which are involved in the rearrangement of the cytoskeleton in stimulated cells, contain the PH domain that binds to PIP3. This binding leads to Akt activation [100]. The phosphorylation of ERK by PI3Kγ also leads to ROS generation, apoptosis inhibition, and chromatin decondensation [101]. Thus, isoform PKC β triggers PI3Kγ pathway, which also activates ERK, contributing to functions like ROS production (Figure 2).
PI3K participates in the intracellular ROS production by the NADPH oxidase complex [34,42,102,103]. Class III PI3K phosphorylates PtdIns to PtdIns3P (PI3P), described to regulate lysosomal trafficking and to control autophagy. Interestingly, PI3P contributes to the first phase of intracellular ROS production by binding to the subunit p40phox of the NADPH oxidase [104,105]. Class II PI3Ks also phosphorylate PtdIns to PtdIns3P, however their function has been little explored [96]. Class I and II PI3K have an intrinsic relation to ROS production, which includes the later activation of proteins such as Akt, although the Akt2 isoform is PI3K-independent [39]. Akt also phosphorylates mTORC, inhibits anti-apoptotic proteins and activates the NF-κB complex [100]. However, more studies are necessary to elucidate the functions of Class II PI3K.
Furthermore, class I PI3Ks activity is modulated by families of phosphoinositide phosphatases that include the tumor suppressor protein phosphatase and tensin homolog deleted from chromosome 10 (PTEN), SH2-containing inositol polyphosphate 5-phosphatase 1 (SHIP) and inositol-3,4-bisphosphate 4-phosphatase (INPP4A/4B) [99]. PTEN is a phospholipid phosphatase that dephosphorylates PIP3 and converts it to PIP2, antagonizing the PI3K activity. Thus, PTEN modulates various cell functions through dephosphorylation including the regulation of cell growth, adhesion, chemotaxis, and migration [106]. Besides that, PTEN participates in ROS production [107], because PIP2 is cleaved by phospholipase C (PLC) into DAG and inositol-triphosphate (IP3) that participate in PKC activation and the release of intracellular calcium, respectively [19] (Video S1). Interestingly, Teimourian and Moghanloo (2015) showed that PTEN also contributes to NETs production [108]. They showed that PTEN down-regulation in neutrophils differentiated from HL-60 cells, activated by PMA, reduced 26% of NETs production. In addition, it was observed that PTEN depletion leads to an increase in PIP3 levels and improves ROS production [99]. Such depletion impairs PIP2 formation, therefore impacting the production of DAG and IP3 by phospholipase C (PLC) [109]. As a consequence, the calcium release induced by IP3 and PKC activation induced by DAG would also be downregulated, likely decreasing NETs formation. On the other hand, the reason for the increased PIP3 levels not compensating for NET production remains an open question.

The Relation between Calcium and PMA Activation

Calcium is an intracellular signaling ion important for neutrophil adhesion and ROS production [110]. However, there is disagreement in the literature about PMA activated calcium influx and efflux levels and its contribution to ROS and NETs formation. Shen et al. (1999), using fluorescence spectrophotometry, showed that PMA increases intracellular calcium by less than 20% in the first seconds after stimulation [111]. Although PMA activates endoplasmic reticulum stress [112], some authors do not consider the influx and efflux of calcium by PMA to be significant, rather only a minimal oscillation in calcium levels early after stimulation [80,112,113,114]. The occurrence of this oscillation in calcium levels together with ROS production suggests its involvement in NETosis [83]. It has been shown that the depletion of calcium stores by thapsigargin affects ROS production [115,116], while reduction of ROS by chelating extracellular calcium with EGTA is not observed [115,117].
Furthermore, calcium ion is associated with phospholipases A2 (PLA2) and D (PLD) activity, possibly involved in ROS production [118,119]. PLA2 is phosphorylated by PKC and ERK1. When activated, it cleaves membrane phospholipids to produce arachidonic acid, which is released by neutrophils activated by PMA [120,121]. Meanwhile, phospholipase D, which hydrolyzes phospholipids to phosphatidic acid (PA) and choline [122], is phosphorylated by nPKC, leading to its activation in response to PMA [123]. PLD is regulated by ADP-ribosylation factor (ARF-1) [124], which also regulates actin assembly allowing vesicle fusion [125].
The decrease of intracellular calcium impairs PLA2 and PLD [122,126], while extracellular calcium has no effect on PLA2 [126]. However, there is a study suggesting that PLA2 is not involved in NADPH oxidase activity, but influences ROS by affecting granule fusion [119].
On the other hand, Gupta et al. (2014) and Suh et al. (2001) showed that chelation of extracellular calcium reduces ROS by approximately 60% [127,128]. Moreover, Gupta et al. (2014) showed that for efficient NETosis to occur there must be a mobilization of intra-and extracellular calcium pools [127]. Accordingly, Parker et al. (2012) and Kenny et al. (2017) showed that the intracellular calcium chelator, BAPTA-AM, reduced ROS-dependent NETs in PMA stimulation [2,129]. Conflictingly, Douda et al. (2015) suggest that the absence of extracellular calcium had no effect on NETs release based on cells cultured without calcium [130]. Although calcium could contribute to ROS production by PLA2 and PLD, the literature suggests that intracellular calcium is not decisive to ROS production activated by PMA, because it does not induce a significant intracellular calcium mobilization. Meanwhile, the participation of extracellular calcium and the contribution mechanisms of both stores to NETs formation remain unclear, and it is still necessary to identify whether PMA activation is expressively affected by the extracellular calcium concentration.

4. Neutrophil Extracellular Trap (NET) Formation

NETs are a mechanism used by neutrophils, in which they release DNA and proteins to promote the extracellular capture and death of microorganisms. However, they are considered double-edged swords for being associated with the cause and severity of some diseases, such as thrombosis, autoimmune diseases, and cancers. Recently, it was also associated with the severe acute respiratory syndrome observed in COVID-19 [131,132]. A better understanding of the pathways related to NETs is needed so that more effective therapeutic approaches to these diseases can be developed [133].
NETs is a process that occurs independently of phagocytosis [1,134,135]. Currently, NETs are classified as suicidal or vital. Suicidal NETs are ROS-dependent and vital NETs can be triggered by ROS-dependent or ROS independent mechanisms [133,136]. PMA is able to induce ROS-dependent suicidal NETs [1,84]. Suicidal NETosis is a death process that requires 3–4 h for the complete formation of NETs, although some cells already show formation within 30–60 min after activation [84,127,137,138]. In this mechanism, NADPH oxidase is activated via PKC/Raf/MERK/ERK, that also leads to the activation of the enzyme peptidylarginine deaminase 4 (PAD4), which modifies histones, converting arginine (Arg) into citrulline (Cit). Such modification promotes the loss of positive charges, preventing their interaction with DNA, leading to chromatin decondensation [139,140,141].
Hydrogen peroxide is also responsible for activating and dissociating a complex of proteins present in azurophilic granules (inhibition of such proteases by SerpinB1 limits NETs formation) [142] Then, H2O2, together with MPO, activates NE, which degrades F-actin allowing the translocation of NE to the nucleus [143], promoting a partial degradation of histones [144]. In addition, F-actin glutathionylation triggered by ROS also leads to actin degradation [145], promoting structural changes that, together with the mechanisms described above, culminate in chromatin swelling and decondensation [2,83,146] (Figure 2). This allows NETs to be released via nuclear budding or vesicle formation in vital NETs [81] or via rupture of the nuclear membrane and finally of the plasma membrane in suicidal NETs. Recently, the nuclear component lamin B was described to be disassembled after phosphorylation by PKC α resulting in nuclear rupture [147].
Even though PI3K inhibition does not reduce NETs formation [148], this pathway participates in NETs formation by means of its contribution to ROS production. There is still disagreement about whether the Akt pathway contributes to NETs formation. Starting from evidence that PMA induces Akt phosphorylation [84], Douda et al. (2014) suggested that ROS-dependent NETosis in PMA (25 nM) stimulated neutrophils requires Akt activation, using both Akt and NOX2 inhibitor assays [149]. On the other hand, DeSouza-Vieira et al. (2016) rejected that hypothesis, as they found that the Akt inhibitor (124005) had no effect on NETs release when cells were activated with 100 nM PMA [101]. The different methods of these studies do not allow us to do an appropriate comparison, thus, more studies are needed to elucidate this.

4.1. Protein Arginine Deiminase 4 (PAD4)

The activation of PAD4 induced by PMA leads to chromatin decondensation [84,146,150]. The chromatin structure is organized with two of each histones H3, H4, H2B and H2At in the DNA to form a nucleosome core particle, and a linker formed by DNA and H1 [151]. Peptidylarginine deiminases (PADs) are a family of enzymes that are located in the nucleus and modify histones by converting arginine (Arg) to citrulline (Cit) [140,141]. This decreases the positive charges along the histone termini, reducing electrostatic interactions with DNA, and thus allowing chromatin to assume an extended conformation [141].
There are five highly conserved PADs in mammalians, namely PAD1, 2, 3, 4 and 6 [151]. Neutrophils express high levels of PAD4 [152] which mediates chromatin decondensation and NETs formation [153], ejecting 15–25-nm chromatin fibers [1]. Li et al. (2010) showed that mouse neutrophils activated by PMA have 48.5% of histone citrullination [154], while Holmes et al. (2019) identified approximately 20% in humans [155]. Other evidence in the literature shows further divergence whether NETs induced by PMA are citrullination-dependent or independent. Tatsiy e McDonald (2018) showed that GSK484, a specific inhibitor of PAD4, abrogated the citrullination of histone H3, impairing NETs formation [84] and Hawez et al. (2019) detected the PMA-induced increase in PAD4 and in citrullination of histone H3 by western blotting [156]. Oppositely, Neeli and Radic (2013) and Konig and Andrade (2016) did not detect citrullination of histone H3 by western blotting after 2 h of PMA stimulus [140,150]. Neeli and Radic (2013) demonstrated the inhibition of PAD4 by PKC α and its activation by PKC ζ isoforms, but the dominant isoform after PMA stimulation is the alpha, coherent with the lack of citrullination in the NETs [150]. In that regard, Kenny et al. (2017) did not observe impairment on NETs by inhibition of PAD2 (Cl-amidine) and PAD4 (BB-Cl-amidine and TDFA) [129]. In 2018, de Bont et al. showed that citrullination activated by PMA was detected in NETs after inhibition of serine proteases [83]. They suggested that serine proteases promote the cleavage of citrullinated histones, possibly explaining the previous disagreement. Based on these results, and on the results discussed below about PAD4 citrullination of NF-κB p65 (see the topic NF-κB), we suggest that PAD4 may influence NET formation by at least two pathways, based on histone H3 citrullination or IL-1β release and gasdermin cleavage triggered by activity of serine proteases. This second pathway possibly explains the presence of NETs in the absence of citrullinated histones, as the same proteases might cleave the citrullinated histones (see Gasdermin D topic) (Video S1).

4.2. Autophagy

Autophagy is a catabolic process that plays important roles in preventing damage to cells, promoting cell survival during nutrient deprivation and oxidative stress [157,158,159]. During phagocytosis, ROS produced by NADPH oxidase or the activation of AMPK generated by ATP depletion leads to the activation of autophagy [160,161].
The induction of autophagy occurs via autophagy related (ATG) proteins. There is a cooperation between two protein complexes, one is composed by a serine/threonine protein kinase (ULK1) interacting with ATG101 and ATG13. The other complex contains PI3K and ATG14 [162]. PI3K catalyzes the production of PIP3, which binds LC3B [163]. LC3B is then translocated to the autophagosome membrane [164] in the maturation process. Further protein interactions, such as with Rubicon and Rab7 stimulate the fusion of the autophagosome with endosomes/lysosomes [161,165].
The canonical autophagy pathway is modulated by the activation of mTORC1 to phosphorylate ULK1 [160], while the induction of autophagy by phagocytosis is dependent on lysosome-associated membrane protein (LAP) [161].
Autophagy participates in some functions of the neutrophil, such as degranulation and ROS production [164]. Some studies showed that neutrophils activated by PMA release NETs with the involvement of autophagy as well as ROS [166], as the inhibition of either autophagy or ROS production prevented chromatin decondensation [167]. These studies used PI3K inhibition that leads to impairment of the NETs, but it also inhibits ROS production. Thus, the role of autophagy in NETs was not quite clear. However, Park et al. (2017) observed an increase of LC3I and LC3II induced by PMA, and the use of autophagy inhibitors eliminated what the authors named an autophagy priming for NETs [168]. Besides that, recently Germic et al. (2017) showed that Atg5 knockdown does not affect NETs production in neutrophils activated by GM-CSF, C5a, LPS and PMA, elucidating that autophagy is not mandatory for NETs production [169].

4.3. NF-κB

Nuclear factor-kappa B (NF-κB) is a family of transcription factors that consists of five proteins that share the amino-terminal Rel homology domain (RHD) [170]. The five proteins p65 (RelA), RelB, c-Rel, p105/p50 (NF-κB1), and p100/52 (NF-κB2) can associate among themselves, forming homo- or heterodimers in a more transcriptionally active structure. The p50/65 heterodimer is the most abundant form, being found in almost all cell types [171].
NF-κB can be activated in both the canonical and non-canonical pathways. In the canonical pathway, microbial products and proinflammatory cytokines trigger the activation of p65-containing complexes and it requires the IKKβ and IKKγ kinases. On the other hand, in the non-canonical pathway, which probably works together with the canonical, different cytokines trigger the activation of p52-containing complexes. Both pathways lead to the translocation of different NF-κB heterodimers to the nucleus [172,173].
In neutrophils, NF-κB promotes cell adhesion, inflammation and apoptosis. Contrary to the LPS stimulation, which inhibits NF-κB in neutrophils, leading to apoptosis [174], PMA stimulus activates the canonical pathway, translocating NF-κB p50/p65 to the nucleus [175]. p65 leads to NETs formation, while its inhibition decreases NETs [176]. Interestingly, during the inhibition of NF-κB, neutrophil survival could be sustained by p38 MAPK activation after autophosphorylation [177]. The p65 could be phosphorylated in serine (Ser), threonine (Thr), or tyrosine (Tyr) sites by kinases such as PKAc, MSK-1/MSK-2 and Pin1 [172]. In addition, p38 MAPK can regulate the NF-κB pathway [178]. Furthermore, PAD4 can citrullinate NF-κB p65 to promote its nuclear translocation and mediate IL-1β production (Figure 2) [179], which also induces NETs formation and enhances ROS production [180]. It is known that IL-1β can be released by caspase-dependent or independent pathways [181]. In the caspase-dependent pathway, there is activation via the canonical inflammasome pathway, while the non-canonical pathway, which is activated by PMA [182], is independent of the inflammasome. Both pathways allow for the cleavage of gasdermin D [183,184].

4.4. Gasdermin D

Recently, some studies showed the importance of yet another mechanism for NETs formation, the gasdermin D protein. Gasdermin D (GSDMD) has the N-terminal and the C-terminal domains separated by a binder portion. It is known to promote a type of cell death, called pyroptosis [183,184]. Gasdermin cleavage leaves the portion GSDMD-C in the cytosol, while the GSDMD-N product is oligomerized in the cell membrane forming pores [185]. In neutrophils activated by PMA, gasdermin was cleaved by ROS-activated serine proteases such as NE [182], allowing the release of IL-1β [181]. As mentioned above, IL-1β triggers NETs formation [180,186] and is involved in the process of recruiting other neutrophils to the infection site in response to large pathogens that cannot be phagocytosed [187], also entrapped in NETs [135]. PMA-activated neutrophils tend to be short-lived cells [188], due to apoptosis, pyroptosis [189] and NETosis, which is the main death mechanism [1,188,189,190,191], even though they utilize autophagy mechanisms [167]. However, Desai et al. (2016) showed that inhibition of necroptosis (RIPK1 and MLKL) and knockout of RIPK3 reduces NETs formation [192]. On the contrary, Amini et al. (2016) using knockout of RIPK3 [193], and Remijsen et al. (2011) who inhibited RIPK1 by necrostatin-1, did not observe effects on NETs formation [167]. Thus, gasdermin D-induced pore formation can be considered a common point among ROS production, NET formation, IL-1β release and NF-κB p65 pathways, although the mechanism for the latter is not quite clear in the literature yet. It could be activated by p38 MAPK and suffer modification by PAD4, to then participate in NETs formation by gasdermin D activation. However, it remains to be elucidated if this is actually a cascade, therefore connecting the upstream and downstream proteins, or if the pathways of gasdermin D cleavage and IL-1β release by NF-κB are distinct and concurrently performed.

5. Conclusions

Even though PMA is not an endogenous or microbial activator, it allows for the understanding of different pathways of neutrophil activation. PMA crosses the cell membrane independently of receptors and activates PKC α, β and δ. PKC δ phosphorylates MARCKS that leads to the activation of the PI3K pathway, while PKC β phosphorylates PI3K and MAPK pathways. Moreover, PI3Kγ activates ERK, and these pathways work together, all leading to the phosphorylation and translocation of components of the NADPH oxidase complex. The regulatory crosstalk between ROS production and cytoskeleton remodeling still requires further investigation, as these processes might be hubs for integrative responses of the neutrophil.
Much more detail is provided in the literature regarding the PKC, MAPK and PI3K pathways, raising possible interpretations, such as a regulatory mechanism balancing ROS and NETs production. However, some aspects are still open to elucidation, such as the role of calcium release and the Akt contribution to NETs production.
PKC ζ contributes to the citrullination of histones promoted by PAD4. However, we found disagreement in the literature regarding whether NETs are dependent or not on histone citrullination. Interestingly, PAD4 participates in other functions, such as NF-κB activation and the release of IL-1β. Moreover, IL-1β triggers the cleavage of gasdermin D by serine proteases such as elastase, leading to both pore formation and NETs formation. Although it is known that IL-1β recruits other neutrophils and is involved in the process of NETs formation, the pathways involved are not fully elucidated.
In all, even the effects of one of the most studied neutrophil activators still has a number of controversial or unclear aspects that we highlighted in this first present review.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cells11182889/s1, Video S1: Details of major signaling pathways in PMA-activated neutrophils.

Author Contributions

Conceptualization, W.F. and H.L.D.; methodology, H.L.D. and W.F.; formal analysis, H.L.D. and W.F.; investigation, H.L.D., W.A.A.S. and W.F.; resources, W.F. and M.S.C.; data curation, H.L.D., W.A.A.S., M.S.C. and W.F.; writing—original draft preparation, H.L.D. and W.F.; writing—review and editing, H.L.D., W.A.A.S., M.S.C. and W.F.; supervision, W.F. and M.S.C.; project administration, W.F.; funding acquisition, W.F. and M.S.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by CNPq (Conselho Nacional de Desenvolvimento Científico e Tecnológico), CAPES (Coordenação de Aperfeiçoamento de Pessoal de Nível Superior), FAP-DF (Fundação de Apoio à Pesquisa do Distrito Federal), FINEP (Financiadora de Estudos e Projetos), FINATEC (Fundação de Empreendimentos Científicos e Tecnológicos) and UnB (University of Brasilia).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Brinkmann, V.; Reichard, U.; Goosmann, C.; Fauler, B.; Uhlemann, Y.; Weiss, D.S.; Weinrauch, Y.; Zychlinsky, A. Neutrophil Extracellular Traps Kill Bacteria. Science 2004, 303, 1532–1535. [Google Scholar] [CrossRef] [PubMed]
  2. Parker, H.; Dragunow, M.; Hampton, M.B.; Kettle, A.J.; Winterbourn, C.C. Requirements for NADPH Oxidase and Myeloperoxidase in Neutrophil Extracellular Trap Formation Differ Depending on the Stimulus. J. Leukoc. Biol. 2012, 92, 841–849. [Google Scholar] [CrossRef]
  3. Kolaczkowska, E.; Kubes, P. Neutrophil Recruitment and Function in Health and Inflammation. Nat. Rev. Immunol. 2013, 13, 159–175. [Google Scholar] [CrossRef] [PubMed]
  4. Heit, B.; Tavener, S.; Raharjo, E.; Kubes, P. An Intracellular Signaling Hierarchy Determines Direction of Migration in Opposing Chemotactic Gradients. J. Cell Biol. 2002, 159, 91–102. [Google Scholar] [CrossRef]
  5. Aquino, E.N.; Neves, A.C.D.; Santos, K.C.; Uribe, C.E.; Souza, P.E.N.; Correa, J.R.; Castro, M.S.; Fontes, W. Proteomic Analysis of Neutrophil Priming by PAF. Protein Pept. Lett. 2016, 23, 142–151. [Google Scholar] [CrossRef]
  6. Gera, N.; Swanson, K.D.; Jin, T. β-Arrestin 1-Dependent Regulation of Rap2 Is Required for FMLP-Stimulated Chemotaxis in Neutrophil-like HL-60 Cells. J. Leukoc. Biol. 2017, 101, 239–251. [Google Scholar] [CrossRef]
  7. Goel, G.; Makkar, H.P.S.; Francis, G.; Becker, K. Phorbol Esters: Structure, Biological Activity, and Toxicity in Animals. Int. J. Toxicol. 2007, 26, 279–288. [Google Scholar] [CrossRef]
  8. Pagani, A.; Gaeta, S.; Savchenko, A.I.; Williams, C.M.; Appendino, G. An Improved Preparation of Phorbol from Croton Oil. Beilstein J. Org. Chem. 2017, 13, 1361–1367. [Google Scholar] [CrossRef]
  9. Steinberg, S.F. Structural Basis of Protein Kinase C Isoform Function. Physiol. Rev. 2008, 88, 1341–1378. [Google Scholar] [CrossRef]
  10. Fontayne, A.; Dang, P.M.; El Benna, J. Phosphorylation of P47 Phox Sites by PKC R, ΙΙ, δ, and: Effect on Binding to P22 Phox and on NADPH Oxidase Activation. Society 2002, 24, 7743–7750. [Google Scholar]
  11. Dekker, L.V.; Leitges, M.; Altshuler, G.; Mistry, N.; Mcdermott, A.; Roes, J.; Segal, A.W. Protein Kinase C-β Contributes to NADPH Oxidase Activation in Neutrophils. Biochem. J. 2000, 347, 285–289. [Google Scholar] [CrossRef]
  12. Brown, G.E.; Stewart, M.Q.; Liu, H.; Ha, V.L.; Yaffe, M.B. A Novel Assay System Implicates PtdIns(3,4)P2, PtdIns(3)P, and PKCδ in Intracellular Production of Reactive Oxygen Species by the NADPH Oxidase. Mol. Cell 2003, 11, 35–47. [Google Scholar] [CrossRef]
  13. Laudanna, C.; Mochly-Rosen, D.; Liron, T.; Constantin, G.; Butcher, E.C. Evidence of ζ Protein Kinase C Involvement in Polymorphonuclear Neutrophil Integrin-Dependent Adhesion and Chemotaxis. J. Biol. Chem. 1998, 273, 30306–30315. [Google Scholar] [CrossRef] [PubMed]
  14. Niggli, V.; Djafarzadeh, S.; Keller, H. Stimulus-Induced Selective Association of Actin-Associated Proteins (α-Actinin) and Protein Kinase C Isoforms with the Cytoskeleton of Human Neutrophils. Exp. Cell Res. 1999, 250, 558–568. [Google Scholar] [CrossRef]
  15. Dang, P.M.C.; Rais, S.; Hakim, J.; Périanin, A. Redistribution of Protein Kinase C Isoforms in Human Neutrophils Stimulated by Formyl Peptides and Phorbol Myristate Acetate. Biochem. Biophys. Res. Commun. 1995, 212, 664–672. [Google Scholar] [CrossRef]
  16. Hattori, H.; Subramanian, K.K.; Sakai, J.; Jia, Y.; Li, Y.; Porter, T.F.; Loison, F.; Sarraj, B.; Kasorn, A.; Jo, H.; et al. Small-Molecule Screen Identifies Reactive Oxygen Species as Key Regulators of Neutrophil Chemotaxis. Proc. Natl. Acad. Sci. USA 2010, 107, 3546. [Google Scholar] [CrossRef]
  17. Sakai, J.; Li, J.; Subramanian, K.K.; Mondal, S.; Bajrami, B.; Hattori, H.; Jia, Y.; Dickinson, B.C.; Zhong, J.; Ye, K.; et al. Reactive Oxygen Species-Induced Actin Glutathionylation Controls Actin Dynamics in Neutrophils. Immunity 2012, 37, 1037–1049. [Google Scholar] [CrossRef]
  18. Fontes, W.; Sousa, M.V.; Aragão, J.B.; Morhy, L. Determination of the Amino Acid Sequence of the Plant Cytolysin Enterolobin. Arch. Biochem. Biophys. 1997, 347, 201–207. [Google Scholar] [CrossRef]
  19. Belambri, S.A.; Rolas, L.; Raad, H.; Hurtado-Nedelec, M.; Dang, P.M.C.; El-Benna, J. NADPH Oxidase Activation in Neutrophils: Role of the Phosphorylation of Its Subunits. Eur. J. Clin. Investig. 2018, 48, e12951. [Google Scholar] [CrossRef]
  20. Bertram, A.; Ley, K. Protein Kinase C Isoforms in Neutrophil Adhesion and Activation. Arch. Immunol. Ther. Exp. 2011, 59, 79–87. [Google Scholar] [CrossRef]
  21. Sheats, M.K.; Sung, E.J.; Adler, K.B.; Jones, S.L. In Vitro Neutrophil Migration Requires Protein Kinase C-Delta (δ-PKC)-Mediated Myristoylated Alanine-Rich C-Kinase Substrate (MARCKS) Phosphorylation. Inflammation 2015, 38, 1126–1141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Christiansen, N.O.; Borregaard, N. Translocation of Protein Kinase C to Subcellular Fractions of Human Neutrophils. Scand. J. Immunol. 1989, 29, 409–416. [Google Scholar] [CrossRef] [PubMed]
  23. Nixon, J.B.; McPhail, L.C. Protein Kinase C (PKC) Isoforms Translocate to Triton-Insoluble Fractions in Stimulated Human Neutrophils: Correlation of Conventional PKC with Activation of NADPH Oxidase. J. Immunol. 1999, 163, 4574–4582. [Google Scholar]
  24. Dang, P.M.C.; Hakim, J.; Périanin, A. Immunochemical Identification and Translocation of Protein Kinase C Zeta in Human Neutrophils. FEBS Lett. 1994, 349, 338–342. [Google Scholar] [CrossRef]
  25. Dang, P.M.-C.; Fontayne, A.; Hakim, J.; El Benna, J.; Périanin, A. Protein Kinase C ζ Phosphorylates a Subset of Selective Sites of the NADPH Oxidase Component P47 Phox and Participates in Formyl Peptide-Mediated Neutrophil Respiratory Burst. J. Immunol. 2001, 166, 1206–1213. [Google Scholar] [CrossRef]
  26. Kato, M.; Yamaguchi, T.; Tachibana, A.; Kimura, H. Differential Role of an Atypical Protein Kinase C, PKC ζ, in Regulation of Human Eosinophil and Neutrophil Functions. Int. Arch. Allergy Immunol. 2005, 137, 27–34. [Google Scholar] [CrossRef]
  27. Eckert, R.E.; Neuder, L.E.; Park, J.; Adler, K.B.; Jones, S.L. Myristoylated Alanine-Rich C-Kinase Substrate (MARCKS) Protein Regulation of Human Neutrophil Migration. Am. J. Respir. Cell Mol. Biol. 2010, 42, 586–594. [Google Scholar] [CrossRef]
  28. Rorvig, S.; Ostergaard, O.; Heegaard, N.H.H.; Borregaard, N. Proteome Profiling of Human Neutrophil Granule Subsets, Secretory Vesicles, and Cell Membrane: Correlation with Transcriptome Profiling of Neutrophil Precursors. J. Leukoc. Biol. 2013, 94, 711–721. [Google Scholar] [CrossRef]
  29. De Castro, M.S.; Fontes, W.; Morhy, L.; Bloch, C., Jr. Complete Amino Acid Sequences of Two γ-Thionins from Maize (Zea Mays L.) Seeds. Protein Pept. Lett. 1996, 3, 267–274. [Google Scholar] [CrossRef]
  30. Lominadze, G.; Powell, D.W.; Luerman, G.C.; Link, A.J.; Ward, R.A.; McLeish, K.R. Proteomic Analysis of Human Neutrophil Granules. Mol. Cell. Proteom. 2005, 4, 1503–1521. [Google Scholar] [CrossRef]
  31. Rorvig, S.; Honore, C.; Larsson, L.-I.; Ohlsson, S.; Pedersen, C.C.; Jacobsen, L.C.; Cowland, J.B.; Garred, P.; Borregaard, N. Ficolin-1 Is Present in a Highly Mobilizable Subset of Human Neutrophil Granules and Associates with the Cell Surface after Stimulation with FMLP. J. Leukoc. Biol. 2009, 86, 1439–1449. [Google Scholar] [CrossRef] [PubMed]
  32. Borregaard, N.; Kjeldsen, L.; Rygaard, K.; Bastholm, L.; Nielsen, M.H.; Sengelev, H.; Bjerrum, O.W.; Johnsent, A.H. Stimulus-Dependent Secretion of Plasma Proteins from Human Neutrophils. J. Clin. Investig. 1992, 90, 86–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Karlsson, A.; Nixon, J.B.; McPhail, L.C. Phorbol Myristate Acetate Induces Neutrophil NADPH-Oxidase Activity by Two Separate Signal Transduction Pathways: Dependent or Independent of Phosphatidylinositol 3-Kinase. J. Leukoc. Biol. 2000, 67, 396–404. [Google Scholar] [CrossRef] [PubMed]
  34. Degroote, R.L.; Weigand, M.; Hauck, S.M.; Deeg, C.A. IL8 and PMA Trigger the Regulation of Different Biological Processes in Granulocyte Activation. Front. Immunol. 2020, 10, 3064. [Google Scholar] [CrossRef] [PubMed]
  35. Chou, W.-H.; Choi, D.-S.; Zhang, H.; Mu, D.; Mcmahon, T.; Kharazia, V.N.; Lowell, C.A.; Ferriero, D.M.; Messing, R.O. Neutrophil Protein Kinase Cδ as a Mediator of Stroke-Reperfusion Injury. J. Clin. Investig. 2004, 114, 49–56. [Google Scholar] [CrossRef]
  36. Takashi, S.; Park, J.; Fang, S.; Koyama, S.; Parikh, I.; Adler, K.B. A Peptide against the N-Terminus of Myristoylated Alanme-Rich C Kinase Substrate Inhibits Degranulation of Human Leukocytes in Vitro. Am. J. Respir. Cell Mol. Biol. 2006, 34, 647–652. [Google Scholar] [CrossRef]
  37. Mócsai, A.; Ligeti, E.; Lowell, C.A.; Berton, G. Adhesion-Dependent Degranulation of Neutrophils Requires the Src Family Kinases Fgr and Hck. J. Immunol. 1999, 162, 1120–1126. [Google Scholar]
  38. Chen, J.; Tang, H.; Hay, N.; Xu, J.; Ye, R.D. Akt Isoforms Differentially Regulate Neutrophil Functions. Blood 2010, 115, 4237–4246. [Google Scholar] [CrossRef]
  39. Abdel-Latif, D.; Steward, M.; Lacy, P. Neutrophil Primary Granule Release and Maximal Superoxide Generation Depend on Rac2 in a Common Signalling Pathway. Can. J. Physiol. Pharmacol. 2005, 83, 69–75. [Google Scholar] [CrossRef]
  40. Azevedo, E.P.; Rochael, N.C.; Guimarães-Costa, A.B.; De Souza-Vieira, T.S.; Ganilho, J.; Saraiva, E.M.; Palhano, F.L.; Foguel, D. A Metabolic Shift toward Pentose Phosphate Pathway Is Necessary for Amyloid Fibril- and Phorbol 12-Myristate 13-Acetate-Induced Neutrophil Extracellular Trap (NET) Formation. J. Biol. Chem. 2015, 290, 22174–22183. [Google Scholar] [CrossRef]
  41. Ambruso, D.R.; Cusack, N.; Thurman, G. NADPH Oxidase Activity of Neutrophil Specific Granules: Requirements for Cytosolic Components and Evidence of Assembly during Cell Activation. Mol. Genet. Metab. 2004, 81, 313–321. [Google Scholar] [CrossRef] [PubMed]
  42. Dahlgren, C.; Karlsson, A.; Bylund, J. Intracellular Neutrophil Oxidants: From Laboratory Curiosity to Clinical Reality. J. Immunol. 2019, 202, 3127–3134. [Google Scholar] [CrossRef] [PubMed]
  43. Björnsdottir, H.; Welin, A.; Dahlgren, C.; Karlsson, A.; Bylund, J. Quantification of Heterotypic Granule Fusion in Human Neutrophils by Imaging Flow Cytometry. Data Br. 2016, 6, 386–393. [Google Scholar] [CrossRef]
  44. Borregaard, N.; Heiple, J.M.; Simons, E.R.; Clark, R.A. Subcellular Localization of the B-Cytochrome Component of the Human Neutrophil Microbicidal Oxidase: Translocation during Activation. J. Cell Biol. 1983, 97, 52–61. [Google Scholar] [CrossRef]
  45. Clark, R.A.; Volpp, B.D.; Leidal, K.G.; Nauseef, W.M. Two Cytosolic Components of the Human Neutrophil Respiratory Burst Oxidase Translocate to the Plasma Membrane during Cell Activation. J. Clin. Investig. 1990, 85, 714–721. [Google Scholar] [CrossRef]
  46. Shaw, S.K.; Ma, S.; Kim, M.B.; Rao, R.M.; Hartman, C.U.; Froio, R.M.; Yang, L.; Jones, T.; Liu, Y.; Nusrat, A.; et al. Coordinated Redistribution of Leukocyte LFA-1 and Endothelial Cell ICAM-1 Accompany Neutrophil Transmigration. J. Exp. Med. 2004, 200, 1571–1580. [Google Scholar] [CrossRef]
  47. Barnard, J.W.; Biro, M.G.; Lo, S.K.; Ohno, S.; Carozza, M.A.; Moyle, M.; Soule, H.R.; Malik, A.B. Neutrophil Inhibitory Factor Prevents Neutrophil-Dependent Lung Injury. J. Immunol. 1995, 155, 4876–4881. [Google Scholar]
  48. Campos, V.; Melo, R.C.N.; Silva, L.P.; Aquino, E.N.; Castro, M.S.; Fontes, W. Characterization of Neutrophil Adhesion to Different Titanium Surfaces. Bull. Mater. Sci. 2014, 37, 157–166. [Google Scholar] [CrossRef]
  49. Videm, V.; Strand, E. Changes in Neutrophil Surface-Receptor Expression after Stimulation with FMLP, Endotoxin, Interleukin-8 and Activated Complement Compared to Degranulation. Scand. J. Immunol. 2004, 59, 25–33. [Google Scholar] [CrossRef]
  50. Peng, S.; Chen, S.B.; Li, L.D.; Tong, C.F.; Li, N.; Lü, S.Q.; Long, M. Impact of Real-Time Shedding on Binding Kinetics of Membrane-Remaining l-Selectin to PSGL-1. Am. J. Physiol.-Cell Physiol. 2019, 316, C678–C689. [Google Scholar] [CrossRef]
  51. Alam, C.A.S.; Seed, M.P.; Freemantle, C.; Brown, J.; Perretti, M.; Carrier, M.; Divwedi, A.; West, D.C.; Gustafson, S.; Colville-Nash, P.R.; et al. The Inhibition of Neutrophil-Endothelial Cell Adhesion by Hyaluronan Independent of CD44. Inflammopharmacology 2005, 12, 535–550. [Google Scholar] [CrossRef] [PubMed]
  52. Davenpeck, K.L.; Brummet, M.E.; Hudson, S.A.; Mayer, R.J.; Bochner, B.S. Activation of Human Leukocytes Reduces Surface P-Selectin Glycoprotein Ligand-1 (PSGL-1, CD162) and Adhesion to P-Selectin In Vitro. J. Immunol. 2000, 165, 2764–2772. [Google Scholar] [CrossRef]
  53. Lofgren, R.; Ng-Sikorski, J.; Sjolander, A.; Andersson, T. Β2 Integrin Engagement Triggers Actin Polymerization and Phosphatidylinositol Trisphosphate Formation in Non-Adherent Human Neutrophils. J. Cell Biol. 1993, 123, 1597–1605. [Google Scholar] [CrossRef]
  54. Bengtsson, T.; Orselius, K.; Wetterö, J. Role of the Actin Cytoskeleton during Respiratory Burst in Chemoattractant-Stimulated Neutrophils. Cell Biol. Int. 2006, 30, 154–163. [Google Scholar] [CrossRef]
  55. Shen, Y.C.; Chen, C.F.; Wang, S.Y.; Sung, Y.J. Impediment to Calcium Influx and Reactive Oxygen Production Accounts for the Inhibition of Neutrophil Mac-1 up-Regulation and Adhesion by Tetrandrine. Mol. Pharmacol. 1999, 55, 186–193. [Google Scholar] [CrossRef]
  56. Matsumoto, M.; Shigeta, A.; Miyasaka, M.; Hirata, T. CD43 Plays Both Antiadhesive and Proadhesive Roles in Neutrophil Rolling in a Context-Dependent Manner. J. Immunol. 2008, 181, 3628–3635. [Google Scholar] [CrossRef] [PubMed]
  57. Rieu, P.; Porteu, F.; Bessou, G.; Lesavre, P.; Halbwachs-Mecarelli, L. Human Neutrophils Release Their Major Membrane Sialoprotein, Leukosialin (CD43), during Cell Activation. Eur. J. Immunol. 1992, 22, 3021–3026. [Google Scholar] [CrossRef]
  58. Remold-O’Donnell, E.; Parent, D. Downregulation of Neutrophil CD43 by Opsonized Zymosan. Blood 1995, 85, 337–342. [Google Scholar] [CrossRef] [PubMed]
  59. Mambole, A.; Baruch, D.; Nusbaum, P.; Bigot, S.; Suzuki, M.; Lesavre, P.; Fukuda, M.; Halbwachs-Mecarelli, L. The Cleavage of Neutrophil Leukosialin (CD43) by Cathepsin G Releases Its Extracellular Domain and Triggers Its Intramembrane Proteolysis by Presenilin/Gamma-Secretase. J. Biol. Chem. 2008, 283, 23627–23635. [Google Scholar] [CrossRef]
  60. Sheppard, F.R.; Kelher, M.R.; Moore, E.E.; McLaughlin, N.J.D.; Banerjee, A.; Silliman, C.C. Structural Organization of the Neutrophil NADPH Oxidase: Phosphorylation and Translocation during Priming and Activation. J. Leukoc. Biol. 2005, 78, 1025–1042. [Google Scholar] [CrossRef]
  61. Nguyen, G.T.; Green, E.R.; Mecsas, J. Neutrophils to the ROScue: Mechanisms of NADPH Oxidase Activation and Bacterial Resistance. Front. Cell. Infect. Microbiol. 2017, 7, 373. [Google Scholar] [CrossRef] [PubMed]
  62. Grogan, A.; Reeves, E.; Keep, N.; Wientjes, F.; Totty, N.F.; Burlingame, A.L.; Justin Hsuan, J.; Segal, A.W. Cytosolic Phox Proteins Interact with and Regulate the Assembly of Coronin in Neutrophils. J. Cell Sci. 1997, 110, 3071–3081. [Google Scholar] [CrossRef] [PubMed]
  63. Morimatsu, T.; Kawagoshi, A.; Yoshida, K.; Tamura, M. Actin Enhances the Activation of Human Neutrophil NADPH Oxidase in a Cell-Free System. Biochem. Biophys. Res. Commun. 1997, 230, 206–210. [Google Scholar] [CrossRef] [PubMed]
  64. El Benna, J.; Dang, P.M.C.; Andrieu, V.; Vergnaud, S.; Dewas, C.; Cachia, O.; Fay, M.; Morel, F.; Chollet-Martin, S.; Hakim, J.; et al. P40(Phox) Associates with the Neutrophil Triton X-100-Insoluble Cytoskeletal Fraction and PMA-Activated Membrane Skeleton: A Comparative Study with P67(Phox) and P47(Phox). J. Leukoc. Biol. 1999, 66, 1014–1020. [Google Scholar] [CrossRef] [PubMed]
  65. Hattori, H.; Subramanian, K.K.; Sakai, J.; Luo, H.R. Reactive Oxygen Species as Signaling Molecules in Neutrophil Chemotaxis. Commun. Integr. Biol. 2010, 3, 278. [Google Scholar] [CrossRef]
  66. Keller, H.U.; Niggli, V. The PKC-Inhibitor RO 31-8220 Selectively Suppresses PMA- and Diacylglycerol-Induced Fluid Pinocytosis and Actin Polymerization in PMNS. Biochem. Biophys. Res. Commun. 1993, 194, 1111–1116. [Google Scholar] [CrossRef]
  67. Downey, G.P.; Chan, C.K.; Lea, P.; Takai, A.; Grinstein, S. Phorbol Ester-Induced Actin Assembly in Neutrophils: Role of Protein Kinase C. J. Cell Biol. 1992, 116, 695–706. [Google Scholar] [CrossRef]
  68. Hendey, B.; Zhu, C.L.; Greenstein, S. Fas Activation Opposes PMA-Stimulated Changes in the Localization of PKCδ: A Mechanism for Reducing Neutrophil Adhesion to Endothelial Cells. J. Leukoc. Biol. 2002, 71, 863–870. [Google Scholar] [CrossRef]
  69. Belambri, S.A.; Hurtado-Nedelec, M.; Senator, A.; Makni-Maalej, K.; Fay, M.; Gougerot-Pocidalo, M.-A.; Marie, J.-C.; Dang, P.M.-C.; El-Benna, J. Phosphorylation of P47phox Is Required for Receptor-Mediated NADPH Oxidase/NOX2 Activation in Epstein-Barr Virus-Transformed Human B Lymphocytes. Am. J. Blood Res. 2012, 2, 187–193. [Google Scholar]
  70. Bouin, A.P.; Grandvaux, N.; Vignais, P.V.; Fuchs, A. P40(Phox) Is Phosphorylated on Threonine 154 and Serine 315 during Activation of the Phagocyte NADPH Oxidase: Implication of a Protein Kinase C- Type Kinase in the Phosphorylation Process. J. Biol. Chem. 1998, 273, 30097–30103. [Google Scholar] [CrossRef]
  71. Dang, P.M.C.; Morel, F.; Gougerot-Pocidalo, M.A.; El Benna, J. Phosphorylation of the NADPH Oxidase Component P67PHOX by ERK2 and P38MAPK: Selectivity of Phosphorylated Sites and Existence of an Intramolecular Regulatory Domain in the Tetratricopeptide-Rich Region. Biochemistry 2003, 42, 4520–4526. [Google Scholar] [CrossRef] [PubMed]
  72. Raad, H.; Paclet, M.H.; Boussetta, T.; Kroviarski, Y.; Morel, F.; Quinn, M.T.; Gougerot-Pocidalo, M.A.; Dang, P.M.C.; El-Benna, J. Regulation of the Phagocyte NADPH Oxidase Activity: Phosphorylation of Gp91phox/NOX2 by Protein Kinase C Enhances Its Diaphorase Activity and Binding to Rac2, P67phox, and P47phox. FASEB J. 2009, 23, 1011–1022. [Google Scholar] [CrossRef] [PubMed]
  73. Regier, D.S.; Greene, D.G.; Sergeant, S.; Jesaitis, A.J.; McPhail, L.C. Phosphorylation of P22(Phox) Is Mediated by Phospholipase D-Dependent and -Independent Mechanisms. Correlation of NADPH Oxidase Activity and P22(Phox) Phosphorylation. J. Biol. Chem. 2000, 275, 28406–28412. [Google Scholar] [CrossRef] [PubMed]
  74. Majumdar, S.; Rossi, M.W.; Fujiki, T.; Phillips, W.A.; Disa, S.; Queen, C.F.; Johnston, R.B., Jr.; Rosen, O.M.; Corkey, B.E.; Korchak, H. Protein Kinase C Isotypes and Signaling in Neutrophils. J. Biol. Chem. 1991, 266, 9285–9294. [Google Scholar] [CrossRef]
  75. Kelher, M.R.; McLaughlin, N.J.D.; Banerjee, A.; Elzi, D.J.; Gamboni, F.; Khan, S.Y.; Meng, X.; Mitra, S.; Silliman, C.C. LysoPCs Induce Hck- and PKCδ-Mediated Activation of PKCγ Causing P47 Phox Phosphorylation and Membrane Translocation in Neutrophils. J. Leukoc. Biol. 2016, 101, 261–273. [Google Scholar] [CrossRef]
  76. DeCoursey, T.E.; Ligeti, E. Regulation and Termination of NADPH Oxidase Activity. Cell. Mol. Life Sci. 2005, 62, 2173–2193. [Google Scholar] [CrossRef]
  77. Dusi, S.; Della Bianca, V.; Grzeskowiak, M.; Rossi, F. Relationship between Phosphorylation and Translocation to the Plasma Membrane of P47phox and P67phox and Activation of the NADPH Oxidase in Normal and Ca(2+)-Depleted Human Neutrophils. Biochem. J. 1993, 290, 173–178. [Google Scholar] [CrossRef]
  78. Stolk, J.; Hiltermann, T.J.; Dijkman, J.H.; Verhoeven, A.J. Characteristics of the Inhibition of NADPH Oxidase Activation in Neutrophils by Apocynin, a Methoxy-Substituted Catechol. Am. J. Respir. Cell Mol. Biol. 1994, 11, 95–102. [Google Scholar] [CrossRef]
  79. Soroush, F.; Tang, Y.; Guglielmo, K.; Engelmann, A.; Liverani, E.; Patel, A.; Langston, J.; Sun, S.; Kunapuli, S.; Kiani, M.F.; et al. Protein Kinase C-Delta (PKCδ) Tyrosine Phosphorylation Is a Critical Regulator of Neutrophil-Endothelial Cell Interaction in Inflammation. Shock 2019, 51, 538–547. [Google Scholar] [CrossRef]
  80. Rabani, R.; Cossette, C.; Graham, F.; Powell, W.S. Protein Kinase C Activates NAD Kinase in Human Neutrophils. Free Radic. Biol. Med. 2020, 161, 50–59. [Google Scholar] [CrossRef]
  81. Pilsczek, F.H.; Salina, D.; Poon, K.K.H.; Fahey, C.; Yipp, B.G.; Sibley, C.D.; Robbins, S.M.; Green, F.H.Y.; Surette, M.G.; Sugai, M.; et al. A Novel Mechanism of Rapid Nuclear Neutrophil Extracellular Trap Formation in Response to Staphylococcus Aureus Florian. J. Immunol. 2010, 185, 7413–7425. [Google Scholar] [CrossRef] [PubMed]
  82. Lim, M.B.H.; Kuiper, J.W.P.; Katchky, A.; Goldberg, H.; Glogauer, M. Rac2 Is Required for the Formation of Neutrophil Extracellular Traps. J. Leukoc. Biol. 2011, 90, 771–776. [Google Scholar] [CrossRef]
  83. de Bont, C.M.; Koopman, W.J.H.; Boelens, W.C.; Pruijn, G.J.M. Stimulus-Dependent Chromatin Dynamics, Citrullination, Calcium Signalling and ROS Production during NET Formation. Biochim. Biophys. Acta-Mol. Cell Res. 2018, 1865, 1621–1629. [Google Scholar] [CrossRef]
  84. Tatsiy, O.; McDonald, P.P. Physiological Stimuli Induce PAD4-Dependent, ROS-Independent NETosis, with Early and Late Events Controlled by Discrete Signaling Pathways. Front. Immunol. 2018, 9, 2036. [Google Scholar] [CrossRef] [PubMed]
  85. Widmann, C.; Gibson, S.; Jarpe, M.B.; Johnson, G.L. Mitogen-Activated Protein Kinase: Conservation of a Three-Kinase Module From Yeast to Human. Physiol. Rev. 1999, 79, 143–180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Zhang, W.; Liu, H.T. MAPK Signal Pathways in the Regulation of Cell Proliferation in Mammalian Cells. Cell Res. 2002, 12, 9–18. [Google Scholar] [CrossRef]
  87. Chiang, C.H.; Chuang, C.H.; Liu, S.L. Apocynin Attenuates Ischemia-Reperfusion Lung Injury in an Isolated and Perfused Rat Lung Model. Transl. Res. 2011, 158, 17–29. [Google Scholar] [CrossRef]
  88. Mouzaoui, S.; Djerdjouri, B.; Makhezer, N.; Kroviarski, Y.; El-Benna, J.; Dang, P.M.C. Tumor Necrosis Factor-α-Induced Colitis Increases NADPH Oxidase 1 Expression, Oxidative Stress, and Neutrophil Recruitment in the Colon: Preventive Effect of Apocynin. Mediat. Inflamm. 2014, 2014, 312484. [Google Scholar] [CrossRef]
  89. Arumugam, S.; Girish Subbiah, K.; Kemparaju, K.; Thirunavukkarasu, C. Neutrophil Extracellular Traps in Acrolein Promoted Hepatic Ischemia Reperfusion Injury: Therapeutic Potential of NOX2 and P38MAPK Inhibitors. J. Cell. Physiol. 2017, 233, 3244–3261. [Google Scholar] [CrossRef]
  90. Dang, P.M.C.; Raad, H.; Derkawi, R.A.; Boussetta, T.; Paclet, M.H.; Belambri, S.A.; Makni-Maalej, K.; Kroviarski, Y.; Morel, F.; Gougerot-Pocidalo, M.A.; et al. The NADPH Oxidase Cytosolic Component P67phox Is Constitutively Phosphorylated in Human Neutrophils: Regulation by a Protein Tyrosine Kinase, MEK1/2 and Phosphatases 1/2A. Biochem. Pharmacol. 2011, 82, 1145–1152. [Google Scholar] [CrossRef]
  91. Dang, P.M.C.; Stensballe, A.; Boussetta, T.; Raad, H.; Dewas, C.; Kroviarski, Y.; Hayem, G.; Jensen, O.N.; Gougerot-Pocidalo, M.A.; El-Benna, J. A Specific P47phox-Serine Phosphorylated by Convergent MAPKs Mediates Neutrophil NADPH Oxidase Priming at Inflammatory Sites. J. Clin. Investig. 2006, 116, 2033–2043. [Google Scholar] [CrossRef] [PubMed]
  92. Schönwaßer, D.C.; Palmer, R.H.; Herget, T.; Parker, P.J. P42 MAPK Phosphorylates 80 KDa MARCKS at Ser-113. FEBS Lett. 1996, 395, 1–5. [Google Scholar] [CrossRef] [PubMed]
  93. Arbuzova, A.; Schmitz, A.A.P.; Vergères, G. Cross-Talk Unfolded: MARCKS Proteins. Biochem. J. 2002, 362, 1–12. [Google Scholar] [CrossRef] [PubMed]
  94. Ziemba, B.P.; Burke, J.E.; Masson, G.; Williams, R.L.; Falke, J.J. Regulation of PI3K by PKC and MARCKS: Single-Molecule Analysis of a Reconstituted Signaling Pathway. Biophys. J. 2016, 110, 1811–1825. [Google Scholar] [CrossRef]
  95. Takami, M.; Terry, V.; Petruzzelli, L. Signaling Pathways Involved in IL-8-Dependent Activation of Adhesion Through Mac-1. J. Immunol. 2002, 168, 4559–4566. [Google Scholar] [CrossRef] [Green Version]
  96. Hawkins, P.T.; Stephens, L.R.; Suire, S.; Wilson, M. PI3K Signaling in Neutrophils. In Current Topics in Microbiology and Immunology; Springer: Berlin/Heidelberg, Germany, 2010; pp. 183–202. ISBN 978-3-642-13663-4. [Google Scholar]
  97. Puri, K.D.; Doggett, T.A.; Huang, C.Y.; Douangpanya, J.; Hayflick, J.S.; Turner, M.; Penninger, J.; Diacovo, T.G. The Role of Endothelial PI3Kγ Activity in Neutrophil Trafficking. Blood 2005, 106, 150–157. [Google Scholar] [CrossRef]
  98. Walser, R.; Burke, J.E.; Gogvadze, E.; Bohnacker, T.; Zhang, X.; Hess, D.; Küenzi, P.; Leitges, M.; Hirsch, E.; Williams, R.L.; et al. PKCβ Phosphorylates PI3Kγ to Activate It and Release It from GPCR Control. PLoS Biol. 2013, 11, e1001587. [Google Scholar] [CrossRef]
  99. Luo, H.R.; Mondal, S. Molecular Control of PtdIns(3,4,5)P3 Signaling in Neutrophils. EMBO Rep. 2015, 16, 149–163. [Google Scholar] [CrossRef]
  100. Hemmings, B.A.; Restuccia, D.F. PI3K-PKB/Akt Pathway. Cold Spring Harb. Lab. 2016, 4, a011189. [Google Scholar]
  101. DeSouza-Vieira, T.; Guimarães-Costa, A.; Rochael, N.C.; Lira, M.N.; Nascimento, M.T.; de Lima-Gomez, P.S.; Mariante, R.M.; Persechini, P.M.; Saraiva, E.M. Neutrophil Extracellular Traps Release Induced by Leishmania: Role of PI3Kγ, ERK, PI3Kσ, PKC, and [Ca 2+ ]. J. Leukoc. Biol. 2016, 100, 801–810. [Google Scholar] [CrossRef]
  102. Li, Z.; Jiang, H.; Xie, W.; Zuchuan, Z.; Smrcka, A.V.; Wu, D. Roles of PLC-β 2 and -Β3 and PI3Kγ in Chemoattractant- Mediated Signal Transduction. Science 2000, 287, 1046–1049. [Google Scholar] [CrossRef] [PubMed]
  103. Wrann, C.D.; Tabriz, N.A.; Barkhausen, T.; Klos, A.; van Griensven, M.; Pape, H.C.; Kendoff, D.O.; Guo, R.; Ward, P.A.; Krettek, C.; et al. The Phosphatidylinositol 3-Kinase Signaling Pathway Exerts Protective Effects during Sepsis by Controlling C5a-Mediated Activation of Innate Immune Functions. J. Immunol. 2007, 178, 5940–5948. [Google Scholar] [CrossRef] [PubMed]
  104. Ellson, C.; Davidson, K.; Anderson, K.; Stephens, L.R.; Hawkins, P.T. PtdIns3P Binding to the PX Domain of P40phox Is a Physiological Signal in NADPH Oxidase Activation. EMBO J. 2006, 25, 4468–4478. [Google Scholar] [CrossRef] [PubMed]
  105. Ellson, C.D.; Gobert-Gosse, S.; Anderson, K.E.; Davidson, K.; Erdjument-Bromage, H.; Tempst, P.; Thuring, J.W.; Cooper, M.A.; Lim, Z.Y.; Holmes, A.B.; et al. Ptdlns(3)P Regulates the Neutrophil Oxidase Complex by Binding to the PX Domain of P40phox. Nat. Cell Biol. 2001, 3, 679–682. [Google Scholar] [CrossRef] [PubMed]
  106. Yamada, K.M.; Araki, M. Tumor Suppressor PTEN: Modulator of Cell Signaling, Growth, Migration and Apoptosis. J. Cell Sci. 2001, 114, 2375–2382. [Google Scholar] [CrossRef]
  107. Subramanian, K.K.; Jia, Y.; Zhu, D.; Simms, B.T.; Jo, H.; Hattori, H.; You, J.; Mizgerd, J.P.; Luo, H.R. Tumor Suppressor PTEN Is a Physiologic Suppressor of Chemoattractant- Mediated Neutrophil Functions. Blood 2007, 109, 4028–4037. [Google Scholar] [CrossRef] [Green Version]
  108. Teimourian, S.; Moghanloo, E. Role of PTEN in Neutrophil Extracellular Trap Formation. Mol. Immunol. 2015, 66, 319–324. [Google Scholar] [CrossRef]
  109. Antal, C.E.; Newton, A.C. Spatiotemporal Dynamics of Phosphorylation in Lipid Second Messenger Signaling. Mol. Cell. Proteomics 2013, 12, 3498–3508. [Google Scholar] [CrossRef]
  110. Immler, R.; Simon, S.I.; Sperandio, M. Calcium Signalling and Related Ion Channels in Neutrophil Recruitment and Function. Eur. J. Clin. Investig. 2018, 48, e12964. [Google Scholar] [CrossRef]
  111. Essin, K.; Salanova, B.; Kettritz, R.; Sausbier, M.; Luft, F.C.; Kraus, D.; Bohn, E.; Autenrieth, I.B.; Peschel, A.; Ruth, P.; et al. Large-Conductance Calcium-Activated Potassium Channel Activity Is Absent in Human and Mouse Neutrophils and Is Not Required for Innate Immunity. Am. J. Physiol.-Cell Physiol. 2007, 293, 45–54. [Google Scholar] [CrossRef]
  112. Berg, K.; Ericsson, M.; Lindgren, M.; Gustafsson, H. A High Precision Method for Quantitative Measurements of Reactive Oxygen Species in Frozen Biopsies. PLoS ONE 2014, 9, e90964. [Google Scholar] [CrossRef]
  113. Mahomed, A.G.; Anderson, R. Activation of Human Neutrophils with Chemotactic Peptide, Opsonized Zymosan and the Calcium Ionophore A23187, but Not with a Phorbol Ester, Is Accompanied by Efflux and Store-Operated Influx of Calcium. Inflammation 2000, 24, 559–569. [Google Scholar] [CrossRef] [PubMed]
  114. Caimi, G.; Ferrara, F.; Montana, M.; Meli, F.; Canino, B.; Carollo, C.; Lo Presti, R. Acute Ischemic Stroke: Polymorphonuclear Leukocyte Membrane Fluidity and Cytosolic Ca2+ Concentration at Baseline and after Chemotactic Activation. Stroke 2000, 31, 1578–1582. [Google Scholar] [CrossRef]
  115. Bei, L.; Hu, T.; Qian, Z.M.; Shen, X. Extracellular Ca2+ Regulates the Respiratory Burst of Human Neutrophils. Biochim. Biophys. Acta-Mol. Cell Res. 1998, 1404, 475–483. [Google Scholar] [CrossRef]
  116. Schenten, V.; Melchior, C.; Steinckwich, N.; Tschirhart, E.J.; Brechard, S. Sphingosine Kinases Regulate NOX2 Activity via P38 MAPK-Dependent Translocation of S100A8/A9. J. Leukoc. Biol. 2011, 89, 587–596. [Google Scholar] [CrossRef] [PubMed]
  117. Dusi, S.; Donini, M.; Rossi, F. Tyrosine Phosphorylation and Activation of NADPH Oxidase in Human Neutrophils: A Possible Role for MAP Kinases and for a 75 KDa Protein. Biochem. J. 1994, 304, 243–250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Hu, T.; Liu, Z.; Shen, X. Roles of Phospholipase D in Phorbol Myristate Acetate-Stimulated Neutrophil Respiratory Burst. J. Cell. Mol. Med. 2011, 15, 647–653. [Google Scholar] [CrossRef]
  119. Björnsdottir, H.; Granfeldt, D.; Welin, A.; Bylund, J.; Karlsson, A. Inhibition of Phospholipase A2 Abrogates Intracellular Processing of NADPH-Oxidase Derived Reactive Oxygen Species in Human Neutrophils. Exp. Cell Res. 2013, 319, 761–774. [Google Scholar] [CrossRef]
  120. Hazan, I.; Dana, R.; Granot, Y.; Levy, R. Cytosolic Phospholipase A2 and Its Mode of Activation in Human Neutrophils by Opsonized Zymosan: Correlation between 42/44 KDa Mitogen-Activated Protein Kinase, Cytosolic Phospholipase A2 and NADPH Oxidase. Biochem. J. 1997, 326, 867–876. [Google Scholar] [CrossRef]
  121. Syrbu, S.I.; Waterman, W.H.; Molski, T.F.P.; Nagarkatti, D.; Hajjar, J.J.; Sha’afi, R.I. Phosphorylation of Cytosolic Phospholipase A2 and the Release of Arachidonic Acid in Human Neutrophils. J. Immunol. 1999, 162, 2334–2340. [Google Scholar]
  122. Bougoin, S.; Poubelle, P.E.; Liao, N.W.; Emezawa, K.; Borgeat, P.; Naccache, P.H. Granulocyte-Macrophage Colony-Stimulating Factor Primes Phospholipase D Activity in Human Nerutrophils in Vitro: Role of Calcium, G-Proteins and Tyrosine Kinases. Cell. Signal. 1992, 4, 487–500. [Google Scholar] [CrossRef]
  123. Lopez, I.; Burns, D.J.; Lambeth, J.D. Regulation of Phospholipase D by Protein Kinase C in Human Neutrophils. Conventional Isoforms of Protein Kinase C Phosphorylate a Phospholipase D- Related Component in the Plasma Membrane. J. Biol. Chem. 1995, 270, 19465–19472. [Google Scholar] [CrossRef]
  124. Whatmore, J.; Cronin, P.; Cockcroft, S. ARF1-Regulated Phospholipase D in Human Neutrophils Is Enhanced by PMA and MgATP. FEBS Lett. 1994, 352, 113–117. [Google Scholar] [CrossRef]
  125. D’Souza-Schorey, C.; Chavrier, P. ARF Proteins: Roles in Membrane Traffic and Beyond. Nat. Rev. Mol. Cell Biol. 2006, 7, 347–358. [Google Scholar] [CrossRef]
  126. Zhang, H.; Garlichs, C.D.; Mügge, A.; Daniel, W.G. Role of Ca2+-ATPase Inhibitors in Activation of Cytosolic Phospholipase A2 in Human Polymorphonuclear Neutrophils. Eur. J. Pharmacol. 1999, 364, 229–237. [Google Scholar] [CrossRef]
  127. Gupta, A.K.; Giaglis, S.; Hasler, P.; Hahn, S. Efficient Neutrophil Extracellular Trap Induction Requires Mobilization of Both Intracellular and Extracellular Calcium Pools and Is Modulated by Cyclosporine A. PLoS ONE 2014, 9, 19–21. [Google Scholar] [CrossRef] [PubMed]
  128. Suh, B.-C.; Kim, J.-S.; Namgung, U.; Ha, H.; Kim, K.-T. P2X7 Nucleotide Receptor Mediation of Membrane Pore Formation and Superoxide Generation in Human Promyelocytes and Neutrophils. J. Immunol. 2001, 166, 6754–6763. [Google Scholar] [CrossRef] [Green Version]
  129. Kenny, E.F.; Herzig, A.; Kruger, R.; Muth, A.; Mondal, S.; Thompson, P.R.; Brinkmann, V.; von Bernuth, H.; Zychlinsky, A. Diverse Stimuli Engage Different Neutrophil Extracellular Trap Pathways Elaine. Elife 2017, 6, e24437. [Google Scholar] [CrossRef]
  130. Douda, D.N.; Khan, M.A.; Grasemann, H.; Palaniyar, N. SK3 Channel and Mitochondrial ROS Mediate NADPH Oxidase-Independent NETosis Induced by Calcium Influx. Proc. Natl. Acad. Sci. USA 2015, 112, 2817–2822. [Google Scholar] [CrossRef]
  131. Veras, F.P.; Pontelli, M.C.; Silva, C.M.; Toller-Kawahisa, J.E.; de Lima, M.; Nascimento, D.C.; Schneider, A.H.; Caetité, D.; Tavares, L.A.; Paiva, I.M.; et al. SARS-CoV-2-Triggered Neutrophil Extracellular Traps Mediate COVID-19 Pathology. J. Exp. Med. 2020, 217, e20201129. [Google Scholar] [CrossRef]
  132. Ackermann, M.; Anders, H.J.; Bilyy, R.; Bowlin, G.L.; Daniel, C.; De Lorenzo, R.; Egeblad, M.; Henneck, T.; Hidalgo, A.; Hoffmann, M.; et al. Patients with COVID-19: In the Dark-NETs of Neutrophils. Cell Death Differ. 2021, 28, 3125–3139. [Google Scholar] [CrossRef]
  133. Brinkmann, V. Neutrophil Extracellular Traps in the Second Decade. J. Innate Immun. 2018, 10, 414–421. [Google Scholar] [CrossRef] [PubMed]
  134. Urban, C.F.; Ermert, D.; Schmid, M.; Abu-Abed, U.; Goosmann, C.; Nacken, W.; Brinkmann, V.; Jungblut, P.R.; Zychlinsky, A. Neutrophil Extracellular Traps Contain Calprotectin, a Cytosolic Protein Complex Involved in Host Defense against Candida Albicans. PLoS Pathog. 2009, 5, e1000639. [Google Scholar] [CrossRef] [PubMed]
  135. Branzk, N.; Lubojemska, A.; Hardison, S.E.; Wang, Q.; Maximiliano, G.; Brown, G.D.; Papayannopoulos, V. Neutrophils Sense Microbial Size and Selectively Release Neutrophil Extracellular Traps in Response to Large Pathogens. Nat. Immunol. 2014, 15, 1017–1025. [Google Scholar] [CrossRef]
  136. Yousefi, S.; Mihalache, C.; Kozlowski, E.; Schmid, I.; Simon, H.U. Viable Neutrophils Release Mitochondrial DNA to Form Neutrophil Extracellular Traps. Cell Death Differ. 2009, 16, 1438–1444. [Google Scholar] [CrossRef] [PubMed]
  137. Chatfield, S.M.; Grebe, K.; Whitehead, L.W.; Rogers, K.L.; Nebl, T.; Murphy, J.M.; Wicks, I.P. Monosodium Urate Crystals Generate Nuclease-Resistant Neutrophil Extracellular Traps via a Distinct Molecular Pathway. J. Immunol. 2018, 200, 1801–1816. [Google Scholar] [CrossRef]
  138. Della Coletta, A.M.; Bachiega, T.F.; de Quaglia e Silva, J.C.; Victoriano de Campos Soares, Â.M.; De Faveri, J.; Marques, S.A.; Marques, M.E.A.; Ximenes, V.F.; Dias-Melicio, L.A. Neutrophil Extracellular Traps Identification in Tegumentary Lesions of Patients with Paracoccidioidomycosis and Different Patterns of NETs Generation In Vitro. PLoS Negl. Trop. Dis. 2015, 9, e0004037. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Delgado-Rizo, V.; Martínez-Guzmán, M.A.; Iñiguez-Gutierrez, L.; Garcia-Orozco, A.; Alvarado-Navarro, A.; Fafutis-Morris, M. Neutrophil Extracellular Traps and Its Implications in Inflammation: An Overview. Front. Immunol. 2017, 8, 81. [Google Scholar] [CrossRef]
  140. Konig, M.F.; Andrade, F. A Critical Reappraisal of Neutrophil Extracellular Traps and NETosis Mimics Based on Differential Requirements for Protein Citrullination. Front. Immunol. 2016, 7, 461. [Google Scholar] [CrossRef]
  141. Wang, Y.; Wysocka, J.; Sayegh, J.; Lee, Y.; Perlin, J.R.; Leonelli, L.; Sonbuchner, L.S.; Mcdonald, C.H.; Cook, R.G.; Roeder, R.C.; et al. Human PAD4 Regulates Histone Arginine Methylation Levels via Demethylimination. Science 2004, 306, 279–283. [Google Scholar] [CrossRef]
  142. Farley, K.; Stolley, J.M.; Zhao, P.; Cooley, J.; Remold-O’Donnell, E. A SerpinB1 Regulatory Mechanism Is Essential for Restricting Neutrophil Extracellular Trap Generation. J. Immunol. 2012, 189, 4574–4581. [Google Scholar] [CrossRef] [PubMed]
  143. Metzler, K.D.; Goosmann, C.; Lubojemska, A.; Zychlinsky, A.; Papayannopoulos, V. Myeloperoxidase-Containing Complex Regulates Neutrophil Elastase Release and Actin Dynamics during NETosis. Cell Rep. 2014, 8, 883–896. [Google Scholar] [CrossRef]
  144. Papayannopoulos, V.; Metzler, K.D.; Hakkim, A.; Zychlinsky, A. Neutrophil Elastase and Myeloperoxidase Regulate the Formation of Neutrophil Extracellular Traps. J. Cell Biol. 2010, 191, 677–691. [Google Scholar] [CrossRef] [PubMed]
  145. Stojkov, D.; Amini, P.; Oberson, K.; Sokollik, C.; Duppenthaler, A.; Simon, H.U.; Yousefi, S. ROS and Glutathionylation Balance Cytoskeletal Dynamics in Neutrophil Extracellular Trap Formation. J. Cell Biol. 2017, 216, 4073–4090. [Google Scholar] [CrossRef] [PubMed]
  146. Neubert, E.; Meyer, D.; Rocca, F.; Günay, G.; Kwaczala-Tessmann, A.; Grandke, J.; Senger-Sander, S.; Geisler, C.; Egner, A.; Schön, M.P.; et al. Chromatin Swelling Drives Neutrophil Extracellular Trap Release. Nat. Commun. 2018, 9, 3767. [Google Scholar] [CrossRef] [PubMed]
  147. Li, Y.; Li, M.; Weigel, B.; Mall, M.; Werth, V.P.; Liu, M.-L. Nuclear Envelope Rupture and NET Formation Is Driven by PKCα-mediated Lamin B Disassembly. EMBO Rep. 2020, 21, e48779. [Google Scholar] [CrossRef]
  148. Xie, K.; Varatnitskaya, M.; Maghnouj, A.; Bader, V.; Winklhofer, K.F.; Hahn, S.; Leichert, L.I. Activation Leads to a Significant Shift in the Intracellular Redox Homeostasis of Neutrophil-like Cells. Redox Biol. 2020, 28, 101344. [Google Scholar] [CrossRef]
  149. Douda, D.N.; Yip, L.; Khan, M.A.; Grasemann, H.; Palaniyar, N. To the Editor: Akt Is Essential to Induce NADPH-Dependent NETosis and to Switch the Neutrophil Death to Apoptosis. Blood 2014, 123, 597–600. [Google Scholar] [CrossRef]
  150. Neeli, I.; Radic, M. Opposition between PKC Isoforms Regulates Histone Deimination and Neutrophil Extracellular Chromatin Release. Front. Immunol. 2013, 4, 38. [Google Scholar] [CrossRef]
  151. Wang, S.; Wang, Y. Peptidylarginine Deiminases in Citrullination, Gene Regulation, Health and Pathogenesis. Biochim. Biophys. Acta 2013, 1829, 1126–1135. [Google Scholar] [CrossRef]
  152. Nakashima, K.; Hagiwara, T.; Yamada, M. Nuclear Localization of Peptidylarginine Deiminase V and Histone Deimination in Granulocytes. J. Biol. Chem. 2002, 277, 49562–49568. [Google Scholar] [CrossRef] [PubMed]
  153. Wang, Y.; Li, M.; Stadler, S.; Correll, S.; Li, P.; Wang, D.; Hayama, R.; Leonelli, L.; Han, H.; Grigoryev, S.A.; et al. Histone Hypercitrullination Mediates Chromatin Decondensation and Neutrophil Extracellular Trap Formation. J. Cell Biol. 2009, 184, 205–213. [Google Scholar] [CrossRef] [PubMed]
  154. Li, P.; Li, M.; Lindberg, M.R.; Kennett, M.J.; Xiong, N.; Wang, Y. PAD4 Is Essential for Antibacterial Innate Immunity Mediated by Neutrophil Extracellular Traps. J. Exp. Med. 2010, 207, 1853–1862. [Google Scholar] [CrossRef] [PubMed]
  155. Holmes, C.L.; Shim, D.; Kernien, J.; Johnson, C.J.; Nett, J.E.; Shelef, M.A. Insight into Neutrophil Extracellular Traps through Systematic Evaluation of Citrullination and Peptidylarginine Deiminases. J. Immunol. Res. 2019, 2019, 2160192. [Google Scholar] [CrossRef] [PubMed]
  156. Hawez, A.; Al-Haidari, A.; Madhi, R.; Rahman, M.; Thorlacius, H. MiR-155 Regulates PAD4-Dependent Formation of Neutrophil Extracellular Traps. Front. Immunol. 2019, 10, 2462. [Google Scholar] [CrossRef]
  157. Chargui, A.; Cesaro, A.; Mimouna, S.; Fareh, M.; Brest, P. Subversion of Autophagy in Adherent Invasive Escherichia Coli-Infected Neutrophils Induces Inflammation and Cell Death. PLoS ONE 2012, 7, e51727. [Google Scholar] [CrossRef]
  158. Mitroulis, I.; Kourtzelis, I.; Kambas, K.; Rafail, S.; Chrysanthopoulou, A.; Speletas, M.; Ritis, K. Regulation of the Autophagic Machinery in Human Neutrophils. Eur. J. Immunol. 2010, 40, 1461–1472. [Google Scholar] [CrossRef]
  159. Narni-Mancinelli, E.; Soudja, S.M.; Crozat, K.; Dalod, M.; Gounon, P. Inflammatory Monocytes and Neutrophils Are Licensed to Kill during Memory Responses In Vivo. PLoS Pathog. 2011, 7, 1002457. [Google Scholar] [CrossRef]
  160. Kim, J.; Kundu, M.; Viollet, B.; Guan, K.-L. AMPK and MTOR Regulate Autophagy through Direct Phosphorylation of Ulk1. Nat. Cell Biol. 2011, 13, 132–141. [Google Scholar] [CrossRef]
  161. Martinez, J.; Malireddi, R.K.S.; Lu, Q.; Cunha, L.D.; Pelletier, S.; Gingras, S.; Orchard, R.; Guan, J.L.; Tan, H.; Peng, J.; et al. Molecular Characterization of LC3-Associated Phagocytosis Reveals Distinct Roles for Rubicon, NOX2 and Autophagy Proteins. Nat. Cell Biol. 2015, 17, 893–906. [Google Scholar] [CrossRef] [PubMed]
  162. Kim, B.W.; Jin, Y.; Kim, J.; Kim, J.H.; Jung, J.; Kang, S.; Kim, I.Y.; Kim, J.; Cheong, H.; Song, H.K. The C-Terminal Region of ATG101 Bridges ULK1 and PtdIns3K Complex in Autophagy Initiation. Autophagy 2018, 14, 2104–2116. [Google Scholar] [CrossRef] [PubMed]
  163. Hurley, J.H.; Young, L.N. Mechanisms of Autophagy Initiation. Annu. Rev. Biochem. 2017, 86, 225–244. [Google Scholar] [CrossRef] [PubMed]
  164. Skendros, P.; Mitroulis, I.; Ritis, K. Autophagy in Neutrophils: From Granulopoiesis to Neutrophil Extracellular Traps. Front. Cell Dev. Biol. 2018, 6, 109. [Google Scholar] [CrossRef]
  165. Liang, C.; Lee, J.-S.; Inn, K.-S.; Gack, M.U.; Li, Q.; Roberts, E.A.; Vergne, I.; Deretic, V.; Feng, P.; Akazawa, C.; et al. Beclin1-Binding UVRAG Targets the Class C Vps Complex to Coordinate Autophagosome Maturation and Endocytic Trafficking. Nat. Cell Biol. 2008, 10, 776–787. [Google Scholar] [CrossRef]
  166. Cheng, M.L.; Ho, H.Y.; Lin, H.Y.; Lai, Y.C.; Chiu, D.T.Y. Effective NET Formation in Neutrophils from Individuals with G6PD Taiwan-Hakka Is Associated with Enhanced NADP+ Biosynthesis. Free Radic. Res. 2013, 47, 699–709. [Google Scholar] [CrossRef] [PubMed]
  167. Remijsen, Q.; Berghe, T.V.; Wirawan, E.; Asselbergh, B.; Parthoens, E.; De Rycke, R.; Noppen, S.; Delforge, M.; Willems, J.; Vandenabeele, P. Neutrophil Extracellular Trap Cell Death Requires Both Autophagy and Superoxide Generation. Cell Res. 2011, 21, 290–304. [Google Scholar] [CrossRef] [PubMed]
  168. Park, S.Y.; Shrestha, S.; Youn, Y.J.; Kim, J.K.; Kim, S.Y.; Kim, H.J.; Park, S.H.; Ahn, W.G.; Kim, S.; Lee, M.G.; et al. Autophagy Primes Neutrophils for Neutrophil Extracellular Trap Formation during Sepsis. Am. J. Respir. Crit. Care Med. 2017, 196, 577–589. [Google Scholar] [CrossRef]
  169. Germic, N.; Stojkov, D.; Oberson, K.; Yousefi, S.; Simon, H.U. Neither Eosinophils nor Neutrophils Require ATG5-Dependent Autophagy for Extracellular DNA Trap Formation. Immunology 2017, 152, 517–525. [Google Scholar] [CrossRef]
  170. Sen, R.; Baltimore, D. Multiple Nuclear Factors Interact with the Immunoglobulin Enhancer Sequences. Cell 1986, 46, 705–716. [Google Scholar] [CrossRef]
  171. Oeckinghaus, A.; Ghosh, S. The NF-KB Family of Transcription Factors and Its Regulation. Cold Spring Harb. Respect. Biol. 2009, 1, a000034. [Google Scholar] [CrossRef]
  172. Giridharan, S.; Srinivasan, M. Mechanisms of NF-ΚB P65 and Strategies for Therapeutic Manipulation. J. Inflamm. Res. 2018, 11, 407. [Google Scholar] [CrossRef]
  173. Liu, T.; Zhang, L.; Joo, D.; Sun, S.C. NF-ΚB Signaling in Inflammation. Signal Transduct. Target. Ther. 2017, 2, 17023. [Google Scholar] [CrossRef]
  174. Castro-Alcaraz, S.; Miskolci, V.; Kalasapudi, B.; Davidson, D.; Vancurova, I. NF-ΚB Regulation in Human Neutrophils by Nuclear IκBα: Correlation to Apoptosis. J. Immunol. 2002, 169, 3947–3953. [Google Scholar] [CrossRef] [PubMed]
  175. Kim, H.; Seo, J.Y.; Roh, K.H.; Lim, J.W.; Kim, K.H. Suppression of NF-ΚB Activation and Cytokine Production by N-Acetylcysteine in Pancreatic Acinar Cells. Free Radic. Biol. Med. 2000, 29, 674–683. [Google Scholar] [CrossRef]
  176. Lapponi, M.J.; Carestia, A.; Landoni, V.I.; Rivadeneyra, L.; Etulain, J.; Negrotto, S.; Pozner, R.G.; Schattner, M. Regulation of Neutrophil Extracellular Trap Formation by Anti-Inflammatory Drugs. J. Pharmacol. Exp. Ther. 2013, 345, 430–437. [Google Scholar] [CrossRef] [PubMed]
  177. Langereis, J.D.; Raaijmakers, H.A.J.A.; Ulfman, L.H.; Koenderman, L. Abrogation of NF-ΚB Signaling in Human Neutrophils Induces Neutrophil Survival through Sustained P38-MAPK Activation. J. Leukoc. Biol. 2010, 88, 655–664. [Google Scholar] [CrossRef] [PubMed]
  178. Chen, H.-C.; Lin, H.-C.; Liu, C.-Y.; Wang, C.-H.; Hwang, T.; Huang, T.-T.; Lin, C.-H.; Kuo, H.-P. Neutrophil Elastase Induces IL-8 Synthesis by Lung Epithelial Cells via the Mitogen-Activated Protein Kinase Pathway. J. Biomed. Sci. 2004, 11, 49–58. [Google Scholar] [CrossRef]
  179. Sun, B.; Dwivedi, N.; Bechtel, T.J.; Paulsen, J.L.; Muth, A.; Bawadekar, M.; Li, G.; Thompson, P.R.; Shelef, M.A.; Schiffer, C.A.; et al. Citrullination of NF-KB P65 Enhances Its Nuclear Localization and TLR-Induced Expression of IL-1β and TNFα HHS Public Access. Sci. Immunol. 2017, 2, eaal3062. [Google Scholar] [CrossRef]
  180. Keshari, R.S.; Jyoti, A.; Dubey, M.; Kothari, N.; Kohli, M.; Bogra, J.; Barthwal, M.K.; Dikshit, M. Cytokines Induced Neutrophil Extracellular Traps Formation: Implication for the Inflammatory Disease Condition. PLoS ONE 2012, 7, e48111. [Google Scholar] [CrossRef]
  181. Greten, F.R.; Arkan, M.C.; Bollrath, J.; Hsu, L.-C.; Goode, J.; Miething, C.; Goktuna, S.I.; Neuenhahn, M.; Fierer, J.; Paxian, S.; et al. NF-ΚB Is a Negative Regulator of IL-1β Secretion as Revealed by Genetic and Pharmacological Inhibition of IKKβ. Cell 2007, 130, 918–931. [Google Scholar] [CrossRef]
  182. Sollberger, G.; Choidas, A.; Burn, G.L.; Habenberger, P.; Di Lucrezia, R.; Kordes, S.; Menninger, S.; Eickhoff, J.; Nussbaumer, P.; Klebl, B.; et al. Gasdermin D Plays a Vital Role in the Generation of Neutrophil Extracellular Traps. Sci. Immunol. 2018, 3, eaar6689. [Google Scholar] [CrossRef] [PubMed]
  183. Kayagaki, N.; Stowe, I.B.; Lee, B.L.; O’Rourke, K.; Anderson, K.; Warming, S.; Cuellar, T.; Haley, B.; Roose-Girma, M.; Phung, Q.T.; et al. Caspase-11 Cleaves Gasdermin D for Non-Canonical Inflammasome Signalling. Nature 2015, 526, 666–671. [Google Scholar] [CrossRef] [PubMed]
  184. Shi, J.; Zhao, Y.; Wang, K.; Shi, X.; Wang, Y.; Huang, H.; Zhuang, Y.; Cai, T.; Wang, F.; Shao, F. Cleavage of GSDMD by Inflammatory Caspases Determines Pyroptotic Cell Death. Nature 2015, 526, 660–665. [Google Scholar] [CrossRef]
  185. Liu, X.; Zhang, Z.; Ruan, J.; Pan, Y.; Magupalli, V.G.; Wu, H.; Lieberman, J. Inflammasome-Activated Gasdermin D Causes Pyroptosis by Forming Membrane Pores. Nature 2016, 535, 153–158. [Google Scholar] [CrossRef] [PubMed]
  186. Meher, A.K.; Spinosa, M.; Davis, J.P.; Pope, N.; Laubach, V.E.; Su, G.; Serbulea, V.; Leitinger, N.; Ailawadi, G.; Gilbert, R.; et al. Novel Role of IL (Interleukin)-1β in Neutrophil Extracellular Trap Formation and Abdominal Aortic Aneurysms. Arterioscler. Thromb. Vasc. Biol. 2018, 38, 843–853. [Google Scholar] [CrossRef] [PubMed]
  187. Warnatsch, A.; Tsourouktsoglou, T.D.; Branzk, N.; Wang, Q.; Reincke, S.; Herbst, S.; Gutierrez, M.; Papayannopoulos, V. Reactive Oxygen Species Localization Programs Inflammation to Clear Microbes of Different Size. Immunity 2017, 46, 421–432. [Google Scholar] [CrossRef]
  188. Takei, H.; Araki, A.; Watanabe, H.; Ichinose, A.; Sendo, F. Rapid Killing of Human Neutrophils by the Potent Activator Phorbol 12-Myristate 13-Acetate (PMA) Accompanied by Changes Different from Typical Apoptosis or Necrosis. J. Leukoc. Biol. 1996, 59, 229–240. [Google Scholar] [CrossRef]
  189. Luehong, N.; Khaowmek, J.; Wongsawan, K.; Chuammitri, P. Preferential Pattern of Mouse Neutrophil Cell Death in Response to Various Stimulants. Vitr. Cell. Dev. Biol.-Anim. 2017, 53, 513–524. [Google Scholar] [CrossRef]
  190. Fadeel, B.; Åhlin, A.; Henter, J.-I.; Orrenius, S.; Hampton, M.B. Involvement of Caspases in Neutrophil Apoptosis: Regulation by Reactive Oxygen Species. Blood 1998, 92, 4808–4818. [Google Scholar] [CrossRef]
  191. Fuchs, T.A.; Abed, U.; Goosmann, C.; Hurwitz, R.; Schulze, I.; Wahn, V.; Weinrauch, Y.; Brinkmann, V.; Zychlinsky, A. Novel Cell Death Program Leads to Neutrophil Extracellular Traps. J. Cell Biol. 2007, 176, 231–241. [Google Scholar] [CrossRef]
  192. Desai, J.; Kumar, S.V.; Mulay, S.R.; Konrad, L.; Romoli, S.; Schauer, C.; Herrmann, M.; Bilyy, R.; Müller, S.; Popper, B.; et al. PMA and Crystal-Induced Neutrophil Extracellular Trap Formation Involves RIPK1-RIPK3-MLKL Signaling. Eur. J. Immunol. 2016, 46, 223–229. [Google Scholar] [CrossRef]
  193. Amini, P.; Stojkov, D.; Wang, X.; Wicki, S.; Kaufmann, T.; Wong, W.W.L.; Simon, H.U.; Yousefi, S. NET Formation Can Occur Independently of RIPK3 and MLKL Signaling. Eur. J. Immunol. 2016, 46, 178–184. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Distribution of articles containing the terms neutrophil and PMA along the last five decades, indexed to the PubMed library.
Figure 1. Distribution of articles containing the terms neutrophil and PMA along the last five decades, indexed to the PubMed library.
Cells 11 02889 g001
Figure 2. Major signaling pathways in PMA-activated neutrophils. PMA crosses the cell membrane and activates PKC isoforms (α, βI, βII, δ) which translocate to the cell membrane while some PKC β isoforms are already there. PKC δ phosphorylates MARCKS which activates the PI3K pathway allowing the phosphorylation of PIP2 to PIP3. PIP3, in turn, binds to PH domains such as the one found in Akt. PTEN regulates the outcome of PI3K by dephosphorylating PIP3 to PIP2. PIP2 is converted to DAG and IP3 by phospholipase C (PLC), leading to Ca2+ release by IP3 and further PKC activation by DAG. PKC also activates the MAPK pathway. PKC, MEK1/2 and p38 phosphorylate the cytosolic components of NADPH oxidase (p47phox, p67phox and p40phox) which translocate to membranes assembling the NADPH oxidase complex in the plasma membrane and specific granules, enabling ROS formation. Hydrogen peroxide (H2O2) activates and dissociates a complex formed by granule proteins, namely myeloperoxidase (MPO), neutrophil elastase (NE), azurocidin (AZU), cathepsin G (CG), lactoferrin (LTF), proteinase 3 (PR3) and lysozyme (LYZ). NE dissociation degrades F-actin allowing the translocation of NE to the nucleus to degrade histones. Moreover, PIP3 activates PKC ζ, which in turn, activates PAD4, causing the citrullination of histones. Akt, PAD4 and p38 activate NF-κB, releasing IL-1β coordinated with the gasdermin D cleavage by NE, allowing the release of NETs. See more details in the supplementary video.
Figure 2. Major signaling pathways in PMA-activated neutrophils. PMA crosses the cell membrane and activates PKC isoforms (α, βI, βII, δ) which translocate to the cell membrane while some PKC β isoforms are already there. PKC δ phosphorylates MARCKS which activates the PI3K pathway allowing the phosphorylation of PIP2 to PIP3. PIP3, in turn, binds to PH domains such as the one found in Akt. PTEN regulates the outcome of PI3K by dephosphorylating PIP3 to PIP2. PIP2 is converted to DAG and IP3 by phospholipase C (PLC), leading to Ca2+ release by IP3 and further PKC activation by DAG. PKC also activates the MAPK pathway. PKC, MEK1/2 and p38 phosphorylate the cytosolic components of NADPH oxidase (p47phox, p67phox and p40phox) which translocate to membranes assembling the NADPH oxidase complex in the plasma membrane and specific granules, enabling ROS formation. Hydrogen peroxide (H2O2) activates and dissociates a complex formed by granule proteins, namely myeloperoxidase (MPO), neutrophil elastase (NE), azurocidin (AZU), cathepsin G (CG), lactoferrin (LTF), proteinase 3 (PR3) and lysozyme (LYZ). NE dissociation degrades F-actin allowing the translocation of NE to the nucleus to degrade histones. Moreover, PIP3 activates PKC ζ, which in turn, activates PAD4, causing the citrullination of histones. Akt, PAD4 and p38 activate NF-κB, releasing IL-1β coordinated with the gasdermin D cleavage by NE, allowing the release of NETs. See more details in the supplementary video.
Cells 11 02889 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Damascena, H.L.; Silveira, W.A.A.; Castro, M.S.; Fontes, W. Neutrophil Activated by the Famous and Potent PMA (Phorbol Myristate Acetate). Cells 2022, 11, 2889. https://doi.org/10.3390/cells11182889

AMA Style

Damascena HL, Silveira WAA, Castro MS, Fontes W. Neutrophil Activated by the Famous and Potent PMA (Phorbol Myristate Acetate). Cells. 2022; 11(18):2889. https://doi.org/10.3390/cells11182889

Chicago/Turabian Style

Damascena, Hylane Luiz, Wendy Ann Assis Silveira, Mariana S. Castro, and Wagner Fontes. 2022. "Neutrophil Activated by the Famous and Potent PMA (Phorbol Myristate Acetate)" Cells 11, no. 18: 2889. https://doi.org/10.3390/cells11182889

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