Next Article in Journal
Studies in a Murine Granuloma Model of Instilled Carbon Nanotubes: Relevance to Sarcoidosis
Next Article in Special Issue
Outer Membrane Vesicle Production by Helicobacter pylori Represents an Approach for the Delivery of Virulence Factors CagA, VacA and UreA into Human Gastric Adenocarcinoma (AGS) Cells
Previous Article in Journal
Improved CRISPR/Cas9 Tools for the Rapid Metabolic Engineering of Clostridium acetobutylicum
Previous Article in Special Issue
Helicobacter pylori Infection Acts Synergistically with a High-Fat Diet in the Development of a Proinflammatory and Potentially Proatherogenic Endothelial Cell Environment in an Experimental Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Formyl Peptide Receptors in Permanent and Low-Grade Inflammation: Helicobacter pylori Infection as a Model

1
Department of Agriculture Sciences, University of Naples Federico II, Portici, 80055 Naples, Italy
2
Department of Microbiology and Molecular Medicine, University of Geneva Medical School, 1211 Genèva, Switzerland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(7), 3706; https://doi.org/10.3390/ijms22073706
Submission received: 28 February 2021 / Revised: 27 March 2021 / Accepted: 30 March 2021 / Published: 2 April 2021
(This article belongs to the Special Issue Helicobacter: Infection, Diagnosis and Treatment)

Abstract

:
Formyl peptide receptors (FPRs) are cell surface pattern recognition receptors (PRRs), belonging to the chemoattractant G protein-coupled receptors (GPCRs) family. They play a key role in the innate immune system, regulating both the initiation and the resolution of the inflammatory response. FPRs were originally identified as receptors with high binding affinity for bacteria or mitochondria N-formylated peptides. However, they can also bind a variety of structurally different ligands. Among FPRs, formyl peptide receptor-like 1 (FPRL1) is the most versatile, recognizing N-formyl peptides, non-formylated peptides, and synthetic molecules. In addition, according to the ligand nature, FPRL1 can mediate either pro- or anti-inflammatory responses. Hp(2-20), a Helicobacter pylori-derived, non-formylated peptide, is a potent FPRL1 agonist, participating in Helicobacter pylori-induced gastric inflammation, thus contributing to the related site or not-site specific diseases. The aim of this review is to provide insights into the role of FPRs in H. pylori-associated chronic inflammation, which suggests this receptor as potential target to mitigate both microbial and sterile inflammatory diseases.

1. Introduction

The human body is continuously exposed to millions of pathogens, which can enter it through different ways. The immune system, and in particular, innate immunity, has an important role in controlling the infections, representing the front-line defense against invading pathogens [1,2,3].
Innate immunity is mediated by anatomic barriers (skin and mucosa), physiological functions (pH values, temperature, chemical mediators), and cellular elements [4]. Skin and mucosal surfaces—including the gastrointestinal tract, the respiratory tract, the urogenital tract, and ocular surface—are the earliest sites of host–pathogen interactions, separating the body from the external environment [3,5]. This feature makes them directly involved in preventing the entry of invading microorganisms [3], acting as chemical–physical barriers [6]. Nevertheless, mechanical or chemical cleansing mechanisms provided by anatomic barriers can fail, promoting the action of other innate immune elements—specifically, immune cells, which participate to potentiate the protective role of the mucosal surface. The mucosal immune system, indeed, comprises approximately 80% of all the immune cells (macrophages, monocytes, neutrophils, dendritic cells, B and T cells, and natural killer cells) [6,7], which express specific receptors called pattern recognition receptors (PRRs).
Pattern recognition receptors are germline-encoded receptors, evolutionary selected and conserved. They mediate innate immune recognition by detecting conserved molecular structures associated with pathogens or host tissue/cell damage, also called pathogen-associated molecular patterns (PAMPs) [8] or damage-associated molecular patterns (DAMPs), respectively.
The activation of pattern recognition receptors initiates the inflammatory response, leading to cytokine, chemokine, prostaglandin, and leukotriene production, as they are key messengers signaling the occurrence of tissue damage.
Inflammation is a complex biological response initiated by the host against a microbial infection or an injury. The aim of this process consists of protecting the host by removing the dangerous stimulus and restoring the damaged tissue, thus favoring the return to the homeostasis [9]. However, if not controlled and self-limited, the inflammatory response may cause severe tissue damage, increasing the microbial pathogenicity [10,11] and contributing to the development of chronic inflammatory diseases [12].
Formyl peptide receptors (FPRs) are a family of pattern recognition receptors, able to recognize principally N-formyl methionine-containing peptides, derived both from bacteria (PAMPs) and mitochondria (DAMPs) [13]. In addition to pattern recognition receptors, formyl peptide receptors regulate the inflammatory reaction by modulating the host defense with regulatory functions in both the initiation and the resolution of inflammation. This suggests their involvement in physiological as well as pathological conditions, making them an important element of the innate immune system with a critical role in human health.
The present review focuses on formyl peptide receptors of the gastrointestinal mucosal immune system, their role in inflammation associated with Helicobacter pylori infection, and the possibility of targeting them to modulate H. pylori-associated chronic inflammation.

2. Formyl Peptide Receptors

2.1. Formyl Peptide Receptors: Cell Distribution and Classification

Formyl peptide receptors are a family of classical chemoattractant receptors. Human cells present three different isoforms of FPRs: FPR1, FPR2, or formyl peptide receptor-like 1 (FPRL1), and FPR3 or Formyl peptide receptor-like 2 (FPRL2). Each receptor is encoded by a specific gene of the human FPR gene family (FPR1, FPR2, FPR3) clustered on the chromosomal region 19q13.3 [14]. FPR1, FPR2, and FPR3 genes share high level of sequence homology [9], resulting in receptor similarity. Compared to FPR1, FPRL1 has one additional amino acid, for a total of 351 residues, and a sequence homology corresponding to 69% between the two receptors. Meanwhile, compared to FPR1, FPRL2 has two additional amino acids, for a total of 352, and a sequence homology of 56% [14]. Despite the high sequence identity among the three FPRs, they differ in cell expression, role (Table 1), and ligand recognition (Table 2). Phagocytic cells were the first in which FPRs were identified. Monocytes express the three FPR genes, but their expression can change in differentiating monocytes. Specifically, monocytes differentiating in macrophages maintain FPRL1 expression, while the transition from immature to mature dendritic cells determines the loss of FPR1. The role of FPRs in macrophage and dendritic cell differentiation has been widely investigated, demonstrating that WKYMV (an FPR ligand) negatively regulates LPS-induced dendritic cell differentiation [15,16], while Lipoxin-A4, LL-37, and serum amyloid A protein (SAA) (FPRL1 agonists) mediate macrophage polarization to the anti-inflammatory form M2 [16]. These findings suggest that FPRs play a regulatory function in immune cell differentiation and activation. Unlike monocytes, neutrophils express both FPR1 and FPRL1, but they lack the FPRL2 gene [14]. Recently, adaptative immune cells were discovered to express FPRs. FPRL1 was found in human tonsillar follicular helper T cells, Th1 cells, Th2 cells, Th17 cells, and in naïve CD4 T cells (CD3+, CD4+, CD45RA+, CD45RO-, and CCR7+) [16,17].
In addition, FPR expression has also been described in platelets, microglial cells, epithelial cells, hepatocytes, fibroblasts, and endothelial cells, as well as spleen, brain, placenta, liver, and further tissues, demonstrating the almost ubiquitous distribution of these receptors [18]. The wide expression of formyl peptide receptors in immune and non-immune cells, including various organs and tissues, suggests their involvement in multiple biological functions, with inflammation as the core.
Among the three receptors, formyl peptide receptor-like 1 is the most ubiquitous one. It has been found in different non-myeloid–endothelial cells, epithelial cells, as well as the hepatocyte-and tissue-gastrointestinal tract, brain, spleen, female organ tissues, pancreas, and endocrine glands [19,20]. Furthermore, compared to FPR1 and FPRL2, FPRL1 can interact with the highest variety of chemically different ligands, including non-formylated peptides, organic molecules, and lipid mediators [19]. These properties make FPRL1 the most promiscuous formyl peptide receptor, and more generally, G-protein coupled receptor (GPCR), able to elicit different cell responses, with inflammation as the common feature.

2.2. FPRs: G-Protein Coupled Receptors

FPRs belong to the family of G-protein coupled receptors (GPCRs), the largest class of cell surface seven-transmembrane proteins. Conventionally, GPCRs are located on the cell surface. However, they can also be expressed in the cell nucleus [46]. Recent studies have demonstrated the nuclear localization of FPRL1 in cancer cell lines [46], providing a further element confirming the versatility of this receptor.
More than 3% of the human genome encodes 800 GPCRs, mainly involved in regulating physiological processes by responding to endogenous ligands, such as hormones, neurotransmitters, chemokines, and calcium ions, or to exogenous ligands, such as odorant molecules and light photons [47,48]. Like hormones and neurotransmitter receptors, GPCRs regulate the communication between the inside and outside of the cells [49], transducing the extracellular signals in cellular response.
GPCRs are coupled to heterotrimeric G proteins, consisting of Gα, Gβ, and Gγ subunits, which initiate the downstream signaling associated with the receptor. Ligand binding stimulates GPCR conformational change and the activation of the heterotrimeric G proteins, resulting in guanosine triphosphate (GTP)-bound Gα and Gβ-γ subunit dissociation. In the active state, G proteins activate specific enzymes, which in turn generate second messengers, leading to cellular response [50].
Chemiotaxis are a common cellular response initiated by GPCRs, expressed on chemotactic cells. This subfamily of G protein-coupled receptors plays an important role in sensing chemoattractant molecules [51], coordinating cell migration at the damaged site, and initiation of the immune response.
N-formylated peptides are one of the first identified chemotactic stimuli [14], exhibiting high affinity for G protein-coupled receptors [51]. This has been demonstrated by the observation that the pertussis toxin inhibits formyl peptide-induced chemotaxis, altering their binding affinity for the receptor [52]. More specifically, the pertussis toxin acts by inducing ADP-ribosylation of the α subunit of a Gi-heterotrimeric G protein [53], affecting the molecular shifts responsible for the cell signal translation pathway and specifically inhibiting cell migration.
Based on this evidence, FPRs are Gi protein-coupled receptors—belonging to the γ-group of rhodopsin-like receptors—able [54] to recognize the presence of bacteria or host cell damage and promote the immune response by activating the chemiotaxis.

2.3. N-Formylated Peptides: Signal Peptides Detected by FPRs

The capability to distinguish between self and non-self is fundamental for the wellness of the host. The immune system has the essential role of initiating a defensive response against pathogens, by detecting microbial chemical elements: pathogen-associated molecular patterns (PAMPs). Bacterial signal peptides represent an important class of PAMPs, responsible for innate immune system activation by interacting with specific pattern recognition receptors (PRRs). They play a critical role in nascent protein translocation from the cytoplasm to other sites inside or outside the cell [55,56].
Signal peptides at the N-terminal of the newly synthesized proteins are recognized by specific particles responsible for locating the proteins at the correct site. After translocation, signal peptides are removed by specific enzymes and the mature protein is released [56,57]. Prokaryotic and eukaryotic signal peptides present three highly conserved domains, containing the correct information for protein translocation and removal: an N-terminal region, starting with a methionine, a hydrophobic core, and a C-terminal region [58]. However, all prokaryotes start the protein synthesis with an N-formyl methionine [59], resulting in the fact that bacterial signal peptides are “N-formylated peptides”. Mitochondria, as bacterial ancestry [60], also initiate proteins synthesis with a formylated methionine, revealing that, besides bacteria, damaged host cells also release N-formylated peptides as damage-associated molecular patterns (DAMPs) [61].
These peptides thus represent a host danger signal, which induces innate immunity activation via specific pattern recognition receptors, corresponding to formyl peptide receptors—named according to their high affinity for N-formylated peptides.

2.4. More Than N-Formylated Peptides: Other Ligands Detected by FPRL1

FPRs were originally discovered as receptors with high binding affinity for bacteria or mitochondria N-formylated peptides. In addition to N-formyl-methionine-leucyl-phenylalanine (fMLF), the prototype for N-formylated peptides [18,21], a variety of structurally different ligands have been shown to interact with formyl peptide receptors.
Compared to FPR1, FPRL1 exhibits a lower affinity for formylated peptides. On the contrary, FPRL1 shows the greatest ligand promiscuity, also detecting non-formylated peptides, synthetic molecules, and the eicosanoid Lipoxin A4 (LXA4) [20]. WKYMV is a synthetic, non-formylated peptide recognized by FPRL1. Figure 1 shows the interaction between FPRL1 and WKYMV, displaying the FPRL1 binding site. According to the nature of its ligand, FPRL1 modifies its conformation and consequently its biological function, eliciting a pro- or anti-inflammatory response. In addition, anti-inflammatory ligands may cause receptor homodimerization or heterodimerization with other FPRs, resulting in the activation of the pro-resolving pathway p38/MAPKAPK/Hsp27/IL10, and in the resolution of the inflammation [20,62,63]. The ability of FPRL1 to bind ligands, mediating opposite biological effects, makes it a potential target to control inflammation and inflammatory-associated diseases.

3. Formyl Peptide Receptors in Helicobacter pylori Chronic Infection

3.1. Helicobacter pylori and Chronic Inflammatory Response

Helicobacter pylori is a Gram-negative bacterium colonizing the gastric mucosa of over 50% of the population worldwide [64]. The bacterium colonizes the stomach and infects gastric epithelial cells, promoting chronic inflammation, leading to chronic gastritis, which can eventually degenerate in peptic ulcer and then gastric carcinoma [65]. Several studies have reported the relationship between H. pylori infection and gastric cancer, classifying the bacterium as the primary cause of gastric carcinoma [66]. Our recent work demonstrates the capability of H. pylori to interfere also with biological processes outside the stomach, contributing to the development of neurodegenerative diseases, such as Alzheimer, or metabolic pathologies (type 2 diabetes, obesity, and cardiovascular diseases) [67]. Although the mechanisms responsible for the extra-gastric manifestations of H. pylori remain unclear, a plausible explanation may reside in the H. pylori-associated, low-grade inflammatory state. Therefore, the clinical outcome of H. pylori infection is related to the severity of the inflammatory response, influenced by both host characteristics and bacterial virulence factors [65].
H. pylori utilizes various virulence factors responsible for its pathogenicity by inducing different pathways, resulting in cytokine and chemokine release, as well as the production of oxygen/nitrogen species (ROS/RNS) and growth factors [68]. These factors—especially in chronic conditions—contribute to tissue damage and severe disease progression.
Lipopolysaccharide (LPS), peptidoglycan, and the cytotoxic-associated gene A (CagA) are the most studied H. pylori pathogenicity factors [68,69]. The gene CagA encodes the cytotoxic protein CagA, whose intracellular signaling results in the activation of inflammatory genes and in the modification of the cell scaffold, thus promoting neoplastic transformation [70]. LPS and peptidoglycan instead are directly involved in the bacterium’s adhesion to the gastric epithelium [64]. As these structures are unique to the pathogen, known as pathogen-associated molecular patterns (PAMPs), they are recognized by pattern recognition receptors (PRRs) displayed by eukaryotic cells. Specifically, Toll-like receptor 4 (TLR4) binds LPS, while peptidoglycan is recognized by gastric cells through nucleotide oligomerization domain 1 (NOD1) [68].
FPRs are one of the most relevant class of PRRs. They play an important role in H. pylori infection by interacting with Hp(2-20), an H. pylori-released, non-formylated peptide with the greatest affinity for formyl peptide receptor like-1. Hp(2-20) is a chemotactic factor [34], contributing to Helicobacter pylori-induced gastric inflammation [71,72].

3.2. Helicobacter pylori Hp(2-20) Modulated the Host Immune Response by Interacting with FPRL1

Hp(2-20) is a cecropin-like peptide released during H. pylori growth and possessing different functional characteristics. As an antimicrobial peptide (AMP), Hp(2-20) plays an important role in H. pylori gastric mucosa persistence by acting as bactericidal molecule [73], thus conferring an advantage over other microorganisms. Moreover, it orchestrates the host inflammatory response by interacting with FPRs [34]. When H. pylori colonizes the human stomach, the epithelial cells of the gastric mucosa represent the first component of the innate immune response to be encountered [74]. Hp(2-20) participates to activate the innate immune response by interacting with the FPRL1 expressed on epithelial gastric cells. In particular, the interaction between Hp(2-20) and FPRL1 favors H. pylori gastric mucosa inflammation [71] by initiating a cell signaling cascade, which leads to cytokine release, the recruitment of inflammatory cells, and stimulation of NADPH oxidase-dependent superoxide generation [75]. This bacterial peptide also stimulates gastric epithelial cell migration and proliferation, and increases the expression level of vascular endothelial growth factor (VEGF), whose role is to restore the gastric mucosa after induced injury [7,34]. These events, if contained, favor the host defense and the restoration of the homeostasis. However, in the case of H. pylori infection, the inflammatory response is sustained and exacerbated, thus resulting in the worsening of the clinical outcome due to the onset of severe gastric and extra-gastric diseases [67,76]. Interestingly, despite the massive inflammatory response, Helicobacter pylori has developed several strategies to manipulate the host immune system and persist within the gastric mucosa. H. pylori is able to both send and integrate signals from the gastric epithelium, allowing the host and bacteria to become linked in a dynamic equilibrium [77]. The long-term persistence of H. pylori promotes a permanent and low-grade inflammation, which is a risk factor for severe diseases, including cancer.
Given this scenario, Hp(2-20) has a critical role in the pathological processes associated with H. pylori infection by its interaction with FPRL1.

3.3. Hp(2-20) and FPRL1: Intracellular Signalling Cascade

In addition to G-protein coupled receptors, after ligand binding, FPRs undergo a conformational change that makes them able to interact with the heterotrimeric G protein. Upon activation, the G protein α subunit exchanges guanosine diphosphate (GDP) with guanosine triphosphate (GTP) and dissociates from β-γ subunits [78], leading to phospholipase Cβ (PLCβ) and phosphoinositide 3-kinaseγ activation (PI3Kγ) [22]. Both phospholipase Cβ and phosphoinositide 3-kinaseγ mediate signaling events associated with different cellular responses, such as chemiotaxis, reactive oxygen species (ROS) generation, and degranulation.
Phospholipase Cβ hydrolyzes phosphatidylinositol 4,5-biphosphate (PIP2) in diacylglycerol (DAG)—which remains at the membrane level—and inositol trisphosphate (IP3). This latter moves to the endoplasmic reticulum, regulating Ca++ release and increasing cytosolic Ca++ levels. In addition, both of the second messengers activate protein kinase C (PKC), resulting in NADPH oxidase activation and ROS production [78].
Phosphoinositide 3-kinaseγ, instead, mediates the conversion of phosphatidylinositol 4,5-biphosphate (PIP2) in phosphatidylinositol 3,4,5-trisphosphate (PIP3), which in turn activates protein kinase C (PKC). Furthermore, PI3Kγ activates protein kinase B (Akt) [79], directly involved in regulating the transcriptional activity of the nuclear factor kB (NF-kB). Other intracellular effectors in the formyl peptide receptors signaling cascade include phospholipase A2 and D, mitogen-activated protein kinase (MAPK), extracellular signal-regulated kinase (ERK)1/2, c-Jun N-terminal kinase (JNK), and p38 [21,80,81].
The Hp(2-20)–FPRL1-induced signaling pathway promotes the activation of p42/44 MAPK (ERK), Akt, and signal transducer and activator of transcription (STAT) 3 [75,82], all mediators with a central role in the host defense against an injury. Akt and ERK induce the activation of the transcriptional nuclear factor kB (NF-kB) [83,84], which plays a critical role in innate immunity as the primary regulator of inflammatory response by activating cytokine genes and the inflammasome [85]. Signal transducer and activator of transcription (STAT) 3 also has important immunomodulatory properties. However, chronic activation of STAT3, due to the H. pylori persistence, can degenerate in chronic inflammation and severe associated diseases [86]. Furthermore, several studies have shown the common involvement of the MAPK/ERK pathway in cancer progression [87]. Taken together, these traits make Hp(2-20) and formyl peptide receptor-like 1 critical players in H. pylori infection (Figure 2).
Shortly, despite the role of Hp(2-20) in modulating the immune response against H. pylori infection, the chronic stimulation of FPRL1 induced by bacterium persistence activates intracellular responses, resulting in exacerbation of the inflammatory process, thus contributing to the clinical outcome of the infection.

4. Conclusions and Future Perspectives

The FPR family is one of the most important class of cell surface receptors, playing important roles in innate immunity and host defense by mediating key events during the inflammatory response. Nevertheless, detrimental effects can result from FPR activation. In recent years, several studies have demonstrated the link between FPR activation and the pathogenesis of inflammatory diseases, including cancer [13].
Helicobacter pylori infection represents the most frequent cause of gastric carcinoma. It colonizes and infects gastric epithelial cells by promoting chronic inflammation, leading to severe site-specific diseases [88]. In this context, FPRs—specifically FPRL1—play a critical role by interacting with the H. pylori-derived peptide Hp(2-20).
The activation of FPRL1 by Hp(2-20) triggers epithelial cell migration, proliferation, and angiogenesis. Thus, chronic stimulation of FPRL1 by Hp(2-20), resulting in persistent low grade inflammation, makes these functions harmful and Hp(2-20) a recognized risk factor, not only for H. pylori-associated gastric cancer [89], but also for other chronic inflammatory-related diseases [90].
Based on these considerations, FPR inhibition could represent a promising therapeutical strategy for treating various chronic inflammatory pathologies. Of particular interest is FRPL1, expressed by a variety of cells and tissues and able to bind a broad range of structurally different ligands, exerting pro- or anti-inflammatory actions. Thus, FPRL1 represents a potential target to inhibit inflammatory responses by using agonists with anti-inflammatory properties or antagonists.
This drew attention in searching novel natural molecules able to antagonize FPRL1 or elicit anti-inflammatory responses, paving the way for the discovery of new drugs for the treatment of inflammatory diseases.
Such an approach could also be extended to inflammatory disorders associated with viral infections. Our predictive studies (data not shown) indicate that FPRL1—showing an effective interaction with various coronavirus peptides—could also contribute to the coronavirus disease (COVID)-19 pathogenesis.

Author Contributions

Conceptualization, R.C. writing—original draft preparation, P.C.; writing—review and editing, C.M.; visualization, M.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

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. Yatim, K.M.; Lakkis, F.G. A brief journey through the immune system. Clin. J. Am. Soc. Nephrol. 2015, 10, 1274–1281. [Google Scholar] [CrossRef] [Green Version]
  2. Mogensen, T.H. Pathogen recognition and inflammatory signaling in innate immune defenses. Clin. Microbiol. Rev. 2009, 22, 240–273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Basset, C.; Holton, J.; O’Mahony, R.; Roitt, I. Innate immunity and pathogen-host interaction. Vaccine 2003, 21, S12–S23. [Google Scholar] [CrossRef]
  4. Marshall, J.S.; Warrington, R.; Watson, W.; Kim, H.L. An introduction to immunology and immunopathology. Allergy Asthma Clin. Immunol. 2018, 14, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Leoni, G.; Neumann, P.A.; Sumagin, R.; Denning, T.L.; Nusrat, A. Wound repair: Role of immune-epithelial interactions. Mucosal Immunol. 2015, 8, 959–968. [Google Scholar] [CrossRef] [Green Version]
  6. Holmgren, J.; Czerkinsky, C. Mucosal immunity and vaccines. Nat. Med. 2005, 11, S45. [Google Scholar] [CrossRef]
  7. Jeong, Y.S.; Bae, Y.S. Formyl peptide receptors in the mucosal immune system. Exp. Mol. Med. 2020, 52, 1694–1704. [Google Scholar] [CrossRef]
  8. Medzhitov, R. Recognition of microorganisms and activation of the immune response. Nature 2007, 449, 819–826. [Google Scholar] [CrossRef]
  9. Filep, J.G.; Sekheri, M.; El Kebir, D. Targeting formyl peptide receptors to facilitate the resolution of inflammation. Eur. J. Pharmacol. 2018, 833, 339–348. [Google Scholar] [CrossRef]
  10. Barton, G.M. A calculated response: Control of inflammation by the innate immune system. J. Clin. Investig. 2008, 118, 413. [Google Scholar] [CrossRef]
  11. Xiao, T.S. Innate immunity and inflammation. Cell. Mol. Immunol. 2017, 14, 1–3. [Google Scholar] [CrossRef]
  12. Zhong, J.; Shi, G. Editorial: Regulation of Inflammation in Chronic Disease. Front. Immunol. 2019, 10, 737. [Google Scholar] [CrossRef] [Green Version]
  13. Weiß, E.; Kretschmer, D. Formyl-Peptide Receptors in Infection, Inflammation, and Cancer. Trends Immunol. 2018, 39, 815–829. [Google Scholar] [CrossRef]
  14. Ye, R.D.; Boulay, F.; Ji, M.W.; Dahlgren, C.; Gerard, C.; Parmentier, M.; Serhan, C.N.; Murphy, P.M. International union of basic and clinical pharmacology. LXXIII. Nomenclature for the formyl peptide receptor (FPR) family. Pharmacol. Rev. 2009, 61, 119–161. [Google Scholar] [CrossRef] [PubMed]
  15. Kang, H.K.; Lee, H.-Y.; Kim, M.-K.; Park, K.S.; Park, Y.M.; Kwak, J.-Y.; Bae, Y.-S. The Synthetic Peptide Trp-Lys-Tyr-Met-Val-D-Met Inhibits Human Monocyte-Derived Dendritic Cell Maturation via Formyl Peptide Receptor and Formyl Peptide Receptor-Like 2. J. Immunol. 2005, 175, 685–692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Lee, H.Y.; Lee, M.; Bae, Y.-S. Formyl Peptide Receptors in Cellular Differentiation and Inflammatory Diseases. J. Cell. Biochem. 2017, 118, 1300–1307. [Google Scholar] [CrossRef] [PubMed]
  17. Chen, K.; Bao, Z.; Gong, W.; Tang, P.; Yoshimura, T.; Wang, J.M. Regulation of inflammation by members of the formyl-peptide receptor family. J. Autoimmun. 2017, 85, 64–77. [Google Scholar] [CrossRef] [PubMed]
  18. Migeotte, I.; Communi, D.; Parmentier, M. Formyl peptide receptors: A promiscuous subfamily of G protein-coupled receptors controlling immune responses. Cytokine Growth Factor Rev. 2006, 17, 501–519. [Google Scholar] [CrossRef]
  19. He, H.-Q.; Ye, R. The Formyl Peptide Receptors: Diversity of Ligands and Mechanism for Recognition. Molecules 2017, 22, 455. [Google Scholar] [CrossRef]
  20. Krepel, S.A.; Wang, J.M. Chemotactic Ligands that Activate G-Protein-Coupled Formylpeptide Receptors. Int. J. Mol. Sci. 2019, 20, 3426. [Google Scholar] [CrossRef] [Green Version]
  21. Le, Y.; Murphy, P.M.; Wang, J.M. Formyl-peptide receptors revisited. Trends Immunol. 2002, 23, 541–548. [Google Scholar] [CrossRef]
  22. Le, Y.; Oppenheim, J.J.; Wang, J.M. Pleiotropic roles of formyl peptide receptors. Cytokine Growth Factor Rev. 2001, 12, 91–105. [Google Scholar] [CrossRef]
  23. Kim, S.D.; Kim, J.M.; Jo, S.H.; Lee, H.Y.; Lee, S.Y.; Shim, J.W.; Seo, S.-K.; Yun, J.; Bae, Y.-S. Functional Expression of Formyl Peptide Receptor Family in Human NK Cells. J. Immunol. 2009, 183, 5511–5517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Kim, M.K.; Min, D.S.; Park, Y.J.; Kim, J.H.; Ryu, S.H.; Bae, Y.S. Expression and functional role of formyl peptide receptor in human bone marrow-derived mesenchymal stem cells. FEBS Lett. 2007, 581, 1917–1922. [Google Scholar] [CrossRef] [Green Version]
  25. Heo, S.C.; Kwon, Y.W.; Jang, I.H.; Jeong, G.O.; Yoon, J.W.; Kim, C.D.; Kwon, S.M.; Bae, Y.S.; Kim, J.H. WKYMVm-induced activation of formyl peptide receptor 2 stimulates ischemic neovasculogenesis by promoting homing of endothelial colony-forming cells. Stem Cells 2014, 32, 779–790. [Google Scholar] [CrossRef] [PubMed]
  26. Schröder, N.; Schaffrath, A.; Welter, J.A.; Putzka, T.; Griep, A.; Ziegler, P.; Brandt, E.; Samer, S.; Heneka, M.T.; Kaddatz, H.; et al. Inhibition of formyl peptide receptors improves the outcome in a mouse model of Alzheimer disease. J. Neuroinflamm. 2020, 17, 131. [Google Scholar] [CrossRef] [PubMed]
  27. Yu, N.; Zhang, S.; Lu, J.; Li, Y.; Yi, X.; Tang, L.; Su, L.; Ding, Y. Serum amyloid A, an acute phase protein, stimulates proliferative and proinflammatory responses of keratinocytes. Cell Prolif. 2017, 50. [Google Scholar] [CrossRef] [Green Version]
  28. VanCompernolle, S.E.; Clark, K.L.; Rummel, K.A.; Todd, S.C. Expression and Function of Formyl Peptide Receptors on Human Fibroblast Cells. J. Immunol. 2003, 171, 2050–2056. [Google Scholar] [CrossRef] [Green Version]
  29. McCoy, R.; Haviland, D.L.; Molmenti, E.P.; Ziambaras, T.; Wetsel, R.A.; Perlmutter, D.H. N-formylpeptide and complement C5a receptors are expressed in liver cells and mediate hepatic acute phase gene regulation. J. Exp. Med. 1995, 182, 207–217. [Google Scholar] [CrossRef] [Green Version]
  30. Chen, K.; Liu, M.; Liu, Y.; Yoshimura, T.; Shen, W.; Le, Y.; Durum, S.; Gong, W.; Wang, C.; Gao, J.L.; et al. Formylpeptide receptor-2 contributes to colonic epithelial homeostasis, inflammation, and tumorigenesis. J. Clin. Investig. 2013, 123, 1694–1704. [Google Scholar] [CrossRef] [Green Version]
  31. Bufe, B.; Schumann, T.; Kappl, R.; Bogeski, I.; Kummerow, C.; Podgórska, M.; Smola, S.; Hoth, M.; Zufall, F. Recognition of bacterial signal peptides by mammalian formyl peptide receptors: A new mechanism for sensing pathogens. J. Biol. Chem. 2015, 290, 7369–7387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Wenceslau, C.F.; Szasz, T.; McCarthy, C.G.; Baban, B.; NeSmith, E.; Webb, R.C. Mitochondrial N-formyl peptides cause airway contraction and lung neutrophil infiltration via formyl peptide receptor activation. Pulm. Pharmacol. Ther. 2016, 37, 49–56. [Google Scholar] [CrossRef] [PubMed]
  33. Lind, S.; Gabl, M.; Holdfeldt, A.; Mårtensson, J.; Sundqvist, M.; Nishino, K.; Dahlgren, C.; Mukai, H.; Forsman, H. Identification of Residues Critical for FPR2 Activation by the Cryptic Peptide Mitocryptide-2 Originating from the Mitochondrial DNA–Encoded Cytochrome b. J. Immunol. 2019, 202, 2710–2719. [Google Scholar] [CrossRef] [PubMed]
  34. de Paulis, A.; Prevete, N.; Rossi, F.W.; Rivellese, F.; Salerno, F.; Delfino, G.; Liccardo, B.; Avilla, E.; Montuori, N.; Mascolo, M.; et al. Helicobacter pylori Hp(2–20) Promotes Migration and Proliferation of Gastric Epithelial Cells by Interacting with Formyl Peptide Receptors In Vitro and Accelerates Gastric Mucosal Healing In Vivo. J. Immunol. 2009, 183, 3761–3769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Mills, J.S. Peptides derived from HIV-1, HIV-2, Ebola virus, SARS coronavirus and coronavirus 229E exhibit high affinity binding to the formyl peptide receptor. Biochim. Biophys. Acta Mol. Basis Dis. 2006, 1762, 693–703. [Google Scholar] [CrossRef] [Green Version]
  36. Shen, W.; Proost, P.; Li, B.; Gong, W.; Le, Y.; Sargeant, R.; Murphy, P.M.; Van Damme, J.; Wang, J.M. Activation of the chemotactic peptide receptor FPRL1 in monocytes phosphorylates the chemokine receptor CCR5 and attenuates cell responses to selected chemokines. Biochem. Biophys. Res. Commun. 2000, 272, 276–283. [Google Scholar] [CrossRef] [Green Version]
  37. Le, Y.; Jiang, S.; Hu, J.; Gong, W.; Su, S.; Dunlop, N.M.; Shen, W.; Li, B.; Wang, J.M. N36, a synthetic N-terminal heptad repeat domain of the HIV-1 envelope protein gp41, is an activator of human phagocytes. Clin. Immunol. 2000, 96, 236–242. [Google Scholar] [CrossRef] [Green Version]
  38. Bellner, L.; Thorén, F.; Nygren, E.; Liljeqvist, J.-Å.; Karlsson, A.; Eriksson, K. A Proinflammatory Peptide from Herpes Simplex Virus Type 2 Glycoprotein G Affects Neutrophil, Monocyte, and NK Cell Functions. J. Immunol. 2005, 174, 2235–2241. [Google Scholar] [CrossRef] [Green Version]
  39. Lin, Q.; Fang, D.; Hou, X.; Le, Y.; Fang, J.; Wen, F.; Gong, W.; Chen, K.; Wang, J.M.; Su, S.B. HCV Peptide (C5A), an Amphipathic α-Helical Peptide of Hepatitis Virus C, Is an Activator of N-Formyl Peptide Receptor in Human Phagocytes. J. Immunol. 2011, 186, 2087–2094. [Google Scholar] [CrossRef] [Green Version]
  40. Hebeda, C.B.; Sandri, S.; Benis, C.M.; de Paula-Silva, M.; Loiola, R.A.; Reutelingsperger, C.; Perretti, M.; Farsky, S.H.P. Annexin A1/Formyl Peptide Receptor Pathway Controls Uterine Receptivity to the Blastocyst. Cells 2020, 9, 1188. [Google Scholar] [CrossRef]
  41. Guo, Z.; Hu, Q.; Xu, L.; Guo, Z.N.; Ou, Y.; He, Y.; Yin, C.; Sun, X.; Tang, J.; Zhang, J.H. Lipoxin A4 Reduces Inflammation Through Formyl Peptide Receptor 2/p38 MAPK Signaling Pathway in Subarachnoid Hemorrhage Rats. Stroke 2016, 47, 490–497. [Google Scholar] [CrossRef] [Green Version]
  42. Hinrichs, B.H.; Matthews, J.D.; Siuda, D.; O’Leary, M.N.; Wolfarth, A.A.; Saeedi, B.J.; Nusrat, A.; Neish, A.S. Serum Amyloid A1 Is an Epithelial Prorestitutive Factor. Am. J. Pathol. 2018, 188, 937–949. [Google Scholar] [CrossRef] [Green Version]
  43. Kim, S.-H.; Kim, Y.N.; Jang, Y.-S. Cutting Edge: LL-37–Mediated Formyl Peptide Receptor-2 Signaling in Follicular Dendritic Cells Contributes to B Cell Activation in Peyer’s Patch Germinal Centers. J. Immunol. 2016, 198, 629–633. [Google Scholar] [CrossRef] [Green Version]
  44. Christophe, T.; Karlsson, A.; Dugave, C.; Rabiet, M.J.; Boulay, F.; Dahlgren, C. The Synthetic Peptide Trp-Lys-Tyr-Met-Val-Met-NH2 Specifically Activates Neutrophils through FPRL1/Lipoxin A4 Receptors and is an Agonist for the Orphan Monocyte-expressed Chemoattractant Receptor FPRL2. J. Biol. Chem. 2001, 276, 21585–21593. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Schepetkin, I.A.; Khlebnikov, A.I.; Giovannoni, M.P.; Kirpotina, L.N.; Cilibrizzi, A.; Quinn, M.T. Development of Small Molecule Non-peptide Formyl Peptide Receptor (FPR) Ligands and Molecular Modeling of Their Recognition. Curr. Med. Chem. 2014, 21, 1478–1504. [Google Scholar] [CrossRef] [Green Version]
  46. Cattaneo, F.; Parisi, M.; Fioretti, T.; Sarnataro, D.; Esposito, G.; Ammendola, R. Nuclear localization of Formyl-Peptide Receptor 2 in human cancer cells. Arch. Biochem. Biophys. 2016, 603, 10–19. [Google Scholar] [CrossRef] [PubMed]
  47. Pierce, K.L.; Premont, R.T.; Lefkowitz, R.J. Seven-transmembrane receptors. Nat. Rev. Mol. Cell Biol. 2002, 3, 639–650. [Google Scholar] [CrossRef]
  48. Wang, X.; Iyer, A.; Lyons, A.B.; Körner, H.; Wei, W. Emerging roles for G-protein coupled receptors in development and activation of macrophages. Front. Immunol. 2019, 10, 2031. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Rosenbaum, D.M.; Rasmussen, S.G.F.; Kobilka, B.K. The structure and function of G-protein-coupled receptors. Nature 2009, 459, 356–363. [Google Scholar] [CrossRef] [Green Version]
  50. Cattaneo, F.; Guerra, G.; Parisi, M.; De Marinis, M.; Tafuri, D.; Cinelli, M.; Ammendola, R. Cell-Surface Receptors Transactivation Mediated by G Protein-Coupled Receptors. Int. J. Mol. Sci. 2014, 15, 19700–19728. [Google Scholar] [CrossRef] [Green Version]
  51. Sun, L.; Ye, R.D. Role of G protein-coupled receptors in inflammation. Acta Pharmacol. Sin. 2012, 33, 342–350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Gilder, A.S.; Wang, L.; Natali, L.; Karimi-Mostowfi, N.; Brifault, C.; Gonias, S.L. Pertussis toxin is a robust and selective inhibitor of high grade glioma cell migration and invasion. PLoS ONE 2016, 11, e0168418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Locht, C.; Coutte, L.; Mielcarek, N. The ins and outs of pertussis toxin. FEBS J. 2011, 278, 4668–4682. [Google Scholar] [CrossRef] [PubMed]
  54. Zhuang, Y.; Liu, H.; Zhou, X.E.; Verma, R.K.; de Waal, P.W.; Jang, W.; Xu, T.H.; Wang, L.; Meng, X.; Zhao, G.; et al. Structure of formylpeptide receptor 2-Gi complex reveals insights into ligand recognition and signaling. Nat. Commun. 2020, 11, 1–12. [Google Scholar] [CrossRef] [Green Version]
  55. Ivankov, D.N.; Payne, S.H.; Galperin, M.Y.; Bonissone, S.; Pevzner, P.A.; Frishman, D. How many signal peptides are there in bacteria? Environ. Microbiol. 2013, 15, 983–990. [Google Scholar] [CrossRef] [Green Version]
  56. Bufe, B.; Zufall, F. The sensing of bacteria: Emerging principles for the detection of signal sequences by formyl peptide receptors. Biomol. Concepts 2016, 7, 205–214. [Google Scholar] [CrossRef]
  57. Tjalsma, H.; Bolhuis, A.; Jongbloed, J.D.H.; Bron, S.; van Dijl, J.M. Signal Peptide-Dependent Protein Transport inBacillus subtilis: A Genome-Based Survey of the Secretome. Microbiol. Mol. Biol. Rev. 2000, 64, 515–547. [Google Scholar] [CrossRef] [Green Version]
  58. Von Heijne, G. Analysis of the distribution of charged residues in the N-terminal region of signal sequences: Implications for protein export in prokaryotic and eukaryotic cells. EMBO J. 1984, 3, 2315–2318. [Google Scholar] [CrossRef]
  59. Piatkov, K.I.; Vu, T.T.M.; Hwang, C.-S.; Varshavsky, A. Formyl-methionine as a degradation signal at the N-termini of bacterial proteins. Microb. Cell 2015, 2, 376. [Google Scholar] [CrossRef] [Green Version]
  60. Meyer, A.; Laverny, G.; Bernardi, L.; Charles, A.L.; Alsaleh, G.; Pottecher, J.; Sibilia, J.; Geny, B. Mitochondria: An organelle of bacterial origin controlling inflammation. Front. Immunol. 2018, 9, 536. [Google Scholar] [CrossRef]
  61. Zhang, Q.; Raoof, M.; Chen, Y.; Sumi, Y.; Sursal, T.; Junger, W.; Brohi, K.; Itagaki, K.; Hauser, C.J. Circulating mitochondrial DAMPs cause inflammatory responses to injury. Nature 2010, 464, 104–107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Cooray, S.N.; Gobbetti, T.; Montero-Melendez, T.; McArthur, S.; Thompson, D.; Clark, A.J.L.; Flower, R.J.; Perretti, M. Ligand-specific conformational change of the G-protein-coupled receptor ALX/FPR2 determines proresolving functional responses. Proc. Natl. Acad. Sci. USA 2013, 110, 18232–18237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Raabe, C.A.; Gröper, J.; Rescher, U. Biased perspectives on formyl peptide receptors. Biochim. Biophys. Acta Mol. Cell Res. 2019, 1866, 305–316. [Google Scholar] [CrossRef] [PubMed]
  64. Gottesmann, M.; Paraskevopoulou, V.; Mohammed, A.; Falcone, F.H.; Hensel, A. BabA and LPS inhibitors against Helicobacter pylori: Pectins and pectin-like rhamnogalacturonans as adhesion blockers. Appl. Microbiol. Biotechnol. 2020, 104, 351–363. [Google Scholar] [CrossRef]
  65. Cuomo, P.; Papaianni, M.; Fulgione, A.; Guerra, F.; Capparelli, R.; Medaglia, C. An Innovative Approach to Control H. pylori-Induced Persistent Inflammation and Colonization. Microorganisms 2020, 8, 1214. [Google Scholar] [CrossRef] [PubMed]
  66. Polk, D.B.; Peek, R.M. Helicobacter pylori: Gastric cancer and beyond. Nat. Rev. Cancer 2010, 10, 403–414. [Google Scholar] [CrossRef] [Green Version]
  67. Cuomo, P.; Papaianni, M.; Sansone, C.; Iannelli, A.; Iannelli, D.; Medaglia, C.; Paris, D.; Motta, A.; Capparelli, R. An In Vitro Model to Investigate the Role of Helicobacter pylori in Type 2 Diabetes, Obesity, Alzheimer’s Disease and Cardiometabolic Disease. Int. J. Mol. Sci. 2020, 21, 8369. [Google Scholar] [CrossRef]
  68. Lamb, A.; Chen, L.F. Role of the Helicobacter pylori-Induced inflammatory response in the development of gastric cancer. J. Cell. Biochem. 2013, 114, 491–497. [Google Scholar] [CrossRef] [Green Version]
  69. Jimenez-Soto, L.F.; Haas, R. The CagA toxin of Helicobacter pylori: Abundant production but relatively low amount translocated. Sci. Rep. 2016, 6, 1–7. [Google Scholar] [CrossRef] [Green Version]
  70. Hatakeyama, M. Structure and function of helicobacter pylori caga, the first-identified bacterial protein involved in human cancer. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 2017, 93, 196–219. [Google Scholar] [CrossRef] [Green Version]
  71. Bylund, J.; Christophe, T.; Boulay, F.; Nyström, T.; Karlsson, A.; Dahlgren, C. Proinflammatory activity of a cecropin-like antibacterial peptide from Helicobacter pylori. Antimicrob. Agents Chemother. 2001, 45, 1700–1704. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. de Paulis, A.; Prevete, N.; Fiorentino, I.; Walls, A.F.; Curto, M.; Petraroli, A.; Castaldo, V.; Ceppa, P.; Fiocca, R.; Marone, G. Basophils Infiltrate Human Gastric Mucosa at Sites of Helicobacter pylori Infection, and Exhibit Chemotaxis in Response to H. pylori- derived Peptide Hp(2–20). J. Immunol. 2004, 172, 7734–7743. [Google Scholar] [CrossRef] [Green Version]
  73. Capparelli, R.; De Chiara, F.; Nocerino, N.; Montella, R.C.; Iannaccone, M.; Fulgione, A.; Romanelli, A.; Avitabile, C.; Blaiotta, G.; Capuano, F. New perspectives for natural antimicrobial peptides: Application as antinflammatory drugs in a murine model. BMC Immunol. 2012, 13, 61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Peek, R.M.; Fiske, C.; Wilson, K.T. Role of innate immunity in Helicobacter pylori-induced gastric malignancy. Physiol. Rev. 2010, 90, 831–858. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Cattaneo, F.; Parisi, M.; Ammendola, R. Distinct Signaling Cascades Elicited by Different Formyl Peptide Receptor 2 (FPR2) Agonists. Int. J. Mol. Sci. 2013, 14, 7193–7230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Contaldi, F.; Capuano, F.; Fulgione, A.; Cigliano, R.A.; Sanseverino, W.; Iannelli, D.; Medaglia, C.; Capparelli, R. The hypothesis that Helicobacter pylori predisposes to Alzheimer’s disease is biologically plausible. Sci. Rep. 2017, 7, 7817. [Google Scholar] [CrossRef]
  77. Blaser, M.J.; Kirschner, D. The equilibria that allow bacterial persistence in human hosts. Nature 2007, 449, 843–849. [Google Scholar] [CrossRef]
  78. Dorward, D.A.; Lucas, C.D.; Chapman, G.B.; Haslett, C.; Dhaliwal, K.; Rossi, A.G. The role of formylated peptides and formyl peptide receptor 1 in governing neutrophil function during acute inflammation. Am. J. Pathol. 2015, 185, 1172–1184. [Google Scholar] [CrossRef] [Green Version]
  79. Zhang, L.; Wang, G.; Chen, X.; Xue, X.; Guo, Q.; Liu, M.; Zhao, J. Formyl peptide receptors promotes neural differentiation in mouse neural stem cells by ROS generation and regulation of PI3KAKT signaling. Sci. Rep. 2017, 7, 1–16. [Google Scholar] [CrossRef]
  80. Bae, Y.S.; Song, J.Y.; Kim, Y.; He, R.; Ye, R.D.; Kwak, J.Y.; Suh, P.G.; Ryu, S.H. Differential activation of formyl peptide receptor signaling by peptide ligands. Mol. Pharmacol. 2003, 64, 841–847. [Google Scholar] [CrossRef]
  81. Cattaneo, F.; Russo, R.; Castaldo, M.; Chambery, A.; Zollo, C.; Esposito, G.; Pedone, P.V.; Ammendola, R. Phosphoproteomic analysis sheds light on intracellular signaling cascades triggered by Formyl-Peptide Receptor 2. Sci. Rep. 2019, 9, 1–16. [Google Scholar] [CrossRef]
  82. Rossi, F.W.; Prevete, N.; Montuori, N.; Ragno, P.; Selleri, C.; Marone, G.; de Paulis, A. Hp(2-20) peptide of Helicobacter py-lori and the innate immune receptors: Specific role(s) of the formyl peptide receptors. Le Infezioni Medicina 2012, 20 (Suppl. S2), 19–25. [Google Scholar]
  83. Bai, D.; Ueno, L.; Vogt, P.K. Akt-mediated regulation of NFκB and the essentialness of NFκB for the oncogenicity of PI3K and Akt. Int. J. Cancer 2009, 125, 2863–2870. [Google Scholar] [CrossRef] [Green Version]
  84. Seo, J.H.; Lim, J.W.; Kim, H. Differential Role of ERK and p38 on NF-κB Activation in Helicobacter pylori-Infected Gastric Epithelial Cells. J. Cancer Prev. 2013, 18, 346–350. [Google Scholar] [CrossRef] [Green Version]
  85. Liu, T.; Zhang, L.; Joo, D.; Sun, S.C. NF-κB signaling in inflammation. Signal. Transduct. Target. Ther. 2017, 2, 1–9. [Google Scholar] [CrossRef] [Green Version]
  86. Kasembeli, M.; Bharadwaj, U.; Robinson, P.; Tweardy, D. Contribution of STAT3 to Inflammatory and Fibrotic Diseases and Prospects for its Targeting for Treatment. Int. J. Mol. Sci. 2018, 19, 2299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Reddy, K.B.; Nabha, S.M.; Atanaskova, N. Role of MAP kinase in tumor progression and invasion. Cancer Metastasis Rev. 2003, 22, 395–403. [Google Scholar] [CrossRef] [PubMed]
  88. Crowe, S.E. Helicobacter infection, chronic inflammation, and the development of malignancy. Curr. Opin. Gastroenterol. 2005, 21, 32–38. [Google Scholar] [PubMed]
  89. Hou, X.L.; Ji, C.D.; Tang, J.; Wang, Y.; Wang, Y.X.; Xiang, D.F.; Li, H.Q.; Liu, W.; Liu, W.W.; Wang, J.X.; et al. FPR2 promotes invasion and metastasis of gastric cancer cells and predicts the prognosis of patients. Sci. Rep. 2017, 7, 1–11. [Google Scholar] [CrossRef] [PubMed]
  90. Capparelli, R.; Iannelli, D. Genetics of Host Protection against Helicobacter pylori Infections. Int. J. Mol. Sci. 2021, 22, 3192. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the synthetic peptide WKYMV binding pocket for FPRL1.The receptor is shown in cartoon representation colored in cyan, while the ligand is shown in sticks colored in green. The amino acids of the binding site are represented in sticks colored in light pink. WKYMV forms hydrophobic interactions with the amino acids of the extracellular domain (ExD); II, III, IV, VI, VII transmembrane portions (He); II, III, IV extracellular portions (Ex); and hydrogen bonds with E89 (ExII), R205 (ExIII), D281, and N285 (ExIV). PyMOL Molecular Graphic System (Version 1.3 Shrodinger, LLC) was used to represent the WKYMV–FPRL1 interaction (PDBcode: 6LW5).
Figure 1. Schematic representation of the synthetic peptide WKYMV binding pocket for FPRL1.The receptor is shown in cartoon representation colored in cyan, while the ligand is shown in sticks colored in green. The amino acids of the binding site are represented in sticks colored in light pink. WKYMV forms hydrophobic interactions with the amino acids of the extracellular domain (ExD); II, III, IV, VI, VII transmembrane portions (He); II, III, IV extracellular portions (Ex); and hydrogen bonds with E89 (ExII), R205 (ExIII), D281, and N285 (ExIV). PyMOL Molecular Graphic System (Version 1.3 Shrodinger, LLC) was used to represent the WKYMV–FPRL1 interaction (PDBcode: 6LW5).
Ijms 22 03706 g001
Figure 2. Schematic signalling pathway of Hp(2-20)-activated FPRL1. FPRL1 participates in the pathogenesis of H. pylori infection by interacting with Hp(2-20). During Hp(2-20) binding, the G protein α subunit exchanges guanosine diphosphate (GDP) with guanosine triphosphate (GTP) and dissociates from β-γ subunits. This activates various signal transduction events, resulting in oxidative stress, cell migration, inflammation, cell proliferation, and angiogenesis. These events, which are chronically induced, may lead to severe gastric diseases, including cancer. Abbreviations: Akt: protein kinase B; ERK: extracellular signal-regulated kinase; GTP: guanosine triphosphate; MEKK: mitogen-activated protein kinase kinase; NF-kB: nuclear factor kB; PI3Kγ: phosphoinositide 3-kinase gamma; PLCβ: phospholipase Cβ; PKC: protein Kinase C; ROS: reactive oxygen species; STAT3: signal transducer and activator of transcription 3; VEGF: vascular endothelial growth factor.
Figure 2. Schematic signalling pathway of Hp(2-20)-activated FPRL1. FPRL1 participates in the pathogenesis of H. pylori infection by interacting with Hp(2-20). During Hp(2-20) binding, the G protein α subunit exchanges guanosine diphosphate (GDP) with guanosine triphosphate (GTP) and dissociates from β-γ subunits. This activates various signal transduction events, resulting in oxidative stress, cell migration, inflammation, cell proliferation, and angiogenesis. These events, which are chronically induced, may lead to severe gastric diseases, including cancer. Abbreviations: Akt: protein kinase B; ERK: extracellular signal-regulated kinase; GTP: guanosine triphosphate; MEKK: mitogen-activated protein kinase kinase; NF-kB: nuclear factor kB; PI3Kγ: phosphoinositide 3-kinase gamma; PLCβ: phospholipase Cβ; PKC: protein Kinase C; ROS: reactive oxygen species; STAT3: signal transducer and activator of transcription 3; VEGF: vascular endothelial growth factor.
Ijms 22 03706 g002
Table 1. Distribution and roles of formyl peptide receptors (FPRs; FPR1, formyl peptide receptor-like (FPRL)1, and FPRL2) in innate, adaptative, and non-immune cells. The table also reports the formyl peptide receptor expression in organs or tissues, or where it is not well-defined.
Table 1. Distribution and roles of formyl peptide receptors (FPRs; FPR1, formyl peptide receptor-like (FPRL)1, and FPRL2) in innate, adaptative, and non-immune cells. The table also reports the formyl peptide receptor expression in organs or tissues, or where it is not well-defined.
Cells/TissuesFormyl Peptide ReceptorsRoleReferences
Innate Immune Cell
Expression
NeutrophilsFPR1, FPRL1Chemotaxis, phagocytosis, superoxide generation[16,21,22]
Natural killer cellsFPR1, FPRL1Interferonγ production[16,23]
Immature dendritic cellsFPR1, FPRL2Chemotaxis[15,21,22]
Mature dendritic cellsFPRL2Chemotaxis[15,21,22]
MonocytesFPR1, FPRL1, FPRL2Chemotaxis, pro-inflammatory activity[16]
MacrophagesFPR1, FPRL1, FPRL2Chemotaxis, pro-inflammatory activity[16]
Adaptative Immune Cell ExpressionNaïve CD4 T cells (CD3+, CD4+, CD45RA+, CD45RO–, CCR7+)FPRL1Interferon-γ production[16,17]
Th1 cellsFPRL1[16,17]
Th2FPRL1[16,17]
Th17FPRL1[16,17]
Non-Immune Cells, Organ/Tissue ExpressionEpithelial cellsFPRL1Chemotaxis[18,24]
Endothelial cellsFPRL1Chemotaxis, angiogenesis, and cell proliferation[25]
Microglial cellsFPRL1, FPRL2Inflammation and neurogenerative activity[18,26]
KeratinocytesFPRL1Cell proliferation and pro-inflammatory activity[27]
FibroblastsFPRL1Chemotaxis and innate immune response stimulation[28]
AstrocytesFPRL1Inflammation and neurogenerative activity[18,26]
HepatocytesFPRL1Chemotaxis, angiogenesis[18,29]
Intestinal epithelial cellsFPR1, FPRL1Cell proliferation, inflammation, and tumorigenesis[30]
BrainFPRL1Inflammation and neurodegenerative activity[16,17,18]
SpleenFPRL1, FPRL2Innate immune response[16,17,18]
PlacentaFPRL1, FPRL2Innate immune response[16,17,18]
LungFPRL1, FPRL2Innate immune response[16,17,18]
TestisFPRL1Innate immune response[16,17,18]
TracheaFPRL2Innate immune response[16,17,18]
Lymph nodesFPRL2Innate immune response[16,17,18]
Table 2. Formyl peptide receptor (FPR1, FPRL1, and FPRL2) representative ligands. The table summarizes the main differently derived formylated or non-formylated FPR ligands, indicating the origin, selectivity, and general intracellular signaling.
Table 2. Formyl peptide receptor (FPR1, FPRL1, and FPRL2) representative ligands. The table summarizes the main differently derived formylated or non-formylated FPR ligands, indicating the origin, selectivity, and general intracellular signaling.
ClassificationLigandOriginSignalingSelectivityReferences
Formylated Bacterial Peptidesf-MLFE. coliCa++ mobilization, superoxide generationFPR1[14]
f-MKNFKGBacillusCa++ mobilization, superoxide generationFPRL1[31]
f-MGFFISStreptococcusCa++ mobilization, superoxide generationFPR1, FPRL1[31]
f-MAMKKLSalmonellaCa++ mobilization, superoxide generationFPR1[31]
f-MVMKFKHaemophilusCa++ mobilization, superoxide generationFPR1, FPRL1[31]
f-MFIYYCKStaphylococcusCa++ mobilization, superoxide generationFPR1[31]
f-MKKIMLListeriaCa++ mobilization, superoxide generationFPR1, FPRL1[31]
f-MKKNLVClostridiumCa++ mobilization, superoxide generationFPRL1[31]
Formylated Mitochondria Peptidesf-MMYALFMitochondrionSuperoxide generationFPRL1[32]
f-MLKIVMitochondrionCa++ mobilization, ERK activationFPRL1[32]
f-MYFINILTLMitochondrionCa++ mobilization, ERK activationFPRL1[32]
f-MFADRWMitochondrionCa++ mobilization, ERKs activationFPRL1[32]
Mitocryptide-2MitochondrionCa++ mobilization, ERK activationFPRL1[33]
Microbe-Derived Non-Formylated PeptidesHp(2-20)Helicobacter pyloriSuperoxide generation, cell proliferation, Akt and STAT3 activation, VEGFA secretionFPRL1[34]
OC43 Coronavirus proteinOC43 CoronavirusUnknownNot clear[35]
229E Coronavirus protein229E CoronavirusUnknownNot clear[35]
NL36 Coronavirus proteinNL36 CoronavirusUnknownNot clear[35]
spike proteinEbola virusUnknownNot clear[35]
T20/DP178HIV gp41Ca++ mobilizationFPR1[20]
T21/DP107HIV gp41Ca++ mobilizationFPR1, FPRL1[20]
V3 peptideHIV gp120Ca++ mobilization, CCR5 desensitizationFPRL1[36]
N36 peptideHIV gp41Ca++ mobilization, chemokine receptorsn desensitization, NF-kB activationFPRL1[37]
gG-2p20Herpes simplex virusSuperoxide generation, NADPH oxidase activation, apoptosisFPR1[38]
C5a HCV peptideHepatitis C virusCa++ mobilization, degranulation, superoxide generation, MAPK activationFPRL1[39]
Host-Derived MoleculesAnnexin 1HostERK phosphorylation, NF-kB pathwayFPRL1[40]
Lipoxin-A4HostCa++ mobilization, ERKs, PI3K, Akt phosphorylationFPRL1[41]
SAAHostCa++ mobilization, ERKs, JNK and p38MAPK activation, cytokine release, NF-kB and COX2 inductionFPRL1[42]
Aβ-42hostPI3K/Akt pathway activationFPRL1[26]
LL-37hostCa++ mobilization, Bcl-xL expression, caspase-3 inhibition, MAPK and JAK/STAT signalingFPRL1[43]
Synthetic PeptidesWKYMVmsyntheticCa++ mobilization, NADPH oxidase activation, ERK phosphorylation, MAPK and JNK activation, PKC activationFPRL1[44]
Synthetic MoleculesQuinazolinonesSyntheticCa++ mobilization, ERK activationFPRL1[45]
BenzimidazolesSyntheticCa++ mobilizationFPR1[45]
PyrazolonesSyntheticCa++ mobilization, desensitization of chemokine receptorsFPRL1[45]
Pyridazin-3(2H)-onesSyntheticCa++ mobilizationFPR1[45]
Chiral pyridazinesSyntheticCa++ mobilizationFPR1, FPRL1[45]
N-phenylureasSyntheticCa++ mobilizationFPRL1[45]
ERK: extracellular signal-regulated kinase; Akt: protein kinase B; STAT3: signal transducer and activator of transcription 3; NADPH: nicotinamide adenine dinucleotide phosphate; VEGF: vascular endothelial growth factor; Aβ-42: β amyloid protein 42; JNK: c-Jun N-terminal kinase; PKC: protein kinase C; COX2: cyclooxygenase-2; PI3K: phosphoinositide 3-kinase; SAA: serum amyloid protein A; CCR5: C–C chemokine receptor 5; JAK: Janus kinase; MAPK: mitogen-activated protein kinase; NF-kB: nuclear factor kB.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cuomo, P.; Papaianni, M.; Capparelli, R.; Medaglia, C. The Role of Formyl Peptide Receptors in Permanent and Low-Grade Inflammation: Helicobacter pylori Infection as a Model. Int. J. Mol. Sci. 2021, 22, 3706. https://doi.org/10.3390/ijms22073706

AMA Style

Cuomo P, Papaianni M, Capparelli R, Medaglia C. The Role of Formyl Peptide Receptors in Permanent and Low-Grade Inflammation: Helicobacter pylori Infection as a Model. International Journal of Molecular Sciences. 2021; 22(7):3706. https://doi.org/10.3390/ijms22073706

Chicago/Turabian Style

Cuomo, Paola, Marina Papaianni, Rosanna Capparelli, and Chiara Medaglia. 2021. "The Role of Formyl Peptide Receptors in Permanent and Low-Grade Inflammation: Helicobacter pylori Infection as a Model" International Journal of Molecular Sciences 22, no. 7: 3706. https://doi.org/10.3390/ijms22073706

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