Abstract

Chronic inflammation is a component of numerous diseases including autoimmune, metabolic, neurodegenerative, and cancer. The discovery and characterization of specialized pro-resolving mediators (SPMs) critical to the resolution of inflammation, and their cognate G protein-coupled receptors (GPCRs) has led to a significant increase in the understanding of this physiological process. Approximately 20 ligands, including lipoxins, resolvins, maresins, and protectins, and 6 receptors (FPR2/ALX, GPR32, GPR18, chemerin1, BLT1, and GPR37) have been identified highlighting the complex and multilayered nature of resolution. Therapeutic efforts in targeting these receptors have proved challenging, with very few ligands apparently progressing through to preclinical or clinical development. To date, some knowledge gaps remain in the understanding of how the activation of these receptors, and their downstream signaling, results in efficient resolution via apoptosis, phagocytosis, and efferocytosis of polymorphonuclear leukocytes (mainly neutrophils) and macrophages. SPMs bind and activate multiple receptors (ligand poly-pharmacology), while most receptors are activated by multiple ligands (receptor pleiotropy). In addition, allosteric binding sites have been identified signifying the capacity of more than one ligand to bind simultaneously. These fundamental characteristics of SPM receptors enable alternative targeting strategies to be considered, including biased signaling and allosteric modulation. This review describes those ligands and receptors involved in the resolution of inflammation, and highlights the most recent clinical trial results. Furthermore, we describe alternative mechanisms by which these SPM receptors could be targeted, paving the way for the identification of new therapeutics, perhaps with greater efficacy and fidelity.
Keywords: specialized pro-resolving mediators, G protein-coupled receptor, biased signaling, allostery, resolution of inflammation
Excessive inflammation has been widely recognized as a ubiquitous component of various chronic diseases including asthma, cardiovascular, neurodegenerative and many autoimmune diseases, including rheumatoid arthritis, systemic lupus erythematosus (SLE), and chronic granulomatous disease (CGD). More recently, a link between inflammation and metabolic diseases, including obesity, nonalcoholic steatohepatitis, and fibrosis has emerged.1,2 Defined by the four cardinal characteristics of rubor (redness), calor (warmth), tumor (swelling), and dolor (pain), acute inflammation is the body’s natural defense response intended to provide protection from injury and external pathogens.3−5 However, evidence strongly suggests that uncontrolled chronic inflammation leads to the progression of significant pathophysiology.6
Under healthy conditions, acute inflammatory responses are self-limited and resolve independently by the complex and collaborative actions of immune cells and diverse cellular mediators. Chronic inflammation is ultimately the result of the imbalance between the inflammatory response and the pro-resolving activity. The severity of the outcome of acute inflammation is heavily dependent on the efficacy of resolution.4 Indeed, it is suggested that chronic inflammation may be a result of “frustrated resolution” where the initial acute inflammation is not adequately resolved, leading to a defective immune response.3,7
Traditionally, conventional anti-inflammatory therapies have targeted a reduction, or nullification, of the inflammatory response, but they are typically associated with many undesired side effects.8 For example, nonsteroidal anti-inflammatory drugs (NSAIDs) and selective COX-2 inhibitors cause gastrointestinal complications and renal toxicity.9 Furthermore, some anti-inflammatory drugs require close and extensive monitoring due to their severe immunosuppressive effects, increasing the patient’s risk to opportunistic infections.7,10 Therefore, current anti-inflammatory therapies leave an unmet medical need for the treatment of chronic inflammatory diseases.
With the advanced understanding of inflammation and its processes, pro-resolving strategies, including selectively targeting the G protein-coupled receptor (GPCR) upon which the specialized pro-resolving mediators (SPMs) exert their effects, have been proposed as a new way in targeting chronic inflammation. Here we will discuss alternative mechanisms in targeting these SPM receptors, including the potential of biased agonism and allosteric modulation, which may provide improved efficacy, resulting in greater patient outcomes.
Resolution of Inflammation
Resolution marks the period between clearance of the injurious agent and dead polymorphonuclear leukocytes (PMNs), culminating in the return to homeostasis.3 Traditionally, the period of resolution was postulated to be passive, but now it is appreciated as a complex and active process that is tightly regulated by a wide range of cellular mediators (Figure 1).6,11−13 Once stimulus is removed, various regulatory mechanisms drive the innate immune system to dampen the production of pro-inflammatory mediators, including cytokines, chemokines, eicosanoids, and cell adhesion molecules, preventing interactions with their target receptors. For example, the CXC family of chemokines, which direct neutrophil migration toward the site of inflammation, is cleaved by matrix metalloproteinases (MMP) to end further influx of neutrophils, and thus prevents further unwanted tissue damage.14,15
Figure 1.
Resolution of inflammation. Following insult, injury or infection acute inflammation develops. Edema, followed by polymorphonuclear neutrophils (PMN) infiltration, occurs within minutes to hours, closely followed by the resolution of inflammation by monocytes and macrophages over hours to days. Specialized pro-resolving mediators (SPMs), including lipoxins (LXA4), resolvins (D- and E-series), maresins, and protectins, are biosynthesized to facilitate resolution. Figure generated from our interpretation of multiple reviews and research articles.3,7,35,37,106,109
Efficient neutrophil apoptosis and their clearance by surrounding phagocytes are regarded as the most crucial step in resolution.16 Apoptosis is defined as programmed cell death, intended to prevent the neutrophil from secreting its cytotoxic contents such as reactive oxygen species (ROS) and proteases into the extracellular environment.17,18 Apoptotic neutrophils undergo significant morphological changes including membrane blebbing and cellular shrinkage, resulting in cell detachment and organelle fragmentation.19 Apoptotic neutrophils also secrete various chemoattractants that act as “find-me” signals, including CX3C-chemokine ligand-1 (CX3CL1) and intracellular adhesion molecule 3 (ICAM3).19,20 Furthermore, downregulation of cell-surface molecules such as cluster of differentiation (CD)31, CD46, and CD47 act as “do-not-eat-me” signals, while expressing various “eat-me” signals, including phosphatidylserine (PtdSer) and calreticulin (CRT), that facilitate engulfment by phagocytes and robust efferocytosis.17,21
PtdSer is recognized by membrane receptors, including brain-specific angiogenesis inhibitor (BAI1), stabilin 2, and T-cell immunoglobin mucin domain protein family members (TIM1, TIM3, and TIM4).22−25 Recognition of PtdSer by BAI1 triggers the receptor to signal through ELMO1-DOCK180-RAC complex causing cytoskeletal rearrangement.26 Furthermore, activated stabilin 2 interacts with phosphotyrosine-binding (PTB)-domain-containing engulfment adaptor protein 1 (GULP) and thymosin-β4.27,28 Additionally, various molecules including serum-component milk fat globule-epidermal growth factor 8 (MFG-E8) and protein S, act as “bridging molecules” between PtdSer and efferocytosis receptors including avβ3 integrin, Mer tyrosine kinase (MerTK), and Tryo3-Axl-Mer (TAM).29,30
The expression of CRT is detected by low-density lipoprotein (LDL)-receptor-related protein (CD91) further aiding efferocytosis.31 Phenotype switching of phagocytosing macrophages to “pro-resolution” is initiated by the secretion of anti-inflammatory cytokines interleukin-10 (IL-10) and transforming growth factor-beta (TGF-β) and the downregulation of pro-inflammatory cytokine and chemokine release. These changes culminate in the promotion of tissue repair and wound healing.32,33 Macrophages involved in these processes are removed from the site of inflammation by lymphatic drainage or apoptosis.34 Ultimately, it is this synergistic cooperativity between the different mechanisms that results in efficient efferocytosis.16
In addition to the process described above, the resolution of inflammation is further coordinated by specialized pro-resolving mediators (SPMs). These SPMs promote diverse pro-resolving and anti-inflammatory mechanisms such as assisting in the reduction of PMN migration and the secretion of tissue repair and wound healing molecules.7,35
Specialized Pro-Resolving Mediators
SPMs are lipid mediators that are actively biosynthesised from essential polyunsaturated fatty acids (PUFA) during inflammation (Figure 2).4,36 For comprehensives reviews, including structural elucidation, see refs (37) and (38).
Figure 2.
Specialized pro-resolving mediators (SPMs). Overview of lipoxins, resolvins, maresins, and protectins biosynthesized from omega-3 polyunsaturated fatty acids (PUFA) including docosahexaenoic acid (DHA), eicosapentaenoic acid (EPA), and arachidonic acid (AA).12,13,45,46,49,51−53,55,58,59,63
Lipoxins
Lipoxin A4 (LXA4; 5S,6R,15S-trihydroxy-7E,9E,11Z,13E-eicosatetraenoic acid) and lipoxin B4 (LXB4; 5S,14R,15S-trihydroxy-6E,8Z,10E,12E-eicosatetraenoic acid) were the first lipid SPMs identified.37 They are synthesized from omega-6 (ω-6) arachidonic acid (AA) by lipoxygenase (LOX) conversion and further converted by oxidation of 15-hydroxyperoxyeicosatetraenoic acid (15-HpETE).37 Two routes of LXA4 formation exist, one involving 12-LOX during peripheral blood platelet–leukocyte interactions,38,39 while the second requires a series of LOX that are dependent on the cell type. The conversion is mediated by 5-LOX in neutrophils and 15-LOX in erythrocytes and reticulocytes.37,40,41 Additionally, exogenous aspirin, often used prophylactically in the treatment of inflammation, leads to formation of “aspirin-triggered” (AT) lipoxins, named 15-epi-LX, where acetylation of cyclooxygenase-2 (COX-2) by aspirin leads to the enzyme converting AA into 15R-HETE, serving as a substrate for 5-LOX.42 It is perhaps through this process which highlights aspirin’s significant pro-resolution effects compared to other NSAIDs.4,7 It is interesting to note that statins which are widely used as potent cholesterol-lowering drugs, potentially contribute to the formation of 15-epi-LX by promoting the arachidonate conversion to 15-epi-lipoxin, justifying the clinical benefits and anti-inflammatory effects of statins and aspirin in the treatment of cardiovascular disorders.43In vivo biosynthesis of LXA4 is triggered during the acute inflammation process where the interaction of PMNs with prostaglandin E2 (PGE2) and PGD2 activates 15-lipoxygenase to biosynthesise LXA4.44 In a murine model of peritonitis, the maximal level of LXA4 is typically achieved by 2 h and declines within the first 24 h, corresponding with the increase in PMN infiltration.11
E- and D-Series Resolvins
Resolvins are lipid mediators that are biosynthesized from ω-3 fatty acids, including eicosapentaenoic acid (EPA) and docosahexaenoic acid (DHA), derived from PUFA via enzymatic transformation, resulting in the production of E- and D-series resolvins, respectively.8
E-Series Resolvins
In hypoxic endothelial cells, aspirin-acetylated COX-2 oxygenates 18R-hydro(peroxy)-eicosapentaenoic acid (18R-HEPE). Activated PMNs convert this using 5-LOX into 5S(6)-epoxy-18R-HEPE, which is further hydrolyzed to produce RvE1 (5S,12R,18R-trihydroxy-6Z,8E,10E,14Z,16E-eicosapentaenoic acid).12,13 RvE2 (5S,18R-dihydroxy-6E,8Z,11Z,14Z,16E-eicosapentaenoic acid) is produced by the reduction of 5-LOX and 18R HEPE product, 5S-hydroperoxy,18-hydroxy-EPE in whole blood, while the production by PMNs is enhanced during hypoxia.45 In contrast to RvE1 and RvE2, RvE3 (17R,18S-dihydroxy-5Z,8Z,11Z,13E,15E-eicosapentaenoic acid) is biosynthesised from 18-HEPE via the 12/15-LOX pathway in eosinophils.46 Notably, EPA-derived 18-HEPE and 15-HEPE biosynthesised in hypoxic endothelial cells produce resolution effects by reducing the migration of PMNs, although the activity is much weaker in comparison to RvE1.13 As unacetylated COX-2 are not able to convert and produce 18-HEPE, aspirin, paracetamol, and indomethacin all trigger 18-HEPE production,13 whereas selective COX-2 inhibitors prevent this production. In inflammatory exudates, the precursor 18-HEPE EPA was shown to reach its peak level by 2 h and subsided to base levels within 12 h; however, endogenous RvE1 appeared at a delayed time point, accumulating between 48 and 72 h.47 It is postulated that this result may be dependent on the cellular composition of the exudates utilized; as RvE1 appeared at an earlier time point in exudates of murine skin air pouch model,13 perhaps due to the difference in permeability between peritoneum and skin. RvE2 also appeared in murine peritoneal exudates, stimulated by zymosan A, with a corresponding time point to initial PMN infiltration and declining within 24 h.48
D-Series Resolvins
D-Series resolvins (RvD1-RvD6) are biosynthetically produced from DHA by the action of LOX in PMNs and macrophages.12 DHA is converted to a peroxide intermediate following two LOX steps. Hydrolysis of this product generates RvD1 (7S,8R,17S-trihydroxy-4Z,9E,11E,13Z,15E,19Z-docosahexaenoic acid) and RvD2 (7S,16R,17S-trihydroxy-4Z,8E,10Z,12E,14E,19Z-docosahexaenoic acid), while reduction leads to the production of RvD5 (7S,17S-dihydroxy-4,8,10,13,15,19-docosahexaenoic acid).12,49,50 In the presence of aspirin, COX-2 in hypoxic endothelial cells transforms DHA from PUFA to 13-hydroxy-DHA or 17R-HDHA.12 Activated PMNs further metabolizes these products to AT-RvD1, AT-RvD2 and other AT-D-series resolvins. RvD3 (4S,11R,17S-trihydroxy-5Z,7E,9E,13Z,15E,19Z-docosahexaenoic acid)12,51,52 and RvD4 (4S,5R,17S-trihydroxy-6E,8E,10Z,13Z,15E,19Z-docosahexaenoic acid)53 originate from 4S,5S-epoxy-17S-hydroxy-docosahexaenoic acid, following hydrolysis of 4S-hydroperoxy-17S-hydroxy-docosahexaenoic acid, while RvD6 originates from peroxidase of the same precursor. Interestingly, the in vivo appearance of RvD3 following zymosan challenge in a peritonitis model is significantly increased to peak 48 h postinitiation, while RvD1, RvD2, and RvD5 peak earlier in the resolution phase (∼6–24 h).52 RvD3 is most likely, therefore, produced by a subpopulation of macrophages, high in 15-LOX.52,54In vivo production of RvD4, by an S. aureus dorsal pouch infectious model, appears even later with a sustained release over 72 h, indicating constitutive production and differential regulation.53 RvD6 kinetics remain to be elucidated.
Protectins
Similar to maresins, the biosynthesis of protectins also arises from the formation of epoxide intermediates of DHA. Oxygenation of DHA into 17S-HpDHA via 15-LOX is converted to produce 16S,17S-epoxy-protectin where it is further hydrolyzed into protectin D1 (PD1, 10R,17S-dihydroxy-4Z,7Z,11E,13E,15Z,19Z-docosahexaenoic acid.55 PD1 was found to accumulate in the ipsilateral hemisphere of the brain after focal ischemia resulting in enhanced wound healing and neuroprotection. Due to the location of expression, brain-specific PD1 is known as neuroprotectin D1 (NPD1).56,57 15-LOX action on 17S-HpDHA can also produce 10S, 17S-diHDHA (PDx), an isomer of PD1, which also produces pro-resolving activity. In zymosan A-induced murine peritonitis model, the in vivo production of PD1 was shown to peak at 12 h.11
Maresins
Maresins, or macrophage mediators in resolving inflammation, are a fourth family of SPMs.58 These SPMs are biosynthesised from DHA by macrophages via the action of 12-LOX,59 which catalyzes the oxygenation of DHA into 14-hydroperoxydocosahexaenoic acid (14-HpDHA), followed by the reduction to 13S,14S-epoxy-maresin.49 The 13S-14S-epoxy-maresin is further converted to produce MaR1 (7R,14S-dihydroxy-4Z,8E,10E,12Z,16Z,19Z-docosahexaenoic acid) within human macrophages.60 The complete stereochemistry of MaR1 was confirmed as 7R,14S-dihydroxy-docosa-4Z,8E,10E,12Z,16Z,19Z-hexaenoic acid.61 Conversion of 13S-14S-epoxy-maresin by a soluble epoxide hydrolase results in the formation of MaR2 (13R,14S-dihydroxy-4Z,7Z,9E,11E,16Z,19Z-docosahexaenoic acid).58 The critical role of 12-LOX in the biosynthesis of maresins was demonstrated where 12/15-LOX-deficient mice showed a significant reduction in 14S-HDHA generation in a mouse zymosan A induced model of peritonitis.58In vivo production of MaR1 was discovered utilizing a murine model of acute respiratory distress syndrome, where its early production was detected during platelet–neutrophil interaction, where its level was significantly enhanced within 2 h of reaching a peak at 24 h.62
SPMs Mediate Their Effects through G-Protein-Coupled Receptors (GPCRs)
Extensive studies have demonstrated the pro-resolving activities of SPMs occur as a result of activation of one or more of a family of GPCRs. To date, five key receptors have been identified: FPR2/ALX, chemerin1, and BLT1, and two still classified as orphan receptors, GPR32 and GPR18. Recently another orphan receptor, GPR37, has been suggested as the receptor for PD1.64 Interestingly, more than 140 GPCRs still remain to be paired with their endogenous ligand, indicating the potential for the identification of yet further SPM receptors.65
FPR2/ALX
N-Formyl peptide receptor 2/lipoxin A4 receptor (FPR2/ALX) is predominantly expressed on leukocytes, including neutrophils, monocytes, macrophages, dendritic cells (DC), naïve CD4 T-cells, and Th1, -2, -17 cells.66−69 FPR2/ALX is also expressed on nonimmune cells including endothelial, keratinocytes, and mesenchymal stem cells.66 Furthermore, FPR2/ALX has been identified within the central nervous system, including the hippocampus and prefrontal cortex, indicating a possible role in learning, memory, and balance.70 The surface expression of FPR2/ALX is significantly increased when pro-inflammatory stimuli are present, including tumor necrosis factor-α (TNF-α), platelet-activating factor (PAF), and IL-8.71 Orthologs of FPR2/ALX are present in both human and rodents.72,73 Endogenous and exogenous lipids, peptides, and proteins have been discovered to bind to and activate FPR2/ALX, producing both pro- and anti-inflammatory effects (Table 1).74,75 SPMs from both the lipoxin and resolvin families activate the receptor with high potency, including LXA4, AT-LXA4 (15-epi-LXA4), RvD1, AT-RvD1 (17-epi-RvD1), and annexin A1 (ANXA1). Furthermore, endogenous antagonists for the receptor have also been identified, including serum amyloid A (SAA) and cathelicidin (LL-37; Table 1 and Figure 3).
Table 1. Summary of Published GPCRs and Paired SPMs/Associated Ligands Involved in the Resolution of Inflammationa.
| receptor | expression profile | species ortholog | G-protein coupling and/or β-arrestin | pro-resolving mediators | pKD/pEC50 | refs |
|---|---|---|---|---|---|---|
| FPR2/ALX (ALX, LXA4R, FPRL1, FMLPX, formyl peptide receptor 2) | neutrophils, monocytes, macrophages, immature dendritic cells, T- and B-cells, epithelial cells, microglial cells, astrocytes, hepatocytes, endothelial cells, spleen, lung, kidney, leukocytes | human, mouse and rat | Gi/Go Gq/G11 β-arrestin | LXA4 | 8.8–9.3/12 | (79, 95) |
| AT-LXA4 | (93) | |||||
| RvD1 | 11.9 | (95) | ||||
| AT-RvD1 | 11.1 | (95) | ||||
| ANXA1 | 6.5/5.8–6.1 | (157) | ||||
| Ac2-26 | <6 | (107) | ||||
| SAA | 6.6 | (158) | ||||
| LL-37 | 6.0 | (159) | ||||
| GPR32 (DRV1, RVDR1, resolving D1 receptor) | myeloid cells, human umbilical vein cells, arterial tissue, venous tissue | human | not determined β-arrestin | LXA4 | 9.7 | (95) |
| RvD1 | 11.1 | (95, 100) | ||||
| RvD3 | ∼11 | (53) | ||||
| AT-RvD3 | ∼11 | (53) | ||||
| RvD5 | ∼11 | (111) | ||||
| GPR18 (NAGly receptor) | PMNs, monocytes, macrophages, microglia, BV2 microglial cells | human and mouse | Gi/Go Gq/G11 Gs β-arrestin | NAGly | 8 | (114, 117, 119, 120) |
| RvD2 | 8 | |||||
| Abn-CBD | ||||||
| O-1918 | antagonist | |||||
| chemerin1 (ChemR23, ERV1, RVER1, CMKLR1) | monocytes, macrophages, NK cells, myeloid cells, plasmacytoid dendritic cells, primary adipocytes, endothelial cells, circulating dendritic cells | human and mouse | Gi/Go | Chemerin | 8.3 | (64) |
| RvE1 | 7.9 | (64, 140) | ||||
| RvE2 | 10 | (50) | ||||
| BLT1 (LTB4R) | leukocytes, spleen, thymus, mast cells, monocytes, T- and B-lymphocytes, dendritic cells, endothelial cells | human, mouse, and rat | Gi/Go Gq/G11 β-arrestin | LTB4 | 9.4 | (142) |
| 12R-HETE | 7.5 | (142) | ||||
| RvE1 | 7.35 | (140) | ||||
| RvE2 | 10 | (50, 47) | ||||
| GPR37 (EDNRBL, EDNRLB) | macrophages | human and mouse | Gi/Go | NPD1 (PD1) | 7.6 | (65) |
Nomenclature and values as described in “The Concise Guide to Pharmacology 2017/18”.121
Figure 3.
GPCRs activated by multiple subpopulation pro-resolving mediators. Pairing of SPMs and associate ligands to their cognate receptors have been much explored utilizing various methods including radioligand binding. Early research into SPM pharmacology focused primarily on the ligand activity; however, recent attention has shifted toward further elucidation of the signaling mechanism involved. To date, knowledge gaps remain in the signal transduction pathways leading to increasingly well-described physiological responses. Due to ligand poly-pharmacology and receptor pleiotropy, distinct opportunities to target these receptors via novel mechanisms exist, including biased signaling and allostery. Both mechanisms have been partially described for FPR2/ALX, but due to the vast number of SPMs and GPCRs identified, the combinations could be limitless, clearly highlighting significant opportunities in augmenting the resolution of inflammation. Figure generated from our interpretation of multiple reviews.63,89,95,113,130,140,160
Originally, FPR2 was classified as an FPR receptor due to its activation by the low-affinity endogenous agonist N-formyl-methionyl peptide (fMLP).76,77 Screening of various receptors using radiolabeled [3H]-LXA4 and subsequent GTPase activity, cells transfected with FPR2/ALX cDNA (pINF154) displayed specific and high-affinity LXA4 binding.78 Therefore, the receptor was reclassified as FPR2/ALX as LXA4 displayed the highest affinity of all FPR2/ALX endogenous agonists.79 However, LXA4 binds to multiple receptors including GPR32, FPR3, and the nuclear estrogen receptor.80 Furthermore, LXA4 binds to the transcription factor aryl hydrocarbon receptor (AHR) to induce gene expression of many cytochrome P450 enzymes.81
Binding of LXA4 induces stimulation of monocyte chemotaxis, macrophage differentiation, and efferocytosis.75,82 LXA4 also reduces the adaptive immune response through decreasing memory B-cell antibody production and proliferation.83 LXA4 acts as positive feedback for FPR2/ALX expression by activation of the receptor promotor, thus upregulating FPR2/ALX expression level.84 Various knockout mouse models have been used to demonstrate the interaction of LXA4 with FPR2/ALX. Treatment of Fpr2/Alx–/– mice with LXA4 resulted in the absence of cell recruitment into dorsal air pouches inflamed with IL-1β, while LXA4-induced resolution of ischemia-reperfusion (IR) injury was reduced in these knockout mice.85,86 Furthermore, isolated Fpr2/Alx–/– macrophages failed to effectively clear apoptotic PMNs in response to LXA4.75 Finally, chemical inhibition of LXA4-mediated resolution in a mouse model of pneumosepsis was shown with treatment of an FPR2/ALX antagonist BOC-2.87 However, alternative FPR2/ALX agonists were not investigated.
Interestingly, LXA4 binds to the seventh transmembrane (TM) domain and third extracellular loop (EL) of FPR2/ALX.88 Due to this unusual binding location of LXA4 compared to that of agonist-bound crystal structures of many GPCRs where the orthosteric pocket typically resides deep within TM 3 and 5, we might infer that FPR2/ALX is capable of binding multiple ligands simultaneously. In fact, during inflammation, the pro-inflammatory peptide ligand SAA is released and binds to FPR2/ALX to activate ERK1/2 phosphorylation and phosphoinositide 3-kinase (PI3K), while increasing NF-κB expression, resulting in enhanced PMN activation, differentiation, motility, and survival.89 SAA binds in a distinct location to LXA4 and is dependent on interactions with EL1 and 2.89 When cobound LXA4 acts as an allosteric inhibitor of SAA signaling, it thus reduces SAA-mediated PMN activation and survival.90,91 The cooperativity between allosteric interactions is always reciprocal; therefore, binding of SAA inhibits LXA4 affinity and signaling.88
In addition to LXA4, the aspirin-triggered version, AT-LXA4, enhances resolution by downregulating the expression of β2 integrin Mac-1 (CD11b/CD18), attenuating the signaling of myeloperoxidase (MPO) and suppressing human neutrophil apoptosis.92 AT-LXA4 treatment increased phagocytosis of E. coli in a PI3K- and scavenger-receptor-dependent manner, while the FPR2/ALX gene was 6-fold upregulated in colonic biopsies of Crohn’s disease patients.93
RvD1 binds and activates FPR2/ALX with similar affinity and potency as LXA4 (Table 1).94 Binding of RvD1 results in zymosan-mediated macrophage phagocytosis and cyclic adenosine monophosphate (cAMP)/protein kinase A (PKA) signaling and increases dihydronicotinamide-adenine dinucleotide phosphate (NADPH) oxidase activation while decreasing antiapoptotic protein expression.95 Activation of FPR2/ALX by RvD1 also results in the regulation of specific microRNAs (miRNAs).4,96,97 For example, RvD1 selectively upregulates miR-208a in human macrophages overexpressing FPR2/ALX, which culminates in the downregulation of the suppressor of IL-10, programmed cell death protein 4 (PDCD4), therefore increasing IL-10 production.96 In Fpr2/Alx overexpressing mice with peritonitis, RvD1 resulted in potent and anti-inflammatory effects that were attenuated in Fpr2/Alx–/– mice. Furthermore, PMN number and levels of pro-inflammatory mediators were significantly reduced in the knockout animals.71,96 AT-RvD1 treatment also reduced leukocyte recruitment, neutrophil–platelet interaction, and downregulation of pro-inflammatory mediators, including IL-1β, IL-6, IL-8, and TNF-α.98 Interestingly, AT-RvD1 (and the AT versions of RvD3 and LXA4) inhibit tumor growth in an aggressive murine Lewis lung carcinoma (LLC) model.99
Annexin A1 (ANXA1) is a glucocorticoid-regulated protein with high abundance in myeloid cells and is secreted by apoptotic neutrophils to limit their own recruitment.15,100 ANXA1 activates FPR2/ALX with low nanomolar potency, activating downstream signaling of mitogen-activated protein kinase (MAPK) and extracellular receptor kinase (ERK) phosphorylation101 (Table 1). Binding of ANXA1 leads to enhanced neutrophil apoptosis and efferocytosis of these apoptotic neutrophils by macrophages.100,102 Furthermore, ANXA1 modulates neutrophil adhesion and migration thereby selectively inhibiting PMN adhesion to the endothelium and recruitment.103,104
There is some evidence to suggest that FPR2/ALX can form constitutive homodimers in a ligand-dependent manner.74,105 For example, binding of LXA4 and ANXA1-derived peptide Ac2-26 leads to activation of p38MAPK pathway, enhancing IL-10 production.74,106 Conversely, the binding of SAA reduces the formation of these homodimers.7 Evidence suggests that Ac2-26 stimulates receptor heterodimerization of FPR2/ALX with the closely related receptor, FPR1, resulting in the activation of the Jun N-terminal kinase (JNK)–caspase-3 pathway to promote PMN apoptosis.74,91,106
GPR32
GPR32 is predominantly expressed on human PMNs, monocytes, adipose tissue, and vascular endothelial cells.94,107 GPR32 is currently classified as an orphan receptor. It shares 29% sequence identity with FPR2/ALX and is the closest phylogenetically (Table 2 and Figure 4). RvD1 was identified as a potential agonist as its activation of GPR32, where [3H]-RvD1 bound to human leukocytes, and via screening of series of receptors involved in inflammation, RvD1 significantly lowered the TNF-α stimulated NF-αB signaling in GPR32-overexpressing cells.94 Despite the fact that RvD1 shows a higher affinity for GPR32 than FPR2/ALX, the interaction with GPR32 has not been extensively studied71 (Table 1 and Figure 3). Crucially, GPR32 exists as a pseudogene within rodents; therefore, any effect of RvD1 exhibited in animal studies can be surmised to be solely via FPR2/ALX.107,108FPR2/ALX expression is downregulated in tissue-resident human macrophages indicating that RvD1 may act via GPR32 on these cells.109 Furthermore, treatment of GPR32 expressing activated pro-inflammatory macrophages with RvD1 resulted in a change of phenotype to pro-resolving with increased phagocytosis and reduced secretion of pro-inflammatory cytokines.107 Overexpression of GPR32 on human macrophages enhanced the phagocytosis of zymosan in response to RvD1, whereas short hairpin RNA (shRNA) knockdown of GPR32 opposed these effects.94 Moreover, reduced PMN migration was ablated by treatment with an anti-GPR32 antibody.71 Interestingly, however, this effect was not observed with an anti-FPR2/ALX antibody. Increased RvD1 concentration did, however, reverse this effect, consistent with the higher potency of RvD1 on GPR32 versus FPR2/ALX.71 Clearly these data indicate that inhibition of GPR32 does not reduce the effect of RvD1, suggesting a level of redundancy within the system; an effect often seen in chemokine signaling. Of note, RvD3, AT-RvD3, and RvD5 have all been shown to activate GPR32 in a recombinant system of β-arrestin recruitment,52,110 further highlighting the potential redundancy. Using a high-throughput screen approach a selection of RvD1 mimetics were identified that shared the same binding pocket as RvD1 and activated both β-arrestin recruitment and cAMP.111 However, very recently, no activity was shown for the pairing of GPR32 and RvD1, indicating further investigation into this receptor/ligand pairing is warranted.112
Table 2. Percent Sequence Identity between the Six Human SPM GPCRs.
| FPR2/ALX | GPR32 | GPR18 | chemerin1 | BLT1 | GPR37 | |
|---|---|---|---|---|---|---|
| FPR2/ALX | 100 | 29 | 18 | 26 | 23 | 10 |
| GPR32 | 100 | 15 | 26 | 18 | 7 | |
| GPR18 | 100 | 17 | 15 | 8 | ||
| chemerin1 | 100 | 21 | 9 | |||
| BLT1 | 100 | 9 | ||||
| GPR37 | 100 |
Figure 4.
Phylogenetic tree of the six human SPM GPCRs. The tree was generated using iTOL.161 cml1 = chemerin1 and lt4r1 = BLT1. GPR32 was identified as a receptor for RvD1,94 RvD3, AT-RvD3, and RvD5.52,110 Sharing a close phylogenetic identity with GPR32, FPR2/ALX was initially discovered as an fMLP receptor;76,77 However, the receptor was later reclassified as a receptor for LXA4 due high affinity binding of this ligand.68 Chemerin1 has recently been reclassified following the identification of its cognate ligand, chemerin.134−136 BLT1 binds numerous ligands including LTB4, RvE1, and RvE2.50,140,142 GPR18 still remains as an orphan receptor with several endogenous and SPM ligands having been identified, including NAGly117 and RvD2,113 but yet remaining to be confirmed. GPR37 has recently been identified as a receptor for neuroprotection D1 (NPD1).64
GPR18
GPR18 is abundantly expressed on PMNs, monocytes and macrophages.113−115 GPR18 still remains an orphan receptor, although several endogenous ligands have been proposed, including N-arachidonylglycine (NAGly)116 and RvD2113 (Table 1 and Figure 3). NAGly was identified via screening of a lipid library using CHO cells overexpressing GPR18. Both increases in intracellular Ca2+ and inhibition of forskolin-induced cAMP responses were observed, suggesting both Gαi and Gαq coupling.116 However, signaling by NAGly appears to be context-dependent, i.e., dependent on which cell type or assay system is being studied. For example, in glioblastoma multiforme cell lines endogenously expressing GPR18, NAGly failed to induce ERK1/2 phosphorylation (pERK1/2).117 Moreover, confirmation of GPR18-NAGly pairing was observed in a β-arrestin assay using CHO cells overexpressing GPR18, but it was absent when expressed in human embryonic kidney 293 (HEK293) cells. As there is no supporting evidence from animal studies, interpretation of these findings is difficult.113,116,118 In addition to its role in resolution of inflammation, GPR18 also responds to endogenous and synthetic cannabinoid ligands, including N-arachidonoyl ethanolamine (AEA), 2-arachidonoyl glycerol (2-AG), Δ9-tetrahydrocannabinol (Δ9-THC), and arachidonoyl cyclopropylamide (ACPA),119 although, interestingly, GPR18 shows little structural similarity to the cannabinoid receptors CB1, CB2, and GPR55.120
RvD2 was identified as a low nanomolar agonist for GPR18 using a β-arrestin recruitment assay, which was later confirmed using [3H]-RvD2 binding assays.113,121 Perhaps surprisingly, GPR18 has been shown to couple to Gαs, as incubation of human macrophages, with RvD2, resulted in an increase in cAMP.113 Use of mass cytometry (CyTOF) RvD2-GPR18 interaction was found to involve STAT3 and PKA inhibition in RvD2-stimulated phagocytosis.122 In most cases, the GPCRs described in this review and those involved in PMN migration generally couple to Gαi/o perhaps indicting GPR18 as an atypical SPM receptor (Table 1 and Figure 3). RvD2 treated overexpressing GPR18 macrophages had increased phagocytosis of FITC-zymosan and E. coli, while GPR18–/– macrophages showed reduced function.113 RvD2 treatment increased the expression of classical markers of anti-inflammatory macrophages, CD163 and CD206, while upregulating GPR18 expression113,123 Furthermore, RvD2 activation of mouse Gpr18 resulted in reduced LPS- and ATP-stimulated inflammasomes in macrophages, which was abolished by O-1918, a GPR18/GPR55 antagonist.124
In mouse models of zymosan A induced peritonitis and sepsis; RvD2 reduced IL-1β, IL-17, and IL-10 secretion, caspase-1 activity, increased the anti-inflammatory macrophage population, and increased the survival rates of mice by restoring neutrophil movement preventing further tissue injury from neutrophils recruitment.124,125 Many of these effects were abolished in Gpr18–/– mice.113 In patients with sepsis, GPR18 was found to be significantly reduced in PMNs, indicating a role in postinflammation resolution.126 Clarification of GPR18’s role in endocannabinoid signaling and resolution is required. However, with the discovery of CB2 as a regulator of inflammation,127 GPR18 may appear a worthwhile target, especially as an alternative to developing selective cannabinoid agonists.
Chemerin Receptor 1
Chemerin receptor 1 (chemerin1, ChemR23, or ERV1) is expressed on a wide range of immune cells including monocytes, macrophages, NK cells, myeloid cells, and DCs. Furthermore, these receptors have been identified on adipocytes and endothelial cells.128−130 The expression of chemerin1 is more abundant on antigen-presenting cells (APCs) such as macrophages and DCs rather than neutrophils and T-lymphocytes,63,131 but its expression was shown to be upregulated in neutrophils in diseases including type 2 diabetes (T2D).132 Receptor expression is also upregulated by cytokines and toll-like receptor ligands during monocyte differentiation, while it has been suggested that the expression and function of chemerin1 are restricted to pro-inflammatory macrophages.112 Chemerin1 was initially classified as an orphan GPCR with homology with the formyl peptide receptors133 and anaphylatoxin C3a and C5a receptors,131 until it was recently paired with the chemotactic protein chemerin (Table 1 and Figure 3).134−136 Activation of chemerin1 by chemerin results in Gαi/o coupling, leading to the release of intracellular calcium, inhibition of cAMP-, MAPK-, and ERK1/2-mediated signaling.130 Receptor activation also results in the upregulation of phosphatidylinositol (3,4,5)-trisphosphate (PIP3), protein kinase B (Akt), and downregulation of NF-κB signaling.130
In addition to chemerin, RvE1 was identified as a second endogenous agonist via a screening campaign against a panel of GPCRs (Table 1 and Figure 3).63 [3H]-RvE1 was shown to bind to chemerin1 as competition binding with chemerin revealed complete displacement in membranes prepared from human PMNs. The anti-inflammatory effects of RvE1 were identified by chemerin1-mediated attenuation of TNF-α-induced NF-κB. Treatment of macrophages with RvE1 reduced IL-12 levels and DC migration.63 Furthermore, use of the MAPK inhibitor, PD98059, resulted in decreased RvE1-induced phosphorylation signaling and macrophage phagocytosis of apoptotic PMN.137 Chemerin1-overexpressing mice displayed decreased neutrophil inhibition and a significant increase in phagocytosis while reducing neutrophil infiltration.138,139
However, some controversy exists with the pairing of RvE1 and chemerin1. It has been suggested that due to the low expression of chemerin1 on neutrophils the infiltration and apoptosis of these cells must be driven by an alternative mechanism.128 However, this is questionable as other findings revealed the expression level of chemerin1 was found to be markedly enhanced on human neutrophils during inflammation, including poorly controlled T2D.132 Like many resolvins, RvE1 binds to multiple receptors, including the leukotriene receptor, BLT1, albeit with 100-fold lower affinity (Table 1). In Blt1–/– mice RvE1-mediated inhibition of neutrophil recruitment was significantly reduced in zymosan-induced peritonitis.128,140 However, these findings were further complicated as recent data revealed RvE1-enhanced phagocytosis by macrophages was abolished in chemerin1–/– mice.141 Furthermore, RvE2 was shown to be a partial agonist compared to RvE1 in a CHO-chemerin1 β-arrestin recruitment assay.48 However, no further information is available for this ligand, thus warranting further investigation into this potential ligand–receptor pairing.
Leukotriene BLT1
The discovery of BLT1 embarked when its cDNA was cloned as a high-affinity leukotriene B4 (LTB4) receptor using subtraction strategy (Table 1 and Figure 3),142 where LTB4 was shown to induce an increase in intracellular Ca2+. LTB4 is a potent lipid pro-inflammatory chemoattractant and, via BLT1, induces T helper cell chemotaxis and early effector T-cell recruitment.143 BLT1 shares 21% sequence identity with chemerin1 (Table 2 and Figure 4), and although this value is relatively low, BLT1 was also revealed to be a receptor for RvE1, where binding of [3H]-RvE1 to human PMN membranes was displaced by the selective BLT1 antagonist, U-75302. Furthermore, RvE1 inhibited adenylate cyclase activity and induced intracellular calcium mobilization in HEK293 cells overexpressing BLT1, albeit with 100-fold less potency and reduced efficacy than LTB4. Nevertheless, these data suggest that RvE1 acts as a partial agonist that competes with the LTB4-mediated NF-κB activation and calcium mobilization, indicating a role of RvE1 in reducing BLT1 induced inflammation.140
Intravenous administration of RvE1 significantly reduced PMN infiltration in a zymosan A induced peritonitis model in wild-type mice, while this effect was abolished in Blt1–/– mice. However, increased concentrations of RvE1 opposed this effect indicating a BLT1-independent role,140 possibly via an alternative receptor, e.g., chemerin1. However, as the expression level of chemerin1 on PMNs is low, RvE1-mediated PMN infiltration reduction is most likely via BLT1.128,144 Alternatively, an undiscovered receptor may also be responsible for these effects. Activation of BLT1 by RvE1 also acts as a feedback mechanism for other SPMs, including increased production of LXA4 in FPR2/ALX-mediated resolution of allergic lung inflammation.145
RvE2 has been identified as an additional BLT1 agonist (Table 1 and Figure 3);45 a β-arrestin assay revealed RvE2 blocked LTB4-mediated β-arrestin signaling with a similar potency as RvE1 indicating direct competition with LTB4.48 Various pro-resolving roles of RvE2 have been suggested including regulation of PMN infiltration and IL-10 production.48 Furthermore, RvE1 promotes NADPH oxidase-mediated ROS generation via BLT1 receptors, while RvE2 is unable to replicate these data, suggesting alternative roles in regulating ROS generation by neutrophils.146
GPR37
GPR37, or parkin-associated endothelin-like receptor (Pael-R), was originally discovered through genomic library screening, in an attempt to find a novel neuropeptide receptor.147 The receptor is predominantly expressed in the brain and has been linked to neurological disorders including Parkinson’s disease and autism.148,149 Mutations within GPR37 lead to various autism spectrum disorders, modulation of dopamine reuptake, and oligodendrocyte cell differentiation.150−152 Recently the expression of GPR37 has been confirmed in mouse macrophages but not on microglia.64 Neuroprotection D1 (NPD1, PD1) has been identified as a ligand for GPR37 as it induced a significant increase in intracellular calcium in both HEK293 cells overexpressing GPR37 and mouse peritoneal macrophages (Table 1 and Figure 3).64 NPD1 also induced an increase in peritoneal macrophage phagocytosis of zymosan particles, an effect diminished in Gpr37–/– mouse peritoneal macrophages. Gpr37–/– mice also showed enhanced zymosan-induced upregulation of pro-inflammatory cytokines IL-1β and reduced levels of anti-inflammatory cytokines in inflamed skin.64 More recently, GPR37 has also been suggested to be involved in the protection against cell damage and inflammation after ischemic stroke, as Gpr37–/– mice exhibited enhanced cell death and infarct size.153 However, due to the clear role within the central nervous system (CNS), GPR37 may prove a less than desirable therapeutic target.
Activation of Other Receptors by SPMs
Apart from the GPCRs described above, many SPMs activate other receptors including non-GPCRs, for example nuclear receptors. NPD1 activates peroxisome proliferator-activated receptor gamma (PPARγ) where human neuronal-glial (HNG) cells cotransfected with hPPARγ-GAL4 and MH100-tk-luc enhanced the PPARγ transactivation reporter activity in a dose-dependent manner.154 Transcriptional activity of PPARγ was also significantly increased following treatment with 100 nM NPD1.155 RvD1 has been postulated as a ligand for PPARγ where its suppression of IκBα degradation and NF-κB p65 nuclear translocation were partially reversed by the PPARγ inhibitor GW9662 in an LPS-induced actute lung injury model.156 However, this pairing requires further study as RvD1 and RvE1 failed to activate the nuclear receptors including PPARγ and retinoid X receptor-α when coexpressed in HEK293 cells linked to Gal4 DNA-binding domain.94
Development and Clinical Evaluation of SPMs and Synthetic Agonists
SPMs are highly potent and efficacious endogenous ligands that have shown great potential in preclinical animal models. However, due to their delicate physicochemical nature being prone to metabolic inactivation and the complex chemical routes required to synthesize them,49,162,163 developing stable synthetic analogs to replicate the pro-resolving effects has been of interest. While some ligands have proved useful via topical application very few have progressed to clinical trials. RX-10045 (Resolvyx and Celtic Therapeutics), a synthetic analog of RvE1, recently produced dose-dependent and statistically significant improvements in the treatment of dry-eye disease.164 Furthermore, a 0.1% nanomicellar solution of the compound recently completed a phase 2 study for ocular inflammation and pain in cataract surgery, although no results have yet been disclosed (clinicaltrials.gov; NCT02329743). However, RX-10045 is a strong inhibitor of organic cation transporter-1 (OCT-1), suggesting novel formulation strategies may be required to achieve improved clinical efficacy.165 A study using Lipinova, a combination of three DHA metabolites (17-HDHA, 18-HEPE, and 14-HDHA), to investigate the effects of inflammation on patients following orthopedic surgery is currently on hold following recruitment (clinicaltrials.gov; NCT03434236). Finally, a synthetic benzo-RvD1 analog, benzo-diacetylenic-17R-RvD1 methyl ester (BDA-RvD1), showed significantly high potency in reducing neutrophil infiltration in ischemia-reperfusion-initiated second organ injury and stimulating the phagocytosis of zymosan A particles.166
Various stable analogs of LXA4 have also shown promise in trials. Topical 15(R/S)-methyl-LXA4 significantly reduced the severity of eczema with an efficacy similar to topical corticosteroid, Mometasone furoate (Eloson).167 Inhaled 5(S),6(R)-LXA4 methyl ester and BML-111, an LXA4 agonist, displayed efficacy and safety in a pilot study of asthmatic children with acute moderate episodes, although efficacy was reduced compared to that of the standard of care (salbutamol). Nevertheless, these data indicate an inhaled LXA4 analog in combination with an LXA4 receptor agonist may exhibit a novel therapeutic strategy for asthma.168 Benzo-LXA4 analogs have also portrayed diverse pro-resolving properties, where it was shown to reduce PMN infiltration and pro-inflammatory cytokine generation,163 suppress renal fibrosis,169 and attenuate obesity-induced adipose inflammation while restoring the ratio of pro-/anti-inflammatory macrophages.170 Finally, a benzo-fused ring-modified LXA4 analog (BLXA4-ME) is currently being assessed as a topical oral rinse treatment of gingivitis (clinicaltrials.gov; NCT02342691).
In addition to these LXA4 analogs, whose effects are mostly via FPR2/ALX, extensive drug discovery efforts have been ongoing to develop selective and potent FPR2/ALX agonists. One of the first reported FPR2/ALX synthetic agonists was AR234245 from Arena Pharmaceuticals in 2004. Since then multiple pharmaceutical and academic groups have developed FPR2/ALX agonists, including Acadia Pharmaceuticals, Allergan, Amgen, and Daiichi Sankyo, many of which share a common aniline moiety, substituted at the para position; for a comprehensive view on these medicinal chemistry efforts, see ref (106). Since these initial efforts, Actelion Pharmaceuticals have become the leaders in this field with the development of the first-in-class anti-inflammatory agonist ACT-389949. In two phase 1 randomized double-blind studies, ACT-389949 resulted in a dose-dependent, long-lasting internalization of FPR2/ALX in leukocytes. Although promising, the inhibitory effect on leukocyte number appeared to be only transient, ∼2 h postdose, with numbers reverting to baseline shortly thereafter. In addition, both pro- and anti-inflammatory cytokines were upregulated, while no effect was observed in an LPS inhalation model.171 Although safe and well-tolerated, receptor internalization clearly demonstrates the added challenges developing a potent and efficacious agonist and may explain the rapid rise in leukocytes following the initial decrease. No preclinical characterization of ACT-389949 has been reported; however, a recent study using CHO cells expressing FPR2/ALX has shown that ACT-389949 efficiently recruits β-arrestin with low nanomolar potency.172 As β-arrestin recruitment is a hallmark of receptor internalization these findings support those from the phase 1 trials.173
Current State of SPM Drug Discovery and Novel Modes of Targeting SPM GPCRs
It is clearly evident that the pharmacology of these SPM GPCRs is complex, multifaceted and in most cases incomplete. Great efforts have been made in the development of synthetic SPM analogs or small molecule agonists. However, this development has been markedly hampered by a lack of understanding of the mechanism(s) of SPM/ligand–receptor internalization, and signal transduction pathways–the potent tachyphylaxis of ACT-388949 attest to this. At present two key knowledge gaps exist. First, the greatest complexity is the identification of multiple pro-resolving mediators that act on multiple GPCRs. These ligands include not only the endogenous SPMs but also their aspirin-triggered forms. Moreover, the endogenous production of these SPMs is typically transient, which for the D-series resolvins (RvD1–RvD6) appear to be time-dependent during the resolution phase; thus, the kinetic variation in the release can lead to distinct ligand-induced signaling events.174
Second, as described above, multiple binding sites have been identified at FPR2/ALX allowing both lipids (LXA4) and peptides (SAA, LL-37, ANXA1) to bind simultaneously. This allosteric binding can lead to both positive and negative outcomes. For example, binding of SAA results in the release of IL-8 by airway epithelial cells and can be detected in the bronchoalveolar lavage fluid of chronic obstructive pulmonary disease (COPD) patients, while coaddition of LXA4 results in the inhibition of IL-8 release and decreased neutrophil recruitment.175 However, during acute exacerbations, peripheral blood SAA levels are significantly increased, therefore dampening the inhibitory effects of LXA4.175 As many SPMs bind to multiple receptors, allosteric binding sites may be present on all SPM GPCRs. In fact, many of these ligands have been demonstrated to act allosterically on a number of non-SPM GPCRs. For example, treatment of LXA4 in mouse brain led to cannabinoid-like effects, which were efficiently blocked by CB1R antagonists.176 Furthermore, the affinity of a range of cannabinoids, including anandamide, CP55940, and WIN55212-2 were increased in the presence of LXA4, demonstrating clear allosteric cooperativity.176 Similarly, the cathelicidin-related antimicrobial peptide, LL-37, acts as an allosteric modulator of P2X purinoreceptor 7 (P2 × 7) and C-X-C chemokine receptor 2 (CXCR2) receptors.177,178
To date, six SPM GPCRs have been identified, three of which still remain orphan receptors, i.e., receptors for which identification of the endogenous ligand(s) has not been yet officially confirmed. Over 140 orphan GPCRs remain to be deorphanized,66 indicating that perhaps even more SPM receptors could exist, which would be consistent with the vast number of SPMs identified. As such, ligand poly-pharmacology and receptor pleiotropy suggest distinct opportunities to target these receptors via novel mechanisms.
Biased Signaling
Previous dogma was that agonist binding would stabilize a single active conformation of the receptor and initiate the downstream signaling events.179,180 Although each ligand could have different efficacies, each agonist would activate a shared set of signaling transduction pathways.180 However, it is now widely accepted that each ligand is able to stabilize a unique receptor conformation and selectively produce a signal that is “biased” toward a specific transduction pathway.181 The attraction of biased agonists in drug development is 4-fold.182 First, an agonist can be biased toward a favorable pathway, e.g., β-arrestin activation of parathyroid agonists for the treatment of osteoporosis.183 Second, an agonist can be biased away from a detrimental or an undesired pathway that could result in off-target side effects, e.g., β-arrestin activation of CXCR3 reduces T-cell migration and induces inflammation.184 Third, bias can be away from the desired pathway and inhibition of the endogenous agonist, e.g., angiotensin-mediated β-arrestin signaling and inhibition of angiotensin-mediated vasoconstriction in heart failure.185 Finally, where different isoforms of the same receptor exist in discrete locations and display differential G-protein coupling (e.g., CXCR3 and histamine H3R).186,187 Generally, bias is classified if differences are observed in signaling profiles.182 This phenomenon is extremely useful in GPCR drug discovery as the profile of a compound can be optimized to the most opportune signaling pathway. Finally, biased signaling can be affected by the binding kinetics of ligand in question.188 In fact, ligand binding kinetics has now become a more appreciable aspect of drug discovery and should be given due attention.189−191
Biased agonism has been identified for GPCRs where multiple endogenous agonists exist. For example, α-neoendorphin, Met-enkephalin-Arg-Phe, and endomorphin-1, endogenous agonists of the μ-opioid receptor (MOP), display bias away from receptor internalization, β-arrestin recruitment, and ERK1/2 phosphorylation in comparison to the synthetic peptide, DAMGO.192 Meanwhile, angiotensin 1 receptor (AT1R), an agonist AT-(1–7) has been identified as a β-arrestin biased agonist involved in cardioprotection.193 Therefore, given the large number of SPMs that pleiotropically activate multiple GPCRs, it is perhaps surprising that there is little information on biased signaling of the SPM GPCRs. However, some evidence does exist for FPR2/ALX, albeit using synthetic molecules. FPR2-activating pepducin F2 Pal10 appears biased away from β-arrestin in comparison to the FPR2-specific peptide agonist, WKYMVM, which resulted in reduced receptor internalization and impaired neutrophil chemotaxis.194 More recently, Lau-[(I)-Aoc]-[Lys-βNrpe]6-NH2, an analog of the FPR2-selective lipidated α-peptide/β-peptoid agonist Lau-[(S)-Aoc]-[Lys-βNPhe]6-NH2, was developed with a similar bias away from β-arrestin recruitment.195 Moreover, compound 17b, an FPR1/2 biased agonist for the treatment of myocardial infarction, revealed bias away from the detrimental FPR1/2-mediated calcium mobilization but retained ERK1/2 and Akt phosphorylation fundamental for pro-survival.196
Although not strictly biased signaling, ANXA1, Ac2-26, and LXA4 have been shown to induce biased homo and heterodimerization of FPR2/ALX and FPR1, both in recombinant HEK293 assay systems and neutrophils, leading to alterations in the signaling profile.107 Concomitant binding of ANXA1 and Ac2-26 leads to homodimerization of FPR2/ALX resulting in the activation of MAPK and the release of the anti-inflammatory cytokine IL-10, while binding Ac2-26 and LXA4 results in FPR2/ALX-FPR1 heterodimerization, activation of JNK/caspase-3, and neutrophil apoptosis. Formation of this heterodimer resulted in a loss of the inhibitory effects of the endogenous antagonists SAA and LL-37.75,107 However, the existence of physiologically relevant family A GPCR dimers still remains a subject of intense investigation. Evidence suggests only a small proportion of a single receptor population exist as dimers (<20%), and when observed, these events appear to be ligand-specific.197 Although of great interest, no follow up study has been published in the intervening six years since this original discovery, but further investigation into this receptor dimerization is clearly warranted.
Of all the receptors described GPR32 could be the best placed for the observation of biased signaling. Although very little is known about the receptors signaling cascade four of the six D-series resolvin’s bind to GPR32.53 Furthermore, there is evidence that each aspirin-triggered version binds, indicating a vast number of endogenous ligands for one receptor.53 Interestingly, many of these ligands were identified using a β-arrestin recruitment assay where GPR32 was overexpressed,95 indicating that a broader array of pathways would need to be assessed in an attempt to identify bias. However, differences in the receptor internalization have been identified in the absence or presence of the pro-inflammatory cytokine TNF-α, indicating possible inflammation-driven pathway bias.
Allostery
Along with bias, allosteric modulation of GPCRs has become an important addition to the drug discovery landscape. In addition to multiple orthosteric binding sites, where the natural ligand(s) for the receptor binds, additional binding sites have been discovered on many GPCRs. These allosteric binding sites are spatially distinct from the orthosteric ligand-binding pocket, which enables the binding of two ligands simultaneously.198 Binding of these ligands can occur independent of each other, producing their own intrinsic effect. However, when bound concomitantly the ligands exert reciprocal effects, a phenomenon termed cooperativity. Four predominant classes of allosteric modulator have been identified from drug discovery efforts (although potentially many more may exist): (i) positive allosteric modulators (PAM), which enhance the endogenous ligands activity; (ii) ago-PAMs, where the PAM also displays its own intrinsic efficacy; (iii) negative allosteric modulators (NAM), which oppose the endogenous ligand’s activity; and (iv) neutral allosteric ligand (NAL), which binds but has no net effect on the endogenous ligand.199
The benefit of an allosteric modulator compared to an orthosteric ligand is potentially 2-fold. First, generally allosteric binding sites are less conserved than their orthosteric binding site counterparts, therefore enabling the possibility of greater subtype selectivity, if necessary. Second, and perhaps more importantly, the effect of the modulator is limited by the degree of cooperativity and is therefore independent of ligand concentration, enabling greater levels of compound safety.199 In essence, allosteric modulators are able to fine-tune the pharmacological response as desired. This effect enables the spatial and temporal nature of endogenous agonist signaling to be maintained, notably when changing the kinetics of release of certain SPMs during resolution are known.
As of 2018, 17 modulators have been approved or are currently under active clinical development. These modulators have been developed for a wide range of diseases from HIV (Maraviroc; NAM, CCR5) to hypercalcemia (Cinacalcet; PAM, calcium-sensing receptor) and schizophrenia (ASP4345; PAM, dopamine D1).200 This steady growth in the mechanistic understanding of allosteric modulators and the subsequent development of potential therapeutics clearly highlights a distinct opportunity in targeting SPM GPCRs. Although the exact binding sites of most SPMs are still to be elucidated, based on the knowledge gained from FPR2/ALX the probability of similar allosteric interactions occurring would appear high. As described above, GPR18 was originally designated as a cannabinoid receptor due to its binding of cannabinoid ligands, including NAGly and various synthetic molecules.120 Therefore, as LXA4 binds allosterically to CB1R, allosteric modulation of GPR18 may also be a distinct possibility. Moreover, due to the lipid composition of many SPMs access to their binding site may be via the membrane lipid bilayer, as demonstrated by sphingosine-1-phosphate binding to its receptor (S1P1R), AM841 and 2-AG binding to CB2R, and vorapaxar binding to protease-activated receptor-1 (PAR1),201−204 thus potentially allowing simultaneous binding of ligands from the extracellular space. For a recent comprehensive review, see ref (205). Intriguingly, a cryo-EM structure of FPR2/ALX in complex with Gαi has been registered in the Research Collaboratory for Structural Bioinformatics-Protein Data Bank (RCSB-PDB; structure ID 60MM). To date, no description of the ligand(s) used is reported; however, this may reveal the structural confirmation of multiple binding sites.
Biased Allostery
As a composite of the preceding sections, allosteric modulators can in themselves generate biased signaling. For example, the ago component of an ago-PAM can display bias alone, while the PAM component can bias an otherwise unbiased agonist.206,207 Furthermore, cooperativity can alter the bias depending on which endogenous ligand is present, a term known as probe dependency.207 Examples of biased modulators now exist including PAMs of muscarinic acetylcholine receptor 1 (M1R) and metabotropic glutamate receptor 5 (mGlu5R) and NAMs of the calcium-sensing receptor (CaSR) and prostaglandin F2α, (PGF2α).208−212 For comprehensive reviews, see refs (207) and (213).
Application to SPM GPCR Drug Discovery
These novel mechanisms in targeting SPM GPCRs may be even more relevant following the disappointing phase 1 trial of the FPR2/ALX agonist, ACT-389949.167 Due to the compound’s efficient recruitment of β-arrestin,168 internalization of the receptor was persistent, ultimately causing the ligand to act as a functional “antagonist”. Therefore, a biased ligand away from β-arrestin may prove a more suitable candidate. Conversely, there is a growing body of work demonstrating intracellular signaling of receptors where either the initial signaling is prolonged (persistent signaling) or the alternative signaling pathways are activated,214,215 indicating that in some cases receptor internalization could be a valued characteristic or necessity.76 Alternatively, development of an FPR2/ALX PAM to RvD1 could be worth considering, enabling greater control of receptor internalization when only the endogenous agonist is present. In fact, a biased PAM (away from β-arrestin) may be the ideal candidate. However, as the compound produced both pro- and anti-inflammatory cytokine release, other pathways (and/or receptors if dimers are involved) may have been implicit in its reduced clinical efficacy. Finally, as described above, COPD patients suffering from acute exacerbations may benefit from a selective NAM of SAA, typically elevated in peripheral blood, or a PAM of LXA4 (or potentially a combination of both via probe dependency), thus restoring LXA4’s inhibitory effect. Ultimately, due to the vast number of SPMs and GPCRs identified the combinations could be limitless, clearly highlighting significant opportunities in augmenting the resolution of inflammation. Nevertheless, complete pharmacological characterization of the signaling pathways is required, enabling target-specific decisions to be made.
Conclusion
Discovery of specialized pro-resolving mediators and their cognate receptors has led to a greater understanding of the resolution of inflammation. However, over 20 years have passed since the first identification of FPR2/ALX and LXA4, and to date no compound has successfully reached the clinic for any SPM receptor. Although significant strides have been made in identifying the numerous ligand/receptor pairings, the downstream signaling observed is more complex, leading to increasingly well-described physiological responses. The information gained from multiple individual studies has allowed, in part, development of a representation of how resolution may occur;160 nevertheless, many unanswered questions requiring investigation.
With the substantial increase in ligand/receptor function, including biased signaling and allosteric modulation, these new mechanisms add further complexity to an area already rich in challenges. However, these mechanisms should be seen as great opportunities, especially in light of the recently failed FPR2/ALX phase 1 trial. Perhaps the greatest hurdle to overcome is in the complete understanding of receptor signaling, particularly in native immune cells, thus highlighting which ligand and/or pathway may require greater consideration. Nevertheless, with both pharmacological and analytical tools readily available, therapeutics for inflammatory diseases may soon be commonplace.
The authors declare no competing financial interest.
This article is made available for a limited time sponsored by ACS under the ACS Free to Read License, which permits copying and redistribution of the article for non-commercial scholarly purposes.
References
- Riddy D. M.; Cook A. E.; Shackleford D. M.; Pierce T. L.; Mocaer E.; Mannoury la Cour C.; Sors A.; Charman W. N.; Summers R. J.; Sexton P. M.; Christopoulos A.; Langmead C. J. (2019) Drug-receptor kinetics and sigma-1 receptor affinity differentiate clinically evaluated histamine H3 receptor antagonists. Neuropharmacology 144, 244–255. 10.1016/j.neuropharm.2018.10.028. [DOI] [PubMed] [Google Scholar]
- Musso G.; Gambino R.; Cassader M.; Paschetta E.; Sircana A. (2018) Specialized Proresolving Mediators: Enhancing Nonalcoholic Steatohepatitis and Fibrosis Resolution. Trends Pharmacol. Sci. 39, 387–401. 10.1016/j.tips.2018.01.003. [DOI] [PubMed] [Google Scholar]
- Fullerton J. N.; Gilroy D. W. (2016) Resolution of inflammation: a new therapeutic frontier. Nat. Rev. Drug Discovery 15, 551–567. 10.1038/nrd.2016.39. [DOI] [PubMed] [Google Scholar]
- Recchiuti A.; Serhan C. N. (2012) Pro-Resolving Lipid Mediators (SPMs) and Their Actions in Regulating miRNA in Novel Resolution Circuits in Inflammation. Front. Immunol. 3, 298. 10.3389/fimmu.2012.00298. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Serhan C. N.; Chiang N.; Van Dyke T. E. (2008) Resolving inflammation: dual anti-inflammatory and pro-resolution lipid mediators. Nat. Rev. Immunol. 8, 349–361. 10.1038/nri2294. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Serhan C. N.; Savill J. (2005) Resolution of inflammation: the beginning programs the end. Nat. Immunol. 6, 1191–1197. 10.1038/ni1276. [DOI] [PubMed] [Google Scholar]
- Serhan C. N.; Levy B. D. (2018) Resolvins in inflammation: emergence of the pro-resolving superfamily of mediators. J. Clin. Invest. 128, 2657–2669. 10.1172/JCI97943. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alessandri A. L.; Sousa L. P.; Lucas C. D.; Rossi A. G.; Pinho V.; Teixeira M. M. (2013) Resolution of inflammation: mechanisms and opportunity for drug development. Pharmacol. Ther. 139, 189–212. 10.1016/j.pharmthera.2013.04.006. [DOI] [PubMed] [Google Scholar]
- Sostres C.; Gargallo C. J.; Arroyo M. T.; Lanas A. (2010) Adverse effects of non-steroidal anti-inflammatory drugs (NSAIDs, aspirin and coxibs) on upper gastrointestinal tract. Best Pract Res. Clin Gastroenterol 24, 121–132. 10.1016/j.bpg.2009.11.005. [DOI] [PubMed] [Google Scholar]
- Serhan C. N. (2017) Treating inflammation and infection in the 21st century: new hints from decoding resolution mediators and mechanisms. FASEB J. 31, 1273–1288. 10.1096/fj.201601222R. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bannenberg G. L.; Chiang N.; Ariel A.; Arita M.; Tjonahen E.; Gotlinger K. H.; Hong S.; Serhan C. N. (2005) Molecular circuits of resolution: formation and actions of resolvins and protectins. J. Immunol. 174, 4345–4355. 10.4049/jimmunol.174.7.4345. [DOI] [PubMed] [Google Scholar]
- Serhan C. N.; Hong S.; Gronert K.; Colgan S. P.; Devchand P. R.; Mirick G.; Moussignac R. L. (2002) Resolvins: a family of bioactive products of omega-3 fatty acid transformation circuits initiated by aspirin treatment that counter proinflammation signals. J. Exp. Med. 196, 1025–1037. 10.1084/jem.20020760. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Serhan C. N.; Clish C. B.; Brannon J.; Colgan S. P.; Chiang N.; Gronert K. (2000) Novel functional sets of lipid-derived mediators with antiinflammatory actions generated from omega-3 fatty acids via cyclooxygenase 2-nonsteroidal antiinflammatory drugs and transcellular processing. J. Exp. Med. 192, 1197–1204. 10.1084/jem.192.8.1197. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dean R. A.; Cox J. H.; Bellac C. L.; Doucet A.; Starr A. E.; Overall C. M. (2008) Macrophage-specific metalloelastase (MMP-12) truncates and inactivates ELR+ CXC chemokines and generates CCL2, −7, −8, and −13 antagonists: potential role of the macrophage in terminating polymorphonuclear leukocyte influx. Blood 112, 3455–3464. 10.1182/blood-2007-12-129080. [DOI] [PubMed] [Google Scholar]
- Ortega-Gomez A.; Perretti M.; Soehnlein O. (2013) Resolution of inflammation: an integrated view. EMBO Mol. Med. 5, 661–674. 10.1002/emmm.201202382. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Greenlee-Wacker M. C. (2016) Clearance of apoptotic neutrophils and resolution of inflammation. Immunol Rev. 273, 357–370. 10.1111/imr.12453. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dorward D. A.; Rossi A. G.; Dransfield I.; Lucas C. D. (2014) Assessment of neutrophil apoptosis. Methods Mol. Biol. 1124, 159–180. 10.1007/978-1-62703-845-4_10. [DOI] [PubMed] [Google Scholar]
- Fox S.; Leitch A. E.; Duffin R.; Haslett C.; Rossi A. G. (2010) Neutrophil apoptosis: relevance to the innate immune response and inflammatory disease. J. Innate Immun. 2, 216–227. 10.1159/000284367. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Poon I. K.; Lucas C. D.; Rossi A. G.; Ravichandran K. S. (2014) Apoptotic cell clearance: basic biology and therapeutic potential. Nat. Rev. Immunol. 14, 166–180. 10.1038/nri3607. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Duffin R.; Leitch A. E.; Fox S.; Haslett C.; Rossi A. G. (2010) Targeting granulocyte apoptosis: mechanisms, models, and therapies. Immunol. Rev. 236, 28–40. 10.1111/j.1600-065X.2010.00922.x. [DOI] [PubMed] [Google Scholar]
- Fadok V. A.; Voelker D. R.; Campbell P. A.; Cohen J. J.; Bratton D. L.; Henson P. M. (1992) Exposure of phosphatidylserine on the surface of apoptotic lymphocytes triggers specific recognition and removal by macrophages. J. Immunol. 148, 2207–2216. [PubMed] [Google Scholar]
- Kobayashi N.; Karisola P.; Pena-Cruz V.; Dorfman D. M.; Jinushi M.; Umetsu S. E.; Butte M. J.; Nagumo H.; Chernova I.; Zhu B.; Sharpe A. H.; Ito S.; Dranoff G.; Kaplan G. G.; Casasnovas J. M.; Umetsu D. T.; Dekruyff R. H.; Freeman G. J. (2007) TIM-1 and TIM-4 glycoproteins bind phosphatidylserine and mediate uptake of apoptotic cells. Immunity 27, 927–940. 10.1016/j.immuni.2007.11.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Miyanishi M.; Tada K.; Koike M.; Uchiyama Y.; Kitamura T.; Nagata S. (2007) Identification of Tim4 as a phosphatidylserine receptor. Nature 450, 435–439. 10.1038/nature06307. [DOI] [PubMed] [Google Scholar]
- Nakayama M.; Akiba H.; Takeda K.; Kojima Y.; Hashiguchi M.; Azuma M.; Yagita H.; Okumura K. (2009) Tim-3 mediates phagocytosis of apoptotic cells and cross-presentation. Blood 113, 3821–3830. 10.1182/blood-2008-10-185884. [DOI] [PubMed] [Google Scholar]
- Park S. Y.; Jung M. Y.; Kim H. J.; Lee S. J.; Kim S. Y.; Lee B. H.; Kwon T. H.; Park R. W.; Kim I. S. (2008) Rapid cell corpse clearance by stabilin-2, a membrane phosphatidylserine receptor. Cell Death Differ. 15, 192–201. 10.1038/sj.cdd.4402242. [DOI] [PubMed] [Google Scholar]
- Park D.; Tosello-Trampont A. C.; Elliott M. R.; Lu M.; Haney L. B.; Ma Z.; Klibanov A. L.; Mandell J. W.; Ravichandran K. S. (2007) BAI1 is an engulfment receptor for apoptotic cells upstream of the ELMO/Dock180/Rac module. Nature 450, 430–434. 10.1038/nature06329. [DOI] [PubMed] [Google Scholar]
- Lee S. J.; So I. S.; Park S. Y.; Kim I. S. (2008) Thymosin beta4 is involved in stabilin-2-mediated apoptotic cell engulfment. FEBS Lett. 582, 2161–2166. 10.1016/j.febslet.2008.03.058. [DOI] [PubMed] [Google Scholar]
- Park S. Y.; Kang K. B.; Thapa N.; Kim S. Y.; Lee S. J.; Kim I. S. (2008) Requirement of adaptor protein GULP during stabilin-2-mediated cell corpse engulfment. J. Biol. Chem. 283, 10593–10600. 10.1074/jbc.M709105200. [DOI] [PubMed] [Google Scholar]
- Anderson H. A.; Maylock C. A.; Williams J. A.; Paweletz C. P.; Shu H.; Shacter E. (2003) Serum-derived protein S binds to phosphatidylserine and stimulates the phagocytosis of apoptotic cells. Nat. Immunol. 4, 87–91. 10.1038/ni871. [DOI] [PubMed] [Google Scholar]
- Hanayama R.; Tanaka M.; Miwa K.; Shinohara A.; Iwamatsu A.; Nagata S. (2002) Identification of a factor that links apoptotic cells to phagocytes. Nature 417, 182–187. 10.1038/417182a. [DOI] [PubMed] [Google Scholar]
- Gardai S. J.; McPhillips K. A.; Frasch S. C.; Janssen W. J.; Starefeldt A.; Murphy-Ullrich J. E.; Bratton D. L.; Oldenborg P. A.; Michalak M.; Henson P. M. (2005) Cell-surface calreticulin initiates clearance of viable or apoptotic cells through trans-activation of LRP on the phagocyte. Cell 123, 321–334. 10.1016/j.cell.2005.08.032. [DOI] [PubMed] [Google Scholar]
- Martin C. J.; Peters K. N.; Behar S. M. (2014) Macrophages clean up: efferocytosis and microbial control. Curr. Opin. Microbiol. 17, 17–23. 10.1016/j.mib.2013.10.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Prame Kumar K.; Nicholls A. J.; Wong C. H. Y. (2018) Partners in crime: neutrophils and monocytes/macrophages in inflammation and disease. Cell Tissue Res. 371, 551–565. 10.1007/s00441-017-2753-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Watanabe S.; Alexander M.; Misharin A. V.; Budinger G. R. S. (2019) The role of macrophages in the resolution of inflammation. J. Clin. Invest. 129, 2619–2628. 10.1172/JCI124615. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kytikova O.; Novgorodtseva T.; Denisenko Y.; Antonyuk M.; Gvozdenko T. (2019) Pro-Resolving Lipid Mediators in the Pathophysiology of Asthma. Medicina (Kaunas) 55, 284. 10.3390/medicina55060284. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chandrasekharan J. A.; Sharma-Walia N. (2015) Lipoxins: nature’s way to resolve inflammation. J. Inflammation Res. 8, 181–192. 10.2147/JIR.S90380. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chiang N.; Serhan C. N. (2017) Structural elucidation and physiologic functions of specialized pro-resolving mediators and their receptors. Mol. Aspects Med. 58, 114–129. 10.1016/j.mam.2017.03.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Serhan C. N.; Petasis N. A. (2011) Resolvins and protectins in inflammation resolution. Chem. Rev. 111, 5922–5943. 10.1021/cr100396c. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Serhan C. N.; Hamberg M.; Samuelsson B. (1984) Lipoxins: novel series of biologically active compounds formed from arachidonic acid in human leukocytes. Proc. Natl. Acad. Sci. U. S. A. 81, 5335–5339. 10.1073/pnas.81.17.5335. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Serhan C. N.; Sheppard K. A. (1990) Lipoxin formation during human neutrophil-platelet interactions. Evidence for the transformation of leukotriene A4 by platelet 12-lipoxygenase in vitro. J. Clin. Invest. 85, 772–780. 10.1172/JCI114503. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Romano M.; Serhan C. N. (1992) Lipoxin generation by permeabilized human platelets. Biochemistry 31, 8269–8277. 10.1021/bi00150a021. [DOI] [PubMed] [Google Scholar]
- Chiang N.; Arita M.; Serhan C. N. (2005) Anti-inflammatory circuitry: lipoxin, aspirin-triggered lipoxins and their receptor ALX. Prostaglandins, Leukotrienes Essent. Fatty Acids 73, 163–177. 10.1016/j.plefa.2005.05.003. [DOI] [PubMed] [Google Scholar]
- Samuelsson B.; Dahlen S. E.; Lindgren J. A.; Rouzer C. A.; Serhan C. N. (1987) Leukotrienes and lipoxins: structures, biosynthesis, and biological effects. Science 237, 1171–1176. 10.1126/science.2820055. [DOI] [PubMed] [Google Scholar]
- Claria J.; Serhan C. N. (1995) Aspirin triggers previously undescribed bioactive eicosanoids by human endothelial cell-leukocyte interactions. Proc. Natl. Acad. Sci. U. S. A. 92, 9475–9479. 10.1073/pnas.92.21.9475. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Planaguma A.; Pfeffer M. A.; Rubin G.; Croze R.; Uddin M.; Serhan C. N.; Levy B. D. (2010) Lovastatin decreases acute mucosal inflammation via 15-epi-lipoxin A4. Mucosal Immunol. 3, 270–279. 10.1038/mi.2009.141. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Levy B. D.; Clish C. B.; Schmidt B.; Gronert K.; Serhan C. N. (2001) Lipid mediator class switching during acute inflammation: signals in resolution. Nat. Immunol. 2, 612–619. 10.1038/89759. [DOI] [PubMed] [Google Scholar]
- Tjonahen E.; Oh S. F.; Siegelman J.; Elangovan S.; Percarpio K. B.; Hong S.; Arita M.; Serhan C. N. (2006) Resolvin E2: identification and anti-inflammatory actions: pivotal role of human 5-lipoxygenase in resolvin E series biosynthesis. Chem. Biol. 13, 1193–1202. 10.1016/j.chembiol.2006.09.011. [DOI] [PubMed] [Google Scholar]
- Isobe Y.; Arita M.; Matsueda S.; Iwamoto R.; Fujihara T.; Nakanishi H.; Taguchi R.; Masuda K.; Sasaki K.; Urabe D.; Inoue M.; Arai H. (2012) Identification and structure determination of novel anti-inflammatory mediator resolvin E3, 17,18-dihydroxyeicosapentaenoic acid. J. Biol. Chem. 287, 10525–10534. 10.1074/jbc.M112.340612. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hong S.; Porter T. F.; Lu Y.; Oh S. F.; Pillai P. S.; Serhan C. N. (2008) Resolvin E1 metabolome in local inactivation during inflammation-resolution. J. Immunol. 180, 3512–3519. 10.4049/jimmunol.180.5.3512. [DOI] [PubMed] [Google Scholar]
- Oh S. F.; Dona M.; Fredman G.; Krishnamoorthy S.; Irimia D.; Serhan C. N. (2012) Resolvin E2 formation and impact in inflammation resolution. J. Immunol. 188, 4527–4534. 10.4049/jimmunol.1103652. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sun Y. P.; Oh S. F.; Uddin J.; Yang R.; Gotlinger K.; Campbell E.; Colgan S. P.; Petasis N. A.; Serhan C. N. (2007) Resolvin D1 and its aspirin-triggered 17R epimer. Stereochemical assignments, anti-inflammatory properties, and enzymatic inactivation. J. Biol. Chem. 282, 9323–9334. 10.1074/jbc.M609212200. [DOI] [PubMed] [Google Scholar]
- Winkler J. W.; Uddin J.; Serhan C. N.; Petasis N. A. (2013) Stereocontrolled total synthesis of the potent anti-inflammatory and pro-resolving lipid mediator resolvin D3 and its aspirin-triggered 17R-epimer. Org. Lett. 15, 1424–1427. 10.1021/ol400484u. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dalli J.; Winkler J. W.; Colas R. A.; Arnardottir H.; Cheng C. Y.; Chiang N.; Petasis N. A.; Serhan C. N. (2013) Resolvin D3 and aspirin-triggered resolvin D3 are potent immunoresolvents. Chem. Biol. 20, 188–201. 10.1016/j.chembiol.2012.11.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Winkler J. W.; Orr S. K.; Dalli J.; Cheng C. Y.; Sanger J. M.; Chiang N.; Petasis N. A.; Serhan C. N. (2016) Resolvin D4 stereoassignment and its novel actions in host protection and bacterial clearance. Sci. Rep. 6, 18972. 10.1038/srep18972. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stables M. J.; Shah S.; Camon E. B.; Lovering R. C.; Newson J.; Bystrom J.; Farrow S.; Gilroy D. W. (2011) Transcriptomic analyses of murine resolution-phase macrophages. Blood 118, e192–208. 10.1182/blood-2011-04-345330. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Serhan C. N.; Gotlinger K.; Hong S.; Lu Y.; Siegelman J.; Baer T.; Yang R.; Colgan S. P.; Petasis N. A. (2006) Anti-inflammatory actions of neuroprotectin D1/protectin D1 and its natural stereoisomers: assignments of dihydroxy-containing docosatrienes. J. Immunol. 176, 1848–1859. 10.4049/jimmunol.176.3.1848. [DOI] [PubMed] [Google Scholar]
- Mukherjee P. K.; Marcheselli V. L.; Serhan C. N.; Bazan N. G. (2004) Neuroprotectin D1: a docosahexaenoic acid-derived docosatriene protects human retinal pigment epithelial cells from oxidative stress. Proc. Natl. Acad. Sci. U. S. A. 101, 8491–8496. 10.1073/pnas.0402531101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hong S.; Tjonahen E.; Morgan E. L.; Lu Y.; Serhan C. N.; Rowley A. F. (2005) Rainbow trout (Oncorhynchus mykiss) brain cells biosynthesize novel docosahexaenoic acid-derived resolvins and protectins-Mediator lipidomic analysis. Prostaglandins Other Lipid Mediators 78, 107–116. 10.1016/j.prostaglandins.2005.04.004. [DOI] [PubMed] [Google Scholar]
- Serhan C. N.; Yang R.; Martinod K.; Kasuga K.; Pillai P. S.; Porter T. F.; Oh S. F.; Spite M. (2009) Maresins: novel macrophage mediators with potent antiinflammatory and proresolving actions. J. Exp. Med. 206, 15–23. 10.1084/jem.20081880. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Deng B.; Wang C. W.; Arnardottir H. H.; Li Y.; Cheng C. Y.; Dalli J.; Serhan C. N. (2014) Maresin biosynthesis and identification of maresin 2, a new anti-inflammatory and pro-resolving mediator from human macrophages. PLoS One 9, e102362. 10.1371/journal.pone.0102362. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dalli J.; Zhu M.; Vlasenko N. A.; Deng B.; Haeggstrom J. Z.; Petasis N. A.; Serhan C. N. (2013) The novel 13S,14S-epoxy-maresin is converted by human macrophages to maresin 1 (MaR1), inhibits leukotriene A4 hydrolase (LTA4H), and shifts macrophage phenotype. FASEB J. 27, 2573–2583. 10.1096/fj.13-227728. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Serhan C. N.; Dalli J.; Karamnov S.; Choi A.; Park C. K.; Xu Z. Z.; Ji R. R.; Zhu M.; Petasis N. A. (2012) Macrophage proresolving mediator maresin 1 stimulates tissue regeneration and controls pain. FASEB J. 26, 1755–1765. 10.1096/fj.11-201442. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Abdulnour R. E.; Dalli J.; Colby J. K.; Krishnamoorthy N.; Timmons J. Y.; Tan S. H.; Colas R. A.; Petasis N. A.; Serhan C. N.; Levy B. D. (2014) Maresin 1 biosynthesis during platelet-neutrophil interactions is organ-protective. Proc. Natl. Acad. Sci. U. S. A. 111, 16526–16531. 10.1073/pnas.1407123111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arita M.; Bianchini F.; Aliberti J.; Sher A.; Chiang N.; Hong S.; Yang R.; Petasis N. A.; Serhan C. N. (2005) Stereochemical assignment, antiinflammatory properties, and receptor for the omega-3 lipid mediator resolvin E1. J. Exp. Med. 201, 713–722. 10.1084/jem.20042031. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bang S.; Xie Y. K.; Zhang Z. J.; Wang Z.; Xu Z. Z.; Ji R. R. (2018) GPR37 regulates macrophage phagocytosis and resolution of inflammatory pain. J. Clin. Invest. 128, 3568–3582. 10.1172/JCI99888. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tang X. L.; Wang Y.; Li D. L.; Luo J.; Liu M. Y. (2012) Orphan G protein-coupled receptors (GPCRs): biological functions and potential drug targets. Acta Pharmacol. Sin. 33, 363–371. 10.1038/aps.2011.210. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen K.; Bao Z.; Gong W.; Tang P.; Yoshimura T.; Wang J. M. (2017) Regulation of inflammation by members of the formyl-peptide receptor family. J. Autoimmun. 85, 64–77. 10.1016/j.jaut.2017.06.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Devosse T.; Guillabert A.; D’Haene N.; Berton A.; De Nadai P.; Noel S.; Brait M.; Franssen J. D.; Sozzani S.; Salmon I.; Parmentier M. (2009) Formyl peptide receptor-like 2 is expressed and functional in plasmacytoid dendritic cells, tissue-specific macrophage subpopulations, and eosinophils. J. Immunol. 182, 4974–4984. 10.4049/jimmunol.0803128. [DOI] [PubMed] [Google Scholar]
- Fiore S.; Ryeom S. W.; Weller P. F.; Serhan C. N. (1992) Lipoxin recognition sites. Specific binding of labeled lipoxin A4 with human neutrophils. J. Biol. Chem. 267, 16168–16176. [PubMed] [Google Scholar]
- 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. (2009) Functional expression of formyl peptide receptor family in human NK cells. J. Immunol. 183, 5511–5517. 10.4049/jimmunol.0802986. [DOI] [PubMed] [Google Scholar]
- Ho C. F.-Y.; Ismail N. B.; Koh J. K.-Z.; Gunaseelan S.; Low Y.-H.; Ng Y.-K.; Chua J. J.-E.; Ong W.-Y. (2018) Localisation of Formyl-Peptide Receptor 2 in the Rat Central Nervous System and Its Role in Axonal and Dendritic Outgrowth. Neurochem. Res. 43, 1587. 10.1007/s11064-018-2573-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Norling L. V.; Dalli J.; Flower R. J.; Serhan C. N.; Perretti M. (2012) Resolvin D1 limits polymorphonuclear leukocyte recruitment to inflammatory loci: receptor-dependent actions. Arterioscler., Thromb., Vasc. Biol. 32, 1970–1978. 10.1161/ATVBAHA.112.249508. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chiang N.; Takano T.; Arita M.; Watanabe S.; Serhan C. N. (2003) A novel rat lipoxin A4 receptor that is conserved in structure and function. Br. J. Pharmacol. 139, 89–98. 10.1038/sj.bjp.0705220. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takano T.; Fiore S.; Maddox J. F.; Brady H. R.; Petasis N. A.; Serhan C. N. (1997) Aspirin-triggered 15-epi-lipoxin A4 (LXA4) and LXA4 stable analogues are potent inhibitors of acute inflammation: evidence for anti-inflammatory receptors. J. Exp. Med. 185, 1693–1704. 10.1084/jem.185.9.1693. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Filep J. G. (2013) Biasing the lipoxin A4/formyl peptide receptor 2 pushes inflammatory resolution. Proc. Natl. Acad. Sci. U. S. A. 110, 18033–18034. 10.1073/pnas.1317798110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maderna P.; Cottell D. C.; Toivonen T.; Dufton N.; Dalli J.; Perretti M.; Godson C. (2010) FPR2/ALX receptor expression and internalization are critical for lipoxin A4 and annexin-derived peptide-stimulated phagocytosis. FASEB J. 24, 4240–4249. 10.1096/fj.10-159913. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ye R. D.; Cavanagh S. L.; Quehenberger O.; Prossnitz E. R.; Cochrane C. G. (1992) Isolation of a cDNA that encodes a novel granulocyte N-formyl peptide receptor. Biochem. Biophys. Res. Commun. 184, 582–589. 10.1016/0006-291X(92)90629-Y. [DOI] [PubMed] [Google Scholar]
- Murphy P. M.; Ozcelik T.; Kenney R. T.; Tiffany H. L.; McDermott D.; Francke U. (1992) A structural homologue of the N-formyl peptide receptor. Characterization and chromosome mapping of a peptide chemoattractant receptor family. J. Biol. Chem. 267, 7637–7643. [PubMed] [Google Scholar]
- Fiore S.; Maddox J. F.; Perez H. D.; Serhan C. N. (1994) Identification of a human cDNA encoding a functional high affinity lipoxin A4 receptor. J. Exp. Med. 180, 253–260. 10.1084/jem.180.1.253. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brink C.; Dahlen S. E.; Drazen J.; Evans J. F.; Hay D. W.; Nicosia S.; Serhan C. N.; Shimizu T.; Yokomizo T. (2003) International Union of Pharmacology XXXVII. Nomenclature for leukotriene and lipoxin receptors. Pharmacol. Rev. 55, 195–227. 10.1124/pr.55.1.8. [DOI] [PubMed] [Google Scholar]
- Russell R.; Gori I.; Pellegrini C.; Kumar R.; Achtari C.; Canny G. O. (2011) Lipoxin A4 is a novel estrogen receptor modulator. FASEB J. 25, 4326–4337. 10.1096/fj.11-187658. [DOI] [PubMed] [Google Scholar]
- Schaldach C. M.; Riby J.; Bjeldanes L. F. (1999) Lipoxin A4: a new class of ligand for the Ah receptor.(aryl hydrocarbon). Biochemistry 38, 7594. 10.1021/bi982861e. [DOI] [PubMed] [Google Scholar]
- Maddox J. F.; Serhan C. N. (1996) Lipoxin A4 and B4 are potent stimuli for human monocyte migration and adhesion: selective inactivation by dehydrogenation and reduction. J. Exp. Med. 183, 137–146. 10.1084/jem.183.1.137. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ramon S.; Bancos S.; Serhan C. N.; Phipps R. P. (2014) Lipoxin A(4) modulates adaptive immunity by decreasing memory B-cell responses via an ALX/FPR2-dependent mechanism. Eur. J. Immunol. 44, 357–369. 10.1002/eji.201343316. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Simiele F.; Recchiuti A.; Mattoscio D.; De Luca A.; Cianci E.; Franchi S.; Gatta V.; Parolari A.; Werba J. P.; Camera M.; Favaloro B.; Romano M. (2012) Transcriptional regulation of the human FPR2/ALX gene: evidence of a heritable genetic variant that impairs promoter activity. FASEB J. 26, 1323–1333. 10.1096/fj.11-198069. [DOI] [PubMed] [Google Scholar]
- Dufton N.; Hannon R.; Brancaleone V.; Dalli J.; Patel H. B.; Gray M.; D’Acquisto F.; Buckingham J. C.; Perretti M.; Flower R. J. (2010) Anti-inflammatory role of the murine formyl-peptide receptor 2: ligand-specific effects on leukocyte responses and experimental inflammation. J. Immunol. 184, 2611–2619. 10.4049/jimmunol.0903526. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brancaleone V.; Gobbetti T.; Cenac N.; le Faouder P.; Colom B.; Flower R. J.; Vergnolle N.; Nourshargh S.; Perretti M. (2013) A vasculo-protective circuit centered on lipoxin A4 and aspirin-triggered 15-epi-lipoxin A4 operative in murine microcirculation. Blood 122, 608–617. 10.1182/blood-2013-04-496661. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sordi R.; Menezes-de-Lima O. Jr.; Horewicz V.; Scheschowitsch K.; Santos L. F.; Assreuy J. (2013) Dual role of lipoxin A4 in pneumosepsis pathogenesis. Int. Immunopharmacol. 17, 283–292. 10.1016/j.intimp.2013.06.010. [DOI] [PubMed] [Google Scholar]
- Chiang N.; Fierro I. M.; Gronert K.; Serhan C. N. (2000) Activation of lipoxin A(4) receptors by aspirin-triggered lipoxins and select peptides evokes ligand-specific responses in inflammation. J. Exp. Med. 191, 1197–1208. 10.1084/jem.191.7.1197. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bena S.; Brancaleone V.; Wang J. M.; Perretti M.; Flower R. J. (2012) Annexin A1 Interaction with the FPR2/ALX Receptor. J. Biol. Chem. 287, 24690–24697. 10.1074/jbc.M112.377101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Filep J. G.; Sekheri M.; El Kebir D. (2018) Targeting formyl peptide receptors to facilitate the resolution of inflammation. Eur. J. Pharmacol. 833, 339–348. 10.1016/j.ejphar.2018.06.025. [DOI] [PubMed] [Google Scholar]
- Krishnamoorthy N.; Abdulnour R. E.; Walker K. H.; Engstrom B. D.; Levy B. D. (2018) Specialized Proresolving Mediators in Innate and Adaptive Immune Responses in Airway Diseases. Physiol. Rev. 98, 1335–1370. 10.1152/physrev.00026.2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- El Kebir D.; Jozsef L.; Pan W.; Wang L.; Petasis N. A.; Serhan C. N.; Filep J. G. (2009) 15-epi-lipoxin A4 inhibits myeloperoxidase signaling and enhances resolution of acute lung injury. Am. J. Respir. Crit. Care Med. 180, 311–319. 10.1164/rccm.200810-1601OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Prescott D.; McKay D. M. (2011) Aspirin-triggered lipoxin enhances macrophage phagocytosis of bacteria while inhibiting inflammatory cytokine production. Am. J. Physiol Gastrointest Liver Physiol 301, G487–497. 10.1152/ajpgi.00042.2011. [DOI] [PubMed] [Google Scholar]
- Krishnamoorthy S.; Recchiuti A.; Chiang N.; Yacoubian S.; Lee C. H.; Yang R.; Petasis N. A.; Serhan C. N. (2010) Resolvin D1 binds human phagocytes with evidence for proresolving receptors. Proc. Natl. Acad. Sci. U. S. A. 107, 1660–1665. 10.1073/pnas.0907342107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee H. N.; Surh Y. J. (2013) Resolvin D1-mediated NOX2 inactivation rescues macrophages undertaking efferocytosis from oxidative stress-induced apoptosis. Biochem. Pharmacol. 86, 759–769. 10.1016/j.bcp.2013.07.002. [DOI] [PubMed] [Google Scholar]
- Krishnamoorthy S.; Recchiuti A.; Chiang N.; Fredman G.; Serhan C. N. (2012) Resolvin D1 receptor stereoselectivity and regulation of inflammation and proresolving microRNAs. Am. J. Pathol. 180, 2018–2027. 10.1016/j.ajpath.2012.01.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li Y.; Dalli J.; Chiang N.; Baron R. M.; Quintana C.; Serhan C. N. (2013) Plasticity of leukocytic exudates in resolving acute inflammation is regulated by MicroRNA and proresolving mediators. Immunity 39, 885–898. 10.1016/j.immuni.2013.10.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eickmeier O.; Seki H.; Haworth O.; Hilberath J. N.; Gao F.; Uddin M.; Croze R. H.; Carlo T.; Pfeffer M. A.; Levy B. D. (2013) Aspirin-triggered resolvin D1 reduces mucosal inflammation and promotes resolution in a murine model of acute lung injury. Mucosal Immunol. 6, 256–266. 10.1038/mi.2012.66. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gilligan M. M.; Gartung A.; Sulciner M. L.; Norris P. C.; Sukhatme V. P.; Bielenberg D. R.; Huang S.; Kieran M. W.; Serhan C. N.; Panigrahy D. (2019) Aspirin-triggered proresolving mediators stimulate resolution in cancer. Proc. Natl. Acad. Sci. U. S. A. 116, 6292–6297. 10.1073/pnas.1804000116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Scannell M.; Flanagan M. B.; deStefani A.; Wynne K. J.; Cagney G.; Godson C.; Maderna P. (2007) Annexin-1 and peptide derivatives are released by apoptotic cells and stimulate phagocytosis of apoptotic neutrophils by macrophages. J. Immunol. 178, 4595–4605. 10.4049/jimmunol.178.7.4595. [DOI] [PubMed] [Google Scholar]
- Ernst S.; Lange C.; Wilbers A.; Goebeler V.; Gerke V.; Rescher U. (2004) An annexin 1 N-terminal peptide activates leukocytes by triggering different members of the formyl peptide receptor family. J. Immunol. 172, 7669–7676. 10.4049/jimmunol.172.12.7669. [DOI] [PubMed] [Google Scholar]
- Vago J. P.; Nogueira C. R.; Tavares L. P.; Soriani F. M.; Lopes F.; Russo R. C.; Pinho V.; Teixeira M. M.; Sousa L. P. (2012) Annexin A1 modulates natural and glucocorticoid-induced resolution of inflammation by enhancing neutrophil apoptosis. J. Leukocyte Biol. 92, 249–258. 10.1189/jlb.0112008. [DOI] [PubMed] [Google Scholar]
- Sugimoto M. A.; Vago J. P.; Teixeira M. M.; Sousa L. P. (2016) Annexin A1 and the Resolution of Inflammation: Modulation of Neutrophil Recruitment, Apoptosis, and Clearance. J. Immunol. Res. 2016, 8239258. 10.1155/2016/8239258. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dalli J.; Norling L. V.; Renshaw D.; Cooper D.; Leung K. Y.; Perretti M. (2008) Annexin 1 mediates the rapid anti-inflammatory effects of neutrophil-derived microparticles. Blood 112, 2512–2519. 10.1182/blood-2008-02-140533. [DOI] [PubMed] [Google Scholar]
- Corminboeuf O.; Leroy X. (2015) FPR2/ALXR agonists and the resolution of inflammation. J. Med. Chem. 58, 537–559. 10.1021/jm501051x. [DOI] [PubMed] [Google Scholar]
- Cooray S. N.; Gobbetti T.; Montero-Melendez T.; McArthur S.; Thompson D.; Clark A. J.; Flower R. J.; Perretti M. (2013) Ligand-specific conformational change of the G-protein-coupled receptor ALX/FPR2 determines proresolving functional responses. Proc. Natl. Acad. Sci. U. S. A. 110, 18232–18237. 10.1073/pnas.1308253110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schmid M.; Gemperle C.; Rimann N.; Hersberger M. (2016) Resolvin D1 Polarizes Primary Human Macrophages toward a Proresolution Phenotype through GPR32. J. Immunol. 196, 3429–3437. 10.4049/jimmunol.1501701. [DOI] [PubMed] [Google Scholar]
- Sansbury B. E.; Spite M. (2016) Resolution of Acute Inflammation and the Role of Resolvins in Immunity, Thrombosis, and Vascular Biology. Circ. Res. 119, 113–130. 10.1161/CIRCRESAHA.116.307308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Waechter V.; Schmid M.; Herova M.; Weber A.; Gunther V.; Marti-Jaun J.; Wust S.; Rosinger M.; Gemperle C.; Hersberger M. (2012) Characterization of the promoter and the transcriptional regulation of the lipoxin A4 receptor (FPR2/ALX) gene in human monocytes and macrophages. J. Immunol. 188, 1856–1867. 10.4049/jimmunol.1101788. [DOI] [PubMed] [Google Scholar]
- Chiang N.; Fredman G.; Backhed F.; Oh S. F.; Vickery T.; Schmidt B. A.; Serhan C. N. (2012) Infection regulates pro-resolving mediators that lower antibiotic requirements. Nature 484, 524–528. 10.1038/nature11042. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chiang N.; Barnaeva E.; Hu X.; Marugan J.; Southall N.; Ferrer M.; Serhan C. N. (2019) Identification of Chemotype Agonists for Human Resolvin D1 Receptor DRV1 with Pro-Resolving Functions. Cell Chem. Biol. 26, 244–254. 10.1016/j.chembiol.2018.10.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Foster S. R.; Hauser A. S.; Vedel L.; Strachan R. T.; Huang X. P.; Gavin A. C.; Shah S. D.; Nayak A. P.; Haugaard-Kedstrom L. M.; Penn R. B.; Roth B. L.; Brauner-Osborne H.; Gloriam D. E. (2019) Discovery of Human Signaling Systems: Pairing Peptides to G Protein-Coupled Receptors. Cell 179, 895–908.e821. 10.1016/j.cell.2019.10.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chiang N.; Dalli J.; Colas R. A.; Serhan C. N. (2015) Identification of resolvin D2 receptor mediating resolution of infections and organ protection. J. Exp. Med. 212, 1203–1217. 10.1084/jem.20150225. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pirault J.; Back M. (2018) Lipoxin and Resolvin Receptors Transducing the Resolution of Inflammation in Cardiovascular Disease. Front. Pharmacol. 9, 1273. 10.3389/fphar.2018.01273. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang X.; Sumida H.; Cyster J. G. (2014) GPR18 is required for a normal CD8alphaalpha intestinal intraepithelial lymphocyte compartment. J. Exp. Med. 211, 2351–2359. 10.1084/jem.20140646. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kohno M.; Hasegawa H.; Inoue A.; Muraoka M.; Miyazaki T.; Oka K.; Yasukawa M. (2006) Identification of N-arachidonylglycine as the endogenous ligand for orphan G-protein-coupled receptor GPR18. Biochem. Biophys. Res. Commun. 347, 827–832. 10.1016/j.bbrc.2006.06.175. [DOI] [PubMed] [Google Scholar]
- Finlay D. B.; Joseph W. R.; Grimsey N. L.; Glass M. (2016) GPR18 undergoes a high degree of constitutive trafficking but is unresponsive to N-Arachidonoyl Glycine. PeerJ 4, e1835. 10.7717/peerj.1835. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McHugh D.; Hu S. S.; Rimmerman N.; Juknat A.; Vogel Z.; Walker J. M.; Bradshaw H. B. (2010) N-arachidonoyl glycine, an abundant endogenous lipid, potently drives directed cellular migration through GPR18, the putative abnormal cannabidiol receptor. BMC Neurosci. 11, 44. 10.1186/1471-2202-11-44. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McHugh D. (2012) GPR18 in microglia: implications for the CNS and endocannabinoid system signalling. Br. J. Pharmacol. 167, 1575–1582. 10.1111/j.1476-5381.2012.02019.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alexander S. P.; Christopoulos A.; Davenport A. P; Kelly E.; Marrion N. V; Peters J. A.; Faccenda E.; Harding S. D.; Pawson A. J.; Sharman J. L.; Southan C.; Davies J. A. (2017) the Concise Guide to Pharmacology 2017/18: G protein-coupled receptors. Br. J. Pharmacol. 174, S17–S129. 10.1111/bph.13878. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chiang N.; de la Rosa X.; Libreros S.; Serhan C. N. (2017) Novel Resolvin D2 Receptor Axis in Infectious Inflammation. J. Immunol. 198, 842–851. 10.4049/jimmunol.1601650. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zizzo G.; Hilliard B. A.; Monestier M.; Cohen P. L. (2012) Efficient clearance of early apoptotic cells by human macrophages requires M2c polarization and MerTK induction. J. Immunol. 189, 3508–3520. 10.4049/jimmunol.1200662. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lopategi A.; Flores-Costa R.; Rius B.; Lopez-Vicario C.; Alcaraz-Quiles J.; Titos E.; Claria J. (2019) Frontline Science: Specialized proresolving lipid mediators inhibit the priming and activation of the macrophage NLRP3 inflammasome. J. Leukocyte Biol. 105, 25–36. 10.1002/JLB.3HI0517-206RR. [DOI] [PubMed] [Google Scholar]
- Kurihara T.; Jones C. N.; Yu Y. M.; Fischman A. J.; Watada S.; Tompkins R. G.; Fagan S. P.; Irimia D. (2013) Resolvin D2 restores neutrophil directionality and improves survival after burns. FASEB J. 27, 2270–2281. 10.1096/fj.12-219519. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang L.; Qiu C.; Yang L.; Zhang Z.; Zhang Q.; Wang B.; Wang X. (2019) GPR18 expression on PMNs as biomarker for outcome in patient with sepsis. Life Sci. 217, 49–56. 10.1016/j.lfs.2018.11.061. [DOI] [PubMed] [Google Scholar]
- Turcotte C.; Blanchet M.-R.; Laviolette M.; Flamand N. (2016) The CB∧sub 2∧ receptor and its role as a regulator of inflammation. Cell. Mol. Life Sci. 73, 4449–4470. 10.1007/s00018-016-2300-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bondue B.; Wittamer V.; Parmentier M. (2011) Chemerin and its receptors in leukocyte trafficking, inflammation and metabolism. Cytokine Growth Factor Rev. 22, 331–338. 10.1016/j.cytogfr.2011.11.004. [DOI] [PubMed] [Google Scholar]
- Luangsay S.; Wittamer V.; Bondue B.; De Henau O.; Rouger L.; Brait M.; Franssen J. D.; de Nadai P.; Huaux F.; Parmentier M. (2009) Mouse ChemR23 is expressed in dendritic cell subsets and macrophages, and mediates an anti-inflammatory activity of chemerin in a lung disease model. J. Immunol. 183, 6489–6499. 10.4049/jimmunol.0901037. [DOI] [PubMed] [Google Scholar]
- Mariani F.; Roncucci L. (2015) Chemerin/chemR23 axis in inflammation onset and resolution. Inflammation Res. 64, 85–95. 10.1007/s00011-014-0792-7. [DOI] [PubMed] [Google Scholar]
- Samson M.; Edinger A. L.; Stordeur P.; Rucker J.; Verhasselt V.; Sharron M.; Govaerts C.; Mollereau C.; Vassart G.; Doms R. W.; Parmentier M. (1998) ChemR23, a putative chemoattractant receptor, is expressed in monocyte-derived dendritic cells and macrophages and is a coreceptor for SIV and some primary HIV-1 strains. Eur. J. Immunol. 28, 1689–1700. . [DOI] [PubMed] [Google Scholar]
- Freire M. O.; Dalli J.; Serhan C. N.; Van Dyke T. E. (2017) Neutrophil Resolvin E1 Receptor Expression and Function in Type 2 Diabetes. J. Immunol. 198, 718–728. 10.4049/jimmunol.1601543. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gantz I.; Konda Y.; Yang Y. K.; Miller D. E.; Dierick H. A.; Yamada T. (1996) Molecular cloning of a novel receptor (CMKLR1) with homology to the chemotactic factor receptors. Cytogenet. Genome Res. 74, 286–290. 10.1159/000134436. [DOI] [PubMed] [Google Scholar]
- Zabel B. A.; Ohyama T.; Zuniga L.; Kim J. Y.; Johnston B.; Allen S. J.; Guido D. G.; Handel T. M.; Butcher E. C. (2006) Chemokine-like receptor 1 expression by macrophages in vivo: regulation by TGF-beta and TLR ligands. Exp. Hematol. 34, 1106–1114. 10.1016/j.exphem.2006.03.011. [DOI] [PubMed] [Google Scholar]
- Zabel B. A.; Silverio A. M.; Butcher E. C. (2005) Chemokine-like receptor 1 expression and chemerin-directed chemotaxis distinguish plasmacytoid from myeloid dendritic cells in human blood. J. Immunol. 174, 244–251. 10.4049/jimmunol.174.1.244. [DOI] [PubMed] [Google Scholar]
- Meder W.; Wendland M.; Busmann A.; Kutzleb C.; Spodsberg N.; John H.; Richter R.; Schleuder D.; Meyer M.; Forssmann W. G. (2003) Characterization of human circulating TIG2 as a ligand for the orphan receptor ChemR23. FEBS Lett. 555, 495–499. 10.1016/S0014-5793(03)01312-7. [DOI] [PubMed] [Google Scholar]
- Ohira T.; Arita M.; Omori K.; Recchiuti A.; Van Dyke T. E.; Serhan C. N. (2010) Resolvin E1 receptor activation signals phosphorylation and phagocytosis. J. Biol. Chem. 285, 3451–3461. 10.1074/jbc.M109.044131. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gao L.; Faibish D.; Fredman G.; Herrera B. S.; Chiang N.; Serhan C. N.; Van Dyke T. E.; Gyurko R. (2013) Resolvin E1 and chemokine-like receptor 1 mediate bone preservation. J. Immunol. 190, 689–694. 10.4049/jimmunol.1103688. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Herrera B. S.; Hasturk H.; Kantarci A.; Freire M. O.; Nguyen O.; Kansal S.; Van Dyke T. E. (2015) Impact of resolvin E1 on murine neutrophil phagocytosis in type 2 diabetes. Infect. Immun. 83, 792–801. 10.1128/IAI.02444-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arita M.; Ohira T.; Sun Y. P.; Elangovan S.; Chiang N.; Serhan C. N. (2007) Resolvin E1 selectively interacts with leukotriene B4 receptor BLT1 and ChemR23 to regulate inflammation. J. Immunol. 178, 3912–3917. 10.4049/jimmunol.178.6.3912. [DOI] [PubMed] [Google Scholar]
- Laguna-Fernandez A.; Checa A.; Carracedo M.; Artiach G.; Petri M. H.; Baumgartner R.; Forteza M. J.; Jiang X.; Andonova T.; Walker M. E.; Dalli J.; Arnardottir H.; Gistera A.; Thul S.; Wheelock C. E.; Paulsson-Berne G.; Ketelhuth D. F. J.; Hansson G. K.; Back M. (2018) ERV1/ChemR23 Signaling Protects Against Atherosclerosis by Modifying Oxidized Low-Density Lipoprotein Uptake and Phagocytosis in Macrophages. Circulation 138, 1693–1705. 10.1161/CIRCULATIONAHA.117.032801. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yokomizo T.; Izumi T.; Chang K.; Takuwa Y.; Shimizu T. (1997) A G-protein-coupled receptor for leukotriene B4 that mediates chemotaxis. Nature 387, 620–624. 10.1038/42506. [DOI] [PubMed] [Google Scholar]
- Tager A. M.; Luster A. D. (2003) BLT1 and BLT2: the leukotriene B4 receptors. Prostaglandins, Leukotrienes Essent. Fatty Acids 69, 123–134. 10.1016/S0952-3278(03)00073-5. [DOI] [PubMed] [Google Scholar]
- Wittamer V.; Franssen J. D.; Vulcano M.; Mirjolet J. F.; Le Poul E.; Migeotte I.; Brezillon S.; Tyldesley R.; Blanpain C.; Detheux M.; Mantovani A.; Sozzani S.; Vassart G.; Parmentier M.; Communi D. (2003) Specific recruitment of antigen-presenting cells by chemerin, a novel processed ligand from human inflammatory fluids. J. Exp. Med. 198, 977–985. 10.1084/jem.20030382. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Haworth O.; Cernadas M.; Yang R.; Serhan C. N.; Levy B. D. (2008) Resolvin E1 regulates interleukin 23, interferon-gamma and lipoxin A4 to promote the resolution of allergic airway inflammation. Nat. Immunol. 9, 873–879. 10.1038/ni.1627. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Unno Y.; Sato Y.; Fukuda H.; Ishimura K.; Ikeda H.; Watanabe M.; Tansho-Nagakawa S.; Ubagai T.; Shuto S.; Ono Y. (2018) Resolvin E1, but not resolvins E2 and E3, promotes fMLF-induced ROS generation in human neutrophils. FEBS Lett. 592, 2706–2715. 10.1002/1873-3468.13215. [DOI] [PubMed] [Google Scholar]
- Marazziti D.; Golini E.; Gallo A.; Lombardi M. S.; Matteoni R.; Tocchini-Valentini G. P. (1997) Cloning of GPR37, a Gene Located on Chromosome 7 Encoding a Putative G-Protein-Coupled Peptide Receptor, from a Human Frontal Brain EST Library. Genomics 45, 68–77. 10.1006/geno.1997.4900. [DOI] [PubMed] [Google Scholar]
- Donohue P. J.; Shapira H.; Mantey S. A.; Hampton L. L.; Jensen R. T.; Battey J. F. (1998) A human gene encodes a putative G protein-coupled receptor highly expressed in the central nervous system. Mol. Brain Res. 54, 152–160. 10.1016/S0169-328X(97)00336-7. [DOI] [PubMed] [Google Scholar]
- Lopes J. P.; Morato X.; Souza C.; Pinhal C.; Machado N. J.; Canas P. M.; Silva H. B.; Stagljar I.; Gandia J.; Fernandez-Duenas V.; Lujan R.; Cunha R. A.; Ciruela F. (2015) The role of parkinson’s disease-associated receptor GPR37 in the hippocampus: functional interplay with the adenosinergic system. J. Neurochem. 134, 135–146. 10.1111/jnc.13109. [DOI] [PubMed] [Google Scholar]
- Fujita-Jimbo E.; Yu Z. L.; Li H.; Yamagata T.; Mori M.; Momoi T.; Momoi M. Y. (2012) Mutation in Parkinson disease-associated, G-protein-coupled receptor 37 (GPR37/PaelR) is related to autism spectrum disorder. PLoS One 7, e51155. 10.1371/journal.pone.0051155. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Marazziti D.; Mandillo S.; Di Pietro C.; Golini E.; Matteoni R.; Tocchini-Valentini G. P. (2007) GPR37 associates with the dopamine transporter to modulate dopamine uptake and behavioral responses to dopaminergic drugs. Proc. Natl. Acad. Sci. U. S. A. 104, 9846–9851. 10.1073/pnas.0703368104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang H. J.; Vainshtein A.; Maik-Rachline G.; Peles E. (2016) G protein-coupled receptor 37 is a negative regulator of oligodendrocyte differentiation and myelination. Nat. Commun. 7, 10884. 10.1038/ncomms10884. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McCrary M. R.; Jiang M. Q.; Giddens M. M.; Zhang J. Y.; Owino S.; Wei Z. Z.; Zhong W.; Gu X.; Xin H.; Hall R. A.; Wei L.; Yu S. P. (2019) Protective effects of GPR37 via regulation of inflammation and multiple cell death pathways after ischemic stroke in mice. FASEB J. 33, 10680–10691. 10.1096/fj.201900070R. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhao Y.; Calon F.; Julien C.; Winkler J. W.; Petasis N. A.; Lukiw W. J.; Bazan N. G. (2011) Docosahexaenoic acid-derived neuroprotectin D1 induces neuronal survival via secretase- and PPARgamma-mediated mechanisms in Alzheimer’s disease models. PLoS One 6, e15816. 10.1371/journal.pone.0015816. [DOI] [PMC free article] [PubMed] [Google Scholar]
- White P. J.; Mitchell P. L.; Schwab M.; Trottier J.; Kang J. X.; Barbier O.; Marette A. (2015) Transgenic omega-3 PUFA enrichment alters morphology and gene expression profile in adipose tissue of obese mice: Potential role for protectins. Metab., Clin. Exp. 64, 666–676. 10.1016/j.metabol.2015.01.017. [DOI] [PubMed] [Google Scholar]
- Liao Z.; Dong J.; Wu W.; Yang T.; Wang T.; Guo L.; Chen L.; Xu D.; Wen F. (2012) Resolvin D1 attenuates inflammation in lipopolysaccharide-induced acute lung injury through a process involving the PPARgamma/NF-kappaB pathway. Respir Res. 13, 110. 10.1186/1465-9921-13-110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Perretti M.; Leroy X.; Bland E. J.; Montero-Melendez T. (2015) Resolution Pharmacology: Opportunities for Therapeutic Innovation in Inflammation. Trends Pharmacol. Sci. 36, 737–755. 10.1016/j.tips.2015.07.007. [DOI] [PubMed] [Google Scholar]
- Su S. B.; Gong W.; Gao J. L.; Shen W.; Murphy P. M.; Oppenheim J. J.; Wang J. M. (1999) A seven-transmembrane, G protein-coupled receptor, FPRL1, mediates the chemotactic activity of serum amyloid A for human phagocytic cells. J. Exp. Med. 189, 395–402. 10.1084/jem.189.2.395. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang D.; Chen Q.; Schmidt A. P.; Anderson G. M.; Wang J. M.; Wooters J.; Oppenheim J. J.; Chertov O. (2000) LL-37, the neutrophil granule- and epithelial cell-derived cathelicidin, utilizes formyl peptide receptor-like 1 (FPRL1) as a receptor to chemoattract human peripheral blood neutrophils, monocytes, and T cells. J. Exp. Med. 192, 1069. 10.1084/jem.192.7.1069. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arienti S.; Barth N. D.; Dorward D. A.; Rossi A. G.; Dransfield I. (2019) Regulation of Apoptotic Cell Clearance During Resolution of Inflammation. Front. Pharmacol. 10, 891. 10.3389/fphar.2019.00891. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Letunic I.; Bork P. (2019) Interactive Tree Of Life (iTOL) v4: recent updates and new developments. Nucleic Acids Res. 47, W256–W259. 10.1093/nar/gkz239. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arita M.; Oh S. F.; Chonan T.; Hong S.; Elangovan S.; Sun Y. P.; Uddin J.; Petasis N. A.; Serhan C. N. (2006) Metabolic inactivation of resolvin E1 and stabilization of its anti-inflammatory actions. J. Biol. Chem. 281, 22847–22854. 10.1074/jbc.M603766200. [DOI] [PubMed] [Google Scholar]
- Sun Y. P.; Tjonahen E.; Keledjian R.; Zhu M.; Yang R.; Recchiuti A.; Pillai P. S.; Petasis N. A.; Serhan C. N. (2009) Anti-inflammatory and pro-resolving properties of benzo-lipoxin A(4) analogs. Prostaglandins, Leukotrienes Essent. Fatty Acids 81, 357–366. 10.1016/j.plefa.2009.09.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cortina M. S.; Bazan H. E. (2011) Docosahexaenoic acid, protectins and dry eye. Curr. Opin. Clin. Nutr. Metab. Care 14, 132–137. 10.1097/MCO.0b013e328342bb1a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cholkar K.; Trinh H. M.; Vadlapudi A. D.; Wang Z.; Pal D.; Mitra A. K. (2015) Interaction Studies of Resolvin E1 Analog (RX-10045) with Efflux Transporters. J. Ocul. Pharmacol. Ther. 31, 248–255. 10.1089/jop.2014.0144. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Orr S. K.; Colas R. A.; Dalli J.; Chiang N.; Serhan C. N. (2015) Proresolving actions of a new resolvin D1 analog mimetic qualifies as an immunoresolvent. Am. J. Physiol Lung Cell Mol. Physiol 308, L904–911. 10.1152/ajplung.00370.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu S. H.; Chen X. Q.; Liu B.; Wu H. J.; Dong L. (2013) Efficacy and safety of 15(R/S)-methyl-lipoxin A(4) in topical treatment of infantile eczema. Br. J. Dermatol. 168, 172–178. 10.1111/j.1365-2133.2012.11177.x. [DOI] [PubMed] [Google Scholar]
- Kong X.; Wu S. H.; Zhang L.; Chen X. Q. (2017) Pilot application of lipoxin A4 analog and lipoxin A4 receptor agonist in asthmatic children with acute episodes. Exp. Ther. Med. 14, 2284–2290. 10.3892/etm.2017.4787. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Borgeson E.; Docherty N. G.; Murphy M.; Rodgers K.; Ryan A.; O’Sullivan T. P.; Guiry P. J.; Goldschmeding R.; Higgins D. F.; Godson C. (2011) Lipoxin A(4) and benzo-lipoxin A(4) attenuate experimental renal fibrosis. FASEB J. 25, 2967–2979. 10.1096/fj.11-185017. [DOI] [PubMed] [Google Scholar]
- Borgeson E.; Johnson A. M.; Lee Y. S.; Till A.; Syed G. H.; Ali-Shah S. T.; Guiry P. J.; Dalli J.; Colas R. A.; Serhan C. N.; Sharma K.; Godson C. (2015) Lipoxin A4 Attenuates Obesity-Induced Adipose Inflammation and Associated Liver and Kidney Disease. Cell Metab. 22, 125–137. 10.1016/j.cmet.2015.05.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stalder A. K.; Lott D.; Strasser D. S.; Cruz H. G.; Krause A.; Groenen P. M. A.; Dingemanse J. (2017) Biomarker-guided clinical development of the first-in-class anti-inflammatory FPR2/ALX agonist ACT-389949. Br. J. Clin. Pharmacol. 83, 476–486. 10.1111/bcp.13149. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lind S.; Sundqvist M.; Holmdahl R.; Dahlgren C.; Forsman H.; Olofsson P. (2019) Functional and signaling characterization of the neutrophil FPR2 selective agonist Act-389949. Biochem. Pharmacol. 166, 163–173. 10.1016/j.bcp.2019.04.030. [DOI] [PubMed] [Google Scholar]
- Ferguson S. S. G.; Downey W. E.; Colapietro A.-M.; Barak L. S.; Ménard L.; Caron M. G. (1996) Role of β-Arrestin in Mediating Agonist-Promoted G Protein-Coupled Receptor Internalization. Science 271, 363–366. 10.1126/science.271.5247.363. [DOI] [PubMed] [Google Scholar]
- Spite M.; Hellmann J.; Tang Y.; Mathis S. P.; Kosuri M.; Bhatnagar A.; Jala V. R.; Haribabu B. (2011) Deficiency of the Leukotriene B4 Receptor, BLT-1, Protects against Systemic Insulin Resistance in Diet-Induced Obesity. J. Immunol. 187, 1942–1949. 10.4049/jimmunol.1100196. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bozinovski S.; Uddin M.; Vlahos R.; Thompson M.; McQualter J. L.; Merritt A.-S.; Wark P. A. B.; Hutchinson A.; Irving L. B.; Levy B. D.; Anderson G. P. (2012) Serum amyloid A opposes lipoxin A(4) to mediate glucocorticoid refractory lung inflammation in chronic obstructive pulmonary disease. Proc. Natl. Acad. Sci. U. S. A. 109, 935–940. 10.1073/pnas.1109382109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pamplona F. A.; Ferreira J.; Menezes de Lima O. J.; Duarte F. S.; Bento A. F.; Forner S.; Villarinho J. G.; Bellocchio L.; Wotjak C. T.; Lerner R.; Monory K.; Lutz B.; Canetti C.; Matias I.; Calixto J. B.; Marsicano G.; Guimaraes M. Z. P.; Takahashi R. N. (2012) Anti-inflammatory lipoxin A4 is an endogenous allosteric enhancer of CB1 cannabinoid receptor. Proc. Natl. Acad. Sci. U. S. A. 109, 21134–21139. 10.1073/pnas.1202906109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Di Virgilio F.; Giuliani A. L.; Vultaggio-Poma V.; Falzoni S.; Sarti A. C. (2018) Non-nucleotide Agonists Triggering P2 × 7 Receptor Activation and Pore Formation. Front. Pharmacol. 9, 3260. 10.3389/fphar.2018.00039. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang Z.; Cherryholmes G.; Chang F.; Rose D. M.; Schraufstatter I.; Shively J. E. (2009) Evidence that cathelicidin peptide LL-37 may act as a functional ligand for CXCR2 on human neutrophils. Eur. J. Immunol. 39, 3181–3194. 10.1002/eji.200939496. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kenakin T.; Christopoulos A. (2013) Signalling bias in new drug discovery: detection, quantification and therapeutic impact. Nat. Rev. Drug Discovery 12, 205–216. 10.1038/nrd3954. [DOI] [PubMed] [Google Scholar]
- Wisler J. W.; DeWire S. M.; Whalen E. J.; Violin J. D.; Drake M. T.; Ahn S.; Shenoy S. K.; Lefkowitz R. J. (2007) A unique mechanism of beta-blocker action: carvedilol stimulates beta-arrestin signaling. Proc. Natl. Acad. Sci. U. S. A. 104, 16657–16662. 10.1073/pnas.0707936104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Violin J. D.; Crombie A. L.; Soergel D. G.; Lark M. W. (2014) Biased ligands at G-protein-coupled receptors: promise and progress. Trends Pharmacol. Sci. 35, 308–316. 10.1016/j.tips.2014.04.007. [DOI] [PubMed] [Google Scholar]
- Kenakin T. (2015) The Effective Application of Biased Signaling to New Drug Discovery. Mol. Pharmacol. 88, 1055. 10.1124/mol.115.099770. [DOI] [PubMed] [Google Scholar]
- Gesty-Palmer D.; Luttrell L. M. (2011) Biasing” the parathyroid hormone receptor: a novel anabolic approach to increasing bone mass?. Br. J. Pharmacol. 164, 59–67. 10.1111/j.1476-5381.2011.01450.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Smith J. S.; Nicholson L. T.; Suwanpradid J.; Glenn R. A.; Knape N. M.; Alagesan P.; Gundry J. N.; Wehrman T. S.; Atwater A. R.; Gunn M. D.; MacLeod A. S.; Rajagopal S. (2018) Biased agonists of the chemokine receptor CXCR3 differentially control chemotaxis and inflammation. Sci. Signaling 11, eaaq1075. 10.1126/scisignal.aaq1075. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Violin J. D.; DeWire S. M.; Yamashita D.; Rominger D. H.; Nguyen L.; Schiller K.; Whalen E. J.; Gowen M.; Lark M. W. (2010) Selectively engaging beta-arrestins at the angiotensin II type 1 receptor reduces blood pressure and increases cardiac performance. J. Pharmacol. Exp. Ther. 335, 572–579. 10.1124/jpet.110.173005. [DOI] [PubMed] [Google Scholar]
- Berchiche Y. A.; Sakmar T. P. (2016) CXC Chemokine Receptor 3 Alternative Splice Variants Selectively Activate Different Signaling Pathways. Mol. Pharmacol. 90, 483–495. 10.1124/mol.116.105502. [DOI] [PubMed] [Google Scholar]
- Riddy D. M.; Cook A. E.; Diepenhorst N. A.; Bosnyak S.; Brady R.; Mannoury-la-Cour C.; Mocaer E.; Summers R. J.; Charman W. N.; Sexton P. M.; Christopoulos A.; Langmead C. J. (2017) Isoform-Specific Biased Agonism of Histamine H3 Receptor Agonists. Mol. Pharmacol. 91, 87–99. 10.1124/mol.116.106153. [DOI] [PubMed] [Google Scholar]
- Herenbrink C. K.; Sykes D. A.; Donthamsetti P.; Canals M.; Coudrat T.; Shonberg J.; Scammells P. J.; Capuano Ben; Sexton P. M.; Charlton S. J.; Javitch J. A.; Christopoulos A.; Lane J. R. (2016) The role of kinetic context in apparent biased agonism at GPCRs. Nat. Commun. 7, 10842. 10.1038/ncomms10842. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sykes D. A.; Stoddart L. A.; Kilpatrick L. E.; Hill S. J. (2019) Binding kinetics of ligands acting at GPCRs. Mol. Cell. Endocrinol. 485, 9–19. 10.1016/j.mce.2019.01.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Riddy D. M.; Cook A. E.; Shackleford D. M.; Pierce T. L.; Mocaer E.; Mannoury la Cour C.; Sors A.; Charman W. N.; Summers R. J.; Sexton P. M.; Christopoulos A.; Langmead C. J. (2019) Drug-receptor kinetics and sigma-1 receptor affinity differentiate clinically evaluated histamine H3 receptor antagonists. Neuropharmacology 144, 244–255. 10.1016/j.neuropharm.2018.10.028. [DOI] [PubMed] [Google Scholar]
- Riddy D. M.; Valant C.; Rueda P.; Charman W. N.; Sexton P. M.; Summers R. J.; Christopoulos A.; Langmead C. J. (2015) Label-Free Kinetics: Exploiting Functional Hemi-Equilibrium to Derive Rate Constants for Muscarinic Receptor Antagonists. Mol. Pharmacol. 88, 779–790. 10.1124/mol.115.100545. [DOI] [PubMed] [Google Scholar]
- Thompson G. L.; Lane J. R.; Coudrat T.; Sexton P. M.; Christopoulos A.; Canals M. (2015) Biased Agonism of Endogenous Opioid Peptides at the μ-Opioid Receptor. Mol. Pharmacol. 88, 335–346. 10.1124/mol.115.098848. [DOI] [PubMed] [Google Scholar]
- Teixeira L. B.; Parreiras-E-Silva L. T.; Bruder-Nascimento T.; Duarte D. A.; Simões S. C.; Costa R. M.; Rodríguez D. Y.; Ferreira P. A. B.; Silva C. A. A.; Abrao E. P.; Oliveira E. B.; Bouvier M.; Tostes R. C.; Costa-Neto C. M. (2017) Ang-(1–7) is an endogenous β-arrestin-biased agonist of the AT(1) receptor with protective action in cardiac hypertrophy. Sci. Rep. 7, 11903–11903. 10.1038/s41598-017-12074-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gabl M.; Holdfeldt A.; Sundqvist M.; Lomei J.; Dahlgren C.; Forsman H. (2017) FPR2 signaling without beta-arrestin recruitment alters the functional repertoire of neutrophils. Biochem. Pharmacol. 145, 114–122. 10.1016/j.bcp.2017.08.018. [DOI] [PubMed] [Google Scholar]
- Holdfeldt A.; Skovbakke S. L.; Gabl M.; Nielsen C.; Dahlgren C.; Franzyk H.; Forsman H. (2019) Structure–Function Characteristics and Signaling Properties of Lipidated Peptidomimetic FPR2 Agonists: Peptoid Stereochemistry and Residues in the Vicinity of the Headgroup Affect Function. ACS Omega 4, 5968–5982. 10.1021/acsomega.9b00098. [DOI] [Google Scholar]
- Qin C. X.; May L. T.; Li R.; Cao N.; Rosli S.; Deo M.; Alexander A. E.; Horlock D.; Bourke J. E.; Yang Y. H.; Stewart A. G.; Kaye D. M.; Du X.-J.; Sexton P. M.; Christopoulos A.; Gao X.-M.; Ritchie R. H. (2017) Small-molecule-biased formyl peptide receptor agonist compound 17b protects against myocardial ischaemia-reperfusion injury in mice. Nat. Commun. 8, 14232. 10.1038/ncomms14232. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Felce J. H.; Latty S. L.; Knox R. G.; Mattick S. R.; Lui Y.; Lee S. F.; Klenerman D.; Davis S. J. (2017) Receptor Quaternary Organization Explains G Protein-Coupled Receptor Family Structure. Cell Rep. 20, 2654–2665. 10.1016/j.celrep.2017.08.072. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Changeux J.-P.; Christopoulos A. (2016) Allosteric Modulation as a Unifying Mechanism for Receptor Function and Regulation. Cell 166, 1084–1102. 10.1016/j.cell.2016.08.015. [DOI] [PubMed] [Google Scholar]
- Wootten D.; Miller L. J.; Koole C.; Christopoulos A.; Sexton P. M. (2017) Allostery and Biased Agonism at Class B G Protein-Coupled Receptors. Chem. Rev. 117, 111–138. 10.1021/acs.chemrev.6b00049. [DOI] [PubMed] [Google Scholar]
- Langmead C. J. (2018) New Frontiers in G protein-coupled receptor drug discovery. Drug Target Rev. 4, 44–47. [Google Scholar]
- Hanson M. A.; Cherezov V.; Griffith M. T.; Roth C. B.; Jaakola V.-P.; Chien E. Y. T.; Velasquez J.; Kuhn P.; Stevens R. C. (2008) A specific cholesterol binding site is established by the 2.8 A structure of the human beta2-adrenergic receptor. Structure 16, 897–905. 10.1016/j.str.2008.05.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pei Y.; Mercier R. W.; Anday J. K.; Thakur G. A.; Zvonok A. M.; Hurst D.; Reggio P. H.; Janero D. R.; Makriyannis A. (2008) Ligand-binding architecture of human CB2 cannabinoid receptor: evidence for receptor subtype-specific binding motif and modeling GPCR activation. Chem. Biol. 15, 1207–1219. 10.1016/j.chembiol.2008.10.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hurst D. P.; Schmeisser M.; Reggio P. H. (2013) Endogenous lipid activated G protein-coupled receptors: Emerging structural features from crystallography and molecular dynamics simulations. Chem. Phys. Lipids 169, 46–56. 10.1016/j.chemphyslip.2013.01.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bokoch M. P.; Jo H.; Valcourt J. R.; Srinivasan Y.; Pan A. C.; Capponi S.; Grabe M.; Dror R. O.; Shaw D. E.; DeGrado W. F.; Coughlin S. R. (2018) Entry from the Lipid Bilayer: A Possible Pathway for Inhibition of a Peptide G Protein-Coupled Receptor by a Lipophilic Small Molecule. Biochemistry 57, 5748–5758. 10.1021/acs.biochem.8b00577. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Szlenk C. T.; GC J. B.; Natesan S. (2019) Does the lipid bilayer orchestrate access and binding of ligands to transmembrane orthosteric/allosteric sites of GPCRs?. Mol. Pharmacol. 96, 527. 10.1124/mol.118.115113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Edelstein S. J.; Changeux J.-P. (2016) Biased Allostery. Biophys. J. 111, 902–908. 10.1016/j.bpj.2016.07.044. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gentry P. R.; Sexton P. M.; Christopoulos A. (2015) Novel Allosteric Modulators of G Protein-coupled Receptors. J. Biol. Chem. 290, 19478–19488. 10.1074/jbc.R115.662759. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Marlo J. E.; Niswender C. M.; Days E. L.; Bridges T. M.; Xiang Y.; Rodriguez A. L.; Shirey J. K.; Brady A. E.; Nalywajko T.; Luo Q.; Austin C. A.; Williams M. B.; Kim K.; Williams R.; Orton D.; Brown H. A.; Lindsley C. W.; Weaver C. D.; Conn P. J. (2009) Discovery and characterization of novel allosteric potentiators of M1 muscarinic receptors reveals multiple modes of activity. Mol. Pharmacol. 75, 577–588. 10.1124/mol.108.052886. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sengmany K.; Singh J.; Stewart G. D.; Conn P. J.; Christopoulos A.; Gregory K. J. (2017) Biased allosteric agonism and modulation of metabotropic glutamate receptor 5: Implications for optimizing preclinical neuroscience drug discovery. Neuropharmacology 115, 60–72. 10.1016/j.neuropharm.2016.07.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hellyer S. D.; Albold S.; Wang T.; Chen A. N. Y.; May L. T.; Leach K.; Gregory K. J. (2018) Selective” Class C G Protein-Coupled Receptor Modulators Are Neutral or Biased mGlu5 Allosteric Ligands. Mol. Pharmacol. 93, 504–514. 10.1124/mol.117.111518. [DOI] [PubMed] [Google Scholar]
- Cook A. E.; Mistry S. N.; Gregory K. J.; Furness S. G. B.; Sexton P. M.; Scammells P. J.; Conigrave A. D.; Christopoulos A.; Leach K. (2015) Biased allosteric modulation at the CaS receptor engendered by structurally diverse calcimimetics. Br. J. Pharmacol. 172, 185–200. 10.1111/bph.12937. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Goupil E.; Tassy D.; Bourguet C.; Quiniou C.; Wisehart V.; Petrin D.; Le Gouill C.; Devost D.; Zingg H. H.; Bouvier M.; Saragovi H. U.; Chemtob S.; Lubell W. D.; Claing A.; Hébert T. E.; Laporte S. A. (2010) A novel biased allosteric compound inhibitor of parturition selectively impedes the prostaglandin F2alpha-mediated Rho/ROCK signaling pathway. J. Biol. Chem. 285, 25624–25636. 10.1074/jbc.M110.115196. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Burger W. A. C.; Sexton P. M.; Christopoulos A.; Thal D. M. (2018) Toward an understanding of the structural basis of allostery in muscarinic acetylcholine receptors. J. Gen. Physiol. 150, 1360. 10.1085/jgp.201711979. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang W.; Qiao Y.; Li Z. (2018) New Insights into Modes of GPCR Activation. Trends Pharmacol. Sci. 39, 367–386. 10.1016/j.tips.2018.01.001. [DOI] [PubMed] [Google Scholar]
- Jong Y.-J. I.; Harmon S. K.; O’Malley K. L. (2018) Intracellular GPCRs Play Key Roles in Synaptic Plasticity. ACS Chem. Neurosci. 9, 2162–2172. 10.1021/acschemneuro.7b00516. [DOI] [PMC free article] [PubMed] [Google Scholar]




