Phagocytes orchestrate acute inflammation and host defense. Here we carried out lipid mediator (LM) metabololipidomics profiling distinct phagocytes: neutrophils (PMN), apoptotic PMN, and macrophages. Efferocytosis increased specialized pro-resolving mediator (SPM) biosynthesis, including Resolvin D1 (RvD1), RvD2, and RvE2, which were further elevated by PMN microparticles. Apoptotic PMN gave elevated prostaglandin E2, lipoxin B4 and RvE2, whereas zymosan-stimulated PMN showed predominantly leukotriene B4 and 20-OH-leukotriene B4, as well as lipoxin marker 5,15-diHETE. Using deuterium-labeled precursors (d8-arachidonic acid, d5-eicosapentaenoic acid, and d5-docosahexaenoic acid), we found that apoptotic PMN and microparticles contributed to SPM biosynthesis during efferocytosis. M2 macrophages produced SPM including maresin-1 (299 ± 8 vs 45 ± 6 pg/2.5 × 105 cells; P < .01) and lower amounts of leukotriene B4 and prostaglandin than M1. Apoptotic PMN uptake by both macrophage subtypes led to modulation of their LM profiles. Leukotriene B4 was down-regulated in M2 (668 ± 81 vs 351 ± 39 pg/2.5 × 105 cells; P < .01), whereas SPM including lipoxin A4 (977 ± 173 vs 675 ± 167 pg/2.5 × 105 cells; P < .05) were increased. Conversely, uptake of apoptotic PMN by M2 macrophages reduced (∼ 25%) overall LM. Together, these results establish LM signature profiles of human phagocytes and related subpopulations. Moreover, they provide evidence for microparticle regulation of specific endogenous LM during defined stages of the acute inflammatory process and their dynamic changes in human primary phagocytes.

Inflammation is the organisms' response to local injury in vascularized tissues programmed to traffic leukocytes and plasma delivery to an injured site or point of bacterial invasion,1  this protective response when uncontrolled in humans is associated with many widely occurring diseases. These include cardiovascular, metabolic, and the classic inflammatory diseases (ie, arthritis and periodontal disease) along with cancers.2  Nonresolving inflammation is now widely acknowledged as a major driver in most of these diseases.3  Resolution of inflammation and dissipation of the local chemical messengers involved in mounting the innate response were thought to be passively diluted with time at the site, hence stopping further leukocyte recruitment and resolving the exudate or battlefield of inflammation.4,5 

Results from this laboratory indicate that resolution of self-limited inflammatory exudates is a biochemically active process that involves the local and temporal biosynthesis of a new genus of specialized pro-resolving mediators (SPMs) with their novel functions mapped employing resolution indices.5-8  SPMs encompass several families of structurally and chemically distinct mediators. These chemical mediator families include lipoxins biosynthesized from arachidonic acid, E-series resolvins (Rv) from eicosapentaenoic acid (EPA), and docosahexaenoic acid (DHA)-derived D-series resolvins, protectins, and maresins. Each potent bioactive member of these families shares a defining action in resolving local inflammation. By definition, they each limit further neutrophil (PMN) recruitment to the site of injury and/or microbial invasion and enhance macrophage uptake of cellular debris and apoptotic PMN to bring about tissue homeostasis.2  Along with these defining properties, specific SPMs carry out more specialized tasks within programmed resolution; hence, the scope of their individual actions are nonoverlapping and evoked via specific cell surface receptors that are G-protein–coupled receptors.2  A systems approach led to the identification of novel bioactive structures coined resolvins and protectins in murine inflammatory exudates and isolated human cells based on liquid chromatography tandem mass spectrometry (LC-MS-MS)–based lipid mediator lipidomics and tandem assessment of their functions in anti-inflammation and pro-resolution.6,9  The complete stereochemistry and total organic synthesis of several key resolvins, protectins, as well as their aspirin-triggered forms are established.10  These include resolvin D1 (7S, 8R, 17S-trihydroxy-4Z, 9E, 11E, 13Z, 15E, 19Z-docosahexaenoic acid), resolvin D2 (7S, 16R, 17S-trihydroxy-4Z, 8E, 10Z, 12E, 14E, 19Z-DHA), 17R-HDHA (17R-hydroxy-4Z,7Z,10Z,13Z,15E,19Z-DHA), neuroprotectin D1/NPD1 (10R, 17S-dihydroxy-4Z, 7Z, 11E, 13E, 15Z, 19Z-DHA), resolvin E1 (5S, 12R, 18R-trihydroxy-6Z, 8E, 10E, 14Z, 16E-eicosapentaenoic acid), and most recently maresin 1 (7R, 14S-dihydroxy-4Z,8E,10E,12Z,16Z,19Z-DHA).11  In addition to confirming the original structural assignments and potent anti-inflammatory and pro-resolving actions in vivo of resolvins, lipoxins, and maresins,10  new studies from others demonstrate their potent actions in experimental colitis,12  arthritis,13  arthritic pain,14  ocular diseases,15  resolving obesity,16  and diabetes.17  Importantly, synthetic SPMs permitted their identification in other biologic sources, including, for example, human inflammatory responses,18,19  human serum,20  fish,21  and other marine organisms22  and the invertebrate phyla.11 

The initial identification of resolvins and protectins in resolving exudates implicated that multiple cell types are involved in their biosynthesis given the dynamic process ongoing within evolving inflammatory exudates.5,6  With authenticated SPMs and pro-inflammatory lipid mediator (LM), we can now carry out targeted LM metabololipidomics via profiling monitoring > 50 individual mediators and pathway markers from these autacoid pathways in initiation and resolution of inflammation.23  Microparticles (MPs) in human synovium can influence the course of inflammation.24  Because evolving self-limited inflammatory exudates produce functional MPs that also signal to stimulate resolution of inflammation in mice,25  in this report we systematically profiled LM and SPM produced by individual human cell types and MP involved in initiation and gauge their contributions to resolution of inflammatory responses. Employing a targeted LM metabololipidomic approach, we found that apoptotic PMNs possess a pro-resolving LM profile and their uptake stimulated SPM biosynthesis in macrophages, a process that is also regulated by PMN MPs and transcellular biosynthesis.

Materials

Materials include the following: lipopolysaccharide (LPS)–serotype 055B:B5, Ficoll-Histopaque 1077–1 (Sigma-Aldrich); hr-GM-CSF, hr-IFN-γ, hr-IL-4, hr-M-CSF (R&D Systems), phenol red free RPMI 1640 (Invitrogen); pooled human Serum (Lonza); human monocyte isolation kit (StemCell Technologies); mouse anti–human CD80-FITC, anti-CD163–PE, CD206-FITC, CD68-FITC, CD80-PE, CD54-APC, HLA-DR–PE/Cy5.5 (BioLegend); carboxyfluorescein diacetate succinimidyl ester (CFDA; Invitrogen); LC grade solvents (Fisher Scientific); Luna C18 columns (Phenomenex); C18 SPE columns (Waters); fluorescently conjugated zymosan A (Invitrogen); arachidonic acid (AA), EPA, and DHA were purchased from Cayman Chemical along with synthetic standards for LC-MS/MS quantification and deuterated internal standards d8-5S-HETE, d4-leukotriene B4 (LTB4), d5-lipoxin A4 (LXA4) and d4-prostaglandin E2 (PGE2), d8-AA, d5-EPA, and d5-DHA. Fresh human leukocytes were prepared from deidentified peripheral blood of healthy volunteers obtained by venipuncture in accordance with Partners Human Research Committee Protocol number 1999P001297. All participants gave written informed consent. This study was conducted in accordance with the Declaration of Helsinki.

Macrophage cultures

Human PBMCs from deidentified healthy human volunteers from the Children's Hospital Boston blood bank were isolated by density-gradient Ficoll-Histopaque isolation11  and monocytes purified using monocyte isolation kit (StemCell Technologies) yielding a 93%-97% CD14+ population. These were then cultured 7 days in RPMI 1640 (10% human serum). Macrophages were differentiated using 20 ng/mL GM-CSF for 7 days. Macrophage preparations were then incubated with fluorescently labeled anti-CD3 and anti-CD19 antibodies (lymphocyte markers). These macrophage incubations were > 95% CD3-CD19 negative (see supplemental Figure 1, available on the Blood Web site; see the Supplemental Materials link at the top of the online article). M1 macrophages were obtained by incubating isolated monocytes 7 days with 20 ng/mL GM-CSF, and stimulating for 24 hours with 20 ng/mL IFN-γ and 1 ng/mL LPS. M2 macrophages were obtained by incubating monocytes with 20 ng/mL M-CSF 7 days and stimulating for 48 hours with 20 ng/mL IL-4.26  Phenotypic linage of the human monocyte-derived macrophages was assessed by flow cytometry using fluorescently conjugated antibodies in accordance with published criteria.26 

Phagocytosis

Macrophage subtypes were obtained as described in “Macrophage cultures.” Human PMNs were isolated by density-gradient Ficoll-Histopaque isolation; these were then cultured overnight at 5 × 106 cells/mL in Dulbecco PBS (DPBS) to obtain apoptotic neutrophils (> 90% AnxAV+; supplemental Figure 2). In some cases, the cells were then counted and stained using CFDA.11 

Labeled apoptotic PMNs were added (3:1 ratio, PMN to macrophages) to differentiated macrophages in the presence of vehicle or MPs. Incubations (60 minutes, 37°C) were stopped after 1 hour. Subsequently, the cells were washed with PBS to remove nonphagocytosed PMN and extracellular fluorescence quenched using trypan blue (1:50 dilution). The number of phagocytosed cells (ie, intracellular) was determined using a SpectraMax M3 plate reader (Molecular Devices) monitoring fluorescence emission at 525 nm. To quantify the number of phagocytosed PMN, calibration curves were obtained using fluorescently labeled apoptotic PMNs that gave linear curves with an r2 value > 0.9. In separate experiments, apoptotic PMNs were incubated with macrophages with or without MPs for 0 minutes or 60 minutes (37°C).

PMN MPs

Human neutrophil MPs were prepared from peripheral blood PMN.25  Briefly, neutrophils were prepared following density separation by layering on Ficoll-Histopaque 1077-1. Cells were centrifuged (300g, 30 minutes, 4°C) and red blood cells lysed by hypotonic lyses. PMN purity was assessed by flow cytometry after cell staining with fluorescently conjugated antibodies to CD3, CD14, CD16, CD11b with > 99% of the cells negative for CD14 and CD3. This suggests little to no monocytic or lymphocytic contamination (see supplemental Figure 2). Neutrophils were suspended in DPBS (2 × 107 cells/mL), stimulated with formyl-methionyl-leucyl-phenylalanine (1μM; 20 minutes, 37°C). Cells were removed by 2 successive centrifugations (3000g; 10 minutes, 4°C), before pelleting MPs by ultracentrifugation (100 000g; 1 hour, 4°C). PMN MPs were characterized according to their forward and side scatter parameters and found to be > 85% AnxA V+ and > 95% CD66b+ (Figure 1), in accordance with published criteria.25  MPs then were suspended in DPBS+/+ and stained using fluorescently labeled anti-CD66b (specific marker for neutrophil MPs). Subsequently, MP numbers in 100 μL buffer were assessed as a function of CD66b positive events using flow cytometry.

Macrophage and MP coincubations

To assess the impact of MPs on macrophage LM biosynthesis, macrophages (2.5 × 105 cells/2 mL) were incubated with or without 5 × 105 MPs (15 minutes, 37°C, pH 7.45). Incubations were stopped by adding ice-cold MeOH followed by extraction and LM profiles identified by LM metabololipidomics (see “Sample extraction and lipid mediator metabololipidomics”). In designated incubations, macrophages were incubated with pertussis toxin (1 μg/mL, 37°C, overnight) or cholera toxin (1 μg/mL, 37°C, 2 hours) before the addition of MPs to the incubations (15 minutes, 37°C, pH 7.45).

Sample extraction and lipid mediator metabololipidomics

All samples for LC-MS/MS analysis were extracted using SPE columns as described previously.23  Briefly, columns were equilibrated with 1 column volume of methanol and 2 volumes ddH2O. Before extraction, 500 pg of deuterium-labeled internal standards d8-5S-HETE, d4LTB4, d5LXA4, and d4PGE2 were added to facilitate quantification of sample recovery.

Sample supernatants were diluted with 10 volumes of ddH2O, acidified (pH ∼ 3.5) and immediately loaded onto the SPE column. After loading, columns were washed with 1 volume of neutral ddH2O and hexane. Samples were eluted with 6 mL methyl formate and taken to dryness using Speedvac or nitrogen stream. Samples were suspended in methanol/water for LC-MS/MS. The LC-UV-MS/MS system, QTrap 5500 (ABSciex) was equipped with an Agilent HP1100 binary pump and diode-array detector. An Agilent Eclipse Plus C18 column (50 mm × 4.6 mm × 1.8 μm or 100 mm × 4.6 mm × 1.8 μm) was used with a gradient of methanol/water/acetic acid of 60:40:0.01 (volume/volume/volume) to 100:0:0.01 at 0.5-mL/min flow rate. To monitor and quantify the levels of the various LMs, a multiple reaction monitoring (MRM) method was developed with signature ion fragments for each molecule. Identification was conducted using published criteria23  with at least 6 diagnostic ions. Calibration curves were obtained using synthetic and authentic LM mixtures (they included d8-5S-HETE, d4-LTB4, d5-LXA4, d4-PGE2, RvD1, RvD2, RvD5, PD1, MaR1, RvE1, RvE2, LXA4, LXB4, LXA5, PGE2, PGD2, PGF, thromboxane B2 (TXB2), PGE3, PGF, TXB3, LTB4, 17-HDHA, 14-HDHA, 7-HDHA, 4-HDHA, 18-HEPE, 15-HEPE, 12-HEPE, 5-HEPE, 15-HETE, 12-HETE, and 5-HETE) at 12.5, 25, 50, 100 pg. Linear calibration curves for each were obtained with r2 values in the range 0.98-0.99. Quantification was carried out based on peak area of the MRM transition and the linear calibration curve for each compound. In case synthetic or biogenic standard for a given product was not available (ie, LXB5, PGD3), calibration curves for products with similar chromatographic behaviors (ie, tri-, di-, or mono-HETEs) to the analyte of interest were used.

Human PMN incubations

Human PMNs isolated as described in “PMN MPs” were either placed directly in methanol, suspended in DPBS (5 × 106 cell/mL), and stimulated with 0.1 mg zymosan or incubated overnight at 37°C, and apoptosis was assessed using propidium iodide and annexin (AnxA) V staining (see supplemental Figure 2). Incubations were stopped with 2 volumes of cold methanol. In designated experiments, PMNs were plated in 12-well (2 mL/well) with or without AA, EPA, and DHA (each at 100 ng) and incubated for the indicated time intervals. The incubations were stopped with 2 volumes of cold methanol and held at −80°C before solid phase extraction.

LM biosynthesis in MP-leukocyte coincubations

PMNs were incubated with 100 ng each of d8-AA, d5-EPA, or d5-DHA for 5 minutes before addition of formyl-methionyl-leucyl-phenylalanine (1μM, 30 minutes, 37°C). The cell suspensions were centrifuged and MPs were obtained as described in “PMN MPs.” Next, apoptotic PMNs were incubated with d8-AA, d5-EPA, or d5-DHA (100 ng each) for 30 minutes (37°C). To assess the contribution of MPs to transcellular biosynthesis of LM, deuterium-labeled MPs were coincubated with unlabeled PMNs and macrophages (1 hour, 37°C). Incubations were then extracted, products identified and quantitated using LC-MS/MS. Similarly, contribution(s) of apoptotic PMNs were determined using deuterium-labeled precursors in apoptotic PMNs incubated with unlabeled macrophages (2.5 × 105 cells/2 mL) and MPs (5 × 105 MP/incubation; 30 minutes, 37°C, pH 7.45).

Statistics

Results are mean ± SEM. Multiple-group comparisons were made using either paired Student t test or 1-way ANOVA followed by Dunnett or Bonferroni posthoc analysis. The P values < .05 were considered significant.

MPs stimulate efferocytosis of apoptotic PMN

Because PMN MPs exert anti-inflammatory and pro-resolving actions in vivo,25  we assessed their ability to regulate macrophage uptake of apoptotic cells (efferocytosis).27  PMN MPs were obtained from activated PMNs (Figure 1A) and incubated with macrophages before addition of fluorescent apoptotic PMNs (Figure 1), dose-dependently increased macrophage-associated fluorescence (Figure 1B). These results indicate that MPs regulate efferocytosis of apoptotic cells.

Figure 1

PMN MPs enhance macrophage efferocytosis of apoptotic PMN. PMN MPs were obtained after PMN stimulation with formyl-methionyl-leucyl-phenylalanine (1μM). (A) The MP population (left panel) was monitored by flow cytometry using fluorescently conjugated AnxAV (right top panel) and anti-CD66b (right bottom panel). (B) Macrophages were prepared by differentiating peripheral blood monocytes in the presence of GM-CSF for 7 days. These were incubated with MP for 5 minutes before the addition of apoptotic (Apo) PMN. Uptake of the CFDA-labeled Apo PMN (3 × 105 cells/well) was monitored after 1-hour incubation (37°C) by assessing the levels of fluorescence (see “Phagocytosis”). (B) Results are expressed as mean ± SEM (n = 4 distinct cell preparations). *P < .05 vs macrophage plus PMN group. **P < .01 vs macrophage plus Apo PMN group.

Figure 1

PMN MPs enhance macrophage efferocytosis of apoptotic PMN. PMN MPs were obtained after PMN stimulation with formyl-methionyl-leucyl-phenylalanine (1μM). (A) The MP population (left panel) was monitored by flow cytometry using fluorescently conjugated AnxAV (right top panel) and anti-CD66b (right bottom panel). (B) Macrophages were prepared by differentiating peripheral blood monocytes in the presence of GM-CSF for 7 days. These were incubated with MP for 5 minutes before the addition of apoptotic (Apo) PMN. Uptake of the CFDA-labeled Apo PMN (3 × 105 cells/well) was monitored after 1-hour incubation (37°C) by assessing the levels of fluorescence (see “Phagocytosis”). (B) Results are expressed as mean ± SEM (n = 4 distinct cell preparations). *P < .05 vs macrophage plus PMN group. **P < .01 vs macrophage plus Apo PMN group.

Close modal

Apoptotic PMN and MPs stimulate macrophages SPM

Next, we assessed endogenous pro- and anti-inflammatory LM biosynthesis by human macrophages (supplemental Figure 1B) during uptake of apoptotic PMNs using targeted LC-MS/MS metabololipidomics, emphasizing functional lipidomics (see supplemental Table 1 and references within). We identified mediators from both lipoxygenase and cyclooxygenase pathways, including RvD1, MaR1, PGE2, and PGF2α, (Figure 2A). All were identified in accordance with published criteria,23  which included matching retention times, fragmentation patterns, and at least 6 characteristic and diagnostic ions for each as illustrated with results obtained for RvD1 (eg, 375 = M-H; 357 = M-H-H2O; 339 = M-H-2H2O; 331 = M-H-CO2; 313 = M-H-H2O-CO2; 295 = M-H-2H2O-CO2; 287 = 303-H2O; 241 = 277-H2O), RvD2, LXB4, and 5,15-diHETE (Figure 2B and insets for diagnostic ions).

Figure 2

SPM biosynthesis during macrophage efferocytosis of apoptotic PMNs is enhanced by PMN MPs. Macrophages were incubated for 5 minutes with PMN MPs before the addition of apoptotic PMNs (7.5 × 105 cells/well). Cells were incubated for 1 hour at 37°C. At the end of the incubation, ice-cold MeOH was added and the products extracted. (A) Representative MRM traces for the identified LMs. (B) Accompanying MS/MS spectra used for identification. (C) Specific bioactive lipid mediator and precursor/pathway markers where: Q1, M-H (parent ion); and Q3, diagnostic ion in the MS-MS (daughter ion) along with mean ± SEM values for each of the mediators identified. Quantification and values obtained after PMN (3 × 105 PMN) and MP (2 × 105 MP) incubations. The detection limit was ∼ 1 pg. *Below limits. (D) D-series resolvins, protectins, and maresins. (E) Lipoxins. (F) Prostaglandins and thromboxanes. (C-F) Results are expressed as mean ± SEM; n = 4 distinct cell preparations. *P < .05 vs macrophage group. **P < .01 vs macrophage group. #P < .05 vs macrophage plus Apo PMN group.

Figure 2

SPM biosynthesis during macrophage efferocytosis of apoptotic PMNs is enhanced by PMN MPs. Macrophages were incubated for 5 minutes with PMN MPs before the addition of apoptotic PMNs (7.5 × 105 cells/well). Cells were incubated for 1 hour at 37°C. At the end of the incubation, ice-cold MeOH was added and the products extracted. (A) Representative MRM traces for the identified LMs. (B) Accompanying MS/MS spectra used for identification. (C) Specific bioactive lipid mediator and precursor/pathway markers where: Q1, M-H (parent ion); and Q3, diagnostic ion in the MS-MS (daughter ion) along with mean ± SEM values for each of the mediators identified. Quantification and values obtained after PMN (3 × 105 PMN) and MP (2 × 105 MP) incubations. The detection limit was ∼ 1 pg. *Below limits. (D) D-series resolvins, protectins, and maresins. (E) Lipoxins. (F) Prostaglandins and thromboxanes. (C-F) Results are expressed as mean ± SEM; n = 4 distinct cell preparations. *P < .05 vs macrophage group. **P < .01 vs macrophage group. #P < .05 vs macrophage plus Apo PMN group.

Close modal

LM quantification was achieved by MRM of signature ion pairs Q1 (parent ion) to Q3 (characteristic daughter ion; Figure 2C). Macrophage efferocytosis of apoptotic PMNs increased SPM biosynthesis, primarily RvD1 (18 ± 10 vs 127 ± 14 pg/2.5 × 105 cells), RvD2 (28 ± 8 vs 124 ± 12 pg/2.5 × 105 cells), and LXB4 (19 ± 2 vs 44 ± 6 pg/2.5 × 105 cells). This was accompanied by increases in prostanoids, including PGE2 (7 ± 1 vs 44 ± 12 pg/2.5 × 105 cells; Figure 2C). Having found that MPs stimulate efferocytosis, we next assessed their actions on SPM biosynthesis during apoptotic PMN uptake. MPs stimulated macrophage biosynthesis of RvD2 (219 ± 44 vs 124 ± 12 pg/2.5 × 105 cells), LXB4 (84 ± 15 vs 44 ± 6 pg/2.5 × 105 cells), and RvE2 (478 ± 57 vs 199 ± 26 pg/2.5 × 105 cells), to a greater extent than apoptotic PMN, while reducing PGF (259 ± 29 vs 459 ± 54 pg/2.5 × 105 cells) and TXB2 (328 ± 21 vs 456 ± 62 pg/2.5 × 105 cells; Figure 2C).

Temporal SPM biosynthesis contributes to regulating the onset of resolution programs.2  Therefore, we next assessed the cumulative levels for each LM family identified. During human macrophage efferocytosis, there was an increase in DHA-derived (Figure 2D) and AA-derived SPM (Figure 2E) biosynthesis, along with an augmentation in prostanoid production (Figure 2F). MPs further enhanced apoptotic PMN-stimulated SPM biosynthesis (Figure 2D-E) while reducing prostanoid production (Figure 2F). To rule out the possibility that MPs carried mediators in addition to their known ability to carry SPM precursors,25  we profiled the MPs (supplemental Figure 3). MPs did not carry mature SPMs. Hence, we profiled the amounts of LM associated with apoptotic PMNs and MPs in the absence of efferocytosis (supplemental Figure 4). Although specific SPMs and LMs were demonstrable, their values were baseline compared with levels produced via macrophages during efferocytosis.

We next investigated whether PMN MPs were able to stimulate macrophage LM biosynthesis. Incubation of PMN MPs with macrophages led to the biosynthesis of lipoxygenase- and cyclooxygenase-derived LM. These included RvD5, PD1, and PGE2 in accordance with criteria23  that included matching retention times (supplemental Figure 5A) and at least 6 diagnostic ions as illustrated for MaR1, PD1, RvE2, and RvE1 (supplemental Figure 5B). These results are first to demonstrate each of these pro-resolving mediators in human macrophage incubations and their interrelationships.

For example, we observed that PMN MPs led to an increase in MaR1 (13 ± 4 vs 21 ± 2 pg/2.5 × 105 macrophages), PD1 (7 ± 1 vs 28 ± 5 pg/2.5 × 105 macrophages), LXB4 (78 ± 4 vs 248 ± 26 pg/2.5 × 105 macrophages), and RvE1 (14 ± 8 vs 72 ± 9 pg/2.5 × 105 macrophages) production (supplemental Figure 5C). In addition, we found that PMN MPs stimulated macrophage prostanoids, in particular PGE2 (18 ± 2 vs 40 ± 6 pg/2.5 × 105 macrophages) and TXB2 (28 ± 7 vs 63 ± 10 pg/2.5 × 105 macrophages). Cumulative LM metabololipidomics for each LM family identified demonstrated that PMN MPs stimulated biosynthesis of DHA- (supplemental Figure 5D), AA- (supplemental Figure E), and EPA- (supplemental Figure 5F) derived SPM along with prostanoids (supplemental Figure 5G). Of interest, incubation of macrophages with G-protein inhibitors (pertussis toxin and cholera toxin) reduced SPM biosynthesis without altering prostanoid levels (supplemental Figure 5C-G). Together, these results demonstrate that MPs selectively stimulate macrophage SPM production in a G-protein–coupled receptor-dependent manner.

Apoptotic PMN possess a pro-resolving LM profile

Because apoptotic PMNs stimulate macrophage SPMs during efferocytosis, we investigated mediator profiles of apoptotic PMNs compared with zymosan-stimulated PMNs. Using targeted LM metabololipidomics, we identified a profile signature of LM that included RvD1, LXB4, PGD2, and PGE2 (Figure 3A). Identification was achieved as illustrated for RvE2, PD1 (also known as NPD1/PD1 when produced in neural systems7 ), RvD2, and LTB4 (Figure 3B).

Figure 3

Elevated SPM biosynthesis in apoptotic PMNs. PMNs were isolated from peripheral blood of healthy volunteers and placed directly in MeOH (PMN), induced to apoptose (Apo PMN), or stimulated with zymosan (PMN + Zy). After extraction, LM profiles were obtained using LC-MS/MS (see “Sample extraction and lipid mediator metabololipidomics”). (A) Representative MRM chromatograms of the LM identified in the apoptotic or Zy-activated neutrophils. Peak heights represent the relative levels of each LM. Cumulative levels for individual LM families are depicted as a function of color intensity, where color scales (eg, white to blue [Apo PMN = ] or white to tan [Zy PMN = ]) are set from zero to 500 pg per 5 × 106 cells. (B) Accompanying MS/MS spectra used for identification. (C) Lipid mediator and precursor/pathway marker transition where: Q1, M-H (parent ion); and Q3, diagnostic ion in the MS-MS (daughter ion), along with mean ± SEM values for each of the mediators identified. Values are for 5 × 106 PMNs. The detection limit was ∼ 1 pg. *Below limits. Cumulative values: (D) Leukotriene B4 (and 20-OH LTB4 metabolite). (E) D-series resolvins, protectins, and maresins. (F) Lipoxins. (G) E-series resolvins and lipoxins. (H) Prostaglandins and thromboxanes. (C-H) Results are expressed as mean ± SEM; n = 5 distinct cell preparations. *P < .05 vs PMNs. **P < .01 vs PMNs. ***P < .01 vs PMNs. #P < .05 vs Apo PMNs.

Figure 3

Elevated SPM biosynthesis in apoptotic PMNs. PMNs were isolated from peripheral blood of healthy volunteers and placed directly in MeOH (PMN), induced to apoptose (Apo PMN), or stimulated with zymosan (PMN + Zy). After extraction, LM profiles were obtained using LC-MS/MS (see “Sample extraction and lipid mediator metabololipidomics”). (A) Representative MRM chromatograms of the LM identified in the apoptotic or Zy-activated neutrophils. Peak heights represent the relative levels of each LM. Cumulative levels for individual LM families are depicted as a function of color intensity, where color scales (eg, white to blue [Apo PMN = ] or white to tan [Zy PMN = ]) are set from zero to 500 pg per 5 × 106 cells. (B) Accompanying MS/MS spectra used for identification. (C) Lipid mediator and precursor/pathway marker transition where: Q1, M-H (parent ion); and Q3, diagnostic ion in the MS-MS (daughter ion), along with mean ± SEM values for each of the mediators identified. Values are for 5 × 106 PMNs. The detection limit was ∼ 1 pg. *Below limits. Cumulative values: (D) Leukotriene B4 (and 20-OH LTB4 metabolite). (E) D-series resolvins, protectins, and maresins. (F) Lipoxins. (G) E-series resolvins and lipoxins. (H) Prostaglandins and thromboxanes. (C-H) Results are expressed as mean ± SEM; n = 5 distinct cell preparations. *P < .05 vs PMNs. **P < .01 vs PMNs. ***P < .01 vs PMNs. #P < .05 vs Apo PMNs.

Close modal

MRM LM quantification of signature ion pairs demonstrated that AA-derived LMs represented ∼ 85% of the apoptotic PMN AA, EPA, and DHA identified LM metabolome. In these cells, the DHA-derived mediators represented ∼ 6%, whereas EPA-derived mediators amounted to ∼ 11% of the targeted LM. On the other hand, in zymosan-stimulated PMNs, the AA-derived LM constituted ∼ 81% of the targeted AA, EPA, and DHA LM metabolomes (Figure 3A). Color scales represent the magnitude of LM within chromatographic clusters within each profile (see Figure 3 legend).

The comparison of endogenous LM identified in apoptotic versus zymosan-activated PMNs revealed that RvD1 (42 ± 12 vs 5 ± 2 pg/5 × 106 cells) and LXB4 (142 ± 37 vs 14 ± 7 pg/5 × 106 cells) biosynthesis was reduced in zymosan-stimulated cells. Similarly, there was a down-regulation in a number of cyclooxygenase-derived LMs in zymosan-stimulated PMNs, including PGD2 (123 ± 29 vs 82 ± 15 pg/5 × 106 cells), PGE2 (390 ± 75 vs 82 ± 16 pg/5 × 106 cells), and PGF (127 ± 40 vs 52 ± 18 pg/5 × 106 cells). Production of the chemoattractant LTB4 (including its 20-OH P450 metabolite) was elevated in zymosan-stimulated cells as opposed to apoptotic PMNs (Figure 3C-D). For intact PMNs without exposure to agonists, the levels for these mediators were close to or below limits (Figure 3C).

To gauge the effector functions endowed on PMNs by their individual LM profiles, we assessed the cumulative LM amounts for the distinct LM families identified using LM metabololipidomics. We found that apoptotic PMNs produced significantly higher SPMs (Figure 3E-G) and prostanoids (Figure 3H) than zymosan or intact PMNs. Assessment of AA, EPA, and DHA-derived monohydroxy products did not reveal significant differences between apoptotic, zymosan-stimulated, or intact PMNs (data not shown).

To appreciate the sequence of events involved in LM production during human PMN apoptosis, we next identified LM present at distinct stages of PMN progression to apoptosis determined by AnxAV staining (supplemental Figure 6A). Because, during inflammation, edema formation supplies the inflammatory response with extracellular substrate,28  we assessed the time course of LM biosynthesis by apoptotic cells in the presence of extracellular substrate. MRM quantification of identified LM demonstrated a significant increase in SPM biosynthesis by apoptotic cells, including D- and E-series resolvins that reached maximum within 30 minutes (supplemental Figure 6B-D). Prostanoids were also elevated (supplemental Figure 6E), whereas LTB4 (including its 20-OH metabolite) occurred during the initial phase and subsequently declined, both in the presence and absence of substrate (supplemental Figure 6F). Although we cannot rule out that a small proportion of potentially nonapoptotic PMNs that might have been present could have contributed to LM profiles in incubations with apoptotic PMNs, it is more than likely that at the 18-hour interval LM profiles reflect the capacity of apoptotic PMNs to biosynthesize specific LMs.

Efferocytosis: apoptotic PMN and MPs, a nidus for macrophage SPM biosynthesis

Next, we investigated the potential contribution of transcellular biosynthesis, in particular for SPM, during macrophage efferocytosis (Figure 4). For this purpose, we used deuterium-labeled (d) precursors, which allowed for the identification of mediators from substrate/precursors derived from either apoptotic PMNs or MPs. LM derived from d-substrate had a higher mass to charge ratio (m/z) than those biosynthesized from endogenous substrate. Specifically, LM produced from d8-AA were 8 atomic mass units (amu) and those from d5-EPA and d5-DHA were 5 amu greater than LMs biosynthesized from endogenous precursors.

Figure 4

Macrophage phagocytosis of apoptotic PMNs. A nidus for transcellular biosynthesis. MPs or PMNs were preincubated with deuterium-labeled substrate (AA, EPA, and DHA, 100 ng each) before coincubation with macrophages for 1 hour at 37°C. Incubations were stopped with ice-cold MeOH and LM extracted. The levels of deuterium-labeled LM biosynthesized during the phagocytosis of apoptotic neutrophils were assessed by LC-MS/MS. (A) Representative MRM chromatograms of the identified LM. (B) Accompanying MS/MS spectra used for identification. (C) LM and precursor/pathway marker transition Q1-Q3 transitions along with mean ± SEM values for each of the mediators identified. The detection limit was ∼ 1 pg. *Below limits of detection. Cumulative values: (D) D-series resolvins, protectins, and maresins. (E) E-series resolvins and lipoxins. (F) Lipoxins. (G) Prostaglandins and thromboxanes. (C-G) Results are ± SEM; n = 5 distinct cell preparations.

Figure 4

Macrophage phagocytosis of apoptotic PMNs. A nidus for transcellular biosynthesis. MPs or PMNs were preincubated with deuterium-labeled substrate (AA, EPA, and DHA, 100 ng each) before coincubation with macrophages for 1 hour at 37°C. Incubations were stopped with ice-cold MeOH and LM extracted. The levels of deuterium-labeled LM biosynthesized during the phagocytosis of apoptotic neutrophils were assessed by LC-MS/MS. (A) Representative MRM chromatograms of the identified LM. (B) Accompanying MS/MS spectra used for identification. (C) LM and precursor/pathway marker transition Q1-Q3 transitions along with mean ± SEM values for each of the mediators identified. The detection limit was ∼ 1 pg. *Below limits of detection. Cumulative values: (D) D-series resolvins, protectins, and maresins. (E) E-series resolvins and lipoxins. (F) Lipoxins. (G) Prostaglandins and thromboxanes. (C-G) Results are ± SEM; n = 5 distinct cell preparations.

Close modal

We first fortified MPs and apoptotic PMN with d-labeled precursors and assessed levels of d-labeled LM, essential fatty acids (AA, EPA, and DHA), and biosynthetic precursors (including 18-HEPE, 15-HEPE, 15-HETE, 17-HDHA, and 14-HDHA). We identified LM precursors, including d5-DHA, d5-17-HDHA, and d8-AA (supplemental Figure 7A), along with a select number of mediators, including d5-RvD5 and d5-RvE2 in accordance with published criteria.23  MRM quantitation of the identified LM demonstrated that the levels of these d-labeled mediators within MPs, and apoptotic PMNs were near limits of detection (< 5 pg/6 × 105 MPs and < 1 pg/1 × 106 PMNs). We also identified d-labeled biosynthetic precursors that included d5-17-HDHA (16 ± 3 pg/5 × 105 MPs and 18 ± 1 pg/7.5 × 105 apoptotic PMNs) and d-labeled essential fatty acids that included d5-DHA (423 ± 35 pg/5 × 105 MPs and 459 ± 133 pg/7.5 × 105 apoptotic PMNs; supplemental Figure 7B). These results suggest that, although there was a significant enrichment of d-labeled precursors in both MPs and apoptotic PMN, only a minor proportion was further converted to bioactive LM in the absence of macrophages.

We next added either d-labeled MPs or d-labeled apoptotic PMN to macrophages assessing LM transcellular biosynthesis during macrophage efferocytosis. Apoptotic PMN uptake by macrophages led to the biosynthesis of d5-RvD2 and d5-RvD5 from the D-series resolvins, d5-PD1 from the protectins, from the lipoxins d8-LXB4 and d5-RvE2 from the E-series resolvins, whereas d8-PGD2 and d8-PGF were the only prostanoids identified (Figure 4A) and d5-RvD5, d5-RvE2, and d8-LXB4 (Figure 4B) where the atomic mass of the parent molecules was increased by either 5 or 8 amu, respectively.

MRM quantification of d-LM biosynthesized during macrophage efferocytosis suggested that MPs acted as a nexus for d5-RvD2, d5-RvD5, (Figure 4C-D) d8-LXA4 (Figure 4F), and d8-PGE2 and d8-PGF (Figure 4G) production. d5-MaR1 and d8-LXB4 biosynthesis was primarily sustained by precursors derived from within apoptotic PMN. d5-PD1 (Figure 4D) and d5-RvE2 (Figure 4E) biosynthesis were equally reliant on precursors from apoptotic cells and MPs. These results demonstrate that transcellular biosynthesis contributes to LM production during macrophage efferocytosis, and both apoptotic PMN and MPs contribute to and sustain a nidus for LM biosynthesis.

Distinct human macrophage subtypes possess distinct LM profiles

We next sought to assess the LM profiles of distinct human macrophage subtypes. Following LM metabolomics of M1 and M2 macrophages, which were prepared from human peripheral blood monocytes (supplemental Figure 1), we identified 35 distinct LMs that included RvD1, PD1, MaR1, LXA5, LXB5, PGD2, and PGE2 (Figure 5A). These mediators were identified using their diagnostic fragmentation patterns, matching a minimum of 6 characteristic fragments as shown for RvD2, PD1, RvE2, and LXA5 (Figure 5B).

Figure 5

Distinct macrophage subtypes biosynthesize specific lipid mediator profiles. M1 and M2 macrophages were prepared from primary human monocytes after incubation with GM-CSF (20 ng/mL), IFN-γ (20 ng/mL), and LPS (100 ng/mL) to produce M1 or M-CSF (20 ng/mL) and IL-4 (20 ng/mL) to obtain M2. Incubations were stopped with ice-cold MeOH and taken for LC-MS/MS analysis (see “Sample extraction and lipid mediator metabololipidomics”). (A) Representative MRM chromatograms for the identified LM. Peak heights represent the relative levels of each mediator in the different macrophage subtypes. Cumulative levels for each lipid mediator category are represented as a function of color intensity, where color scales (eg, white to green for M1 macrophages and white to purple for M2 macrophages) are set from zero to 35 000 pg per 2.5 × 106 cells. (B) Accompanying MS/MS spectra used for identification. (C) Lipid mediator and precursor/pathway marker transition along with mean ± SEM values for each of the mediators identified. The detection limit was ∼ 1 pg. *Below limits. Cumulative values: (D) D-series resolvins, protectins, and maresins. (E) Lipoxins. (F) E-series resolvins and lipoxins. (G) Prostaglandins and thromboxanes. (H) EPA-derived prostaglandins and thromboxanes. (I) Leukotriene B4 (and 20-OH LTB4 metabolite). (J) Leukotriene B5. (D-J) Results are ± SEM; n = 8 distinct cell preparations. *P < .05 vs M1 group. **P < .01 vs M1 group.

Figure 5

Distinct macrophage subtypes biosynthesize specific lipid mediator profiles. M1 and M2 macrophages were prepared from primary human monocytes after incubation with GM-CSF (20 ng/mL), IFN-γ (20 ng/mL), and LPS (100 ng/mL) to produce M1 or M-CSF (20 ng/mL) and IL-4 (20 ng/mL) to obtain M2. Incubations were stopped with ice-cold MeOH and taken for LC-MS/MS analysis (see “Sample extraction and lipid mediator metabololipidomics”). (A) Representative MRM chromatograms for the identified LM. Peak heights represent the relative levels of each mediator in the different macrophage subtypes. Cumulative levels for each lipid mediator category are represented as a function of color intensity, where color scales (eg, white to green for M1 macrophages and white to purple for M2 macrophages) are set from zero to 35 000 pg per 2.5 × 106 cells. (B) Accompanying MS/MS spectra used for identification. (C) Lipid mediator and precursor/pathway marker transition along with mean ± SEM values for each of the mediators identified. The detection limit was ∼ 1 pg. *Below limits. Cumulative values: (D) D-series resolvins, protectins, and maresins. (E) Lipoxins. (F) E-series resolvins and lipoxins. (G) Prostaglandins and thromboxanes. (H) EPA-derived prostaglandins and thromboxanes. (I) Leukotriene B4 (and 20-OH LTB4 metabolite). (J) Leukotriene B5. (D-J) Results are ± SEM; n = 8 distinct cell preparations. *P < .05 vs M1 group. **P < .01 vs M1 group.

Close modal

LM metabololipidomics of M1 macrophage AA, EPA, and DHA identified LM metabolome we found that AA-derived LM amounted to ∼ 48% of LM identified. On the other hand, LM metabololipidomics of M2 macrophages demonstrated that SPM represented ∼ 50% of the identified AA, EPA, and DHA metabolome, consisting of D-series resolvins, protectins, and maresins (∼ 11%), lipoxins (∼ 15%), and E-series resolvins and lipoxins (∼ 24%; Figure 5A).

Using MRM, we compared the endogenous biosynthesis of individual LM between the 2 macrophage subtypes. In these experiments, we found that biosynthesis of distinct SPM was elevated in M2 compared with M1 macrophage; these included RvD5 (196 ± 26 vs 43 ± 5 pg/2.5 × 105 cells), MaR1 (299 ± 8 vs 45 ± 8 pg/2.5 × 105 cells), PD1 (3442 ± 206 vs 1339 ± 206 pg/2.5 × 105 cells), LXA4 (977 ± 173 vs 675 ± 167 pg/2.5 × 105 cells), LXB4 (11 750 ± 724 vs 7560 ± 659 pg/2.5 × 105 cells), LXB5 (4370 ± 397 vs 2551 ± 859 pg/2.5 × 105 cells), and RvE2 (20 680 ± 4910 vs 10 910 ± 4232 pg/2.5 × 105 cells). Conversely, production of cyclooxygenase derived LM was elevated in M1 macrophages; these included PGE2 (2285 ± 435 vs 1573 ± 490 pg/2.5 × 105 cells), PGF(11 325 ± 428 vs 2108 ± 256 pg/2.5 × 105 cells), TXB2 (550 ± 135 vs 222 ± 73 pg/2.5 × 105 cells), and the EPA-derived prostanoids (Figure 5C).

Having found that different human macrophage subtypes possess distinct LM profiles, we next compared the cumulative amounts of different LM families identified in M1 and M2 macrophages to determine effector functions that these endow on the individual macrophage subtypes. SPM biosynthesis was higher in M2 compared with M1 macrophages (Figure 5D-F), whereas production of prostanoids (Figure 5G-H) and leukotrienes LTB4 and LTB5 (Figure 5I-J) was elevated in M1 macrophages compared with M2.

Because apoptotic PMNs stimulate LM biosynthesis in macrophages, we investigated whether this finding held for different macrophage subtypes. Assessment of LM biosynthesis after apoptotic cell efferocytosis demonstrated that apoptotic PMNs uptake gave SPMs in M1 macrophages (supplemental Figure 8A-C) while reducing prostanoids (supplemental Figure 8D-E) and leukotrienes (supplemental Figure 8F,G). Apoptotic PMNs with M2 macrophages reduced overall LM production, including SPMs (supplemental Figure 8A,C). These results indicate that human M1 and M2 macrophages, as defined,26  possess distinct endogenous LM signature profiles.

In the present report using targeted LM metabololipidomics, we determined the endogenous LM profiles of human phagocytes with distinct stimuli. Efferocytosis was found to stimulate SPM, a process regulated by PMN MPs and dependent, at least in part, on precursors obtained from both apoptotic cells and MPs. Apoptotic cells also displayed a pro-resolving LM signature profile, indicating that these specific LM profiles are key determinants of their distinct effector functions from those of zymosan-stimulated PMNs. This distinction was also found for macrophage subtypes (known as M1 and M226 ) involved with initiation and resolution of inflammation.

PMNs are the first leukocyte responders to an inflammatory insult. On recruitment from vasculature, they orchestrate clearance of the ensuing insult and stimulate either propagation or termination of inflammatory responses via temporal release of soluble mediators.29  SPMs are a genus of lipid mediators that include lipoxins, resolvins, protectins, and maresins; by definition, each shares potent anti-inflammatory and pro-resolving actions demonstrable within the picogram to nanogram range. SPMs were initially identified in resolution of self-limited inflammation where they are formed in vivo at levels commensurate with their activity. Specific SPM members (eg, RvD1) also carry additional actions, such as pain regulation and control of local chemokines.12-15  In human skin blisters (a PMN-driven response), biosynthesis of 15-epi-LXA4 and LXA4 and up-regulation of ALX/FPR2 (LXA4 receptor) are key in mediating the anti-inflammatory and pro-resolving actions of low-dose aspirin.19  Therefore, in the present studies, we systematically investigated LM profiles of human apoptotic PMNs and activated freshly isolated PMNs to determine their contributions, as well as PMN MPs.

A number of arachidonic acid-derived lipid mediators produced via both lipoxygenase and cyclooxygenase pathways, such as the classic prostaglandins and leukotrienes, are implicated in the onset and perpetuation of inflammation.30  Leukotriene B4, a potent pro-inflammatory LM, was the most abundant LM from both ionophore- and zymosan-activated PMN (Figure 6A), reaching levels within its bioactive range (supplemental Table 1, which contains complete stereochemical assignments and names for each specific SPM and LM, and references within), underscoring the pro-inflammatory phenotype elicited by these stimuli with intact human PMN from peripheral blood. Human PMN stimulation also up-regulates expression of cyclooxygenase-231  and, consequently, PGE2 and PGD2, which were identified in both. These mediators also possess pro-inflammatory properties that include enhancing leukocyte recruitment and vascular permeability (supplemental Table 1). In addition, both prostaglandins stimulate lipid mediator class switching and up-regulation of SPM at sites of inflammation.2  In the present studies, apoptotic PMN displayed elevated PG biosynthesis, suggesting that during acute inflammation these cells can initiate LM class switching and the onset of resolution. SPM biosynthesis, in particular that of RvE2 and LXB4, was also a distinct signature of apoptotic cells. This emphasizes the pro-resolving phenotype of apoptotic PMN because both RvE2 and LXB4 exert potent actions dampening leukocyte recruitment and reducing vascular leakage (supplemental Table 1). LM metabololipidomics with intact PMN showed that lipoxins, as verified by the presence of LXB4, occurred in the absence of further stimulation, suggesting that SPM biosynthesis can occur in peripheral blood PMN. These findings are in accord with those from patients with abdominal aortic aneurism, where elevated plasma SPM levels are associated with better postoperative outcomes.32 

Figure 6

PMN and macrophage characteristic lipid mediator constitution. Lipid mediator levels were assessed for (A) ionophore-activated, zymosan-stimulated, apoptotic PMN and intact PMN (B) M1 and M2 macrophages. Results are representative of 4-8 separate cell preparations. For quantitative values, see Figures 3 and 5.

Figure 6

PMN and macrophage characteristic lipid mediator constitution. Lipid mediator levels were assessed for (A) ionophore-activated, zymosan-stimulated, apoptotic PMN and intact PMN (B) M1 and M2 macrophages. Results are representative of 4-8 separate cell preparations. For quantitative values, see Figures 3 and 5.

Close modal

The acute inflammatory response in humans is a dynamic process where cell trafficking to ideally contain microbial challenge resolves to homeostasis, a protective stance.1,2  In models of self-resolving sterile inflammation, LTB4 levels peak in the initiation phase, reaching ∼ 3 ng/mouse exudate.7,33  Prostaglandins, including PGE2, peak later in these systems coinciding with the onset of resolution, the PGE2 levels reaching ∼ 6.5 ng pg/mouse exudate.7  SPM biosynthesis, including RvE2, peaks during the resolution phase with levels for this mediator reaching ∼ 2.5 pg/mouse exudate.34  Similarly, in murine models of live infections, PGE2 levels peak with the onset of resolution, reaching ∼ 1773 pg/mouse exudate.35  Hence, the present findings with isolated human PMN staged in vitro (ie, on activation, progression to apoptosis, and during macrophage efferocytosis) follow the in vivo events and levels of LM in self-contained exudates. For example, the primary mediator produced by zymosan-activated human PMN in vitro was LTB4 ∼ 31.5 pg/1 × 106 PMN (Figure 3), levels that are comparable with those reported in vivo.7,33  With apoptotic human PMN, we found elevated biosynthesis of prostaglandins and SPMs that included PGE2 ∼ 62 pg/1 × 106 PMN and RvE2 ∼ 33.2 pg/1 × 106 PMN (Figure 3). These LM levels are consistent with those reported in vivo in mice during the resolution phase of acute inflammation within contained inflammatory exudates.7,33,35  The present results are the first to demonstrate this dynamic shift with human leukocytes in vitro and their interrelationships in the LM profile that resemble distinct exudate stages observed for self-limited inflammatory responses in vivo. This is of general relevance to inflammatory diseases in humans because it is sometimes difficult to obtain temporal changes in LM profiles during disease states and their resolution.36  Nonetheless, LXA4 levels increase during resolution of glomerulonephritis,37  and PD1 levels are diminished in exhaled breath condensates from asthmatics.38 

Systematic mapping of cellular trafficking in a murine model of self-limited acute inflammation demonstrated that mononuclear cell accumulation at the site of inflammation occurs as part of the events that lead to the onset of resolution.1,2,7  During this phase, macrophages accelerate resolution by clearance of apoptotic cells and cellular debris in a nonphlogistic manner in a process referred to as efferocytosis.27  Herein we determined the endogenous LM profiles produced by human macrophages during efferocytosis. These experiments demonstrated an increase in SPM, specifically LXB4 and RvD1. RvD1 possesses potent actions: at concentrations as low as 10nM, it reduces pro-inflammatory cytokines from peritoneal macrophages and stimulates up-regulation of M2-associated markers.16  Cellular activation leads to release of MPs that elicit pro-inflammatory39  as well as protective actions in vivo.25  Specific subtypes of PMN MPs, carrying SPM precursors, accumulate in self-limited exudates preceding onset of resolution.25  Herein we found that the addition of PMN-derived MPs to human macrophages increased SPM biosynthesis in a G-protein–coupled receptor-dependent manner. This enhances macrophage efferocytosis of apoptotic PMNs that in turn led to a further increase in macrophage LM production, in particular SPMs. These include RvD2, LXB4, and RvE2, which each reached concentrations > 100nM. RvD2, at doses as low as 100nM, reduces pro-inflammatory cytokine production in murine bone marrow macrophages stimulated by LPS and reverses TNF-α–induced leukocyte-endothelial interactions at 1nM. In addition, LXB4 and RvE2 are potent tissue protective SPMs (supplemental Table 1). These findings underscore the importance of efferocytosis in resolution and macrophage SPM biosynthesis during efferocytosis, a process regulated by PMN MPs that can contribute to resolution in humans.

LM biosynthesis involves multiple enzyme-regulated steps that dictate the specific stereochemistry of these local mediators. In some instances, one cell type does not possess all the enzymes necessary for LM biosynthesis.40  Therefore, in a process coined as transcellular biosynthesis, intermediates are donated to a second cell type that is equipped with the necessary enzymes for their conversion to bioactive mediators.40  SPM production, including RvD5, LXB4, and RvE2, during efferocytosis was found to rely, at least in part, on substrate/intermediates (using deuterium label) that can be supplied by both apoptotic PMN and MPs, thus demonstrating that this biosynthetic route is also responsible for SPM production, a key step in stimulating resolution.

Distinct macrophage subtypes exert diverse effector functions in propagation and resolution of the inflammatory response.41  In obesity, for example, amelioration in insulin resistance and adipose tissue inflammation is associated with a macrophage phenotypic switch from M1- to M2-like phenotypes.16  Using targeted LM metabololipidomics, we assessed LM profiles in M1 and M2 demonstrating herein that M1 macrophages, characterized by their higher expression of CD68, CD80, MHC class II and CD54,26  eicosanoid biosynthesis, including TXB2 and LTB4, were elevated. RvD2 biosynthesis, which activates potent antibacterial actions (supplemental Table 1), was also increased in line with the concept that M1-macrophages contribute to protection against bacterial infection.42  Hence, lower levels of individual SPM were obtained for M2 compared with M1 macrophages, perhaps reflecting the higher predisposition of certain mediators (eg, RvE1) to undergo further metabolism. This together with higher expression of inactivating enzymes, such as eicosanoid oxidoreductase/15-prostaglandin dehydrogenase that can further convert LM, may give the apparent lower levels of specific LM by M2 macrophages. It is noteworthy that the SPMs are potent mediators that are bioactive in the low nanomolar to picomolar range. Moreover, recent results demonstrate that D-series resolvins, specifically RvD1 and RvD2, are substrate for further metabolism by dehydrogenation, with RvD1 being more rapidly further metabolized than RvD2.43  SPM biosynthesis, in particular PD1, was higher in M2 macrophages consistent with the finding that IL-4 induces 15-LOX expression in human monocytes.44  Similarly MaR1, recently shown to exert potent pro-resolving and tissue regenerative actions,11  production was also elevated in line with the tissue regenerative properties associated with this macrophage subtype.42 

Transcriptomic analysis of murine macrophages during resolution of acute inflammation suggest that at the onset of resolution there is a switch in macrophage phenotype from M1-like to displaying characteristics of both M1 and M2.45  Resolving macrophages possess 15-LOX type 1, which initiates LX and D-series resolvin biosynthesis,6,45  also expressed in the eye,46  and initiated PD1 by cardiac stem cells.47  Because we found that apoptotic PMN stimulate SPM in human macrophages, we assessed the impact of apoptotic cell uptake by lineage-defined macrophages on their LM profile. Uptake of apoptotic cells by M1 cells led to a decrease in prostanoids and leukotrienes, including LTB4 and LTB5 (supplemental Table 1) with a concomitant enhanced SPM biosynthesis to give a pro-resolving LM profile or signature.

It is now well established that cardiovascular diseases are at least in part driven by inflammation.41  ω-3 fatty acids are now also appreciated for their ability to exert a positive bearing on several intermediate determinants of cardiovascular risk.48  One mechanism involves RvE1, demonstrated to be protective in murine cardiovascular disease.49  In addition, in mice, eosinophils contribute to resolution, producing both LXA4 and PD1.50  Our present findings indicate that human phagocytes, apoptotic PMN, and pro-resolving MPs produce these LM, which can contribute to resolution of inflammation and thus may contribute the beneficial actions noted for ω-3 fatty acids48  via the local biosynthesis of potent mediators.

In conclusion, we systematically assessed the main human phagocyte components of an acute inflammatory response (ie, PMN, MPs, and macrophages), using LM metabololipidomics and defined surface markers. The present results translate to defined human cell types, in vivo findings in resolving exudates in murine systems.5,6  In addition, we demonstrate that apoptotic PMN possess a pro-resolving LM signature profile that activates macrophage SPM during efferocytosis. Both apoptotic PMN and MPs are a nexus for SPM biosynthesis by macrophages when performing efferocytosis. A potential caveat to the present findings is the lack of cell cycle synchronicity during PMN apoptosis in vitro. Nonetheless, this is a widely used procedure to obtain apoptotic PMN in the literature, and our findings demonstrate the dynamic biosynthetic range of these apoptotic PMN that produce both pro-inflammatory and pro-resolving mediators. We also establish that distinct macrophage subtypes produce specific LM signature profiles, where M1 cells display a pro-inflammatory, whereas M2 cells display a pro-resolving LM profile, which are modulated on uptake of apoptotic PMN. Taken together, these findings demonstrate that distinct PMN and macrophage phenotypic subpopulations possess signature LM profiles that endow them with specific effector functions modulated during the dynamic process of inflammation and its timely resolution.

This article contains a data supplement.

The publication costs of this article were defrayed in part by page charge payment. Therefore, and solely to indicate this fact, this article is hereby marked “advertisement” in accordance with 18 USC section 1734.

The authors thank Mary Halm Small for expert assistance in manuscript preparation, Dr Nan Chiang (Center for Experimental Therapeutics and Reperfusion Injury, Brigham and Women's Hospital/Harvard Medical School) for helpful discussions, and Dr Sungwhan Oh (Center for Experimental Therapeutics and Reperfusion Injury, Brigham and Women's Hospital/Harvard Medical School) for initial setup of the LC-MS/MS system I.

This work was supported in part by the National Institutes of Health (grants R01GM038765 and P01GM095467).

National Institutes of Health

Contribution: J.D. designed and carried out experiments, analyzed data, and contributed to manuscript and figure preparations; and C.N.S. carried out overall experimental design, conceived the research plan, and contributed to manuscript and figure preparations.

Conflict-of-interest disclosure: C.N.S. is an inventor on patents (resolvins) assigned to BWH and licensed to Resolvyx Pharmaceuticals. C.N.S. is a scientific founder of Resolvyx Pharmaceuticals and owns equity in the company. The interests of C.N.S. were reviewed and are managed by the Brigham and Women's Hospital and Partners HealthCare in accordance with their conflict-of-interest policies. J.D. declares no competing financial interests.

Correspondence: Charles N. Serhan, Harvard Institutes of Medicine, 77 Avenue Louis Pasteur, HIM 829, Boston, MA 02115; e-mail: cnserhan@zeus.bwh.harvard.edu.

1
Majno
 
G
Joris
 
I
Cells, Tissues, and Disease: Principles of General Pathology
2004
2nd Ed
New York, NY
Oxford: Oxford University Press
2
Serhan
 
CN
Novel resolution mechanisms in acute inflammation: to resolve or not?
Am J Pathol
2010
, vol. 
177
 (pg. 
1576
-
1591
)
3
Nathan
 
C
Ding
 
A
Nonresolving inflammation.
Cell
2010
, vol. 
140
 (pg. 
871
-
882
)
4
Willoughby
 
DA
Moore
 
AR
Colville-Nash
 
PR
Gilroy
 
D
Resolution of inflammation.
Int J Immunopharmacol
2000
, vol. 
22
 (pg. 
1131
-
1135
)
5
Serhan
 
CN
Clish
 
CB
Brannon
 
J
Colgan
 
SP
Chiang
 
N
Gronert
 
K
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
2000
, vol. 
192
 (pg. 
1197
-
1204
)
6
Serhan
 
CN
Hong
 
S
Gronert
 
K
, et al. 
Resolvins: a family of bioactive products of omega-3 fatty acid transformation circuits initiated by aspirin treatment that counter pro-inflammation signals.
J Exp Med
2002
, vol. 
196
 (pg. 
1025
-
1037
)
7
Bannenberg
 
GL
Chiang
 
N
Ariel
 
A
, et al. 
Molecular circuits of resolution: formation and actions of resolvins and protectins.
J Immunol
2005
, vol. 
174
 (pg. 
4345
-
4355
)
8
Schwab
 
JM
Chiang
 
N
Arita
 
M
Serhan
 
CN
Resolvin E1 and protectin D1 activate inflammation-resolution programmes.
Nature
2007
, vol. 
447
 (pg. 
869
-
874
)
9
Lu
 
Y
Hong
 
S
Tjonahen
 
E
Serhan
 
CN
Mediator-lipidomics: databases and search algorithms for PUFA-derived mediators.
J Lipid Res
2005
, vol. 
46
 (pg. 
790
-
802
)
10
Serhan
 
CN
Petasis
 
NA
Resolvins and protectins in inflammation-resolution.
Chem Rev
2011
, vol. 
111
 (pg. 
5922
-
5943
)
11
Serhan
 
CN
Dalli
 
J
Karamnov
 
S
, et al. 
Macrophage pro-resolving mediator maresin 1 stimulates tissue regeneration and controls pain.
FASEB J
2012
, vol. 
26
 (pg. 
1755
-
1765
)
12
Bento
 
AF
Claudino
 
RF
Dutra
 
RC
Marcon
 
R
Calixto
 
JB
Omega-3 fatty acid-derived mediators 17(R)-hydroxy docosahexaenoic acid, aspirin-triggered resolvin D1 and resolvin D2 prevent experimental colitis in mice.
J Immunol
2011
, vol. 
187
 (pg. 
1957
-
1969
)
13
Chan
 
MM
Moore
 
AR
Resolution of inflammation in murine autoimmune arthritis is disrupted by cyclooxygenase-2 inhibition and restored by prostaglandin E2-mediated lipoxin A4 production.
J Immunol
2010
, vol. 
184
 
11
(pg. 
6418
-
6426
)
14
Bang
 
S
Yoo
 
S
Yang
 
TJ
Cho
 
H
Hwang
 
SW
17(R)-resolvin D1 specifically inhibits transient receptor potential ion channel vanilloid 3 leading to peripheral antinociception.
Br J Pharmacol
2012
, vol. 
165
 
3
(pg. 
683
-
692
)
15
Rajasagi
 
NK
Reddy
 
PB
Suryawanshi
 
A
Mulik
 
S
Gjorstrup
 
P
Rouse
 
BT
Controlling herpes simplex virus-induced ocular inflammatory lesions with the lipid-derived mediator resolvin E1.
J Immunol
2011
, vol. 
186
 
3
(pg. 
1735
-
1746
)
16
Titos
 
E
Rius
 
B
Gonzalez-Periz
 
A
, et al. 
Resolvin D1 and its precursor docosahexaenoic acid promote resolution of adipose tissue inflammation by eliciting macrophage polarization toward an M2-like phenotype.
J Immunol
2011
, vol. 
187
 
10
(pg. 
5408
-
5418
)
17
White
 
PJ
Arita
 
M
Taguchi
 
R
Kang
 
JX
Marette
 
A
Transgenic restoration of long-chain n-3 fatty acids in insulin target tissues improves resolution capacity and alleviates obesity-linked inflammation and insulin resistance in high-fat-fed mice.
Diabetes
2010
, vol. 
59
 
12
(pg. 
3066
-
3073
)
18
Morris
 
T
Stables
 
M
Colville-Nash
 
P
, et al. 
Dichotomy in duration and severity of acute inflammatory responses in humans arising from differentially expressed proresolution pathways.
Proc Natl Acad Sci U S A
2010
, vol. 
107
 
19
(pg. 
8842
-
8847
)
19
Morris
 
T
Stables
 
M
Hobbs
 
A
, et al. 
Effects of low-dose aspirin on acute inflammatory responses in humans.
J Immunol
2009
, vol. 
183
 
3
(pg. 
2089
-
2096
)
20
Psychogios
 
N
Hau
 
DD
Peng
 
J
, et al. 
The human serum metabolome.
PLoS One
2011
, vol. 
6
 
2
pg. 
e16957
 
21
Raatz
 
SK
Golovko
 
MY
Brose
 
SA
, et al. 
Baking reduces prostaglandin, resolvin, and hydroxy-fatty acid content of farm-raised Atlantic salmon (Salmo salar).
J Agric Food Chem
2011
, vol. 
59
 
20
(pg. 
11278
-
11286
)
22
Tobin
 
DM
Roca
 
FJ
Oh
 
SF
, et al. 
Host genotype-specific therapies can optimize the inflammatory response to mycobacterial infections.
Cell
2012
, vol. 
148
 
3
(pg. 
434
-
446
)
23
Yang
 
R
Chiang
 
N
Oh
 
SF
Serhan
 
CN
Metabolomics-lipidomics of eicosanoids and docosanoids generated by phagocytes.
Curr Protoc Immunol
2011
 
Chapter 14:Unit 14 26
24
Boilard
 
E
Nigrovic
 
PA
Larabee
 
K
, et al. 
Platelets amplify inflammation in arthritis via collagen-dependent microparticle production.
Science
2010
, vol. 
327
 
5965
(pg. 
580
-
583
)
25
Norling
 
LV
Spite
 
M
Yang
 
R
Flower
 
RJ
Perretti
 
M
Serhan
 
CN
Cutting edge: humanized nano-proresolving medicines mimic inflammation-resolution and enhance wound healing.
J Immunol
2011
, vol. 
186
 
10
(pg. 
5543
-
5547
)
26
Bellora
 
F
Castriconi
 
R
Dondero
 
A
, et al. 
The interaction of human natural killer cells with either unpolarized or polarized macrophages results in different functional outcomes.
Proc Natl Acad Sci U S A
2010
, vol. 
107
 
50
(pg. 
21659
-
21664
)
27
Serhan
 
CN
Savill
 
J
Resolution of inflammation: the beginning programs the end.
Nat Immunol
2005
, vol. 
6
 
12
(pg. 
1191
-
1197
)
28
Kasuga
 
K
Yang
 
R
Porter
 
TF
, et al. 
Rapid appearance of resolvin precursors in inflammatory exudates: novel mechanisms in resolution.
J Immunol
2008
, vol. 
181
 
12
(pg. 
8677
-
8687
)
29
Soehnlein
 
O
Lindbom
 
L
Phagocyte partnership during the onset and resolution of inflammation.
Nat Rev Immunol
2010
, vol. 
10
 
6
(pg. 
427
-
439
)
30
Samuelsson
 
B
An elucidation of the arachidonic acid cascade: discovery of prostaglandins, thromboxane and leukotrienes.
Drugs
1987
, vol. 
33
 
Suppl 1
(pg. 
2
-
9
)
31
Pouliot
 
M
Gilbert
 
C
Borgeat
 
P
, et al. 
Expression and activity of prostaglandin endoperoxide synthase-2 in agonist-activated human neutrophils.
FASEB J
1998
, vol. 
12
 
12
(pg. 
1109
-
1123
)
32
Pillai
 
PS
Leeson
 
S
Porter
 
TF
, et al. 
Chemical mediators of inflammation and resolution in post-operative abdominal aortic aneurysm patients.
Inflammation
2012
, vol. 
35
 (pg. 
98
-
113
)
33
Levy
 
BD
Clish
 
CB
Schmidt
 
B
Gronert
 
K
Serhan
 
CN
Lipid mediator class switching during acute inflammation: signals in resolution.
Nat Immunol
2001
, vol. 
2
 (pg. 
612
-
619
)
34
Oh
 
SF
Dona
 
M
Fredman
 
G
Krishnamoorthy
 
S
Irimia
 
D
Serhan
 
CN
Resolvin E2 formation and impact in inflammation resolution.
J Immunol
2012
, vol. 
188
 (pg. 
4527
-
4534
)
35
Chiang
 
N
Fredman
 
G
Bäckhed
 
F
, et al. 
Infection regulates pro-resolving mediators that lower antibiotic requirements.
Nature
2012
, vol. 
484
 (pg. 
524
-
528
)
36
Ho
 
KJ
Spite
 
M
Owens
 
CD
, et al. 
Aspirin-triggered lipoxin and resolvin E1 modulate vascular smooth muscle phenotype and correlate with peripheral atherosclerosis.
Am J Pathol
2010
, vol. 
177
 (pg. 
2116
-
2123
)
37
Wu
 
S-H
Liao
 
P-Y
Yin
 
P-L
Zhang
 
Y-M
Dong
 
L
Elevated expressions of 15-lipoxygenase and lipoxin A4 in children with acute poststreptococcal glomerulonephritis.
Am J Pathol
2009
, vol. 
174
 (pg. 
115
-
122
)
38
Levy
 
BD
Kohli
 
P
Gotlinger
 
K
, et al. 
Protectin D1 is generated in asthma and dampens airway inflammation and hyper-responsiveness.
J Immunol
2007
, vol. 
178
 (pg. 
496
-
502
)
39
Zimmerman
 
GA
Weyrich
 
AS
Immunology: arsonists in rheumatoid arthritis.
Science
2010
, vol. 
327
 
5965
(pg. 
528
-
529
)
40
Marcus
 
AJ
Transcellular metabolism of eicosanoids.
Prog Hemost Thromb
1986
, vol. 
8
 (pg. 
127
-
142
)
41
Swirski
 
FK
Nahrendorf
 
M
Libby
 
P
The ins and outs of inflammatory cells in atheromata.
Cell Metab
2012
, vol. 
15
 
2
(pg. 
135
-
136
)
42
Tacke
 
F
Randolph
 
GJ
Migratory fate and differentiation of blood monocyte subsets.
Immunobiology
2006
, vol. 
211
 
6
(pg. 
609
-
618
)
43
Clària
 
J
Dalli
 
J
Yacoubian
 
S
Gao
 
F
Serhan
 
CN
Resolvin D1 and resolvin D2 govern local inflammatory tone in obese fat.
J Immunol
2012
, vol. 
189
 
5
(pg. 
2597
-
2605
)
44
Nassar
 
GM
Morrow
 
JD
Roberts
 
LJ
Lakkis
 
FG
Badr
 
KF
Induction of 15-lipoxygenase by interleukin-13 in human blood monocytes.
J Biol Chem
1994
, vol. 
269
 (pg. 
27631
-
27634
)
45
Stables
 
MJ
Shah
 
S
Camon
 
EB
, et al. 
Transcriptomic analyses of murine resolution-phase macrophages.
Blood
2011
, vol. 
118
 
26
(pg. 
e192
-
e208
)
46
Gronert
 
K
Maheshwari
 
N
Khan
 
N
Hassan
 
IR
Dunn
 
M
Schwartzman
 
ML
A role for the mouse 12/15-lipoxygenase pathway in promoting epithelial wound healing and host defense.
J Biol Chem
2005
, vol. 
280
 (pg. 
15267
-
15278
)
47
Yanes
 
O
Clark
 
J
Wong
 
DM
, et al. 
Metabolic oxidation regulates embryonic stem cell differentiation.
Nat Chem Biol
2010
, vol. 
6
 
6
(pg. 
411
-
417
)
48
De Caterina
 
R
n-3 fatty acids in cardiovascular disease.
N Engl J Med
2011
, vol. 
364
 
25
(pg. 
2439
-
2450
)
49
Keyes
 
KT
Ye
 
Y
Lin
 
Y
, et al. 
Resolvin E1 protects the rat heart against reperfusion injury.
Am J Physiol Heart Circ Physiol
2010
, vol. 
299
 (pg. 
H153
-
H164
)
50
Yamada
 
T
Tani
 
Y
Nakanishi
 
H
Taguchi
 
R
Arita
 
M
Arai
 
H
Eosinophils promote resolution of acute periotonitis by producing proresolving mediators in mice.
FASEB J
2011
, vol. 
25
 (pg. 
561
-
568
)
Sign in via your Institution