Bioactive metabolites of docosahexaenoic acid

0 downloads 0 Views 455KB Size Report
Jan 11, 2017 - Maresin 1 (7R,14S-diHDHA, conjugated triene E,E,Z) displays potent anti-inflammatory, pro-resolving, analgesic and pro-healing actions [28 ...
Biochimie 136 (2017) 12e20

Contents lists available at ScienceDirect

Biochimie journal homepage: www.elsevier.com/locate/biochi

Review

Bioactive metabolites of docosahexaenoic acid Ondrej Kuda Department of Adipose Tissue Biology, Institute of Physiology of the Czech Academy of Sciences, Videnska 1083, 14220 Prague 4, Czech Republic

a r t i c l e i n f o

a b s t r a c t

Article history: Received 7 October 2016 Received in revised form 2 January 2017 Accepted 8 January 2017 Available online 11 January 2017

Docosahexaenoic acid (DHA) is an essential fatty acid that is recognized as a beneficial dietary constituent and as a source of the anti-inflammatory specialized proresolving mediators (SPM): resolvins, protectins and maresins. Apart from SPMs, other metabolites of DHA also exert potent biological effects. This article summarizes current knowledge on the metabolic pathways involved in generation of DHA metabolites. Over 70 biologically active metabolites have been described, but are often discussed separately within specific research areas. This review follows DHA metabolism and attempts to integrate the diverse DHA metabolites emphasizing those with identified biological effects. DHA metabolites could be divided into DHA-derived SPMs, DHA epoxides, electrophilic oxo-derivatives (EFOX) of DHA, neuroprostanes, ethanolamines, acylglycerols, docosahexaenoyl amides of amino acids or neurotransmitters, and branched DHA esters of hydroxy fatty acids. These bioactive metabolites have pleiotropic effects that include augmenting energy expenditure, stimulating lipid catabolism, modulating the immune response, helping to resolve inflammation, and promoting wound healing and tissue regeneration. As a result they have been shown to exert many beneficial actions: neuroprotection, anti-hypertension, anti-hyperalgesia, anti-arrhythmia, anti-tumorigenesis etc. Given the chemical structure of DHA, the number and geometry of double bonds, and the panel of enzymes metabolizing DHA, it is also likely that novel bioactive derivatives will be identified in the future. © 2017 Elsevier B.V. and Société Française de Biochimie et Biologie Moléculaire (SFBBM). All rights reserved.

Keywords: DHA Specialized proresolving mediators FAHFA DHEA N-acyl amides Omega-3 PUFA

Contents 1. 2. 3.

4.

5.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 Enzymes metabolizing DHA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 Oxygenated metabolites of DHA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 3.1. Maresins - resolution of inflammation, wound healing, analgesic actions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 3.2. Protectins - resolution of inflammation, neuroprotection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 3.3. Resolvins - resolution of inflammation and wound healing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 3.4. Electrophilic oxo-derivatives (EFOX) of DHA e anti-inflammatory, anti-proliferative effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 3.5. Epoxides e anti-hypertensive, analgesic actions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 3.6. Neuroprostanes e cardioprotection, wound healing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 Conjugates of DHA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 4.1. Ethanolamines and glycerol esters e neural development, immunomodulation, metabolic effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 4.2. Branched fatty acid esters of hydroxy fatty acids (FAHFA) e immunomodulation, resolution of inflammation . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 4.3. N-acyl amides - metabolic regulation, neuroprotection, neurotransmission . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 Supplementary data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

E-mail address: [email protected]. http://dx.doi.org/10.1016/j.biochi.2017.01.002 0300-9084/© 2017 Elsevier B.V. and Société Française de Biochimie et Biologie Moléculaire (SFBBM). All rights reserved.

O. Kuda / Biochimie 136 (2017) 12e20

1. Introduction Naturally occurring long-chain omega-3 polyunsaturated fatty acids (omega-3), namely eicosapentaenoic acid (EPA; 20:5 u-3) and docosahexaenoic acid (DHA; 22:6 u-3), have many beneficial metabolic effects, including prevention of cardiovascular diseases [1], amelioration of non-alcoholic fatty liver disease [2], or lowering hypertriacylglycerolemia [3]. They exert anti-inflammatory and hypolipidemic effects (reviewed in Ref. [4]), while increasing catabolism of lipids via a PPARa-mediated mechanism [4e6]. Although negative effects of omega-3 PUFA on prostate cancer risk [7] and prevention of cardiovascular diseases [8] were published these findings were later revised [9e11]. The discrepancies probably reflect multiple variables that could complicate the outcome of these studies (reviewed in Ref. [4]) and suggest that omega-3 dose and context (medication, disease) have to be considered. Patients who could potentially benefit from omega-3 supplementation are often premedicated with over-the-counter nonsteroidal antiinflammatory drugs, pain relievers, statins or other drugs which could interact with lipid metabolism and eicosanoid/docosanoid production and neutralize the potential of omega-3. In addition there is information suggesting that genetic variants can explain some of the individual variability in therapeutic potential of omega-3 supplementation [12,13]. Despite the occasionally divergent findings, omega-3 are generally considered as food supplements with positive properties. The beneficial effects of DHA are mediated by the fatty acid itself (PPARa ligand [14,15]) as well as by its bioactive metabolites e the specialized proresolving mediators (SPM) and other DHA derivatives. Although DHA triggers signaling through the Free fatty acid receptor 4 (FFAR4, also known as G protein-coupled receptor 120; GPR120) [16e18], recent studies have shown that FFAR4 is not required for the anti-inflammatory and insulin sensitizing effects of omega-3 [19e21] supporting existence of additional DHA receptors. The DHA metabolites have a wide range of actions acting simultaneously at different levels and sites. They activate various cell surface receptors (GPR32, GPR110, N-formyl peptide receptor 2; cannabinoid receptor 1 & 2; transient receptor potential channels; etc.) [22e25], and they also act as ligands to nuclear receptors PPARa/g [26,27]. 2. Enzymes metabolizing DHA Dominant enzymes in DHA metabolism are the 5-, 12-, and 15lipoxygenases (LOX) that form DHA hydroperoxides (HpDHA), which are further metabolized by hydrolases, including soluble epoxide hydrolase (sEH), glutathione S-transferase (GST) and several members of the cytochrome P450 superfamily (epoxydases, u-hydrolases) [22,28]. Cyclooxygenase (COX; prostaglandinendoperoxide synthase) 2 also metabolizes DHA, especially when its active site is acetylated by aspirin [29]. Degradation of active metabolites is carried out via oxidation (15-hydroxyprostaglandin dehydrogenase, 15PGDH), conjugation with glutathione (GSH) or hydrolysis by fatty acid amide hydrolase (FAAH) or sEH [30e32]. Only little is known about the synthesis and degradation of N-acylamides [33].

13

namely protectins, resolvins and maresins, according to their structures and biological effects [22,28]. For further details on SPMs (stereochemistry, organic synthesis, etc), readers are directed to recent reviews [28,35]. See Figs. 1e3 and Supplemental Fig. 1 for a schematic overview of the related pathways. 3.1. Maresins - resolution of inflammation, wound healing, analgesic actions Maresins (Fig. 1) are 12-LOX-derived metabolites of DHA produced mainly by macrophages via 13S,14S-epoxy-DHA intermediate. Maresin 1 (7R,14S-diHDHA, conjugated triene E,E,Z) displays potent anti-inflammatory, pro-resolving, analgesic and pro-healing actions [28,36,37]. 13,14-eMaR (13S,14S-epoxy-DHA, E,E,Z) inhibited leukotriene B4 formation and promoted conversion of M1 macrophages to M2 phenotype [38]. 22-hydroxy-Maresin 1 and 14oxo-Maresin 1 were recently identified in human leukocytes [39]. Maresin 1 younger sibling Maresin 2 (13R,14S-diHDHA, Z,E,E) reduces neutrophil infiltration and enhances efferocytosis as shown in human macrophages [40]. Sequential actions of 12-LOX and 5LOX on DHA generated 7S,14S-diHDHA (E,Z,E), which reduced polymorphonuclear leukocyte (PMNL) infiltration in zymosaninduced peritonitis [36]. The epoxide intermediate could also serve as a substrate for GSTM4 to yield the Maresin conjugates in tissue regeneration (MCTR) family, which are involved in regulation of host responses in clearing infections and promoting regeneration [41,42]. The Maresin family could be further expanded with the maresin-like mediators Maresin-L1 (14S,22-diHDHA) and MaresinL2 (14R,22-diHDHA) that were produced by leukocytes and platelets and are involved in wound healing [43]. Alternative u1oxidation of 14-HDHA by P450 produces 14S,21S-diHDHA; 14R,21R diHDHA; 14S,21R-diHDHA; and 14R,21S-diHDHA that also promoted wound healing [44], or anti-inflammatory 14S,20RdiHDHA [45]. 3.2. Protectins - resolution of inflammation, neuroprotection The most famous member of this family (Neuro)Protectin D1 (PD1; 10R,17S-diHDHA, E,E,Z) exhibits many beneficial effects (extensively reviewed in Refs. [28,35]). DHA is converted by 15-LOX to 16S,17S-epoxy-DHA intermediate that is either transformed to PD1 by enzymatic hydrolysis or reacted with GSH into Protectin conjugates in tissue regeneration (PCTR) family, namely 16glutathionyl-17-HDHA, 16-cysteinylglycinyl- 17-HDHA and 16cysteinyl-17-HDHA, with host-protective, proresolving, and tissue-regenerative actions [32](Fig. 2). Other members of the family are aspirin-triggered PD1 (AT-PD1, E,E,Z) [28,29] and PD1 metabolite 22-OH-PD1 (E,E,Z), which have been shown to exert potent proresolving actions by inhibiting PMNL chemotaxis [46]. Protectin DX (10S,17S-diHDHA, E,Z,E), an isomer of PD1 formed by consecutive sequential actions of two lipoxygenase reactions, belongs to a related family of poxytrins (PUFA oxygenated E,Z,E-conjugated triene) [47,48]. PDX inhibits human blood platelet aggregation [47], ROS production and cyclooxygenase activities in human neutrophils [49] and influenza virus replication [50,51]. 3.3. Resolvins - resolution of inflammation and wound healing

3. Oxygenated metabolites of DHA Besides the direct effects of DHA itself, indirect effects are mediated via its oxidation to potent lipid signaling molecules. DHA can be enzymatically oxidized to mono-, di- and tri-hydroxyDHA (HDHA), epoxy- and oxo-DHA metabolites, or modified by a free radical non-enzymatic mechanisms to neuroprostanes [27,29,30,34]. Many of HDHA derivatives belong to a family of SPMs,

The resolvin family members share 17-hydroxy residue introduced to DHA by 15-LOX (Fig. 2). There are 7 members, RvD1 (conjugated tetraene E,E,Z,E)e RvD6 and 17S-HDHA, showing many favorable biological effects such as stimulation of macrophage phagocytosis or inhibition of the production of pro-inflammatory cytokines [28,52]. Resolvins also participate in wound healing in mice [53,54]. As seen before, the intermediate 7S-hydroperoxy-

14

O. Kuda / Biochimie 136 (2017) 12e20

Fig. 1. Schematic overview of the pathways related to maresins. diHDHA, dihydroxydocosahexaenoic acid; DPEP, dipeptidase; eMar, 13,14-epoxy-maresin; GGT, g-glutamyl transferase; GSH, glutathione; GSTM4, glutathione S-transferase; HpDHA, hydroperoxydocosahexaenoic acid; LOX, lipoxygenase; MCTR, Maresin conjugates in tissue regeneration; P450, cytochrome P450; sEH, soluble epoxide hydrolase. The bolded metabolites have been documented in humans.

17S-hydroxy-DHA could be coupled with GSH to form the Resolvin conjugates in tissue regeneration (RCTR) family [32]. Aspirintriggered (AT-) COX acetylation results in a production of the 17Rhydroxy residue family, called AT-resolvins, complementary to the 17S series, which exhibited anti-hyperalgesic effects in rats [55]. 3.4. Electrophilic oxo-derivatives (EFOX) of DHA e antiinflammatory, anti-proliferative effects EFOXs can be generated via enzymatic and nonenzymatic pathways and often serve as PPARg ligands (Fig. 3). 17-oxo-DHA and 13-oxo-DHA are generated by a COX-2-catalyzed mechanism in activated macrophages. These oxo-derivatives act as PPARg agonists and inhibit pro-inflammatory cytokine and nitric oxide production in macrophages [27]. The metabolite 17-oxo-DHA was later confirmed to be a PPARa/g dual covalent agonist and 4-oxo-DHA a selective PPARg agonist [26]. Moreover, EFOX form GSH conjugates after macrophage activation [27]. Besides COX-2, also 15-PGDH catalyzes the oxidation of HDHA to a,b-unsaturated oxo-DHA, namely 4-, 7-, 10-, 11-, 13-, 14-, 16-, 17-, and 20-oxo-DHA species. For instance, exogenous addition of 14-oxo-DHA to primary alveolar macrophages inhibited LPS-induced pro-inflammatory cytokine mRNA expression [30]. Furthermore, LOX-5 in human neutrophils generates 7-oxo-DHA [56], while 4-oxo-DHA was shown to inhibit cancer cell proliferation [57]. 3.5. Epoxides e anti-hypertensive, analgesic actions DHA may compete with arachidonic acid for conversion by cytochrome P450 (CYP2C/2J isoforms) to form epoxydocosapentaenoic acid (EDP or epoxy-DPA, Fig. 3). Among the family of EDPs [31], the metabolite 19,20-EDP (19,20-epoxy-DPA) seems to be the most significant member. It is a highly active anti-

arrhythmic agent [58], acts as a mediator of the anti-hypertensive effects of DHA in angiotensin-II dependent hypertension [59,60], or inhibits vascular endothelial growth factor (VEGF)- and fibroblast growth factor 2-induced angiogenesis, tumor growth, and metastasis via a VEGF receptor 2-dependent mechanism [61]. High hepatic levels of 19,20-EDP were associated with a reduction of tissue inflammation and lipid peroxidation [62]. A mixture of EPDs (7,8-EDP; 10,11-EDP, 13,14-EDP; 16,17-EDP and 19,20-EDP) exerted anti-hyperalgesic effects in animal models of pain associated with inflammation and diabetic neuropathy [31,63]. 3.6. Neuroprostanes e cardioprotection, wound healing Neuroprostanes (NP) are products of a free radical nonenzymatic peroxidation of DHA (Fig. 3). Based on the mechanism of formation and the stability of precursor radicals, the majority of NPs identified in vivo contains F-type prostane rings (F4-NPs) and belongs to 4-, or 20-series regioisomers, depending on the position of the hydroxyl group [34,64]. Among the F4-NPs, 4(RS)-4-F4tNeuroProstane showed the most active anti-arrhythmic properties [34] and other cardioprotective effects including a reduction of events such as arrhythmias and systolic cardiac failure were assigned to peroxidation products of DHA [65]. Also A4/J4-type NPs contribute to the anti-inflammatory actions of DHA metabolites as 14-A4-NP suppress the inflammatory response in macrophages via inhibition of NF-kB signaling [66]. Furthermore, a series of compounds termed Neuroketals can be generated via the NP pathway. These highly reactive g-ketoaldehydes (Neuroketals) react with primary amines, form lysyl-lactam protein conjugates via a pyrrole intermediate and were detected in human brain [67]. Similarly, when DHA esterified in phospholipids is cleaved by reactive oxygen species into 4-hydroxy-7-oxohept-5-enoic acid (HOHA), this product reacts with amino groups of proteins (lysines) and forms 2-

O. Kuda / Biochimie 136 (2017) 12e20

15

Fig. 2. Schematic overview of the pathways related to protectins and resolvins. AT-, aspirin-triggered-; COX, cyclooxygenase; diHDHA, dihydroxydocosahexaenoic acid; GSH, glutathione; GST, glutathione S-transferase; HpDHA, hydroperoxydocosahexaenoic acid; LOX, lipoxygenase; P450, cytochrome P450; PGDH, hydroxyprostaglandin dehydrogenase; PCTR, Protectin conjugates in tissue regeneration; PD, protectin D; RCTR, Resolvin conjugates in tissue regeneration; RvD, resolvin D; sEH, soluble epoxide hydrolase; triHDHA, trihydroxydocosahexaenoic acid. The bolded metabolites have been documented in humans.

(u-carboxyethyl)pyrrole (CEP) adducts [68]. These CEPs are generated during inflammation, recognized by Toll-like-receptors 2 and CD36, and promote angiogenesis and wound healing [69e71]. Of note, DHA at low concentrations also prevents oxidative damage, while at high concentrations acts as a pro-oxidant [72].

4. Conjugates of DHA Besides oxidation, DHA can be conjugated with amines and alcohols to form amides and esters, respectively. See Fig. 4 for a schematic overview of the related pathways.

16

O. Kuda / Biochimie 136 (2017) 12e20

Fig. 3. Schematic overview of the pathways related to oxygenated metabolites of docosahexaenoic acid (DHA). CEP, 2-(u-carboxyethyl)pyrrole; COX, cyclooxygenase; diHDHA, dihydroxydocosahexaenoic acid; diHDPA, dihydroxydocosapentaenoic acid; DPA, docosapentaenoic acid; GSH, glutathione; GST, glutathione S-transferase; HpDHA, hydroperoxydocosahexaenoic acid; HOHA, 4-hydroxy-7-oxohept-5-enoic acid; LOX, lipoxygenase; P450, cytochrome P450; PGDH, hydroxyprostaglandin dehydrogenase; ROS, reactive oxygen species; sEH, soluble epoxide hydrolase. The bolded metabolites have been documented in humans.

Fig. 4. Schematic overview of the pathways related to docosahexaenoic acid (DHA) conjugates. 13-DHAHLA, 13-(docosahexaenoyloxy)-hydroxylinoleic acid; 14-DHAHDHA, 14(docosahexaenoyloxy)-hydroxydocosahexaenoic acid; DHEA, docosahexaenoyl ethanolamine; DHG, docosahexaenoyl glycerol; diHDHEA, dihydroxy-DHEA; DPA, docosapentaenoic acid; HEDPEA, hydroxy-epoxy-docosapentaenoyl ethanolamine; HpDHEA, hydroperoxy-DHEA; LOX, lipoxygenase; NAPE-PLD, N-acyl phosphatidylethanolamine-specific phospholipase D; NAT, N-acyltransferase; PE, phosphatidylethanolamine; P450, cytochrome P450; The bolded metabolites have been documented in humans.

4.1. Ethanolamines and glycerol esters e neural development, immunomodulation, metabolic effects DHA is highly enriched in the brain and is metabolized to Ndocosahexaenoyl ethanolamine (DHEA, synaptamide). It is probably synthesized from DHA-containing phosphatidylethanolamine (DHA-PE) through acylation of phosphatidylethanolamine and subsequent liberation of DHEA by a specific phospholipase D. DHEA stimulates neurite growth, synaptogenesis, promotes differentiation of neural stem cells, and exhibited enhanced glutamatergic synaptic activity [73,74]. Administration of DHEA following lipopolysaccharide injection significantly reduced neuroinflammatory responses in the murine brain [75]. Its anti-inflammatory, immunomodulatory, and other beneficial metabolic effects were documented in macrophages and in adipose tissue [76e78], and it was also shown to enhance glucose uptake in myocytes [79] and activate PPARa [80]. Furthermore, DHEA can be oxidized to 10,17-

diHDHEA, 15-hydroxy-16,17-epoxy-docosapentaenoyl ethanolamine (15-HEDPEA), 13-HEDPEA, or 17S-DHEA with antiinflammatory and organ-protective effects [81,82]. Only little is known about the biology of docosahexaenoyl glycerol (1- or 2DHG) [83], but it could possibly serve as a DHA pool for 19,20EDP synthesis [60]. 4.2. Branched fatty acid esters of hydroxy fatty acids (FAHFA) e immunomodulation, resolution of inflammation There are currently three members of FAHFA family containing DHA: 9- and 13-DHAHLA in which DHA is esterified at the 9th and 13th carbon of hydroxy linoleic acid, and 14-DHAHDHA where DHA is esterified at the 14th carbon of 14-HDHA. We have recently demonstrated that 13-DHAHLA exerts strong anti-inflammatory and proresolving properties while reducing macrophage activation by lipopolysaccharides and enhancing the phagocytosis of

O. Kuda / Biochimie 136 (2017) 12e20 Table 1 DHA metabolites identified in humans. DHA metabolites identified in humans

Human tissue

serum [110], milk [111], synovial fluid [112], leukocytes [39] Maresin 2, 13,14-epoxy- macrophages [40] maresin MCTRs lymph nodes, serum, plasma [42] Maresin-L1, Maresin-L2 leukocytes, platelets [43] Protectin D1 serum [110], cerebrospinal fluid [113,114], hippocampus [115], embryonic stem cells [116], breath condensates/lungs [117], milk [111] RvD1, RvD2 plasma, lymph tissue [110,118,119], adipose tissue [120,121], milk [111,122], cerebrospinal fluid [113,123] RvD3, RvD4 serum [103,110], milk [111] RvD5 synovial fluid [112] RCTRs, PCTRs spleen, macrophages [32] Neuroprostanes/ brain [67], retina [68], urine [124] Neuroketals/CEPs 19,20-EDP, 16,17-EDP plasma [119] DHEA blood [81,86] DHAHLA serum, adipose tissue [84] Maresin 1

zymosan particles [84]. 4.3. N-acyl amides - metabolic regulation, neuroprotection, neurotransmission Recently, an enzyme PM20D1 catalyzing condensation of fatty acids and amino acid (i.e. Phenylalanines) to N-acyl amides has been discovered [33], shedding light to the origin of these mediators which improve glucose homeostasis and increase energy expenditure. A group of N-acyl amides including N-docosahexaenoyl-Ser, Tyr, Trp, Gly, Pro, Val and Asp has been identified as modulators of vanilloid receptorerelated transient receptor potential (TRPV) channels [23]. The authors also noted that N-docosahexaenoyl amides are significantly more stable than other N-acyl amides and may even act as precursor molecules for the more reactive and fast-lived resolvins [23]. DHA could be also coupled with neurotransmitters to form i) N-docosahexaenoyl dopamine [85,86] that is capable of inducing PPARg-dependent breast cancer cell death [87]; ii) neuroprotective N-docosahexaenoyl serotonin [88] that was shown to inhibit glucagon-like peptide-1 secretion [89]; and iii) N-docosahexaenoyl GABA [23]. 5. Conclusion The purpose of this review was to summarize the knowledge about metabolites of DHA with documented biological activity and to illustrate their diversity. Many of the compounds listed above, especially non-enzymatic and P450 products, have been identified and tested only in-vitro and their structure with respect to chirality and configuration of double bonds was poorly explored. The stereochemistry is especially important for di- and tri-HDHA. Validation of these parameters by analytical methods and in vivo biological models is critical and the need is documented by the substitution of PD1 for PDX isomer [50], where the authors originally didn't differentiate between the isomers of 10,17-diHDHA [48,50,51]. Rigorous characterization of resolving E1 including stereochemical assignment, identification of biosynthetic pathway, receptor, and biological actions in cells, mice and humans illustrates the complexity of the research in the field of lipid mediators [90]. Total organic (stereoselective) synthesis, preferably coupled to biosynthesis, was achieved for Maresin 1 [91], Maresin 2 [92], eMaresin [38],MCTRs [41,42], 14,20- and 14,21-diHDHA [44,93],

17

PD1, AT-PD1, PDX [94e96], 22-OH-PD1 [46], PCTRs [97,98], RvD1 and AT-RvD1 [99], RvD2 [100,101], RvD3 and AT-RvD3 [102], RvD4 [103], RvD5 and RvD6 [104,105], Neuroprostanes [106,107], etc. The interaction between biologists and organic chemists is crucial for full characterization of new bioactive lipids, but a demanding stereoselective synthesis (reviewed in Ref. [108]) complicates the process. Origin of the active metabolites adds another layer of complexity. Compounds synthesized locally via enzymes in small quantities and carrying a biological activity could be linked to a specific tissue or a cell type (e.g. maresins in macrophages), but metabolites formed by non-enzymatic reactions could be formed and act nearly anywhere at various concentrations. Their biological potency follows the same trend e spanning from picomolar concentrations for tightly controlled SPMs to micromolar levels for NPs. The biosynthetic machinery and its regulation in time and space represents the key difference between metabolites fulfilling criteria for autacoids and lipid mediators (e.g. SPMs [109]) and the other bioactive metabolites produced in response to oxidative stress (e.g. neuroprostanes) or lipids involved in general signal transduction. It is important to note that only a subgroup of DHA metabolites (Table 1) was identified in humans, either in tissues, isolated cells or fluids, and that their relevance to human pathology has to be explored in the future. The notion that high intake of omega-3 results in negative or off-target effects suggests that special attention should be paid to the bioavailability of omega-3. Comparison of different forms of omega-3 supplementation (seafood, triacylglycerols, phospholipids, ethyl esters, waxes, etc.) in the context of health and disease (metabolic diseases, cardiovascular and cancer risk factors, etc.) might lead to optimized strategies for each situation. Alternatively, targeted use of bioactive DHA metabolites or their analogs with prolonged metabolic stability could be a therapeutic approach in future. It is well documented that DHA and its metabolites have many beneficial effects on human health. However, it is clear that our knowledge of the receptors mediating direct effects of DHA and those of its diverse metabolites is incomplete. Importantly, given the chemical structure of DHA and the known multiple pathways of its metabolism, novel bioactive derivatives of this important fatty acid could be expected in future. Involvement of stereospecific organic synthesis and validation of biosynthetic pathways will be needed to undoubtedly assign biological effects to a specific metabolite. The obvious challenge will be to translate the beneficial effects of the newly discovered DHA-derived lipid mediators, which are observed at the molecular level, into clinical practice and into the new therapeutic strategies. Acknowledgements This study was supported by grants from the Czech Science Foundation (GA16-04859S) and the Ministry of Health of the Czech Republic (NV16-29182A). The author thanks N.A. Abumrad, M. Rossmeisl and J. Kopecky for critical reading of the manuscript and helpful suggestions. Appendix A. Supplementary data Supplementary data related to this article can be found at http:// dx.doi.org/10.1016/j.biochi.2017.01.002. References [1] D. Mozaffarian, R.N. Lemaitre, I.B. King, X. Song, H. Huang, F.M. Sacks, E.B. Rimm, M. Wang, D.S. Siscovick, Plasma phospholipid long-chain omega-

18

O. Kuda / Biochimie 136 (2017) 12e20

[2]

[3]

[4] [5]

[6]

[7]

[8]

[9]

[10]

[11]

[12]

[13] [14]

[15]

[16]

[17]

[18]

[19]

[20]

[21]

[22]

[23]

[24]

3 fatty acids and total and cause-specific mortality in older adults: a cohort study, Ann. Intern Med. 158 (2013) 515e525. E. Scorletti, L. Bhatia, K.G. McCormick, G.F. Clough, K. Nash, L. Hodson, H.E. Moyses, P.C. Calder, C.D. Byrne, W. Study, Effects of purified eicosapentaenoic and docosahexaenoic acids in nonalcoholic fatty liver disease: results from the Welcome* study, Hepatology 60 (2014) 1211e1221. P.M. Kris-Etherton, W.S. Harris, L.J. Appel, A.H.A.N.C.A.H. Association, Omega3 fatty acids and cardiovascular disease: new recommendations from the American heart association, Arterioscler. Thromb. Vasc. Biol. 23 (2003) 151e152. P. Flachs, M. Rossmeisl, J. Kopecky, The effect of n-3 fatty acids on glucose homeostasis and insulin sensitivity, Physiol. Res. 63 (1) (2014) S93eS118. M. Masoodi, O. Kuda, M. Rossmeisl, P. Flachs, J. Kopecky, Lipid signaling in adipose tissue: connecting inflammation & metabolism, Biochim. Biophys. Acta 1851 (2015) 503e518. P.C. Calder, Marine omega-3 fatty acids and inflammatory processes: effects, mechanisms and clinical relevance, Biochim. Biophys. Acta 1851 (2015) 469e484. T.M. Brasky, A.K. Darke, X. Song, C.M. Tangen, P.J. Goodman, I.M. Thompson, F.L. Meyskens Jr., G.E. Goodman, L.M. Minasian, H.L. Parnes, E.A. Klein, A.R. Kristal, Plasma phospholipid fatty acids and prostate cancer risk in the SELECT trial, J. Natl. Cancer Inst. 105 (2013) 1132e1141. D. Kromhout, E.J. Giltay, J.M. Geleijnse, Alpha Omega Trial, n-3 fatty acids and cardiovascular events after myocardial infarction, N. Engl. J. Med. 363 (2010) 2015e2026. M. Haas-Haseman, Weighing the benefits of fish oil for patients with prostate cancer: a subcohort review from the select trial, J. Adv. Pract. Oncol. 6 (2015) 376e378. D. Kromhout, J.M. Geleijnse, J. de Goede, L.M. Oude Griep, B.J. Mulder, M.J. de Boer, J.W. Deckers, E. Boersma, P.L. Zock, E.J. Giltay, n-3 fatty acids, ventricular arrhythmia-related events, and fatal myocardial infarction in postmyocardial infarction patients with diabetes, Diabetes Care 34 (2011) 2515e2520. M.C. de Oliveira Otto, J.H. Wu, A. Baylin, D. Vaidya, S.S. Rich, M.Y. Tsai, D.R. Jacobs Jr., D. Mozaffarian, Circulating and dietary omega-3 and omega-6 polyunsaturated fatty acids and incidence of CVD in the multi-ethnic study of atherosclerosis, J. Am. Heart Assoc. 2 (2013) e000506. J. Madden, C.M. Williams, P.C. Calder, G. Lietz, E.A. Miles, H. Cordell, J.C. Mathers, A.M. Minihane, The impact of common gene variants on the response of biomarkers of cardiovascular disease (CVD) risk to increased fish oil fatty acids intakes, Annu. Rev. Nutr. 31 (2011) 203e234. D. Corella, J.M. Ordovas, Interactions between dietary n-3 fatty acids and genetic variants and risk of disease, Br. J. Nutr. 107 (2012) S271eS283. Q.N. Diep, R.M. Touyz, E.L. Schiffrin, Docosahexaenoic acid, a peroxisome proliferator-activated receptor-alpha ligand, induces apoptosis in vascular smooth muscle cells by stimulation of p38 mitogen-activated protein kinase, Hypertension 36 (2000) 851e855. J. Zuniga, M. Cancino, F. Medina, P. Varela, R. Vargas, G. Tapia, L.A. Videla, V. Fernandez, N-3 PUFA supplementation triggers PPAR-alpha activation and PPAR-alpha/NF-kappaB interaction: anti-inflammatory implications in liver ischemia-reperfusion injury, PLoS One 6 (2011) e28502. D.Y. Oh, S. Talukdar, E.J. Bae, T. Imamura, H. Morinaga, W. Fan, P. Li, W.J. Lu, S.M. Watkins, J.M. Olefsky, GPR120 is an omega-3 fatty acid receptor mediating potent anti-inflammatory and insulin-sensitizing effects, Cell 142 (2010) 687e698. D.Y. Oh, E. Walenta, T.E. Akiyama, W.S. Lagakos, D. Lackey, A.R. Pessentheiner, R. Sasik, N. Hah, T.J. Chi, J.M. Cox, M.A. Powels, J. Di Salvo, C. Sinz, S.M. Watkins, A.M. Armando, H. Chung, R.M. Evans, O. Quehenberger, J. McNelis, J.G. Bogner-Strauss, J.M. Olefsky, A Gpr120-selective agonist improves insulin resistance and chronic inflammation in obese mice, Nat. Med. 20 (2014) 942e947. K. Mobraten, T.M. Haug, C.R. Kleiveland, T. Lea, Omega-3 and omega-6 PUFAs induce the same GPR120-mediated signalling events, but with different kinetics and intensity in Caco-2 cells, Lipids Health Dis. 12 (2013) 101. M. Bjursell, X. Xu, T. Admyre, G. Bottcher, S. Lundin, R. Nilsson, V.M. Stone, N.G. Morgan, Y.Y. Lam, L.H. Storlien, D. Linden, D.M. Smith, Y.M. Bohlooly, J. Oscarsson, The beneficial effects of n-3 polyunsaturated fatty acids on diet induced obesity and impaired glucose control do not require Gpr120, PLoS One 9 (2014) e114942. S.I. Pærregaard, M. Agerholm, A.K. Serup, T. Ma, B. Kiens, L. Madsen, K. Kristiansen, B.A.H. Jensen, FFAR4 (GPR120) signaling is not required for anti-inflammatory and insulin-sensitizing effects of Omega-3 fatty acids, Mediat. Inflamm. 2016 (2016) 1e12. S.V. Shewale, A.L. Brown, X. Bi, E. Boudyguina, J.K. Sawyer, M.A. AlexanderMiller, J.S. Parks, In vivo activation of leukocyte GPR120/FFAR4 by PUFAs has minimal impact on atherosclerosis in LDL receptor knockout mice, J. Lipid Res. 58 (1) (2017 Jan) 236e246, http://dx.doi.org/10.1194/jlr.M072769. M. Spite, J. Claria, C.N. Serhan, Resolvins, specialized proresolving lipid mediators, and their potential roles in metabolic diseases, Cell Metab. 19 (2014) 21e36. S. Raboune, J.M. Stuart, E. Leishman, S.M. Takacs, B. Rhodes, A. Basnet, E. Jameyfield, D. McHugh, T. Widlanski, H.B. Bradshaw, Novel endogenous Nacyl amides activate TRPV1-4 receptors, BV-2 microglia, and are regulated in brain in an acute model of inflammation, Front. Cell Neurosci. 8 (2014) 195. I. Brown, M.G. Cascio, K.W. Wahle, R. Smoum, R. Mechoulam, R.A. Ross,

[25]

[26]

[27]

[28]

[29]

[30]

[31]

[32]

[33]

[34]

[35] [36]

[37]

[38]

[39]

[40]

[41]

[42]

[43]

[44]

[45]

[46]

R.G. Pertwee, S.D. Heys, Cannabinoid receptor-dependent and -independent anti-proliferative effects of omega-3 ethanolamides in androgen receptorpositive and -negative prostate cancer cell lines, Carcinogenesis 31 (2010) 1584e1591. J.W. Lee, B.X. Huang, H. Kwon, M.A. Rashid, G. Kharebava, A. Desai, S. Patnaik, J. Marugan, H.Y. Kim, Orphan GPR110 (ADGRF1) targeted by N-docosahexaenoylethanolamine in development of neurons and cognitive function, Nat. Commun. 7 (2016) 13123. D. Egawa, T. Itoh, Y. Akiyama, T. Saito, K. Yamamoto, 17-OxoDHA is a PPARalpha/gamma dual covalent modifier and agonist, ACS Chem. Biol. 11 (2016) 2447e2455. A.L. Groeger, C. Cipollina, M.P. Cole, S.R. Woodcock, G. Bonacci, T.K. Rudolph, V. Rudolph, B.A. Freeman, F.J. Schopfer, Cyclooxygenase-2 generates antiinflammatory mediators from omega-3 fatty acids, Nat. Chem. Biol. 6 (2010) 433e441. C.N. Serhan, J. Dalli, R.A. Colas, J.W. Winkler, N. Chiang, Protectins and maresins: new pro-resolving families of mediators in acute inflammation and resolution bioactive metabolome, Biochim. Biophys. Acta 1851 (2015) 397e413. C.N. Serhan, S. Hong, K. Gronert, S.P. Colgan, P.R. Devchand, G. Mirick, R.L. Moussignac, 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 (2002) 1025e1037. S.G. Wendell, F. Golin-Bisello, S. Wenzel, R.W. Sobol, F. Holguin, B.A. Freeman, 15-Hydroxyprostaglandin dehydrogenase generation of electrophilic lipid signaling mediators from hydroxy omega-3 fatty acids, J. Biol. Chem. 290 (2015) 5868e5880. C. Morisseau, B. Inceoglu, K. Schmelzer, H.J. Tsai, S.L. Jinks, C.M. Hegedus, B.D. Hammock, Naturally occurring monoepoxides of eicosapentaenoic acid and docosahexaenoic acid are bioactive antihyperalgesic lipids, J. Lipid Res. 51 (2010) 3481e3490. J. Dalli, S. Ramon, P.C. Norris, R.A. Colas, C.N. Serhan, Novel proresolving and tissue-regenerative resolvin and protectin sulfido-conjugated pathways, FASEB J. 29 (2015) 2120e2136. J.Z. Long, K.J. Svensson, L.A. Bateman, H. Lin, T. Kamenecka, I.A. Lokurkar, J. Lou, R.R. Rao, M.R. Chang, M.P. Jedrychowski, J.A. Paulo, S.P. Gygi, P.R. Griffin, D.K. Nomura, B.M. Spiegelman, The secreted enzyme PM20D1 regulates lipidated amino acid uncouplers of mitochondria, Cell 166 (2016) 424e435. J. Roy, C. Oger, J. Thireau, J. Roussel, O. Mercier-Touzet, D. Faure, E. Pinot, C. Farah, D.F. Taber, J.P. Cristol, J.C. Lee, A. Lacampagne, J.M. Galano, T. Durand, J.Y. Le Guennec, Nonenzymatic lipid mediators, neuroprostanes, exert the antiarrhythmic properties of docosahexaenoic acid, Free Radic. Biol. Med. 86 (2015) 269e278. C.N. Serhan, N. Chiang, Resolution phase lipid mediators of inflammation: agonists of resolution, Curr. Opin. Pharmacol. 13 (2013) 632e640. C.N. Serhan, R. Yang, K. Martinod, K. Kasuga, P.S. Pillai, T.F. Porter, S.F. Oh, M. Spite, Maresins: novel macrophage mediators with potent antiinflammatory and proresolving actions, J. Exp. Med. 206 (2009) 15e23. C.N. Serhan, J. Dalli, S. Karamnov, A. Choi, C.K. Park, Z.Z. Xu, R.R. Ji, M. Zhu, N.A. Petasis, Macrophage proresolving mediator maresin 1 stimulates tissue regeneration and controls pain, FASEB J. 26 (2012) 1755e1765. J. Dalli, M. Zhu, N.A. Vlasenko, B. Deng, J.Z. Haeggstrom, N.A. Petasis, C.N. Serhan, 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 (2013) 2573e2583. R.A. Colas, J. Dalli, N. Chiang, I. Vlasakov, J.M. Sanger, I.R. Riley, C.N. Serhan, Identification and actions of the maresin 1 metabolome in infectious inflammation, J. Immunol. 197 (2016) 4444e4452. B. Deng, C.W. Wang, H.H. Arnardottir, Y. Li, C.Y. Cheng, J. Dalli, C.N. Serhan, Maresin biosynthesis and identification of maresin 2, a new antiinflammatory and pro-resolving mediator from human macrophages, PLoS One 9 (2014) e102362. J. Dalli, J.M. Sanger, A.R. Rodriguez, N. Chiang, B.W. Spur, C.N. Serhan, Identification and actions of a novel third maresin conjugate in tissue regeneration: MCTR3, PLoS One 11 (2016) e0149319. J. Dalli, I. Vlasakov, I.R. Riley, A.R. Rodriguez, B.W. Spur, N.A. Petasis, N. Chiang, C.N. Serhan, Maresin conjugates in tissue regeneration biosynthesis enzymes in human macrophages, Proc. Natl. Acad. Sci. U. S. A. 113 (2016) 12232e12237. S. Hong, Y. Lu, H. Tian, B.V. Alapure, Q. Wang, B.A. Bunnell, J.M. Laborde, Maresin-like lipid mediators are produced by leukocytes and platelets and rescue reparative function of diabetes-impaired macrophages, Chem. Biol. 21 (2014) 1318e1329. Y. Lu, H. Tian, S. Hong, Novel 14,21-dihydroxy-docosahexaenoic acids: structures, formation pathways, and enhancement of wound healing, J. Lipid Res. 51 (2010) 923e932. Y. Yokokura, Y. Isobe, S. Matsueda, R. Iwamoto, T. Goto, T. Yoshioka, D. Urabe, M. Inoue, H. Arai, M. Arita, Identification of 14,20-dihydroxy-docosahexaenoic acid as a novel anti-inflammatory metabolite, J. Biochem. 156 (2014) 315e321. J.E. Tungen, M. Aursnes, A. Vik, S. Ramon, R.A. Colas, J. Dalli, C.N. Serhan, T.V. Hansen, Synthesis and anti-inflammatory and pro-resolving activities of 22-OH-PD1, a monohydroxylated metabolite of protectin D1, J. Nat. Prod. 77 (2014) 2241e2247.

O. Kuda / Biochimie 136 (2017) 12e20 [47] P. Chen, B. Fenet, S. Michaud, N. Tomczyk, E. Vericel, M. Lagarde, M. Guichardant, Full characterization of PDX, a neuroprotectin/protectin D1 isomer, which inhibits blood platelet aggregation, FEBS Lett. 583 (2009) 3478e3484. [48] L. Balas, M. Guichardant, T. Durand, M. Lagarde, Confusion between protectin D1 (PD1) and its isomer protectin DX (PDX). An overview on the dihydroxydocosatrienes described to date, Biochimie 99 (2014) 1e7. [49] M. Liu, T. Boussetta, K. Makni-Maalej, M. Fay, F. Driss, J. El-Benna, M. Lagarde, M. Guichardant, Protectin DX, a double lipoxygenase product of DHA, inhibits both ROS production in human neutrophils and cyclooxygenase activities, Lipids 49 (2014) 49e57. [50] M. Morita, K. Kuba, A. Ichikawa, M. Nakayama, J. Katahira, R. Iwamoto, T. Watanebe, S. Sakabe, T. Daidoji, S. Nakamura, A. Kadowaki, T. Ohto, H. Nakanishi, R. Taguchi, T. Nakaya, M. Murakami, Y. Yoneda, H. Arai, Y. Kawaoka, J.M. Penninger, M. Arita, Y. Imai, The lipid mediator protectin D1 inhibits influenza virus replication and improves severe influenza, Cell 153 (2013) 112e125. [51] Y. Imai, Role of omega-3 PUFA-derived mediators, the protectins, in influenza virus infection, Biochim. Biophys. Acta 1851 (2015) 496e502. [52] S. Ramon, S.F. Baker, J.M. Sahler, N. Kim, E.A. Feldsott, C.N. Serhan, L. Martinez-Sobrido, D.J. Topham, R.P. Phipps, The specialized proresolving mediator 17-HDHA enhances the antibody-mediated immune response against influenza virus: a new class of adjuvant? J. Immunol. 193 (2014) 6031e6040. [53] S. Bohr, S.J. Patel, D. Sarin, D. Irimia, M.L. Yarmush, F. Berthiaume, Resolvin D2 prevents secondary thrombosis and necrosis in a mouse burn wound model, Wound Repair Regen. 21 (2013) 35e43. [54] Y. Tang, M.J. Zhang, J. Hellmann, M. Kosuri, A. Bhatnagar, M. Spite, Proresolution therapy for the treatment of delayed healing of diabetic wounds, Diabetes 62 (2013) 618e627. [55] J.F. Lima-Garcia, R.C. Dutra, K. da Silva, E.M. Motta, M.M. Campos, J.B. Calixto, The precursor of resolvin D series and aspirin-triggered resolvin D1 display anti-hyperalgesic properties in adjuvant-induced arthritis in rats, Br. J. Pharmacol. 164 (2011) 278e293. [56] C. Cipollina, S.R. Salvatore, M.F. Muldoon, B.A. Freeman, F.J. Schopfer, Generation and dietary modulation of anti-inflammatory electrophilic omega-3 fatty acid derivatives, PLoS One 9 (2014) e94836. [57] T.J. Pogash, K. El-Bayoumy, S. Amin, K. Gowda, R.L. de Cicco, M. Barton, Y. Su, I.H. Russo, J.A. Himmelberger, M. Slifker, A. Manni, J. Russo, Oxidized derivative of docosahexaenoic acid preferentially inhibit cell proliferation in triple negative over luminal breast cancer cells, In Vitro Cell. Dev. Biol. Anim. 51 (2015) 121e127. [58] C. Arnold, M. Markovic, K. Blossey, G. Wallukat, R. Fischer, R. Dechend, A. Konkel, C. von Schacky, F.C. Luft, D.N. Muller, M. Rothe, W.H. Schunck, Arachidonic acid-metabolizing cytochrome P450 enzymes are targets of u-3 fatty acids, J. Biol. Chem. 285 (2010) 32720e32733. [59] A. Ulu, K.S. Stephen Lee, C. Miyabe, J. Yang, B.G. Hammock, H. Dong, B.D. Hammock, An omega-3 epoxide of docosahexaenoic acid lowers blood pressure in angiotensin-II-dependent hypertension, J. Cardiovasc Pharmacol. 64 (2014) 87e99. [60] C. Morin, S. Fortin, E. Rousseau, 19,20-EpDPE, a bioactive CYP450 metabolite of DHA monoacyglyceride, decreases Ca(2)(þ) sensitivity in human pulmonary arteries, Am. J. Physiol. Heart Circ. Physiol. 301 (2011) H1311eH1318. [61] G. Zhang, D. Panigrahy, L.M. Mahakian, J. Yang, J.Y. Liu, K.S. Stephen Lee, H.I. Wettersten, A. Ulu, X. Hu, S. Tam, S.H. Hwang, E.S. Ingham, M.W. Kieran, R.H. Weiss, K.W. Ferrara, B.D. Hammock, Epoxy metabolites of docosahexaenoic acid (DHA) inhibit angiogenesis, tumor growth, and metastasis, Proc. Natl. Acad. Sci. U. S. A. 110 (2013) 6530e6535. [62] C. Lopez-Vicario, J. Alcaraz-Quiles, V. Garcia-Alonso, B. Rius, S.H. Hwang, E. Titos, A. Lopategi, B.D. Hammock, V. Arroyo, J. Claria, Inhibition of soluble epoxide hydrolase modulates inflammation and autophagy in obese adipose tissue and liver: role for omega-3 epoxides, Proc. Natl. Acad. Sci. U. S. A. 112 (2015) 536e541. [63] K. Wagner, K.S. Lee, J. Yang, B.D. Hammock, Epoxy fatty acids mediate analgesia in murine diabetic neuropathy, Eur. J. Pain. (2016 Sep 15), http:// dx.doi.org/10.1002/ejp.939 (Epub ahead of print). [64] H. Yin, E.S. Musiek, L. Gao, N.A. Porter, J.D. Morrow, Regiochemistry of neuroprostanes generated from the peroxidation of docosahexaenoic acid in vitro and in vivo, J. Biol. Chem. 280 (2005) 26600e26611. [65] S. Jude, S. Bedut, S. Roger, M. Pinault, P. Champeroux, E. White, J.Y. Le Guennec, Peroxidation of docosahexaenoic acid is responsible for its effects on I TO and I SS in rat ventricular myocytes, Br. J. Pharmacol. 139 (2003) 816e822. [66] E.S. Musiek, J.D. Brooks, M. Joo, E. Brunoldi, A. Porta, G. Zanoni, G. Vidari, T.S. Blackwell, T.J. Montine, G.L. Milne, B. McLaughlin, J.D. Morrow, Electrophilic cyclopentenone neuroprostanes are anti-inflammatory mediators formed from the peroxidation of the omega-3 polyunsaturated fatty acid docosahexaenoic acid, J. Biol. Chem. 283 (2008) 19927e19935. [67] N. Bernoud-Hubac, S.S. Davies, O. Boutaud, T.J. Montine, L.J. Roberts 2nd, Formation of highly reactive gamma-ketoaldehydes (neuroketals) as products of the neuroprostane pathway, J. Biol. Chem. 276 (2001) 30964e30970. [68] X. Gu, S.G. Meer, M. Miyagi, M.E. Rayborn, J.G. Hollyfield, J.W. Crabb, R.G. Salomon, Carboxyethylpyrrole protein adducts and autoantibodies, biomarkers for age-related macular degeneration, J. Biol. Chem. 278 (2003) 42027e42035.

19

[69] X.Z. West, N.L. Malinin, A.A. Merkulova, M. Tischenko, B.A. Kerr, E.C. Borden, E.A. Podrez, R.G. Salomon, T.V. Byzova, Oxidative stress induces angiogenesis by activating TLR2 with novel endogenous ligands, Nature 467 (2010) 972e976. [70] Y.W. Kim, V.P. Yakubenko, X.Z. West, G.B. Gugiu, K. Renganathan, S. Biswas, D. Gao, J.W. Crabb, R.G. Salomon, E.A. Podrez, T.V. Byzova, Receptor-mediated mechanism controlling tissue levels of bioactive lipid oxidation products, Circ. Res. 117 (2015) 321e332. [71] M.Y. Pepino, O. Kuda, D. Samovski, N.A. Abumrad, Structure-function of CD36 and importance of fatty acid signal transduction in fat metabolism, Annu. Rev. Nutr. 34 (2014) 281e303. [72] E. Vericel, A. Polette, S. Bacot, C. Calzada, M. Lagarde, Pro- and antioxidant activities of docosahexaenoic acid on human blood platelets, J. Thromb. Haemost. 1 (2003) 566e572. [73] M.A. Rashid, M. Katakura, G. Kharebava, K. Kevala, H.Y. Kim, N-Docosahexaenoylethanolamine is a potent neurogenic factor for neural stem cell differentiation, J. Neurochem. 125 (2013) 869e884. [74] H.Y. Kim, H.S. Moon, D. Cao, J. Lee, K. Kevala, S.B. Jun, D.M. Lovinger, M. Akbar, B.X. Huang, N-Docosahexaenoylethanolamide promotes development of hippocampal neurons, Biochem. J. 435 (2011) 327e336. [75] T. Park, H. Chen, K. Kevala, J.W. Lee, H.Y. Kim, N-Docosahexaenoylethanolamine ameliorates LPS-induced neuroinflammation via cAMP/PKAdependent signaling, J. Neuroinflam. 13 (2016) 284. [76] J. Meijerink, P. Plastina, J.P. Vincken, M. Poland, M. Attya, M. Balvers, H. Gruppen, B. Gabriele, R.F. Witkamp, The ethanolamide metabolite of DHA, docosahexaenoylethanolamine, shows immunomodulating effects in mouse peritoneal and RAW264.7 macrophages: evidence for a new link between fish oil and inflammation, Br. J. Nutr. 105 (2011) 1798e1807. [77] J. Meijerink, M. Poland, M.G. Balvers, P. Plastina, C. Lute, J. Dwarkasing, K. van Norren, R.F. Witkamp, Inhibition of COX-2-mediated eicosanoid production plays a major role in the anti-inflammatory effects of the endocannabinoid N-docosahexaenoylethanolamine (DHEA) in macrophages, Br. J. Pharmacol. 172 (2015) 24e37. [78] M. Rossmeisl, Z.M. Jilkova, O. Kuda, T. Jelenik, D. Medrikova, B. Stankova, B. Kristinsson, G.G. Haraldsson, H. Svensen, I. Stoknes, P. Sjovall, Y. Magnusson, M.G. Balvers, K.C. Verhoeckx, E. Tvrzicka, M. Bryhn, J. Kopecky, Metabolic effects of n-3 PUFA as phospholipids are superior to triglycerides in mice fed a high-fat diet: possible role of endocannabinoids, PLoS One 7 (2012) e38834. [79] J. Kim, M.E. Carlson, B.A. Watkins, Docosahexaenoyl ethanolamide improves glucose uptake and alters endocannabinoid system gene expression in proliferating and differentiating C2C12 myoblasts, Front. Physiol. 5 (2014) 100. [80] A. Artmann, G. Petersen, L.I. Hellgren, J. Boberg, C. Skonberg, C. Nellemann, S.H. Hansen, H.S. Hansen, Influence of dietary fatty acids on endocannabinoid and N-acylethanolamine levels in rat brain, liver and small intestine, Biochim. Biophys. Acta 1781 (2008) 200e212. [81] R. Yang, G. Fredman, S. Krishnamoorthy, N. Agrawal, D. Irimia, D. Piomelli, C.N. Serhan, Decoding functional metabolomics with docosahexaenoyl ethanolamide (DHEA) identifies novel bioactive signals, J. Biol. Chem. 286 (2011) 31532e31541. [82] M. Shinohara, V. Mirakaj, C.N. Serhan, Functional metabolomics reveals novel active products in the DHA metabolome, Front. Immunol. 3 (2012) 81. [83] T. Bisogno, I. Delton-Vandenbroucke, A. Milone, M. Lagarde, V. Di Marzo, Biosynthesis and inactivation of N-arachidonoylethanolamine (anandamide) and N-docosahexaenoylethanolamine in bovine retina, Arch. Biochem. Biophys. 370 (1999) 300e307. [84] O. Kuda, M. Brezinova, M. Rombaldova, B. Slavikova, M. Posta, P. Beier, P. Janovska, J. Veleba, J. Kopecky Jr., E. Kudova, T. Pelikanova, J. Kopecky, Docosahexaenoic acid-derived fatty acid esters of hydroxy fatty acids (FAHFAs) with anti-inflammatory properties, Diabetes 65 (2016) 2580e2590. [85] V.E. Shashoua, G.W. Hesse, N-docosahexaenoyl, 3 hydroxytyramine: a dopaminergic compound that penetrates the blood-brain barrier and suppresses appetite, Life Sci. 58 (1996) 1347e1357. [86] J. Meijerink, M. Balvers, R. Witkamp, N-Acyl amines of docosahexaenoic acid and other n-3 polyunsatured fatty acids - from fishy endocannabinoids to potential leads, Br. J. Pharmacol. 169 (2013) 772e783. [87] D. Rovito, C. Giordano, P. Plastina, I. Barone, F. De Amicis, L. Mauro, P. Rizza, M. Lanzino, S. Catalano, D. Bonofiglio, S. Ando, Omega-3 DHA- and EPAdopamine conjugates induce PPARgamma-dependent breast cancer cell death through autophagy and apoptosis, Biochim. Biophys. Acta 1850 (2015) 2185e2195. [88] M.C. Jin, J.M. Yoo, D.E. Sok, M.R. Kim, Neuroprotective effect of N-acyl 5hydroxytryptamines on glutamate-induced cytotoxicity in HT-22 cells, Neurochem. Res. 39 (2014) 2440e2451. [89] K.C. Verhoeckx, T. Voortman, M.G. Balvers, H.F. Hendriks, M. Wortelboer H, R.F. Witkamp, Presence, formation and putative biological activities of N-acyl serotonins, a novel class of fatty-acid derived mediators, in the intestinal tract, Biochim. Biophys. Acta 1811 (2011) 578e586. [90] M. Arita, F. Bianchini, J. Aliberti, A. Sher, N. Chiang, S. Hong, R. Yang, N.A. Petasis, C.N. Serhan, Stereochemical assignment, antiinflammatory properties, and receptor for the omega-3 lipid mediator resolvin E1, J. Exp. Med. 201 (2005) 713e722. [91] A.R. Rodriguez, B.W. Spur, Total synthesis of the macrophage derived anti-

20

[92]

[93]

[94]

[95]

[96]

[97]

[98]

[99]

[100]

[101]

[102]

[103]

[104] [105]

[106]

[107]

[108] [109]

O. Kuda / Biochimie 136 (2017) 12e20 inflammatory lipid mediator Maresin 1, Tetrahedron Lett. 53 (2012) 4169e4172. A.R. Rodriguez, B.W. Spur, First total synthesis of the macrophage derived anti-inflammatory and pro-resolving lipid mediator Maresin 2, Tetrahedron Lett. 56 (2015) 256e259. T. Goto, D. Urabe, K. Masuda, Y. Isobe, M. Arita, M. Inoue, Total synthesis of four stereoisomers of (4Z,7Z,10Z,12E,16Z,18E)-14,20-dihydroxy4,7,10,12,16,18-docosahexaenoic acid and their anti-inflammatory activities, J. Org. Chem. 80 (2015) 7713e7726. N.A. Petasis, R. Yang, J.W. Winkler, M. Zhu, J. Uddin, N.G. Bazan, C.N. Serhan, Stereocontrolled total synthesis of neuroprotectin D1/protectin D1 and its aspirin-triggered stereoisomer, Tetrahedron Lett. 53 (2012) 1695e1698. M. Aursnes, J.E. Tungen, A. Vik, J. Dalli, T.V. Hansen, Stereoselective synthesis of protectin D1: a potent anti-inflammatory and proresolving lipid mediator, Org. Biomol. Chem. 12 (2014) 432e437. C.N. Serhan, K. Gotlinger, S. Hong, Y. Lu, J. Siegelman, T. Baer, R. Yang, S.P. Colgan, N.A. Petasis, Anti-inflammatory actions of neuroprotectin D1/ protectin D1 and its natural stereoisomers: assignments of dihydroxycontaining docosatrienes, J. Immunol. 176 (2006) 1848e1859. S. Ramon, J. Dalli, J.M. Sanger, J.W. Winkler, M. Aursnes, J.E. Tungen, T.V. Hansen, C.N. Serhan, The protectin PCTR1 is produced by human M2 macrophages and enhances resolution of infectious inflammation, Am. J. Pathol. 186 (2016) 962e973. A.R. Rodriguez, B.W. Spur, Total synthesis of pro-resolving and tissueregenerative Protectin sulfido-conjugates, Tetrahedron Lett. 56 (2015) 5811e5815. Y.P. Sun, S.F. Oh, J. Uddin, R. Yang, K. Gotlinger, E. Campbell, S.P. Colgan, N.A. Petasis, C.N. Serhan, Resolvin D1 and its aspirin-triggered 17R epimer. Stereochemical assignments, anti-inflammatory properties, and enzymatic inactivation, J. Biol. Chem. 282 (2007) 9323e9334. M. Spite, L.V. Norling, L. Summers, R. Yang, D. Cooper, N.A. Petasis, R.J. Flower, M. Perretti, C.N. Serhan, Resolvin D2 is a potent regulator of leukocytes and controls microbial sepsis, Nature 461 (2009) 1287e1291. A.R. Rodrıguez, B.W. Spur, First total synthesis of 7(S),16(R),17(S)-Resolvin D2, a potent anti-inflammatory lipid mediator, Tetrahedron Lett. 45 (2004) 8717e8720. J.W. Winkler, J. Uddin, C.N. Serhan, N.A. Petasis, Stereocontrolled total synthesis of the potent anti-inflammatory and pro-resolving lipid mediator resolvin D3 and its aspirin-triggered 17R-epimer, Org. Lett. 15 (2013) 1424e1427. J.W. Winkler, S.K. Orr, J. Dalli, C.Y. Cheng, J.M. Sanger, N. Chiang, N.A. Petasis, C.N. Serhan, Resolvin D4 stereoassignment and its novel actions in host protection and bacterial clearance, Sci. Rep. 6 (2016) 18972. A.R. Rodriguez, B.W. Spur, First total synthesis of the anti-inflammatory lipid mediator Resolvin D6, Tetrahedron Lett. 53 (2012) 86e89. A.R. Rodriguez, B.W. Spur, First total synthesis of 7(S),17(S)-resolvin D5, a potent anti-inflammatory docosanoid, Tetrahedron Lett. 46 (2005) 3623e3627. A. Porta, M. Pasi, E. Brunoldi, G. Zanoni, G. Vidari, Biology and chemistry of neuroprostanes. First total synthesis of 17-A4-NeuroP: validation of a convergent strategy to a number of cyclopentenone neuroprostanes, Chem. Phys. Lipids 174 (2013) 64e74. G. Zanoni, E.M. Brunoldi, A. Porta, G. Vidari, Asymmetric synthesis of 14-A4tneuroprostane: hunting for a suitable biomarker for neurodegenerative diseases, J. Org. Chem. 72 (2007) 9698e9703. L. Balas, T. Durand, Dihydroxylated E, E,Z-docosatrienes. An overview of their synthesis and biological significance, Prog. Lipid Res. 61 (2016) 1e18. C.N. Serhan, J.Z. Haeggstrom, C.C. Leslie, Lipid mediator networks in cell signaling: update and impact of cytokines, FASEB J. 10 (1996) 1147e1158. ́

[110] R.A. Colas, M. Shinohara, J. Dalli, N. Chiang, C.N. Serhan, Identification and signature profiles for pro-resolving and inflammatory lipid mediators in human tissue, Am. J. Physiol. Cell Physiol. 307 (2014) C39eC54. [111] H. Arnardottir, S.K. Orr, J. Dalli, C.N. Serhan, Human milk proresolving mediators stimulate resolution of acute inflammation, Mucosal Immunol. 9 (2016) 757e766. [112] M. Giera, A. Ioan-Facsinay, R. Toes, F. Gao, J. Dalli, A.M. Deelder, C.N. Serhan, O.A. Mayboroda, Lipid and lipid mediator profiling of human synovial fluid in rheumatoid arthritis patients by means of LC-MS/MS, Biochim. Biophys. Acta 1821 (2012) 1415e1424. [113] H. Pruss, B. Rosche, A.B. Sullivan, B. Brommer, O. Wengert, K. Gronert, J.M. Schwab, Proresolution lipid mediators in multiple sclerosis - differential, disease severity-dependent synthesis - a clinical pilot trial, PLoS One 8 (2013) e55859. [114] S. Hong, K. Gronert, P.R. Devchand, R.L. Moussignac, C.N. Serhan, Novel docosatrienes and 17S-resolvins generated from docosahexaenoic acid in murine brain, human blood, and glial cells. Autacoids in anti-inflammation, J. Biol. Chem. 278 (2003) 14677e14687. [115] W.J. Lukiw, J.G. Cui, V.L. Marcheselli, M. Bodker, A. Botkjaer, K. Gotlinger, C.N. Serhan, N.G. Bazan, A role for docosahexaenoic acid-derived neuroprotectin D1 in neural cell survival and Alzheimer disease, J. Clin. Invest 115 (2005) 2774e2783. [116] O. Yanes, J. Clark, D.M. Wong, G.J. Patti, A. Sanchez-Ruiz, H.P. Benton, S.A. Trauger, C. Desponts, S. Ding, G. Siuzdak, Metabolic oxidation regulates embryonic stem cell differentiation, Nat. Chem. Biol. 6 (2010) 411e417. [117] B.D. Levy, P. Kohli, K. Gotlinger, O. Haworth, S. Hong, S. Kazani, E. Israel, K.J. Haley, C.N. Serhan, Protectin D1 is generated in asthma and dampens airway inflammation and hyperresponsiveness, J. Immunol. 178 (2007) 496e502. [118] E. Mas, K.D. Croft, P. Zahra, A. Barden, T.A. Mori, Resolvins D1, D2, and other mediators of self-limited resolution of inflammation in human blood following n-3 fatty acid supplementation, Clin. Chem. 58 (2012) 1476e1484. [119] N. Psychogios, D.D. Hau, J. Peng, A.C. Guo, R. Mandal, S. Bouatra, I. Sinelnikov, R. Krishnamurthy, R. Eisner, B. Gautam, N. Young, J. Xia, C. Knox, E. Dong, P. Huang, Z. Hollander, T.L. Pedersen, S.R. Smith, F. Bamforth, R. Greiner, B. McManus, J.W. Newman, T. Goodfriend, D.S. Wishart, The human serum metabolome, PLoS One 6 (2011) e16957. [120] J. Claria, J. Dalli, S. Yacoubian, F. Gao, C.N. Serhan, Resolvin D1 and resolvin D2 govern local inflammatory tone in obese fat, J. Immunol. 189 (2012) 2597e2605. [121] J. Claria, B.T. Nguyen, A.L. Madenci, C.K. Ozaki, C.N. Serhan, Diversity of lipid mediators in human adipose tissue depots, Am. J. Physiol. Cell Physiol. 304 (2013) C1141eC1149. [122] G.A. Weiss, H. Troxler, G. Klinke, D. Rogler, C. Braegger, M. Hersberger, High levels of anti-inflammatory and pro-resolving lipid mediators lipoxins and resolvins and declining docosahexaenoic acid levels in human milk during the first month of lactation, Lipids Health Dis. 12 (2013) 89. [123] X. Wang, M. Zhu, E. Hjorth, V. Cortes-Toro, H. Eyjolfsdottir, C. Graff, I. Nennesmo, J. Palmblad, M. Eriksdotter, K. Sambamurti, J.M. Fitzgerald, C.N. Serhan, A.C. Granholm, M. Schultzberg, Resolution of inflammation is altered in Alzheimer's disease, Alzheimers Dement. 11 (40-50) (2015) e41e42. [124] S. Medina, I.D. Miguel-Elizaga, C. Oger, J.M. Galano, T. Durand, M. MartinezVillanueva, M.L. Castillo, I. Villegas-Martinez, F. Ferreres, P. Martinez-Hernandez, A. Gil-Izquierdo, Dihomo-isoprostanes-nonenzymatic metabolites of AdA-are higher in epileptic patients compared to healthy individuals by a new ultrahigh pressure liquid chromatography-triple quadrupole-tandem mass spectrometry method, Free Radic. Biol. Med. 79 (2015) 154e163.