BBAMCB-57702; No. of pages: 16; 4C: 3, 4, 9 Biochimica et Biophysica Acta xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Biochimica et Biophysica Acta journal homepage: www.elsevier.com/locate/bbalip

1

Review

4Q1

Mojgan Masoodi a,⁎, Ondrej Kuda b, Martin Rossmeisl b, Pavel Flachs b, Jan Kopecky b,⁎⁎

5 6

a

7

a r t i c l e

8 9 10 11 12

Article history: Received 24 July 2014 Received in revised form 25 September 2014 Accepted 28 September 2014 Available online xxxx

13 14 15 16 17 18

Keywords: Healthy adipocyte Futile substrate cycle Macrophage Lipid mediator Omega-3 Calorie restriction

i n f o

R O

Nestlé Institute of Health Sciences SA, EPFL Innovation Park, bâtiment H, 1015 Lausanne, Switzerland Department of Adipose Tissue Biology, Institute of Physiology Academy of Sciences of the Czech Republic, v.v.i., Prague, Czech Republic

a b s t r a c t

T

E

D

P

Obesity-associated low-grade inflammation of white adipose tissue (WAT) contributes to development of insulin resistance and other disorders. Accumulation of immune cells, especially macrophages, and macrophage polarization from M2 to M1 state, affect intrinsic WAT signaling, namely anti-inflammatory and proinflammatory cytokines, fatty acids (FA), and lipid mediators derived from both n− 6 and n− 3 long-chain PUFA such as (i) arachidonic acid (AA)-derived eicosanoids and endocannabinoids, and (ii) specialized pro-resolving lipid mediators including resolvins derived from both eicosapentaenoic (EPA) and docosahexaenoic acid (DHA), lipoxins (AA metabolites), protectins and maresins (DHA metabolites). In this respect, potential differences in modulating adipocyte metabolism by various lipid mediators formed by inflammatory M1 macrophages typical of obese state, and non-inflammatory M2 macrophages typical of lean state remain to be established. Studies in mice suggest that (i) transient accumulation of M2 macrophages could be essential for the control of tissue FA levels during activation of lipolysis, (ii) a currently unidentified M2 macrophage-borne signaling molecule(s) could inhibit lipolysis and re-esterification of lipolyzed FA back to triacylglycerols (TAG/FA cycle), and (iii) the egress of M2 macrophages from rebuilt WAT and removal of the negative feedback regulation could allow for a full unmasking of metabolic activities of adipocytes. Thus, M2 macrophages could support remodeling of WAT to a tissue containing metabolically flexible adipocytes endowed with a high capacity of both TAG/FA cycling and oxidative phosphorylation. This situation could be exemplified by a combined intervention using mild calorie restriction and dietary supplementation with EPA/DHA, which enhances the 7formation of “healthy” adipocytes. © 2014 Published by Elsevier B.V.

E

C

b

O

F

3

Lipid signaling in adipose tissue: Connecting inflammation & metabolism☆

2

R

36 37 38 42 40 39

N C O

R

41

U

Abbreviations: 14-HDoHE, 14-hydroxydocosahexaenoic acid; 17-HDoHE, 17-hydroxydocosahexaenoic acid; 15d-PGJ2, 15-deoxy-Δ12, 14-prostaglandin J2; 2-AG, 2arachidonoylglycerol; AA, arachidonic acid; ACC, acetyl-CoA carboxylase; AEA, anandamide (arachidonoylethanolamide); AMPK, AMP-activated protein kinase; AQP7, aquaporin 7; ATGL, desnutrin/adipose triglyceride lipase; ATMs, adipose tissue macrophages; BAT, brown adipose tissue; β-AR, β-adrenergic receptor; CB1 or CB2, cannabinoid receptors type 1 or 2; CCL2, monocyte chemotactic protein-1 (also called MCP1); CLS, crown-like structures; COX, cyclooxygenases; CPT-1, carnitine palmitoyltransferase 1; DAG, diacylglycerol; DAGK, diacylglycerol kinase; DGAT1, acyl-CoA:diacylglycerol acyltransferase; DGL, diacylglycerol lipase; DHA, docosahexaenoic acid; DHEA, docosahexaenoylethanolamide; ECS, endocannabinoid system; EP, receptor for PGE2; EP3, prostaglandin E receptor 3; EPA, eicosapentaenoic acid; EPEA, eicosapentaenoylethanolamide; FA, fatty acids; FAAH, fatty acid amide hydrolase; FAS, fatty acid synthase; FAT/ CD36, FA translocase CD36; FP, receptor for PGF2α; GLUT4, glucose transporter type 4; GPR120, G protein-coupled receptor 120; HF, high-fat; HSL, hormone-sensitive lipase; HX, hepoxilin; ChREBP, carbohydrate response-element binding protein; IP, receptor for PGI2; IR, insulin resistance; IRS-1, insulin receptor substrate-1; LA, linoleic acid; LOX, lipoxygenase; LPL, lipoprotein lipase; LT, leukotriene; MGL, monoglyceride lipase; NAPE-PLD, N-acyl phosphatidylethanolamine-specific phospholipase D; NAT, N-acyl transferase; NE, neutrophil elastase; NFkB, proinflammatory transcription factor NFκB; NP, natriuretic peptides; NPRA, natriuretic peptide receptor A; OEA, oleoylethanolamide; omega-3 PL, oils containing omega-3 PUFA as phospholipids; omega-3PUFA,long-chainn-3 PUFA;omega-3 TAG,oilscontainingomega-3 PUFA astriacylglycerols; OXPHOS, oxidative phosphorylation;p38 MAPK, p38α mitogen-activatedprotein kinase; P450, cytochrome P450; PA, phosphatidic acid; PC, phosphatidylcholine; PD, protectin; PDH, pyruvate dehydrogenase; PDK4, pyruvate dehydrogenase 4; PE, phosphatidylethanolamine; PEPCK, phosphoenolpyruvate carboxykinase; PG, prostaglandin; PGC-1β, the PPARγ coactivator PGC-1β; PGE2, prostaglandin E2; PGF2, prostaglandin F2α; PIP, prostaglandin I2 receptor; PIP2, phosphatidylinositol biphosphate; PKA, protein kinase A; PKB, protein kinase B; PKG, protein kinase G; PLC, phospholipase C; PPAR, peroxisome proliferator-activated receptor; PRRs, pattern-recognition receptors; RBP4, retinol-binding protein 4; RvD1, resolvin D1; SREBP-1c, sterol regulatory element-binding protein 1c; SVF, stromal vascular fraction from adipose tissue; TAG, triacylglycerols; TAG/FA cycle, futile substrate cycle based on lipolysis of intracellular triacylglycerols and re-esterification of fatty acids; TRPV1, transient receptor potential vaniloid type-1; TX, thromboxane; WAT, white adipose tissue ☆ This article is part of a Special Issue entitled Oxygenated metabolism of PUFA: analysis and biological relevance. ⁎ Corresponding author. Tel.: +41 216326156. ⁎⁎ Correspondence to: J. Kopecky, Department of Adipose Tissue Biology, Institute of Physiology of the Academy of Sciences of the Czech Republic, v.v.i., Videnska 1083, 14220 Prague 4, Czech Republic. Tel.: +420 241062554; fax: +420 241062599. E-mail addresses: [email protected] (M. Masoodi), [email protected] (J. Kopecky).

http://dx.doi.org/10.1016/j.bbalip.2014.09.023 1388-1981/© 2014 Published by Elsevier B.V.

Please cite this article as: M. Masoodi, et al., Lipid signaling in adipose tissue: Connecting inflammation & metabolism, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbalip.2014.09.023

20 21 22 23 24 25 26 27 28 29 30 19 31 32 33 34 35

M. Masoodi et al. / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

63

86 87

Adipose tissue is distributed throughout the body in several discrete depots located mainly in the subcutaneous region (subcutaneous fat) or in the thorax and abdominal cavities (visceral fat). Areas composed mainly of white adipocytes are considered WAT, whereas areas mainly composed of brown adipocytes are considered BAT [1]. Different cell morphologies correspond to different functions, as white adipocytes are filled with one lipid droplet (unilocullar adipocytes) and equipped with a small cytosolic compartment, while brown adipocytes are multilocullar and rich in large mitochondria. In addition to the classical WAT and BAT adipocytes, a distinct type of thermogenic inducible adipocytes, so called brite/beige adipocytes also exist [2–4]. A plausible concept of adipose tissue organ as a continuum of both WAT and BAT has been coined by Cinti and colleagues [5,6]. WAT, as the most plastic organ among the metabolically relevant tissues, can represent 5–60% of total body weight [7,8]. Fat mass reflects energy balance, however, adipocyte number is very static in adult humans and independent of body weight fluctuations, even in response to massive weight loss. This indicates that the number of adipocytes is set during childhood and adolescence [8], while only approximately 10% of fat cells are renewed annually in adult humans [9]. Accumulation of visceral fat, which characterizes upper-body obesity, is known to correlate with metabolic syndrome [10]. This is typical for men, who, unlike women, accumulate fat centrally in both subcutaneous and visceral depots [11], but do not show high rates of subcutaneous fat accumulation unless morbidly obese [8].

88

2.1. Secretory functions of adipose tissue

89 90

WAT generates a number of signals, which include cytokines, hormones, growth factors, complement factors and matrix proteins that not only affect the neighboring cells but also target other peripheral tissues as well as the brain. Thus, WAT-derived signals could influence various processes including food intake, energy expenditure, metabolism, immunity and blood pressure. Two major adipokines (adipocytokines) are leptin and adiponectin, although there are probably hundreds of other adipokines and cytokines produced in WAT with some of them exerting only autocrine and paracrine, but not systemic effects (reviewed in [12,13]). The involvement of several adipokines and WAT-borne cytokines in control of (WAT) inflammation as well as insulin sensitivity are mentioned in the text below. In this context, it is especially retinol-binding protein 4 (RBP4), a retinol transporter secreted from adipocytes, which is of relevant importance [14], as its secretion is increased in obesity while it also activates innate

68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85

91 92 93 94 95 96 97 98 99 100 101 102 103

C

66 67

E

64 65

R

59 60

R

57 58

O

55 56

C

53 54

N

51 52

U

49 50

107

Intrinsic metabolism of WAT (Fig. 1) is of key importance for the whole organism with respect to storage of energy in triacylglycerols (TAG) during the postprandial state, as well as the mobilization of energy reserves during fasting or exercise and control of circulating FA levels (reviewed in [17]). Lipolysis of TAG from intracellular lipid droplets (see below) in adipocytes is under complex hormonal control and it is mediated by several co-operating enzymes (reviewed in [18, 19]; see Fig. 1 and Section 2.3). These include: (i) desnutrin/adipose triglyceride lipase (ATGL), an enzyme that catalyzes the initial step in TAG hydrolysis and the rate-limiting hydrolysis of TAG [20], (ii) hormone-sensitive lipase (HSL), which catalyzes the hydrolysis of TAG into diacylglycerols, and then into monoacylglycerols, and finally (iii) monoglyceride lipase (MGL). Besides its metabolic role, MGL also has an important function in the degradation of endocannabinoids ([19, 21]; see Section 4.3 and Fig. 2). Lipolysis is associated with in situ re-esterification of a part of lipolyzed FA back into TAG (TAG/FA cycle; [22–25]), in which acyl-CoA:diacylglycerol acyltransferase 1 (DGAT1) is specially involved, while DGAT2 is linked to endogenous FA synthesis and esterification [26]. Re-esterification requires glyceroneogenesis, i.e. a process in which pyruvate serves as the major precursor for glycerol 3-phosphate formation via phosphoenolpyruvate carboxykinase (PEPCK; [17,27]). The TAG/FA substrate cycling in WAT enables fine tuning of opposite metabolic fluxes, it is essential for buffering of plasma FA levels and marks “healthy” adipocytes (reviewed in [17]; see Section 5). Stimulation of this mechanism by peroxisome proliferatoractivated receptor-γ (PPARγ; product of the Pparg gene; also known as Nr1c3; see Section 2.3) agonists could contribute to the insulinsensitizing effects of these compounds (reviewed in [17]). Energy required for this futile substrate cycle [28] could be covered either by β-oxidation of FA in mitochondria, or by glycolysis/glucose oxidation, and also via energy equivalents produced during de novo FA synthesis [17]. It has been described only recently (reviewed in [19,29]) that the cycle of FA re-esterification and hydrolysis is also required for the formation of the lipolytic products, namely FA and diacylglycerols, acting as regulatory ligands of nuclear receptors (FA acting directly, and diacyl glycerols as their further metabolites); in fact FA must be always activated by the passage through TAG/FA cycle to act as lipid mediators. Thus, in analogy to the situation in cardiac muscle, where ATGL-derived molecule(s) is/are essential for the induction PPARα target-genes (see Section 2.3) and mitochondrial function [30], this mechanism could be also important for the induction of mitochondrial oxidative capacity in WAT. All the enzymes involved in lipolysis, including some of their regulatory proteins such as perilipin, as well as DGAT1, which is involved in FA re-esterification (see above) are located on lipid droplets (also referred to as adiposomes) in adipocytes. These organelles are delimited by a single phospholipid membrane that separates TAG and other lipids inside droplets from the aqueous phase and forms a matrix for all the lipid droplet-associated proteins. Lipid droplets interact with other cell organelles such as peroxisomes, mitochondria and endoplasmic reticulum (reviewed in [19,26,31]). The interaction with endoplasmic reticulum is essential for the formation of TAG, in which DGAT2 localized to both endoplasmic reticulum and lipid droplets is involved [31]. WAT is also an important site of de novo FA synthesis (de novo lipogenesis), namely in rodents, but also in humans, where up to 40% of whole-body de novo FA synthesis from glucose may take place in WAT [32]. Importantly, glucose transporter type 4 (GLUT4)-mediated

108 109

F

2. Adipose tissue

48

2.2. Systemic impact of WAT metabolism

O

62

46 47

R O

61

Obesity, i.e. excessive accumulation of white adipose tissue (WAT), is associated with severe metabolic diseases (metabolic syndrome) and represents one of the key problems of health care systems in affluent societies. Obesity is accompanied by chronic low-grade inflammation of WAT, which, in turn may affect adipocyte metabolism. Consequently, obesity-associated changes in WAT are consistent with an emerging concept that immune and metabolic systems are interconnected. This review focuses on: (i) WAT rather than thermogenic brown adipose tissue (BAT), since it is becoming increasingly apparent that besides energy expenditure in BAT, WAT metabolism may also have important systemic consequences on both the control of circulating fatty acids (FA) levels and insulin sensitivity; (ii) the roles of various lipid molecules in WAT, which are involved in autocrine (intracellular) and paracrine (intercellular) signaling, regulating metabolism of fat cells and reflecting inflammatory status of the tissue; and (iii) mutual interactions between various cell types contained in WAT, specially adipocytes and macrophages, as one of the keys to the integrated control of inflammatory status and metabolism in this tissue.

P

44 45

immunity and inflammatory responses in WAT (ref. [15]; see 104 Section 3.1). In contrast to WAT, the secretory functions of BAT are 105 poorly characterized [16]. 106

D

1. Introduction

T

43

E

2

Please cite this article as: M. Masoodi, et al., Lipid signaling in adipose tissue: Connecting inflammation & metabolism, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbalip.2014.09.023

110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167

M. Masoodi et al. / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

Catecholamines

Chylomicron -TAG VLDL -TAG

PGI2 EP3

InsR β-AR

NP

PIR NPRA

AEA FA

PGE2

Insulin 2-AG

LPL

3

CB1 TRPV1

Omega-3 PUFA

PKG

PKA

PKB

FAT/CD36

GPR120

Glucose HSL

GLUT4 Glucose GLUT1

ATGL

AMPK

Lipolysis

2-AG FA

MGL

AA AQP7 Glycerol

FA

(for hepac gluconeogenesis)

DGAT

F

FAS

Re-esterificaon

FA synthesis

COX/LOX

Lipid droplet

Glycerol-3P

O

ACC

CPT1 PEPCK

Acetyl-CoA PDK4

ATF2 PGC-1α/β α/β

PDH

PPARγγ

Glucose

PC

8-HETE HEPEs OEA

PPARα

ChREBP

Pyruvate SREBP1

Mitochondria

P

Oxaloacetate

R O

p38 MAPK

β-oxidaon

PGJ2 Δ12-PGJ2 15d-PGJ2

D

Nucleus

ChREBP ACC, FAS, GLUT4 FA synthesis

PPARα CPT1 β-oxidaon

PPARγ PEPCK, CD36, PDK4, PC Lipogenesis Adipogenesis

169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185

uptake of glucose into adipocytes, and FA synthesis associated with it, are linked to whole body insulin sensitivity. It has been hypothesized that some de novo formed lipid(s) in adipocytes could support insulin signaling in other tissues (reviewed in [33,34]). A novel (β) isoform of carbohydrate responsive-element binding protein (ChREBPβ, also referred to as MLX interacting protein-like) serves as the major determinant of FA synthesis in WAT and represents a predictor of insulin sensitivity (ref. [34]; see below). Similarly as in most other cells, mitochondria represent the main site of oxygen consumption and ATP production also in white adipocytes. The mitochondrial content of mature white adipocytes is several-fold higher than in preadipocytes and depends on the anatomical location of fat depot (reviewed in [17]). Mitochondria play a critical regulatory role in major metabolic pathways, i.e. de novo FA synthesis, glyceroneogenesis, lipolysis and FA re-esterification in adipocytes (ref. [17]; see above). Obesity and insulin resistance (IR) are associated with a low capacity of both mitochondrial oxidative phosphorylation [17,35,36] and de novo FA synthesis [37] in WAT.

U

168

N C O

R

R

E

C

T

E

Fig. 1. Pathways involved in regulation of adipocyte metabolism. As explained in the main text (see Sections 2.2 and 2.3.), the central node of adipocyte metabolism is lipolysis of triacylglycerols (TAG) and re-esterification of fatty acids (FA) in cytoplasmic lipid droplets (i.e. TAG/FA cycle). The associated metabolic pathways are de novo lipogenesis, glyceroneogenesis via phosphoenolpyruvate carboxykinase (PEPCK), and β-oxidation of FA in mitochondria. The induction of TAG/FA (FA liberated from TAG during lipolysis may be released from adipocyte and taken back by FA translocase CD36; FAT/CD36) requires ATP, leading to lowering of cellular energy status and activation of AMP-activated protein kinase (AMPK), which could stimulate uptake FA by increasing intravascular lipoprotein lipase (LPL) activity and by modulating intracellular TAG lipolysis and re-esterification of FA via a complex mechanism (reviewed in [17]). AMPK activity is decreased by protein kinase B (PKB) and at the early stages of adrenergic stimulation, also by protein kinase A (PKA). AMPK phosphorylates acetyl-CoA carboxylase (ACC), which leads to a decrease in de novo lipogenesis activity and to decrease in cellular malonyl-CoA levels. Malonyl Co-A is an inhibitor of carnitine palmitoyltransferase 1 (CPT-1) activity, the rate-limiting step in β-oxidation. Prolonged activation of AMPK could also decrease glucose oxidation by increasing peroxisome proliferator-activated receptor-γ (PPARγ)/ PPARγ coactivator/1 (PGC-1) transcriptional activity and subsequent induction of pyruvate dehydrogenase 4 (PDK4). Phosphorylation/inhibition of pyruvate dehydrogenase (PDH) by PDK4 enables pyruvate to be used for glyceroneogenesis [203]. Both, PKA and protein kinase G (PKG) induce lipolysis by direct phosphorylation of hormone-sensitive lipase (HSL) and through phosphorylation of perilipin 1 on lipid droplets and they also activate desnutrin/adipose triglyceride lipase (ATGL). PKA and PKG can phosphorylate/activate p38α mitogen-activated protein kinase (p38 MAPK; ref. [204]). Thus, in response to β-agonist or natriuretic peptides (NP), which bind to β-adrenergic receptor (β-AR) and natriuretic peptide receptor A (NPRA), respectively, p38 MAPK phosphorylates the transcriptional regulators ATF2 and PGC-1α, leading to the induction of expression of PPARα and PPARγ target genes (pink box and black box on the bottom). PKA- but not PKG-mediated lipolysis is inhibited by insulin [33]. Endogenous PPARs ligands are depicted in rectangular callouts on the right. Carbohydrate responsive-element binding protein (ChREBP; also called MLX interacting protein-like) is a transcription factor activated by glucose; its target genes are involved in the pathways of glucose and lipid metabolism. ChREBP is inactivated by both PKA and AMPK. In WAT, glucose-mediated activation of canonical ChREBPα isoform induces expression of ChREBPβ, as an important step in the induction of lipogenesis and increase in insulin sensitivity via adipocyte-derived signaling molecules [34]. 2-AG, 2-arachidonoylglycerol; AA, arachidonic acid; AEA, arachidonoylethanolamide; AQP7, aquaporin 7; COX, cyclooxygenase; CB1, cannabinoid receptor type 1; EP3, prostaglandin E receptor 3; FAS, fatty acid synthase; GPR120, G protein-coupled receptor 120; GLUT-*, glucose transporter; InsR, insulin receptor; LOX, lipoxygenase; MGL, monoacylglycerol lipase; PC, pyruvate carboxylase; PIR, prostaglandin I2 receptor; OEA, oleoylethanolamide; PG*, prostaglandin; SREBP1, sterol regulatory element-binding protein 1; TRPV1, transient receptor potential vaniloid type-1.

Even a small shift in the balance of activities in the major metabolic pathways contributing to TAG/FA cycle, i.e. lipolysis and FA re-esterification, as well as in the associated metabolic fluxes (see above) could affect adiposity. Notably, activity of the TAG/FA cycle also plays a crucial role in transformation of WAT to BAT [6]. The TAG/FA cycle in WAT could be responsible for about 2–3% of resting metabolic rate in obese humans [17]. Total WAT contribution to resting metabolic rate in lean human subjects is approximately 5% and it doubles in obesity [23].

186 187 188 189 190 191 192 193 194

2.3. Intracellular pathways engaged in the regulation of adipocyte 195 metabolism 196 The key molecular players that participate in regulation of adipocyte metabolism are shown in Fig. 1. Regarding the topic of this review, i.e. the interactions between the components of metabolism and the immune system, nuclear receptors from the family of PPARs play the key role. These receptors serve as ligand-dependent transcription

Please cite this article as: M. Masoodi, et al., Lipid signaling in adipose tissue: Connecting inflammation & metabolism, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbalip.2014.09.023

197 198 199 200 201

M. Masoodi et al. / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

R1-acyl: -20:4 n-6 -20:5 n-3 -22:6 n-3 9-HEPE -18:1 n-9 -16:0

8-HEPE RvE1

PE

PhL

NAT N-R1-acyl-PE

FAAH 5-HEPE

PTGS PGF2α PGFS

EPEA

OEA PGF2α-EA

Ethanolamides

12-LOX

PGI2

DHEA

FAAH

AEA

AA

FAAH

5-LOX Endocannabinoids

12-LOX

DHA

MGL

2-AG

PLA2

5-HETE

HSL

12-HETE

P450 lyso-PA ATGL

DAGK

HSL

PLA1

PLC

PDX RvD2

TAG

RvD1

n.e.

PIP2

EETs

PA DHETs

8-HETE

HXA3

12-HpETE

HXB3

15-HETE

15-HpETE LXA4

D

sn2-acyl: -20:4 n-6

15-LOX

P

17-HDoHE

R O

DGL

15-LOX

1,2-DAG PD1

Cys-LTs

12-LOX

PhL 14-HDoHE

LTB4

O

Mar1

6k-PGF1α

PTGIS

PEA 12-HEPE

15d-PGJ2

PLA2

N-R1-acyl-EA

EPA

5-LOX

n.e.

TXAS NAPE-PLD

P450 COX-2

Δ12-PGJ2

PGES

PLA2 18-HEPE

PGE2

TXA2

PGD2

PhL

n.e.

n.e.

R1,R2-PC

F

4

209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230

R

R

O

207 208

C

205 206

N

204

factors that control expression of metabolic genes, as well as mitochondrial biogenesis, and they also exert anti-inflammatory effects (reviewed in [38]). Thus PPARs bind as heterodimers with retinoid X receptor to specific regulatory sequences of the target genes, also in the interactions with several regulatory proteins and other transcription factors. Various endogenous lipids have been identified as their activating ligands, including FA, oxylipins, endocannabinoids and phospholipids (see Fig. 1; reviewed in [38]; see Section 4.1). PPARs family members show some tissue-specific differences in their expression, PPARα and PPARγ being abundant predominantly in the liver and adipose tissue, respectively, and PPARβ/δ in many tissues. These differences reflect the preferential involvement of both PPARα and PPARβ/δ in lipid catabolism, and the key regulatory role of PPARγ in adipocyte differentiation as well as in lipogenesis in these cells, respectively (reviewed in [39,40]). In WAT, the PPARγ2 isoform is predominantly expressed [41], located upstream from two major transcription factors controlling the activity of metabolic genes, i.e. sterol regulatory element-binding protein 1c (SREBP-1c) and ChREBP (reviewed in [33]; see Fig. 1). Transcriptional PPARγ coactivator PGC-1β plays a major role in the control of mitochondrial biogenesis in WAT [42]. The anti-inflammatory effects of PPARγ activation in WAT are mediated by multiple mechanisms, including physical interaction of PPARγ with the pro-inflammatory transcription factor NFκB (reviewed in [38]), as well as the prevention of CDK5-mediated phosphorylation of PPARγ at serine 273, which could elicit antidiabetic action [43,44]. The anti-inflammatory role of PPARγ is tightly linked to systemic insulin sensitivity, as documented by the effects of anti-diabetic drugs acting as specific PPARγ agonists, namely thiazolidinediones. Indeed, WAT-specific upregulation of PPARγ confers strong antidiabetic effects [45].

U

202 203

E

C

T

E

Fig. 2. Overview of lipid mediators in WAT. Production of lipid mediators begins on the plasma membrane, where phospholipids (violet) are cleaved by phospholipase A2 to yield free arachidonic, eicosapentaenoic or docosahexaenoic acids, which are subsequently metabolized into either eicosanoids (orange, yellow, green) or docosanoids (blue) by a variety of enzymes. Endocannabinoids (red) come either from membrane phospholipids via N-acyl-phosphatidylethanolamine intermediate or from 1,2-diacylglycerol, product of PIP2, PA or TAG metabolism, in case of 2-AG. 2-AG, 2-arachidonoyl glycerol; AA, arachidonic acid; AEA, arachidonoylethanolamide; ATGL, adipose triglyceride lipase; COX, cyclooxygenase; DAG, diacylglycerol; DAGK, diacylglycerol kinase; DGL, diacylglycerol lipase; DHEA, docosahexaenoylethanolamide; −EA, −ethanolamide; EPEA, eicosapentaenoylethanolamide; FAAH, fatty acid amide hydrolase; HSL, hormone-sensitive lipase; HX*, hepoxilin; LOX, lipoxygenase; LT*, leukotriene; MGL, monoacylglycerol lipase; n.e., non-enzymatic oxidation; NAPE-PLD, N-acyl phosphatidylethanolamine-specific phospholipase D; NAT, N-acyl transferase; OEA, oleoylethanolamide; P450, cytochrome P450; PA, phosphatidic acid; PC, phosphatidylcholine; PE, phosphatidylethanolamine; PEA, palmitoylethanolamide; PG*, prostaglandin; PIP2, phosphatidylinositol biphosphate; PLC, phospholipase C; TX*, thromboxane. For nomenclature of lipid mediators see LIPIDMAPS.org.

3. The importance of obesity-associated inflammation

231

Although obesity is one of the major risk factors for type 2 diabetes, not all obese subjects become diabetic. Obesity-associated low-grade inflammation is now recognized as an important cause of obesityinduced IR [46,47], also in accordance with the general notion that immune and metabolic systems are interconnected [38,48,49]. The inflammatory process, including activation of the innate immune system (macrophages, monocytes, neutrophils), is triggered by WAT expansion and hypoxia. However, molecular mechanism that underlies this process is not fully understood. Adipose tissue macrophages (ATMs) recognize pathogen-associated molecular patterns through expression of pattern-recognition receptors (PRRs) such as Toll-like and Nod-like receptors [50]. Saturated FA or their metabolites have been suggested to induce inflammation by activating Toll-like receptor 4 [51], resulting in downstream activation of several serine kinases such as IkB kinase and JNK [52,53]. In concert with the action of proinflammatory cytokines (see Section 3.1; refs [53,54]), these signaling events converge to enhance the activities of NFκB, and p38α mitogen-activated protein kinase (p38 MAPK) pathways, leading to the release of further inflammatory mediators and impairment of insulin signaling via serine phosphorylation of insulin receptor substrate-1 (IRS-1; reviewed in [55]). Activation of Toll-like receptor 4 by saturated FA, as well as by independently acting pro-inflammatory cytokines, can lead to upregulation of the ceramide biosynthesis pathway. The correlation between increased ceramide levels in plasma and IR has been demonstrated. Ceramides decrease insulin sensitivity and glucose transport via activation of protein kinase B (PKB; also known as Akt) dephosphorylation by

232

Please cite this article as: M. Masoodi, et al., Lipid signaling in adipose tissue: Connecting inflammation & metabolism, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbalip.2014.09.023

233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258

M. Masoodi et al. / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322

F

O

R O

278 279

P

276 277

3.1. Macrophages — major players mediating inflammation in adipose tissue WAT in lean subjects contains so called “alternatively activated” ATMs (i.e. M2 form ATMs) that produce anti-inflammatory cytokines such as IL-4, IL-10 and IL-13, which are marked by the expression of arginase-1 and several other genes. These ATMs promote WAT remodeling and clearance of apoptotic cells (efferocytosis), and support systemic insulin sensitivity (reviewed in [38,70,71]). In obesity, ATMs contribute significantly to increased production of inflammatory markers (see below) that play pivotal roles in the induction of systemic IR, glucose intolerance, and type 2 diabetes, while adiponectin secretion is decreased [72,73]. Accordingly, obesity is accompanied by macrophage polarization from the M2 form to the proinflammatory (classically activated) M1 form of ATMs [74,75]. M1 macrophages secrete proinflammatory cytokines such as TNF-α, IL-6, IL-1β and IL-12 and express lectin Lgals3 (also called as Mac-2), which mediates macrophage phagocytic and inflammatory responses. ATMs can comprise up to 40% of the cells in obese WAT, while their numbers correlates with the degree of obesity [76,77] and represent the strongest predictor of type 2 diabetes development in obese patients [78]. Macrophages may infiltrate WAT in response to chemokines, such as monocyte chemotactic protein-1 (CCL2; also called MCP1), which in contrast to its name is predominantly released from adipocytes as part of a scavenger function in response to hypoxia and adipocyte necrosis. Targeted deletion of the CCL2 gene reduced macrophage accumulation and inflammation in WAT, as well as IR associated with obesity [79]. In addition, over-expression of CCL2 gene in WAT induces macrophage recruitment and IR [79,80]. Increased content of ATMs in WAT during obesity-associated inflammation reflects both post-mitotic differentiation of bone-marrow-derived monocytes infiltrating the tissue [76], as well as local proliferation of macrophages [81,82]. The monocyte recruitment from blood stream is linked to WAT infiltration by M1 ATMs, whereas proliferation of M2 ATMs occurs on site ([81,82]; see below). Macrophages aggregate around dead adipocytes and form syncytia, often referred to as “crown-like structures” (CLS), while scavenging residual adipocyte lipids and ultimately forming multi-nucleate giant cells; thus adipocyte necrotic death could represent the primary stimulus driving ATM accumulation [83]. Although M2 and M1 forms are currently the two defined statuses of macrophages based on in vitro stimulation, macrophage polarization is

323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348

D

274 275

3.2. Involvement of various cells of immune system in macrophage 349 polarization 350

E

272 273

more dynamic in vivo and changes upon their local environment. RBP4 is the key endogenous protein in WAT (see Section 2.1), which is engaged in the control of ATM polarization. It activates both resident ATMs to induce their M1 form, as well as dendritic cells, leading to WAT inflammation and systemic IR [15]. The polarization switching is also influenced by dietary lipid composition. The dietary FA (see above) and FA-derived lipid mediators can influence cell signaling between adipocytes and macrophages in WAT and play an important role in the generation of the adipokine profile. For example in HF diet-induced obese mice, ATMs are polarized toward proinflammatory M1 form whereas treating obese mice with long-chain n− 3 PUFA (omega-3 PUFA) could drive the polarization toward M2 ATMs [71]. Omega-3 PUFA exhibit their anti-inflammatory effects via signaling through G protein-coupled receptor 120 (GPR120), resulting in improved insulin sensitivity in obese mice [71,84]. In contrast, its dysfunctions leads to obesity [85], documenting once again the tight link between the immune system and metabolism. In addition, docosahexaenoic acid (C22:6n−3, DHA)-derived lipid mediators such as resolvin D1 (RvD1) have been reported to decrease ATM accumulation, shifting macrophage polarization toward M2 form and improving insulin sensitivity in obese mice [86,87]. Moreover, production of (ATMs-derived) inflammatory mediators is highly dependent on the type of WAT depot (see [88] and Section 4.3). M2 ATMs exert high activity of mitochondrial oxidative phosphorylation, in agreement with their lasting role in tissue remodeling, while M1 ATMs rely on anaerobic glycolysis ([89]; see Section 5).

T

270 271

C

268 269

E

266 267

R

265

R

263 264

N C O

261 262

protein phosphatase 2a [56,57]. Of note in this regard, pleiotropic beneficial effects of adiponectin are initiated by stimulation of ceramidase activity and enhancement of ceramide catabolism [58]. Ceramides are also capable of activating the NLRP3 inflammasome, a transient complex of proteins responsible for activation of inflammatory processes [59–61]. Nalp3 (NLR family, pyrin domain containing protein 3, encoded by the NLRP3 gene), adaptor molecule ASC and caspase-1, components of the inflammasome, are preferentially expressed in ATMs. The NLRP3 inflammasome has been recently identified as an important contributor to obesity-induced inflammation and IR [55,62] and its activation is mediated by sensing mitochondrial dysfunction through the direct binding of Nalp3 to the mitochondrial lipid cardiolipin ([63]; see also Section 6). Recent studies have reported reduction in weight gain, fat mass and IR in Nlrp3−/−, Casp1−/− and Asc−/− mice with high-fat (HF) diet-induced obesity [60,64,65], demonstrating the important role of inflammasome in regulating insulin signaling. Although proinflammatory mediators play an important role in the initiation of the inflammatory response, chronic low-grade inflammation is also a result of inappropriate resolution capacity and uncontrolled inflammatory response. That resolution of inflammation is an active process regulated by the so-called “specialized pro-resolving lipid mediators” has been discovered by the Serhan's laboratory (reviewed in [66–69]; see Section 4.1).

U

259 260

5

ATM recruitment to WAT and their activation state are influenced by other immune cells in the tissue such as T cells and B cells (i.e. subpopulations of lymphocytes), eosinophils and mast cells. T cells play an important role in obesity-induced inflammation, at least in part, via influencing macrophage infiltration and polarization. Several subsets of T cells are present, like T helper cells, which can be classified into two groups (Th1 and Th2) based on their inflammatory properties (in analogy with ATMs); although both populations express CD4+, Th1 cells produce proinflammatory mediators while Th2 cells produce antiinflammatory mediators. Regulatory T cells secrete anti-inflammatory mediators and induce macrophage polarization toward M2. In the obese state, the number of proinflammatory T cells, CD4+ Th1 and CD8+ effector T cells increase, while anti-inflammatory Th2 cells are reduced in adipose tissue [90–92]. CD8+ T cells have essential roles in the initiation and maintenance of WAT inflammation, recruitment and differentiation of ATMs and systemic IR. The infiltration of WAT by these cells is an early event during the development of HF diet-induced obesity [91]. Regarding granulocytes in WAT, eosinophils decrease in their number during obesity-induced inflammation. Eosinophils are the major IL-4 expressing cells in WAT and contribute to polarization of ATMs toward M2 through an IL-4/IL-13-dependent process (refs. [93–95]; see Section 5). Neutrophils are involved in the initiation of the inflammatory response by contributing to ATMs recruitment and their number increases in obese state. Neutrophils secrete proteases such as neutrophil elastase (NE). Serum NE activity was elevated in dietary-obese mice [96], while the deletion of NE in these mice resulted in reduced inflammation and the content of neutrophils/macrophages in WAT [97]. Serum NE activity was also increased in human obese subjects and was correlated with leptin resistance [96]. The number of mast cells also increased in WAT of obese humans and mice. In addition, it was demonstrated that genetically-induced deficiency of mast cells resulted in reduced body weight gain and the levels of inflammatory cytokines while improving glucose homeostasis [98]. Mast cells could contribute to adipocyte differentiation by producing prostaglandins and directly promoting adipogenesis via production of 15-deoxy-

Please cite this article as: M. Masoodi, et al., Lipid signaling in adipose tissue: Connecting inflammation & metabolism, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbalip.2014.09.023

351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386

6 Table 1 White adipose tissue-borne lipid mediators.

t1:19 t1:20 t1:21 t1:22 t1:23 t1:24 t1:25 t1:26 t1:27 t1:28 t1:29 t1:30 t1:31

t1:36 t1:37 t1:38 t1:39 t1:40 t1:41 t1:42

t1:43 t1:44 t1:45 t1:46

Modulation of adipogenesis

[205]

H

Subcutaneous

Endo

Stimulation of leptin release

[105,101]

Modulation of adipogenesis Modulation of lipolysis

[108] [104]

M

Epididymal

Endo Endo, Exo Endo

Modulation of lipolysis

[100]

H, R M

Subcutaneous Intraabdominal

Endo Endo, Exo

Modulation of lipolysis Shift the differentiation of mesenchymal progenitors toward a brown adipocyte phenotype, induction of UCP1 in WAT

[103,206] [207,208]

M,R,H R

Perivascular

Exo Endo

Adipocyte differentiation Source of vasoactive prostaglandins

[209] [144]

M

Subcutaneous

Endo

Modulation of lipolysis

[210]

Epididymal Subcutaneous

Exo Exo Endo Endo Endo Exo

Increase of glucose transport Inhibition of adipogenesis Inhibition of adipogenesis Modulation of adipogenesis and MCP-1 expression Modulation of adipogenesis Production of Macrophage inhibitory cytokine-1

[109] [211,212] [213] [114] [110] [214]

M

Epididymal

Exo Exo Endo Exo Exo Endo

[215] [216] [111] [217] [116] [159]

M

Epididymal

H, M

Subcutaneous, Endo epididymal Epididymal Endo

Decrease of leptin production Stimulation of adipogenesis Stimulation of adipogenesis Stimulation of adipocyte differentiation Stimulation of adipogenesis Induction of lipid catabolism, anti-inflammatory Stimulation of MCP-1, IL-6, and TNF-α secretion Chemoattractant for macrophages in obesity Lipolysis may stimulate production of CysLTs Upregulation of IL-6, TNF-α, MCP-1

[123]

Induction of ER stress

[219]

Adipose tissue inflammation in obese humans

[119]

Exo Exo

Induction of inflammation and IR Modulation of adipocyte differentiation, decrease of TNF-α, MCP-1, and increase of adiponectin levels

[122,121] [220]

Exo

Stimulation of PPARγ, induction of adipogenesis

[221]

[87]

M H

M R

Perigonadal

M H

M

Exo

Exo

Epididymal

Endo, Exo Subcutaneous, Endo omental

Visceral

M

O

R

F

Endo

R

t1:35

O

t1:34

C

t1:33

LOX pathway

N

t1:32

Subcutaneous

R O

t1:18

H

P

t1:9 t1:10 t1:11 t1:12 t1:13 t1:14 t1:15 t1:16 t1:17

Reference

D

t1:7 t1:8

Adipocytes, SVF Adipose tissue, adipocytes 3T3-L1 PGE2, PGI2 Adipocytes, SVF Adipocytes, SVF, 3T3-L1 Adipocytes Adipose tissue, adipocytes, 3T3-L1 SVF, Ob1771 PGE2,PGI2, TXA2 Adipose tissue PGD2-related Adipose tissue PGF2a 3T3-L1 3T3-L1 Ob1771 d15-PGJ2 & PGJ2 3T3-L1 series Adipocytes Adipocytes, 3T3-L1 3T3-L1 NIH-3T3 3T3-L1 C3H10T1/2 d15-PGJ2, 8-HETE 3T3-L1 d15-PGJ2, PD1/ Adipose PDX tissue LTB4 Adipocytes, SVF LTB4, CysLTs Adipocytes, 3T3-L1 LTC4, CysLTs Adipocytes, SVF 12-HETE, 5-HETE, Adipocytes, LTB4 SVF, 3T3-L1 12/15-LOX related, Adipocytes, 3T3-L1 12-HETE, 12HpETE Adipose tissue, adipocytes, SVF 3T3-L1 EET, DHET Adipose tissue, adipocytes 3T3-F442A 5-HEPE, 8-HEPE, 9-HEPE, 12-HEPE and 18-HEPE RvD1, RvD2 Adipocytes, SVF, macrophage

Source Reported action of mediator

E

t1:5 t1:6

PGE2

Species Tissue

T

COX pathway

Model

C

t1:4

Lipid mediator

E

t1:3

R

t1:1 t1:2

M. Masoodi et al. / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

Adipose tissue, SVF

M

Epididymal

Exo

t1:49 t1:50

Adipose tissue, adipocytes Adipose tissue Adipose tissue Adipose tissue

H, M

Epididymal

Exo

H

Subcutaneous

Endo

M

Epididymal

Exo

M

Epididymal

Endo

Prevention of obesity-linked inflammation and IR

t1:54 t1:55

U

t1:47 t1:48

Downregulation of IL-6, MCP-1, and TNF-α, ROS; upregulation of IL-10, CD206, arginase 1, resistin-like molecule a, and chitinase-3 like protein, stimulation of nonphlogistic phagocytosis; shift M1 N M2 macrophage phenotype Increase of adiponectin production and insulin sensitivity; decrease of IL-6 production and macrophage infiltration; shift M1 N M2 macrophage phenotype Increase of adiponectin level; decrease of leptin, TNF-α, MCP-1, IL-6, and IL-1b levels Enhancement of resolution capacity in fat depot Increase of adiponectin levels and insulin sensitivity

t1:51 t1:52 t1:53

RvD1, RvD2, PD1, LXA4, RvE1 etc. PD1, RvD1, 17HDoHE, PD1, 17-HDoHE, 18-HEPE

M

Epididymal

Endo, Exo

Please cite this article as: M. Masoodi, et al., Lipid signaling in adipose tissue: Connecting inflammation & metabolism, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbalip.2014.09.023

[120] [218]

[122]

[86]

[124]

[126] [222] [223]

M. Masoodi et al. / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

7

Table 1 (continued) Model

Species Tissue

Reported action Source of mediator

Reference

Epididymal

Endo

Induction of lipid catabolism, anti-inflammatory

[159]

M

Perinodal

Exo

Resolution of inflammation

[224]

t1:58 t1:59 t1:60

omega 3 and omega 6 oxylipins LXA4

Endo

Modulation of inflammatory status in adipose tissue

[225]

Perigonadal

Exo

Decrease of IL-6 and increase of IL-10 expression; increase of insulin sensitivity, anti-inflammatory

[226]

t1:61

HXA3, HXB3

Adipose tissue Adipose tissue/ leukocytes Adipose tissue Adipose tissue, adipocytes, SVF 3T3-L1

M

t1:57

PD1/PDX, d15PGJ2 PD1, RvE1, LXA4

Stimulation of adipogenesis

[117]

Epididymal

Endo, Exo Exo Endo

Visceral

Endo

Decrease of inflammation

Epididymal

Endo

Decrease of inflammation

[138]

Endo, Exo Endo

Reduction of IL-6 by DHEA levels

[193]

Reduction of endocannabinoid tonus by Krill

[192]

Exo

Increased AMPK-α phosphorylation and carnitine palmitoyltransferase 1 transcription in adipose tissue, suggesting an increase in ATP-producing catabolic pathway; PEA also polarized adipose tissue macrophages to M2 lean phenotype Impairment of glucose tolerance and inhibition of insulin stimulated glucose uptake Stimulation of lipolysis Enhancement of β-adrenergic-mediated Thermogenesis; switch from white to brown adipocyte phenotype Inhibition of adipogenesis

t1:67 t1:68

Adipose tissue Adipose tissue

M

t1:69

AEA, 2-AG, PEA, OEA PEA

t1:70

OEA

Adipocytes

R

t1:71 t1:72 t1:73

PGF2a-EA

Inguinal, epididymal

R

Adipocytes R Adipose R tissue 3T3-L1, H, M preadipocytes

Epididymal

Endo

Epididymal Epididymal

Exo Exo

O

R O

AEA, 2-AG, EPEA, DHEA DHEA, EPEA

Increase of glucose transport activity Induction of IR

P

t1:66

3T3-L1 3T3-L1, M adipose tissue Adipose R tissue Adipose M tissue 3T3-L1

D

AA, LOX products Endocannabinoids AEA, 2-AG and related compounds AEA, 2-AG

M

E

t1:62 t1:63 t1:64 t1:65

M

Exo

T

t1:56

F

Lipid mediator

[118] [135] [191]

[136]

[227] [228] [229] [137]

Adipocytes, collagenase-liberated adipocytes from adipose tissue; SVF, stromal vascular fraction from adipose tissue (containing immune cells); H, human; M, mouse; R, rat; endo, endogenous source of the mediator (produced by the tissue itself); Endo, endogenous mediator; Exo, exogenous application of the mediator (artificial stimulation); IR, insulin resistance.

387 388

Δ12,14-prostaglandin J2 (15d-PGJ2; see Section 4.1) or indirectly through the regulation of adiponectin [99].

389

4. Lipid signaling in adipose tissue

390 391

397

Besides the cytokines, a large family of endogenous lipid mediators contributes to the immune and metabolic state of WAT. Thus, while traditional eicosanoids/prostanoids as well as endocannabinoids are involved in the control of metabolism and inflammation, omega-3 PUFA-derived mediators exert mostly anti-inflammatory effects. Moreover, other lipids active in cell signaling, such as sphingolipids (ceramides, sphingosine-1-phosphate), phospholipids (lysophosphatidic acid, platelet-activating factor) or steroids, can affect WAT metabolism.

398

4.1. Eicosanoids and specialized pro-resolving lipid mediators

399

Eicosanoids, produced via cyclooxygenases, lipoxygenases, and cytochrome P450 pathways, are potent local mediators of signal transduction and modulate the inflammatory response as well as metabolism of WAT. Although immune cells are the main producers of eicosanoids, adipocytes can also synthesize the main prostanoids and leukotrienes and express eicosanoid receptors. Thus, members of both adipocyte and immune cell lineages are able to communicate using lipid mediators as membrane or nuclear receptor ligands (see also Section 2.3 and Fig. 1). Selected lipid mediators produced in WAT are summarized in Table 1 and Fig. 2. Prostanoids are important for differentiation of adipocytes and regulation of lipolysis in an autocrine and paracrine manner.

395 396

400 401 402 403 404 405 406 407 408 409 410

E

R

R

N C O

394

U

392 393

C

t1:74 t1:75

They are produced by the action of cyclooxygenases (prostaglandin-endoperoxide synthases; COX) 1 and 2, which is a dominant pathway in generating eicosanoids in WAT. The best characterized prostanoid is prostaglandin E2 (PGE2). Endogenous PGE2 produced within WAT was shown to regulate lipolysis [100–103], where it works as an antilipolytic agent [104], although its effect is minor as compared to noradrenergic-mediated stimulation of lipolysis (see Fig. 1). It is still unknown which cell types contribute the most to the prostanoid pool in WAT, but in the human tissue, PGE2 is produced by adipocytes and other cell types [101,105,106]. Together with prostaglandin F2α (PGF2α), PGE2 also inhibits adipogenesis, which was documented in 3T3-L1 adipocytes [107,108]. PGF2α may enhance glucose transport in 3T3-L1 adipocytes by stimulating GLUT1 expression [109]. Cyclopentenone prostaglandins (prostaglandin J2 series, including PGJ2, Δ12-PGJ2 and 15d-PGJ2) act as ligands of PPARγ, the master regulator of adipogenesis. 15d-PGJ2 was shown to be important for differentiation and maturation of adipocytes [110–112], where it interferes in inducible synthesis of PGE2 and PGF2α [113], and in the production of adipokines linked to inflammation [114]. Adipose tissue expresses both PPARα and PPARγ isoforms (see Section 2.3 and Fig. 1), and although not yet demonstrated in WAT yet, various eicosanoids (e.g. 8,9-EET, 11,12-EET, LTB4, 9-HODE, 13-HODE etc.) have been found to activate PPARs in other tissues and cells (reviewed in [115]). Yu et al. tested activation of PPAR family members by eicosanoids; prostanoids, including prostaglandins PGA1 and 2, PGD1 and 2, and PGJ2-series [116], had an effect in 3T3-L1 cells. Of note, 8(S)-HETE and hepoxilins, lipoxygenase products and activators of PPARα and PPARγ, respectively, were able to induce

Please cite this article as: M. Masoodi, et al., Lipid signaling in adipose tissue: Connecting inflammation & metabolism, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbalip.2014.09.023

411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438

464 465 466 467 468 469 470 471 472 473

478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502

C

462 463

E

460 461

R

458 459

R

456 457

O

454 455

C

452 453

N

450 451

U

448 449

4.3. Lipid mediator-related enzymes and receptors in adipose tissue

511

WAT is able to express several enzymes metabolizing PUFA into active lipid mediators. Prostanoid biosynthesis is driven by COX and both of its isoforms, COX-1, the constitutive form, and COX-2, the inducible form, have been identified in this tissue [108,140]. Production of PGE2 is controlled by microsomal prostaglandin E2 synthase-1 and − 2, or by its cytosolic homologues [141]. Production of PGD2 and subsequent J-series is controlled by PGD synthase of which two types have been identified including lipocalintype PGD synthase and hematopoietic PGD synthase [142,143]. Perivascular WAT is able to produce vasoactive thromboxane A2 via thromboxane A2 synthase [144]. Recently, also prostaglandin F2α synthase was identified in 3T3-L1 cells [145]. The expression of membrane receptors for PGE2, PGF2α and PGI2 (i.e. EP, FP, and IP receptors, respectively) was explored in both preadipocytes and mouse mature adipocytes [146], and a crucial role for EP3 in modulation of lipolysis was identified ([146,147]; see Fig. 1). It has been demonstrated by Horillo et al. that WAT could express all the enzymes necessary for leukotriene biosynthesis: 5-lipoxygenase, 5-lipoxygenase activating protein, leukotriene A4 hydrolase, and leukotriene C4 synthase including leukotriene receptors: BLT1, BLT2, CysLT1, and CysLT2 [120]. In fact, the whole panel of lipoxygenases, differentially expressed during adipogenesis, was detected in WAT of mice and humans [148], and signs of depot specialization was identified in human 15-lipoxygenase, where selective expression was found in omental but not in subcutaneous fat [119]. The presence of enzymes for the production (N-acyl-phosphatidylethanolamine-specific phospholipase D, diacylglycerol lipase; see Fig. 2) and degradation (FA amide hydrolase) of endocannabinoids and its main receptor CB1 was demonstrated in adipocytes and WAT [130,149–151]. Also other enzymes of monoacylglycerol metabolism (e.g. MGL) are involved in the synthesis and degradation of endocannabinoids ([19,21]; see Fig. 1).

512 513

5. Mutual links between adipocytes and macrophages in WAT

544

Despite the fact that immune cells and adipocytes represent a functional unit, the mechanistic links between the two major cell types involved, ATMs and adipocytes, remain largely unknown. In obesity, visceral WAT compared to subcutaneous WAT accumulates more inflammatory ATMs, produces more proinflammatory mediators, and is more metabolically active, which includes both basal and stimulated lipolysis (reviewed in [152,153]). Relatively high rates of basal lipolysis in obesity could induce WAT inflammation and accumulation of proinflammatory M1 ATMs by activating PRRs [123,127]. In accordance with this, it has been demonstrated that endogenous oils derived from human WAT could induce neutrophil and M1-macrophage recruitment into the peritoneum [154]. In addition, sex-dependent differences also exist, with female rodents showing lower propensity to obesity-associated ATM accumulation as well as development of metabolic syndrome [153]. In contrast to the traditional view of the association between the obesity-driven ATM accumulation and inflammation in WAT, it has been demonstrated by Kosteli et al. [152] that weight loss and adrenergically-activated lipolysis promote a dynamic immune response in murine gonadal WAT, with an initial transient increase in ATM

545

F

Endocannabinoids, which are arachidonic acid (AA; C20:4n−6)-derivatives, represent another important class of lipid mediators involved in the regulation of WAT metabolism as well as whole-body energy homeostasis. The endocannabinoid system (ECS) was identified in the early 1990s during investigations into the mechanism of action of the major cannabisderived psychotropic compound, Δ9-tetrahydrocannabinol. The classical endocannabinoids, such as anandamide (arachidonoylethanolamide; AEA) and 2-arachidonoylglycerol (2-AG; see Fig. 2), exert cannabimimetic activities, partly through binding and activating cannabinoid receptors type 1 or 2 (CB1 or CB2), as well as through transient receptor potential vaniloid type-1 (TRPV1; ref. [21]). These endocannabinoids, their anabolic and catabolic enzymes, together with their receptors constitute the ECS (for review see ref. [129]). Besides regulating the basic processes such as food intake, the ECS participates in the control of lipid and glucose metabolism, and its dysregulation (mostly over-activity) in obesity might contribute to excessive fat accumulation and related metabolic disturbances [130,131]. The most convincing evidence linking a dysregulated ECS with obesity, type 2 diabetes, and hepatosteatosis came from studies carried out in animals (for instance, see [132–134]). Endocannabinoids and acyl-glycerols are important modulators of inflammation and insulin sensitivity of WAT (see Section 6). Of the less known mediators besides AEA and 2-AG, oleoylethanolamide inhibits insulin action and impairs glucose uptake in adipocytes [135], while palmitoleoylethanolamide increases β-oxidation of FA in response to phosphorylation/activation of AMP-activated protein kinase (AMPK) leading to increased carnitine palmitoyltransferase 1 (CPT-1) gene transcription in WAT (see Fig. 1), together with polarization

446 447

O

477

445

R O

4.2. Endocannabinoids

443 444

503 504

P

476

441 442

of ATMs to M2 type [136]. Prostaglandin F2 ethanolamide, similarly to its precursor AEA, inhibits adipogenesis in mouse 3T3-L1 or human preadipocytes [137]. Furthermore, the anti-inflammatory properties of omega-3 PUFA-derived endocannabinoids, namely eicosapentaenoylethanolamide and docosahexaenoylethanolamide, were documented in mouse WAT [138]. Part of the anti-inflammatory effects of endocannabinoids results from their interactions with PPARs [139].

T

474 475

differentiation of 3T3-L1 preadipocytes [116,117]. Another evidence links lipoxygenase metabolites and activation of PPARγ to glucose uptake in 3T3-L1 adipocytes [118]. Products of lipoxygenase activity, especially of the 12/15lipoxygenase and 5-lipoxygenase pathways, are important for regulation of both adipogenesis [116–118] and WAT inflammation [119–121]. In obese Zucker rats, 12-HETE, 5-HETE and LTB4 contribute to inflammatory state of visceral adipose tissue [122] and 12/15 lipoxygenase products 12(S)-HETE and 12(S)HpETE impair insulin signaling in 3T3-L1 cells [121]. Production of LTC4 by macrophages previously stimulated by FA from adipocyte lipolysis has been linked to IR of adipose tissue [123]. In contrast, several pro-resolving lipid mediators (reviewed in [66–69]), including those derived from eicosapentaenoic acid (C20:5n− 3, EPA; E-series resolvins, RvE1, RvE2, and RvE3) and from DHA (D-series resolvins, and protectins; PD; see also ref. [68]), as well as lipoxins (LXA4, 15-epi-LXA4), are involved in resolution of WAT inflammation. Indeed, the administration of RvD1 [124] and 17-hydroxydocosahexaenoic acid (17-HDoHE) [125], to obese diabetic mice improved glucose tolerance and insulin sensitivity, while reducing ATM accumulation and expression of inflammatory cytokines in WAT. Recently, a wide range of lipid mediators including RvD1, RvD2, PD1, lipoxin A4, and 17-HDoHE, 18-HEPE and 14-hydroxydocosahexaenoic acid (14-HDoHE) were identified in human subcutaneous WAT, while the levels of PD1 and 17-HDoHE decreased in patients with peripheral vascular disease [126]. In addition, an impairment in pro-resolving mediators production such as RvD1, PD1 and 17-HDoHE in inflamed WAT from obese mice has been reported [124]. This deficiency can be the result of impaired bioavailability of their respective precursors, i.e. omega-3 PUFA [125,127], or acceleration in their inactivation and clearance. This is further supported by findings that a key enzyme in metabolic conversion of pro-resolvins, 15-PG-dehydrogenase/eicosanoid oxidoreductase, is up-regulated in WAT under obese state [124]. Of note, some results previously assigned to PD1 might be related to its isomer protectin DX (PDX) [68]. PDX, also known to be produced in WAT, alleviated IR in diabetic db/db mice, but did not resolve WAT inflammation [128].

D

439 440

M. Masoodi et al. / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

E

8

Please cite this article as: M. Masoodi, et al., Lipid signaling in adipose tissue: Connecting inflammation & metabolism, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbalip.2014.09.023

505 506 507 508 509 510

514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543

546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564

M. Masoodi et al. / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601

F

O

R O

584 585

P

582 583

D

580 581

Lipolysis and TAG/FA cycle acvaon in adipocytes

β -AR

FA

Furthermore, it has been demonstrated by Kosteli et al. that ATM depletion results in increased lipolysis in WAT, suggesting that M2 ATMs could release antilipolytic factor(s) of unknown chemical nature [152]. We extend here (see Fig. 3) the above model of Kosteli et al. by stressing the importance of the TAG/FA cycle in adipocytes. Its activity is tightly linked with lipolysis, thus affecting the storage of TAG and adiposity, as well as the levels of circulating free FA and systemic lipid metabolism (see Section 2.2). That M2 ATMs cause TAG/FA cycle activity to change in parallel to the lipolysis cycle has been documented by the simultaneous upregulation of ATGL (a marker of stimulated lipolysis) and DGAT1 (linked to FA re-esterification [26]) when ATMs are depleted [152]. We have observed in mice fed HF diet [159] that activity of mitochondrial oxidative phosphorylation in adipocytes was induced in an additive manner in response to a combined treatment using mild calorie restriction and dietary omega-3 PUFA (see Section 6). At the same time, ATM content and systemic inflammation decreased, while the levels of WAT anti-inflammatory lipid mediators increased. These changes that occurred in gonadal but not in subcutaneous WAT, could contribute to reduced accumulation of body fat under these treatment conditions. The induction of oxidative phosphorylation in adipocytes was associated with activation of the TAG/FA cycle, depending on the complex interplay of the intracellular regulatory pathways including PPARα, PPARγ, AMPK, ECS, and G-protein coupled receptors (see Sections 2.3 and 4.3). Our results suggest that adipocyte endowed with a high capacity of both oxidative phosphorylation and TAG/FA cycle represent a “healthy” and metabolically flexible cell type, which is associated with lean phenotype (for details, see ref. [17]). Based on the above data we hypothesize (see Fig. 3) that induction of pro-resolving lipid mediators in WAT, both in adipocytes and ATMs, for instance in response to omega-3 PUFA supplementation, could augment the “healthy” metabolic phenotype in adipocytes during the activation of lipolysis. This would reflect an egress of M2 ATMs that results in their reduced tissue content and the removal of the inhibitory effect of M2 ATMs on the adipocyte lipolysis and TAG/FA activity (see Fig. 3). This model is also supported by the fact that M2 ATMs are

E

578 579

T

576 577

C

574 575

E

572 573

R

571

R

569 570

N C O

567 568

recruitment. These ATMs express markers that are typical of the anti-inflammatory M2 state. Conversely, manipulations reducing lipolysis decreased ATM accumulation. These observations in animals are consistent with findings in humans demonstrating that a short-term low-calorie diet induces accumulation of ATMs of M2 phenotype [155]. It has been hypothesized that the short-term activation of lipolysis in WAT could lead to recruitment of M2 ATMs depending on lipolysis-released FA from adipocytes [152]. It remains unclear why long-lasting accumulation of FA released from hypertrophic adipocytes could induce accumulation of M1 ATMs, while acute stimulation of lipolysis in WAT could result in the accumulation of M2 ATMs. It can be speculated that the differential effect of FA accumulated in WAT in response to the increased lipolysis (basal and stimulated), could reflect either different FA species accumulated under these conditions [156,157] or a formation of PPARα agonists during the stimulated lipolysis [30], which could also affect the interplay between the immune and metabolic systems ([38,48,49]; see Section 3). Very recently, the role of lipolysis of TAG in lipid droplets mediated by lysosomal acid lipase (but not by either ATGL or HSL; see Section 2.2) in the induction of M2 ATMs has been demonstrated [89]. The emerging role of adipocyte-borne extracellular vesicles in the paracrine signaling between adipocytes and ATMs and the polarization of ATMs should be further explored [158]. M2 ATMs, which accumulate transiently in WAT during stimulation of lipolysis could buffer locally released FA by incorporating them into intracellular TAG [152]. Indeed, M2-macrophages exhibit high activity of mitochondrial oxidative phosphorylation [38,49], i.e. they can efficiently form ATP, which is required for lipogenesis, while enhanced β-oxidation drives the ATP formation and helps to further reduce FA levels. As suggested by O'Neil and Hardie [49], the switch between the M2-linked oxidative metabolism and the glycolytic metabolism typical of M1 macrophages [89] could be regulated by AMPK; the more energy-efficient oxidative metabolism of M2 ATMs could be essential for long-term tissue repair, while the glycolytic M1-linked metabolism could support rapid metabolic responses during inflammation. Thus both AMPK activation and short-term starvation could reflect an ancient starvation pathway depressing energy-costly inflammatory response.

↑ OXPHOS ↑ FA oxidaon

(B)

β-AR

(A)

WAT remodeling ↑ Lipolysis ↑ TAG/FA

Lipid mediator(s) ?

↑ OXPHOS ↑ FA oxidaon

U

565 566

9

β-AR β-AR

FA

FA Resoluon of inflammaon

Macrophage lipid uptake ↑ TAG synthesis ATMs recruitment

↑ OXPHOS ↑ FA oxidaon

(C)

An-inflammatory and pro-resolving lipid mediators (15d-PGJ2, PD1/PDX) Fig. 3. Involvement of ATMs in response of WAT to the combination treatment using calorie restriction and omega-3 PUFA. Prolonged mild calorie restriction, i.e. a modest stimulation of βadrenergically mediated lipolysis (red triangle) and TAG/FA activity (red circular arrows) in adipocytes, results in increased extracellular free fatty acids levels (A) and recruitment of ATMs of M2 phenotype. ATMs can buffer local increases in free fatty acids concentrations through their uptake and incorporation into intracellular triacylglycerols, while depressing lipolysis and TAG/FA activity in adipocytes via an unknown secreted factor (B). Pro-resolving lipid mediators formed from omega-3 PUFA (C) accelerate a decrease in ATM content, thus eliminating the basis for the existence of above mentioned negative regulatory loop (B) and allowing for a full manifestation of both the lipolytic and TAG/FA activity in adipocytes. The capacity of these biochemical pathways as well as mitochondrial oxidative phosphorylation is enhanced in response to stimulation of PPARα-, PPARγ-, and AMPK-signaling and suppression of the activity of the ECS in adipocytes in response to the combination treatment. Thus, resolution of inflammation supports flexible healthy metabolic phenotype of adipocytes (for details, see ref. [17]). This scheme represents a modified version of Fig. 10 of the original article by Kosteli et al. [152]. β-AR, β-adrenergic receptor; OXPHOS, oxidative phosphorylation.

Please cite this article as: M. Masoodi, et al., Lipid signaling in adipose tissue: Connecting inflammation & metabolism, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbalip.2014.09.023

602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638

669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702

C

667 668

E

665 666

R

663 664

R

661 662

O

654 655

C

652 653

N

650 651

U

648 649

F

IR seems to be a pathogenic link between obesity and associated metabolic dysfunctions. In this respect, hypertrophic WAT expansion in obesity associated with the influx of immune cells [160] triggers a cascade of inflammatory responses that negatively affect insulin signaling and glucose uptake in insulin sensitive tissues, thus contributing to systemic IR (for review see [55]). The quantity and type of dietary fat have lasting and powerful effects on obesity-associated IR and inflammation. For instance, it has been shown in healthy as well as obese subjects that various inflammatory markers were lower in individuals consuming the so called Mediterranean diet, whose principal source of fat is olive oil (rich in monounsaturated FA) and which is also rich in fish (reviewed in [161]). These data are also supported by the multicenter KANWU study of healthy subjects receiving various isoenergetic diets for 3 months [162], showing that saturated FA-rich diet impaired while monounsaturated FA-rich diet improved insulin sensitivity when total fat intake did not exceed 37% of energy. Besides proinflammatory cytokines such as IL-1β, IL-6 and TNF-α, also saturated FA are known to activate inflammatory signaling pathways, in this case through Toll-like receptor 4, thus initiating a cascade of events leading to IR (reviewed in [55]; see Section 3). Conversely, proinflammatory cytokines such as IL-1β also upregulate serine palmitoyltransferase [163], i.e. an enzyme involved in the biosynthesis of ceramide that is capable of antagonizing insulin signaling in the muscle [164]. At the same time, ceramides can activate NLRP3-caspase 1 inflammasome complex, which is involved in the cleavage of pro-IL-1β precursor and generation of active IL-1β ([55]; see Section 3). These results suggest that targeting nutrient-sensitive inflammatory pathways by using dietary interventions aimed at specific FA/lipid composition may antagonize inflammation and consequently IR. There has been a lot of progress in our understanding how dietary omega-3 PUFA, namely DHA and EPA might reduce the incidence of cardiovascular disease (for review, see [165]), augment the efficacy of lipid-lowering drugs [166], or even help to reduce adiposity in humans [167–169], reduce hepatic steatosis [170], and contribute to prevention of IR (reviewed in [171]). Furthermore, large amount of experimental evidence comes from rodent studies, showing that dietary omega-3 PUFA can prevent HF diet-induced obesity, IR, dyslipidemia and WAT inflammation [159,172–176], while they can also revert some of the obesity-associated pathologies [175,177]. However, the omega-3 PUFA's effects on glycemic control and insulin sensitivity in diabetic human subjects are usually weak or even negative [178], although this group of patients might still benefit from omega-3 PUFA supplementation when its protective effect against cardiovascular disease is considered [171,179]. Moreover, the combinations of omega-3 PUFA

646 647

O

659 660

645

R O

6. Role of dietary lipid intake and lipid signaling

643 644

P

658

641 642

with various anti-diabetic drugs or mild calorie restriction (see Section 5) might hold great promise especially in the treatment of diabetic patients [159,175]. An important part of omega-3 PUFA's effects on obesity-associated metabolic disorders and cardiovascular disease resides in their strong anti-inflammatory action in various tissues including WAT as well as in modulation of eicosanoid production (see the reviews [180–182]; see Section 4 and 5). In general, the beneficial effects of omega-3 PUFA, especially under obesogenic conditions, are associated with a profound modulation of WAT metabolism and function (reviewed in [180]). Thus, dietary intake of omega-3 PUFA stimulates the expression and secretion of insulin-sensitizing hormone adiponectin in WAT of HF diet-fed mice [183,184] as well as in humans [185] while inducing mitochondrial biogenesis and FA oxidation in adipocytes [186], i.e. the effects specific to abdominal WAT depot. Adiponectin exerts its beneficial metabolic effects (e.g. increased FA oxidation and glucose uptake) in part by stimulating AMPK activity in target tissues such as liver and skeletal muscle [187], which is likely the mechanism that contributes to the insulin-sensitizing effects of omega-3 PUFA, at least in the liver [176]. Moreover, adiponectin improves insulin sensitivity through catabolism of ceramides (see Section 3). A new direction in the study of alternative approaches for the treatment of obesity-related inflammation and metabolic dysfunction has been recently investigated, with a focus on the modulation of ECS by dietary omega-3 PUFA. It is known that stimulation of the ECS activity in WAT, it promotes adipogenesis, lipid accumulation, and reduces mitochondrial biogenesis, while its blockade, e.g. by CB1 antagonist rimonabant, results in elevated lipolysis, mitochondrial biogenesis, increased adiponectin production, and reduced inflammation, in part also through the effect on WAT ATMs and possibly other cell types present in WAT (reviewed in [188]). Although the downregulation of the elevated ECS tone in obesity by rimonabant produced desired metabolic effects, its use in the clinics has been halted due to psychiatric side effects [189]. On the other hand, selectively blocking the peripheral CB receptors through nutritional interventions based on the modification of dietary FA composition (e.g. omega-6/omega-3 PUFA ratio) and subsequent changes in the concentrations of various lipid-derived ligands including endocannabinoids might represent an optimal solution (reviewed in [129] and more recently by [190]). In this respect, Alvheim et al. [133] showed in mice that by increasing dietary content of linoleic acid (LA; C18:2n− 6), i.e. the precursor for the synthesis of AA, from 1% to 8% of energy, (i) tissue levels of AA as well as of 2-AG and AEA were elevated, which resulted in the development of obesity, and (ii) adipogenic effect of LA could be prevented by consuming sufficient amounts of EPA and DHA to reduce the AA-phospholipid pool and normalize the levels of endocannabinoids. In accordance with these data, dietary omega-3 PUFA have been shown to modulate the levels of endocannabinoids in various organs including WAT of rodents with genetic or dietary obesity [134,138, 191,192]. Interestingly, Rossmeisl et al. [138] showed in obesity-prone C57BL/6 mice fed a corn oil-based high-fat diet (rich in LA), that besides decreasing 2-AG and AEA in abdominal WAT, dietary omega-3 PUFA also elevated tissue levels of N-acyl ethanolamines DHEA and EPEA, the endocannbinoid-like molecules derived from DHA and EPA, respectively, which are known to be produced by adipocytes and possess antiinflammatory properties [193]. Furthermore, omega-3 PUFA-induced changes in plasma profile of various endocannabinoids were similar to those observed in abdominal WAT, raising a question to what extent WAT and especially abdominal WAT determines systemic levels of various endocannabinoids. Further evidence suggests that the efficiency of dietary omega-3 PUFA to modulate tissue levels of endocannabinoids partly depends on the lipid form of omega-3 PUFA supplementation. Thus, dietary omega-3 PUFA administered as marine phospholipids (omega-3 PL) either in the form of krill oil [191,192,194] or isolated from fish meat [138] were able to efficiently modify tissue endocannabinoid levels in

T

656 657

involved in active remodeling of WAT [69], while perhaps helping the formation of the “healthy” adipocytes (see above). It is based in large on our observations in mice fed the HF-diet, in which the positive effects of the combined treatment using omega-3 PUFA and calorie restriction on WAT metabolism was only detected in the gonadal but not subcutaneous WAT. This is consistent with a relatively high accumulation of ATMs in the former depot under obesity-promoting conditions (see ref. [17]). The mutual functional links between the immune cells and adipocytes in WAT, mediated possibly by both cytokines and lipid mediators, require further studies. Parenthetically, very recent studies in mice demonstrated the importance of M2 ATM recruitment for the induction of thermogenic brite/beige cells in WAT, depending possibly on the production of catecholamines in M2 ATMs [94,95]. This recruitment is triggered by cytokines released from tissue eosinophils (refs. [93–95]; see Section 3.2). However, this process could be only transient and rather specific for the subcutaneous WAT containing brite/beige precursor cells, since these cells are virtually absent in abdominal WAT of rodents [2–4].

D

639 640

M. Masoodi et al. / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

E

10

Please cite this article as: M. Masoodi, et al., Lipid signaling in adipose tissue: Connecting inflammation & metabolism, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbalip.2014.09.023

703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 Q2 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768

M. Masoodi et al. / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

790 791 792 793 794 795 796 797 798 799 800 801

806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832

C

788 789

E

786 787

R

Acknowledgements

863

Supported by the EU FP7 project DIABAT (HEALTH-F2-2011-278373), by the Czech Science Foundation (13-00871S), by the Ministry of Education, Youth and Sports (LH14040) and Nestlé Institute of Health Sciences.

864

References

868

F

784 785

R

782 783

N C O

780 781

U

778 779

O

Systemic effects of WAT not only reflect the role of this tissue in serving as an energy buffer, but also point to its involvement in the control of circulating free FA levels and insulin sensitivity via secretion of various adipokines, cytokines, and lipid mediators. All the major functions of WAT are closely related and regulated by hormones and by the activity of nervous system. Both metabolic and secretory functions of WAT largely depend on adipocytes themselves, since these cells have the remarkable ability to change their volume as well as their numbers depending on systemic requirements. This kind of tissue plasticity involves not only adipocytes, but also other WAT cell types, such as immune cells and the cells comprising microcirculation inside the tissue. It is becoming increasingly apparent that various lipid mediators formed both in adipocytes and immune cells, namely in the ATMs, are involved not only in endocrine but also in paracrine and autocrine signaling within WAT. Besides classical eicosanoids, prostanoids and endocannabinoids, also specialized pro-resolving lipid mediators are formed in WAT, as also evidenced by the presence of their synthesizing enzymes in the tissue, and participate in various physiological as well as pathophysiological processes, including tissue responses to activated lipolysis and metabolic inflammation in obesity, respectively. At the same time, lipid mediators within WAT together with adipokines and cytokines are likely involved in the control of tissue remodeling, which is the key feature underlying tissue plasticity. Under physiological conditions such as during activation of lipolysis in WAT of lean individuals, M2 ATMs help to control extracellular levels of FA by a negative feedback regulation of lipolysis and TAG/FA activity. The nature of this anti-lipolytic factor(s) remains unclear; however, it could be also represented by some lipid mediator(s). Moreover, M2

776 777

R O

804 805

775

833 834

P

7. Conclusions

773 774

ATMs could contribute to remodeling of WAT under the conditions that lead to “healthy” adipocyte formation, i.e. lean, metabolically flexible adipocytes that are endowed with high TAG/FA cycling and mitochondrial oxidative phosphorylation activity. This situation could be exemplified by a combined intervention using mild calorie restriction and dietary supplementation with EPA/DHA, and possibly potentiated by lower intake of LA, which enhances the formation of “healthy” adipocytes. The mechanisms involved include increased formation of pro-resolving lipid mediators partly derived from omega-3 PUFA, the activation of lipid catabolism by a PPARαmediated mechanism, the decreased activity of ECS, and the activation of AMPK and adrenergically-stimulated lipolysis in WAT, all of which are involved in the formation of metabolically flexible adipocytes. On the other hand, when lipid supply exceeds the storage capacity of fat cells, e.g. in WAT atrophy or obesity, M1 ATMs infiltrate the tissue to remove dead adipocytes while secreting pro-inflammatory molecules; in this respect, pro-resolving lipid mediators are essential for the resolution of inflammation by stimulating the egress of ATMs and the return of WAT to a physiological state. In spite of the recent substantial advances in the characterization of various lipid mediators in WAT, especially under conditions of metabolic inflammation, our understanding of their involvement in the mutual interactions between adipocytes and ATMs is insufficient. We are limited by a scarcity of appropriate in vitro models, as well as the short life of many of these signaling molecules. Nevertheless, the ability to modulate the formation of pro-resolving lipid mediators, e.g. in response to dietary supplementation with omega-3 PUFA (which would also lead to the suppression of ECS activity), or even in response to a topical administration of specific lipid mediators, opens new avenues for the treatment of many diseases linked to the state of chronic inflammation.

D

803

771 772

T

802

genetically obese fa/fa rats [191] or in dietary obese mice [138,192,194]. Furthermore, when directly compared with fish oils containing omega-3 as triacylglycerols (omega-3 TAG), omega-3 PL were more effective in reducing the levels of 2-AG and AEA in visceral WAT [138, 191] and the levels of AEA in the liver and heart [191]. Dietary omega-3 PL also ameliorated dyslipidemia and markedly reduced hepatic steatosis in obese rodents [138,194,195]. Thus, it seems that supplementation of omega-3 PL can exert stronger metabolic effects when compared with omega-3 TAG, possibly due to: (i) improved bioavailability of DHA and especially EPA, as demonstrated in rodents [138] as well as in humans [196–198]; (ii) stronger modulation of the ECS (see above); (iii) regulation of cellular metabolism by yet unidentified omega-3 PL species possibly functioning as ligands to specific nuclear receptors [199,200], which are also involved in the transcriptional regulation of lipid metabolism. Although the superior effects of omega-3 PL have been linked primarily to improved DHA and/or EPA bioavailability (see above and an excellent review published recently by Schuchardt and Hahn [201]), it is still unclear whether there are also other mechanisms involved. Similarly to omega-3 TAG, a relatively high potency has been demonstrated in the case of omega-3 PUFA contained in the extract from the zooplankton Calanus finmarchicus (Calanus oil), especially when the effect on diet-induced obesity and obesity-related disorders in mice is considered [202]. Thus, anti-inflammatory nutritional interventions with omega-3 PUFA, especially those in the form of omega-3 PL, seem to represent an efficient tool also for manipulating the activity of the ECS in various tissues including WAT. The displacement of AA by DHA and EPA in the relevant membrane phospholipid fractions in response to dietary omega-3 PUFA supplementation has been shown to be associated with decreased production of AA-derived EC molecules. However, the precise role of omega-3 PUFA-induced changes in the ECS tone in various tissues, especially in WAT, and their contribution to improved metabolic function under obesity-related conditions remains to be fully elucidated.

E

769 770

11

[1] A. Smorlesi, A. Frontini, A. Giordano, S. Cinti, The adipose organ: white-brown adipocyte plasticity and metabolic inflammation, Obes. Rev. 13 (Suppl. 2) (2012) 83–96. [2] B. Cousin, S. Cinti, M. Morroni, S. Raimbault, D. Ricquier, L. Penicaud, L. Casteilla, Occurrence of brown adipocytes in rat white adipose tissue: molecular and morphological characterization, J. Cell Sci. 103 (1992) 931–942. [3] C. Guerra, R.A. Koza, H. Yamashita, K.W. King, L.P. Kozak, Emergence of brown adipocytes in white fat in mice is under genetic control. Effects on body weight and adiposity, J. Clin. Investig. 102 (1998) 412–420. [4] T.B. Walden, I.R. Hansen, J.A. Timmons, B. Cannon, J. Nedergaard, Recruited vs. nonrecruited molecular signatures of brown, “brite,” and white adipose tissues, American Journal of Physiology-Endocrinology and, Metabolism 302 (2012) E19–E31. [5] S. Cinti, The adipose organ, Prostaglandins Leukot. Essent. Fatty Acids 73 (2005) 9–15. [6] D. Barneda, A. Frontini, S. Cinti, M. Christian, Dynamic changes in lipid dropletassociated proteins in the “browning” of white adipose tissues, Biochim. Biophys. Acta 1831 (2013) 924–933. [7] M.J. Lee, Y. Wu, S.K. Fried, Adipose tissue heterogeneity: implication of depot differences in adipose tissue for obesity complications, Mol. Aspects Med. 34 (2013) 1–11. [8] A.H. Kissebah, G.R. Krakower, Regional adiposity and morbidity, Physiol. Rev. 74 (1994) 761–811. [9] K.L. Spalding, E. Arner, P.O. Westermark, S. Bernard, B.A. Buchholz, O. Bergmann, L. Blomqvist, J. Hoffstedt, E. Naslund, T. Britton, H. Concha, M. Hassan, M. Ryden, J. Frisen, P. Arner, Dynamics of fat cell turnover in humans, Nature 453 (2008) 783–787. [10] J.P. Despres, I. Lemieux, Abdominal obesity and metabolic syndrome, Nature 444 (2006) 881–887. [11] G. Enzi, M. Gasparo, P.R. Biondetti, D. Fiore, M. Semisa, F. Zurlo, Subcutaneous and visceral fat distribution according to sex, age and overweight, evaluated by computed tomography, Am. J. Clin. Nutr. 44 (1986) 739–746. [12] M.E. Trujillo, P.E. Scherer, Adipose tissue-derived factors: impact on health and disease, Endocr. Rev. 27 (2006) 762–778.

Please cite this article as: M. Masoodi, et al., Lipid signaling in adipose tissue: Connecting inflammation & metabolism, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbalip.2014.09.023

835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 862

865 866 867 Q3

869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 898 899 900 901 902

D

P

R O

O

F

[42] N. Enguix, R. Pardo, A. Gonzalez, V.M. Lopez, R. Simo, A. Kralli, J.A. Villena, Mice lacking PGC-1beta in adipose tissues reveal a dissociation between mitochondrial dysfunction and insulin resistance, Mol. Metab. 2 (2013) 215–226. [43] J.H. Choi, A.S. Banks, J.L. Estall, S. Kajimura, P. Bostrom, D. Laznik, J.L. Ruas, M.J. Chalmers, T.M. Kamenecka, M. Bluher, P.R. Griffin, B.M. Spiegelman, Anti-diabetic drugs inhibit obesity-linked phosphorylation of PPARgamma by Cdk5, Nature 466 (2010) 451–456. [44] P. Li, W. Fan, J. Xu, M. Lu, H. Yamamoto, J. Auwerx, D.D. Sears, S. Talukdar, D. Oh, A. Chen, G. Bandyopadhyay, M. Scadeng, J.M. Ofrecio, S. Nalbandian, J.M. Olefsky, Adipocyte NCoR knockout decreases PPARgamma phosphorylation and enhances PPARgamma activity and insulin sensitivity, Cell 147 (2011) 815–826. [45] S. Sugii, P. Olson, D.D. Sears, M. Saberi, A.R. Atkins, G.D. Barish, S.H. Hong, G.L. Castro, Y.Q. Yin, M.C. Nelson, G. Hsiao, D.R. Greaves, M. Downes, R.T. Yu, J.M. Olefsky, R.M. Evans, PPARgamma activation in adipocytes is sufficient for systemic insulin sensitization, Proc. Natl. Acad. Sci. U. S. A. 106 (2009) 22504–22509. [46] S.E. Shoelson, L. Herrero, A. Naaz, Obesity, inflammation, and insulin resistance, Gastroenterology 132 (2007) 2169–2180. [47] J.M. Olefsky, C.K. Glass, Macrophages, inflammation, and insulin resistance, Annu. Rev. Physiol. 72 (2010) 219–246. [48] A. Iyer, D.P. Fairlie, J.B. Prins, B.D. Hammock, L. Brown, Inflammatory lipid mediators in adipocyte function and obesity, Nat. Rev. Endocrinol. 6 (2010) 71–82. [49] L.A. O'Neill, D.G. Hardie, Metabolism of inflammation limited by AMPK and pseudo-starvation, Nature 493 (2013) 346–355. [50] G.M. Barton, J.C. Kagan, A cell biological view of Toll-like receptor function: regulation through compartmentalization, Nat. Rev. Immunol. 9 (2009) 535–542. [51] H. Shi, M.V. Kokoeva, K. Inouye, I. Tzameli, H. Yin, J.S. Flier, TLR4 links innate immunity and fatty acid-induced insulin resistance, J. Clin. Invest. 116 (2006) 3015–3025. [52] S.E. Shoelson, J. Lee, M. Yuan, Inflammation and the IKK beta/I kappa B/NF-kappa B axis in obesity- and diet-induced insulin resistance, Int. J. Obes. Relat. Metab. Disord. 27 (Suppl. 3) (2003) S49–S52. [53] M.C. Arkan, A.L. Hevener, F.R. Greten, S. Maeda, Z.W. Li, J.M. Long, A. WynshawBoris, G. Poli, J. Olefsky, M. Karin, IKK-beta links inflammation to obesity-induced insulin resistance, Nat. Med. 11 (2005) 191–198. [54] D. Cai, M. Yuan, D.F. Frantz, P.A. Melendez, L. Hansen, J. Lee, S.E. Shoelson, Local and systemic insulin resistance resulting from hepatic activation of IKK-beta and NF-kappaB, Nat. Med. 11 (2005) 183–190. [55] M.A. McArdle, O.M. Finucane, R.M. Connaughton, A.M. McMorrow, H.M. Roche, Mechanisms of obesity-induced inflammation and insulin resistance: insights into the emerging role of nutritional strategies, Front. Endocrinol. (Lausanne) 4 (2013) 52. [56] R. Dobrowsky, C. Kamibayashi, M. Mumby, Y. Hannun, Caramide activates heterotrimeric protein phosphatase 2A, J. Biol. Chem. 268 (1993) 15523–15530. [57] J.M. Haus, S.R. Kashyap, T. Kasumov, R. Zhang, K.R. Kelly, R.A. Defronzo, J.P. Kirwan, Plasma ceramides are elevated in obese subjects with type 2 diabetes and correlate with the severity of insulin resistance, Diabetes 58 (2009) 337–343. [58] W.L. Holland, R.A. Miller, Z.V. Wang, K. Sun, B.M. Barth, H.H. Bui, K.E. Davis, B.T. Bikman, N. Halberg, J.M. Rutkowski, M.R. Wade, V.M. Tenorio, M.S. Kuo, J.T. Brozinick, B.B. Zhang, M.J. Birnbaum, S.A. Summers, P.E. Scherer, Receptormediated activation of ceramidase activity initiates the pleiotropic actions of adiponectin, Nat. Med. 17 (2011) 55–63. [59] H. Wen, D. Gris, Y. Lei, S. Jha, L. Zhang, M.T. Huang, W.J. Brickey, J.P. Ting, Fatty acid-induced NLRP3-ASC inflammasome activation interferes with insulin signaling, Nat. Immunol. 12 (2011) 408–415. [60] B. Vandanmagsar, Y.H. Youm, A. Ravussin, J.E. Galgani, K. Stadler, R.L. Mynatt, E. Ravussin, J.M. Stephens, V.D. Dixit, The NLRP3 inflammasome instigates obesityinduced inflammation and insulin resistance, Nat. Med. 17 (2011) 179–188. [61] J.A. Chavez, S.A. Summers, A ceramide-centric view of insulin resistance, Cell Metab. 15 (2012) 585–594. [62] B. Vandanmagsar, Y.H. Youm, A. Ravussin, J.E. Galgani, K. Stadler, R.L. Mynatt, E. Ravussin, J.M. Stephens, V.D. Dixit, The NLRP3 inflammasome instigates obesity-induced inflammation and insulin resistance, Nat. Med. 17 (2011) (179-U214). [63] S.S. Iyer, Q. He, J.R. Janczy, E.I. Elliott, Z. Zhong, A.K. Olivier, J.J. Sadler, V. KnepperAdrian, R.Z. Han, L. Qiao, S.C. Eisenbarth, W.M. Nauseef, S.L. Cassel, F.S. Sutterwala, Mitochondrial cardiolipin Is required for Nlrp3 inflammasome activation, Immunity 39 (2013) 311–323. [64] R. Stienstra, L.A. Joosten, T. Koenen, B. van Tits, J.A. van Diepen, S.A. van den Berg, P.C. Rensen, P.J. Voshol, G. Fantuzzi, A. Hijmans, S. Kersten, M. Muller, W.B. van den Berg, N. van Rooijen, M. Wabitsch, B.J. Kullberg, J.W. van der Meer, T. Kanneganti, C.J. Tack, M.G. Netea, The inflammasome-mediated caspase-1 activation controls adipocyte differentiation and insulin sensitivity, Cell Metab. 12 (2010) 593–605. [65] R. Stienstra, J.A. van Diepen, C.J. Tack, M.H. Zaki, F.L. van de Veerdonk, D. Perera, G.A. Neale, G.J. Hooiveld, A. Hijmans, I. Vroegrijk, S. van den Berg, J. Romijn, P.C. Rensen, L.A. Joosten, M.G. Netea, T.D. Kanneganti, Inflammasome is a central player in the induction of obesity and insulin resistance, Proc. Natl. Acad. Sci. U. S. A. 108 (2011) 15324–15329. [66] A. Recchiuti, C.N. Serhan, Pro-Resolving Lipid Mediators (SPMs) and their actions in regulating miRNA in novel resolution circuits in inflammation, Front. Immunol. 3 (2012) 298. [67] E. Titos, J. Claria, Omega-3-derived mediators counteract obesity-induced adipose tissue inflammation, Prostaglandins Other Lipid Mediat. 107 (2013) 77–84. [68] 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) 1–7.

N

C

O

R

R

E

C

T

[13] H. Sell, C. Habich, J. Eckel, Adaptive immunity in obesity and insulin resistance, Nat. Rev. Endocrinol. 8 (2012) 709–716. [14] T.E. Graham, Q. Yang, M. Bluher, A. Hammarstedt, T.P. Ciaraldi, R.R. Henry, C.J. Wason, A. Oberbach, P.A. Jansson, U. Smith, B.B. Kahn, Retinol-binding protein 4 and insulin resistance in lean, obese and diabetic subjects, N. Engl. J. Med. 354 (2006) 2552–2563. [15] P.M. Moraes-Vieira, M.M. Yore, P.M. Dwyer, I. Syed, P. Aryal, B.B. Kahn, RBP4 activates antigen-presenting cells, leading to adipose tissue inflammation and systemic insulin resistance, Cell Metab. 19 (2014) 512–526. [16] J. Villarroya, R. Cereijo, F. Villarroya, An endocrine role for brown adipose tissue? Am. J. Physiol. Endocrinol. Metab. 305 (2013) E567–E572. [17] P. Flachs, M. Rossmeisl, O. Kuda, J. Kopecky, Stimulation of mitochondrial oxidative capacity in white fat independent of UCP1: a key to lean phenotype, Biochim. Biophys. Acta 1831 (2013) 986–1003. [18] A. Girousse, D. Langin, Adipocyte lipases and lipid droplet-associated proteins: insight from transgenic mouse models, Int. J. Obes. (Lond) 36 (2012) 581–594. [19] R. Zechner, R. Zimmermann, T.O. Eichmann, S.D. Kohlwein, G. Haemmerle, A. Lass, F. Madeo, FAT SIGNALS — lipases and lipolysis in lipid metabolism and signaling, Cell Metab. 15 (2012) 279–291. [20] G. Haemmerle, A. Lass, R. Zimmermann, G. Gorkiewicz, C. Meyer, J. Rozman, G. Heldmaier, R. Maier, C. Theussl, S. Eder, D. Kratky, E.F. Wagner, M. Klingenspor, G. Hoefler, R. Zechner, Defective lipolysis and altered energy metabolism in mice lacking adipose triglyceride lipase, Science 312 (2006) 734–737. [21] B. Spoto, F. Fezza, G. Parlongo, N. Battista, E. Sgro, V. Gasperi, C. Zoccali, M. Maccarrone, Human adipose tissue binds and metabolizes the endocannabinoids anandamide and 2-arachidonoylglycerol, Biochimie 88 (2006) 1889–1897. [22] B.R. Martin, R.M. Denton, The intracellular localization of enzymes in white-adipose-tissue fat-cells and permeability properties of fat-cell mitochondria, Biochem. J. 117 (1970) 861–877. [23] H. Bottcher, P. Furst, Decreased white fat cell thermogenesis in obese individuals, Int. J. Obes. 21 (1997) 439–444. [24] T. Wang, Y. Zang, W. Ling, B.E. Corkey, W. Guo, Metabolic partitioning of endogenous fatty acid in adipocytes, Obes. Res. 11 (2003) 880–887. [25] V.A. Hammond, D.G. Johnston, Substrate cycling between triglyceride and fatty acid in human adipocytes, Metabolism 36 (1987) 308–313. [26] C.L. Yen, S.J. Stone, S. Koliwad, C. Harris, R.V. Farese Jr., Thematic review series: glycerolipids. DGAT enzymes and triacylglycerol biosynthesis, J. Lipid Res. 49 (2008) 2283–2301. [27] C.K. Nye, R.W. Hanson, S.C. Kalhan, Glyceroneogenesis is the dominant pathway for triglyceride glycerol synthesis in vivo in the rat, J. Biol. Chem. 283 (2008) 27565–27574. [28] E.A. Newsholme, B. Crabtree, Substrate cycles: their metabolic energy and thermic consequences in man, Biochem. Soc. Symp. 43 (1976) 183–205. [29] E.P. Mottillo, P. Balasubramanian, Y.H. Lee, C. Weng, E.E. Kershaw, J.G. Granneman, Coupling of lipolysis and de novo lipogenesis in brown, beige, and white adipose tissues during chronic beta3-adrenergic receptor activation, J. Lipid Res. (2014), http://dx.doi.org/10.1194/jlr.M050005 (in press). [30] G. Haemmerle, T. Moustafa, G. Woelkart, S. Buttner, A. Schmidt, T. van de Weijer, M. Hesselink, D. Jaeger, P.C. Kienesberger, K. Zierler, R. Schreiber, T. Eichmann, D. Kolb, P. Kotzbeck, M. Schweiger, M. Kumari, S. Eder, G. Schoiswohl, N. Wongsiriroj, N.M. Pollak, F.P. Radner, K. Preiss-Landl, T. Kolbe, T. Rulicke, B. Pieske, M. Trauner, A. Lass, R. Zimmermann, G. Hoefler, S. Cinti, E.E. Kershaw, P. Schrauwen, F. Madeo, B. Mayer, R. Zechner, ATGL-mediated fat catabolism regulates cardiac mitochondrial function via PPAR-alpha and PGC-1, Nat. Med. 17 (2011) 1076–1085. [31] T.C. Walther, R.V. Farese Jr., The life of lipid droplets, Biochim. Biophys. Acta 1791 (2009) 459–466. [32] C. Chascione, D.H. Elwyn, M. Davila, K.M. Gil, J. Askanazi, J.M. Kinney, Effect of carbohydrate intake on de novo lipogenesis in human adipose tissue, Am. J. Physiol. 253 (1987) E664–E669. [33] M.P. Czech, M. Tencerova, D.J. Pedersen, M. Aouadi, Insulin signalling mechanisms for triacylglycerol storage, Diabetologia 56 (2013) 949–964. [34] M.A. Herman, O.D. Peroni, J. Villoria, M.R. Schon, N.A. Abumrad, M. Bluher, S. Klein, B.B. Kahn, A novel ChREBP isoform in adipose tissue regulates systemic glucose metabolism, Nature 484 (2012) 333–338. [35] C.M. Kusminski, P.E. Scherer, Mitochondrial dysfunction in white adipose tissue, Trends Endocrinol. Metab. 23 (2012) 435–443. [36] J. Naukkarinen, S. Heinonen, A. Hakkarainen, J. Lundbom, K. Vuolteenaho, L. Saarinen, S. Hautaniemi, A. Rodriguez, G. Fruhbeck, P. Pajunen, T. Hyotylainen, M. Oresic, E. Moilanen, A. Suomalainen, N. Lundbom, J. Kaprio, A. Rissanen, K.H. Pietilainen, Characterising metabolically healthy obesity in weight-discordant monozygotic twins, Diabetologia 57 (2014) 167–176. [37] F. Diraison, E. Dusserre, H. Vidal, M. Sothier, M. Beylot, Increased hepatic lipogenesis but decreased expression of lipogenic gene in adipose tissue in human obesity, Am. J. Physiol. Endocrinol. Metab. 282 (2002) E46–E51. [38] W. Wahli, L. Michalik, PPARs at the crossroads of lipid signaling and inflammation, Trends Endocrinol. Metab. 23 (2012) 351–363. [39] P. Escher, O. Braissant, S. Basu-Modak, L. Michalik, W. Wahli, B. Desvergne, Rat PPARs: quantitative analysis in adult rat tissues and regulation in fasting and refeeding, Endocrinology 142 (2001) 4195–4202. [40] Y.X. Wang, C.H. Lee, S. Tiep, R.T. Yu, J. Ham, H. Kang, R.M. Evans, Peroxisomeproliferator-activated receptor delta activates fat metabolism to prevent obesity, Cell 113 (2003) 159–170. [41] P. Tontonoz, E. Hu, B.M. Spiegelman, Stimulation of adipogenesis in fibroblasts by PPARgama2, a lipid-activated transcription factor, Cell 79 (1994) 1147–1156.

U

903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943 944 945 946 947 Q4 948 949 950 951 952 953 954 955 956 957 958 959 960 961 962 963 964 965 966 967 968 969 970 971 972 973 974 975 976 977 978 979 980 981 982 983 984 985 986 987 988

M. Masoodi et al. / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

E

12

Please cite this article as: M. Masoodi, et al., Lipid signaling in adipose tissue: Connecting inflammation & metabolism, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbalip.2014.09.023

989 990 991 992 993 994 995 996 997 998 999 1000 1001 1002 1003 1004 1005 1006 1007 1008 1009 1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 1034 1035 1036 1037 1038 1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052 1053 1054 1055 1056 1057 1058 1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 1069 1070 1071 1072 1073 1074

M. Masoodi et al. / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

N C O

R

R

E

C

D

P

R O

O

F

[92] S. Winer, Y. Chan, G. Paltser, D. Truong, H. Tsui, J. Bahrami, R. Dorfman, Y. Wang, J. Zielenski, F. Mastronardi, Y. Maezawa, D.J. Drucker, E. Engleman, D. Winer, H.M. Dosch, Normalization of obesity-associated insulin resistance through immunotherapy, Nat. Med. 15 (2009) 921–929. [93] D. Wu, A.B. Molofsky, H.E. Liang, R.R. Ricardo-Gonzalez, H.A. Jouihan, J.K. Bando, A. Chawla, R.M. Locksley, Eosinophils sustain adipose alternatively activated macrophages associated with glucose homeostasis, Science 332 (2011) 243–247. [94] R.R. Rao, J.Z. Long, J.P. White, K.J. Svensson, J. Lou, I. Lokurkar, M.P. Jedrychowski, J.L. Ruas, C.D. Wrann, J.C. Lo, D.M. Camera, J. Lachey, S. Gygi, J. Seehra, J.A. Hawley, B.M. Spiegelman, Meteorin-like is a hormone that regulates immune-adipose interactions to increase beige fat thermogenesis, Cell 157 (2014) 1279–1291. [95] Y. Qiu, K.D. Nguyen, J.I. Odegaard, X. Cui, X. Tian, R.M. Locksley, R.D. Palmiter, A. Chawla, Eosinophils and type 2 cytokine signaling in macrophages orchestrate development of functional beige fat, Cell 157 (2014) 1292–1308. [96] V. Mansuy-Aubert, Q.L. Zhou, X. Xie, Z. Gong, J.Y. Huang, A.R. Khan, G. Aubert, K. Candelaria, S. Thomas, D.J. Shin, S. Booth, S.M. Baig, A. Bilal, D. Hwang, H. Zhang, R. Lovell-Badge, S.R. Smith, F.R. Awan, Z.Y. Jiang, Imbalance between neutrophil elastase and its inhibitor alpha1-antitrypsin in obesity alters insulin sensitivity, inflammation, and energy expenditure, Cell Metab. 17 (2013) 534–548. [97] S. Talukdar, D.Y. Oh, G. Bandyopadhyay, D. Li, J. Xu, J. McNelis, M. Lu, P. Li, Q. Yan, Y. Zhu, J. Ofrecio, M. Lin, M.B. Brenner, J.M. Olefsky, Neutrophils mediate insulin resistance in mice fed a high-fat diet through secreted elastase, Nat. Med. 18 (2012) 1407–1412. [98] J. Liu, A. Divoux, J. Sun, J. Zhang, K. Clement, J.N. Glickman, G.K. Sukhova, P.J. Wolters, J. Du, C.Z. Gorgun, A. Doria, P. Libby, R.S. Blumberg, B.B. Kahn, G.S. Hotamisligil, G.P. Shi, Genetic deficiency and pharmacological stabilization of mast cells reduce diet-induced obesity and diabetes in mice, Nat. Med. 15 (2009) 940–945. [99] A. Tanaka, Y. Nomura, A. Matsuda, K. Ohmori, H. Matsuda, Mast cells function as an alternative modulator of adipogenesis through 15-deoxy-delta-12, 14-prostaglandin J2, Am. J. Physiol. Cell Physiol. 301 (2011) C1360–C1367. [100] K. Jaworski, M. Ahmadian, R.E. Duncan, E. Sarkadi-Nagy, K.A. Varady, M.K. Hellerstein, H.Y. Lee, V.T. Samuel, G.I. Shulman, K.H. Kim, S. de Val, C. Kang, H.S. Sul, AdPLA ablation increases lipolysis and prevents obesity induced by high-fat feeding or leptin deficiency, Nat. Med. 15 (2009) 159–168. [101] J.N. Fain, C.W. Leffler, S.W. Bahouth, Eicosanoids as endogenous regulators of leptin release and lipolysis by mouse adipose tissue in primary culture, J. Lipid Res. 41 (2000) 1689–1694. [102] B.B. Fredholm, P. Hedqvist, Role of pre- and postjunctional inhibition by prostaglandin E2 of lipolysis induced by sympathetic nerve stimulation in dog subcutaneous adipose tissue in situ, Br. J. Pharmacol. 47 (1973) 711–718. [103] B. Richelsen, Factors regulating the production of prostaglandin E2 and prostacyclin (prostaglandin I2) in rat and human adipocytes, Biochem. J. 247 (1987) 389–394. [104] K. Chatzipanteli, S. Rudolph, L. Axelrod, Coordinate control of lipolysis by prostaglandin E2 and prostacyclin in rat adipose tissue, Diabetes 41 (1992) 927–935. [105] J.N. Fain, C.W. Leffler, G.S. Cowan Jr., C. Buffington, L. Pouncey, S.W. Bahouth, Stimulation of leptin release by arachidonic acid and prostaglandin E(2) in adipose tissue from obese humans, Metabolism 50 (2001) 921–928. [106] B. Richelsen, J.D. Borglum, S.S. Sorensen, Biosynthetic capacity and regulatory aspects of prostaglandin E2 formation in adipocytes, Mol. Cell. Endocrinol. 85 (1992) 73–81. [107] M.K. Mater, D. Pan, W.G. Bergen, D.B. Jump, Arachidonic acid inhibits lipogenic gene expression in 3T3-L1 adipocytes through a prostanoid pathway, J. Lipid Res. 39 (1998) 1327–1334. [108] H. Yan, A. Kermouni, M. Abdel-Hafez, D.C. Lau, Role of cyclooxygenases COX-1 and COX-2 in modulating adipogenesis in 3T3-L1 cells, J. Lipid Res. 44 (2003) 424–429. [109] G.Y. Chiou, J.C. Fong, Prostaglandin F2alpha increases glucose transport in 3T3-L1 adipocytes through enhanced GLUT1 expression by a protein kinase C-dependent pathway, Cell. Signal. 16 (2004) 415–421. [110] S. Ghoshal, D.B. Trivedi, G.A. Graf, C.D. Loftin, Cyclooxygenase-2 deficiency attenuates adipose tissue differentiation and inflammation in mice, J. Biol. Chem. 286 (2011) 889–898. [111] M.A. Mazid, A.A. Chowdhury, K. Nagao, K. Nishimura, M. Jisaka, T. Nagaya, K. Yokota, Endogenous 15-deoxy-Delta(12,14)-prostaglandin J(2) synthesized by adipocytes during maturation phase contributes to upregulation of fat storage, FEBS Lett. 580 (2006) 6885–6890. [112] I.J. Lodhi, L. Yin, A.P. Jensen-Urstad, K. Funai, T. Coleman, J.H. Baird, M.K. El Ramahi, B. Razani, H. Song, F. Fu-Hsu, J. Turk, C.F. Semenkovich, Inhibiting adipose tissue lipogenesis reprograms thermogenesis and PPARgamma activation to decrease diet-induced obesity, Cell Metab. 16 (2012) 189–201. [113] A.A. Chowdhury, M.S. Rahman, K. Nishimura, M. Jisaka, T. Nagaya, T. Ishikawa, F. Shono, K. Yokota, 15-Deoxy-Delta(12,14)-prostaglandin J(2) interferes inducible synthesis of prostaglandins E(2) and F(2alpha) that suppress subsequent adipogenesis program in cultured preadipocytes, Prostaglandins Other Lipid Mediat. 95 (2011) 53–62. [114] M.S. Hossain, K. Nishimura, M. Jisaka, T. Nagaya, K. Yokota, Prostaglandin J2 series induces the gene expression of monocyte chemoattractant protein-1 during the maturation phase of cultured adipocytes, Gene 502 (2012) 138–141. [115] J.P. Hardwick, K. Eckman, Y.K. Lee, M.A. Abdelmegeed, A. Esterle, W.M. Chilian, J.Y. Chiang, B.J. Song, Eicosanoids in metabolic syndrome, Adv. Pharmacol. 66 (2013) 157–266. [116] K. Yu, W. Bayona, C.B. Kallen, H.P. Harding, C.P. Ravera, G. McMahon, M. Brown, M.A. Lazar, Differential activation of peroxisome proliferator-activated receptors by eicosanoids, J. Biol. Chem. 270 (1995) 23975–23983.

E

T

[69] M. Spite, J. Claria, C.N. Serhan, Resolvins, specialized proresolving lipid mediators, and their potential roles in metabolic diseases, Cell Metab. 19 (2014) 21–36. [70] C.N. Lumeng, J.B. Delproposto, D.J. Westcott, A.R. Saltiel, Phenotypic switching of adipose tissue macrophages with obesity is generated by spatiotemporal differences in macrophage subtypes, Diabetes 57 (2008) 3239–3246. [71] D.Y. Oh, S. Talukdar, E.J. Bae, T. Imamura, H. Morinaga, W.Q. Fan, P.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) 687–698. [72] G.S. Hotamisligil, Inflammation and metabolic disorders, Nature 444 (2006) 860–867. [73] A. Guilherme, J.V. Virbasius, V. Puri, M.P. Czech, Adipocyte dysfunctions linking obesity to insulin resistance and type 2 diabetes, Nat. Rev. Mol. Cell Biol. 9 (2008) 367–377. [74] C.N. Lumeng, S.M. Deyoung, J.L. Bodzin, A.R. Saltiel, Increased inflammatory properties of adipose tissue macrophages recruited during diet-induced obesity, Diabetes 56 (2007) 16–23. [75] C.N. Lumeng, J.L. Bodzin, A.R. Saltiel, Obesity induces a phenotypic switch in adipose tissue macrophage polarization, J. Clin. Invest. 117 (2007) 175–184. [76] S.P. Weisberg, D. McCann, M. Desai, M. Rosenbaum, R.L. Leibel, A.W. Ferrante, Obesity is associated with macrophage accumulation in adipose tissue, J. Clin. Investig. 112 (2003) 1796–1808. [77] H. Xu, G.T. Barnes, Q. Yang, G. Tan, D. Yang, C.J. Chou, J. Sole, A. Nichols, J.S. Ross, L.A. Tartaglia, H. Chen, Chronic inflammation in fat plays a crucial role in the development of obesity-related insulin resistance, J. Clin. Invest. 112 (2003) 1821–1830. [78] N. Kloting, M. Fasshauer, A. Dietrich, P. Kovacs, M.R. Schon, M. Kern, M. Stumvoll, M. Bluher, Insulin-sensitive obesity, Am. J. Physiol. Endocrinol. Metab. 299 (2010) E506–E515. [79] H. Kanda, S. Tateya, Y. Tamori, K. Kotani, K. Hiasa, R. Kitazawa, S. Kitazawa, H. Miyachi, S. Maeda, K. Egashira, M. Kasuga, MCP-1 contributes to macrophage infiltration into adipose tissue, insulin resistance, and hepatic steatosis in obesity, J. Clin. Invest. 116 (2006) 1494–1505. [80] N. Kamei, K. Tobe, R. Suzuki, M. Ohsugi, T. Watanabe, N. Kubota, N. OhtsukaKowatari, K. Kumagai, K. Sakamoto, M. Kobayashi, T. Yamauchi, K. Ueki, Y. Oishi, S. Nishimura, I. Manabe, H. Hashimoto, Y. Ohnishi, H. Ogata, K. Tokuyama, M. Tsunoda, T. Ide, K. Murakami, R. Nagai, T. Kadowaki, Overexpression of monocyte chemoattractant protein-1 in adipose tissues causes macrophage recruitment and insulin resistance, J. Biol. Chem. 281 (2006) 26602–26614. [81] S.U. Amano, J.L. Cohen, P. Vangala, M. Tencerova, S.M. Nicoloro, J.C. Yawe, Y. Shen, M.P. Czech, M. Aouadi, Local proliferation of macrophages contributes to obesityassociated adipose tissue inflammation, Cell Metab. 19 (2014) 162–171. [82] J. Haase, U. Weyer, K. Immig, N. Kloting, M. Bluher, J. Eilers, I. Bechmann, M. Gericke, Local proliferation of macrophages in adipose tissue during obesityinduced inflammation, Diabetologia 57 (2014) 562–571. [83] S. Cinti, G. Mitchell, G. Barbatelli, I. Murano, E. Ceresi, E. Faloia, S. Wang, M. Fortier, A.S. Greenberg, M.S. Obin, Adipocyte death defines macrophage localization and function in adipose tissue of obese mice and humans, J. Lipid Res. 46 (2005) 2347–2355. [84] Y. Oh da, 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. BognerStrauss, J.M. Olefsky, A Gpr120-selective agonist improves insulin resistance and chronic inflammation in obese mice, Nat. Med. 20 (2014) 942–947. [85] A. Ichimura, A. Hirasawa, O. Poulain-Godefroy, A. Bonnefond, T. Hara, L. Yengo, I. Kimura, A. Leloire, N. Liu, K. Iida, H. Choquet, P. Besnard, C. Lecoeur, S. Vivequin, K. Ayukawa, M. Takeuchi, K. Ozawa, M. Tauber, C. Maffeis, A. Morandi, R. Buzzetti, P. Elliott, A. Pouta, M.R. Jarvelin, A. Korner, W. Kiess, M. Pigeyre, R. Caiazzo, H.W. Van, G.L. Van, F. Horber, B. Balkau, C. Levy-Marchal, K. Rouskas, A. Kouvatsi, J. Hebebrand, A. Hinney, A. Scherag, F. Pattou, D. Meyre, T.A. Koshimizu, I. Wolowczuk, G. Tsujimoto, P. Froguel, Dysfunction of lipid sensor GPR120 leads to obesity in both mouse and human, Nature 483 (2012) 350–354. [86] J. Hellmann, Y. Tang, M. Kosuri, A. Bhatnagar, M. Spite, Resolvin D1 decreases adipose tissue macrophage accumulation and improves insulin sensitivity in obese-diabetic mice, FASEB J. 25 (2011) 2399–2407. [87] E. Titos, B. Rius, A. Gonzalez-Periz, C. Lopez-Vicario, E. Moran-Salvador, M. Martinez-Clemente, V. Arroyo, J. Claria, Resolvin D1 and its precursor docosahexaenoic acid promote resolution of adipose tissue inflammation by eliciting macrophage polarization toward an M2-like phenotype, J. Immunol. 187 (2011) 5408–5418. [88] R.W. O'Rourke, M.D. Metcalf, A.E. White, A. Madala, B.R. Winters, I.I. Maizlin, B.A. Jobe, C.T. Roberts Jr., M.K. Slifka, D.L. Marks, Depot-specific differences in inflammatory mediators and a role for NK cells and IFN-gamma in inflammation in human adipose tissue, Int. J. Obes. 33 (2009) 978–990. [89] S.C. Huang, B. Everts, Y. Ivanova, D. O'Sullivan, M. Nascimento, A.M. Smith, W. Beatty, L. Love-Gregory, W.Y. Lam, C.M. O'Neill, C. Yan, H. Du, N.A. Abumrad, J.F. Urban Jr., M.N. Artyomov, E.L. Pearce, E.J. Pearce, Cell-intrinsic lysosomal lipolysis is essential for alternative activation of macrophages, Nat. Immunol. 15 (2014) 846–855. [90] M. Feuerer, L. Herrero, D. Cipolletta, A. Naaz, J. Wong, A. Nayer, J. Lee, A.B. Goldfine, C. Benoist, S. Shoelson, D. Mathis, Lean, but not obese, fat is enriched for a unique population of regulatory T cells that affect metabolic parameters, Nat. Med. 15 (2009) 930–939. [91] S. Nishimura, I. Manabe, M. Nagasaki, K. Eto, H. Yamashita, M. Ohsugi, M. Otsu, K. Hara, K. Ueki, S. Sugiura, K. Yoshimura, T. Kadowaki, R. Nagai, CD8+ effector T cells contribute to macrophage recruitment and adipose tissue inflammation in obesity, Nat. Med. 15 (2009) 914–920.

U

1075 1076 1077 1078 1079 1080 1081 1082 1083 1084 1085 1086 1087 1088 1089 1090 1091 1092 1093 1094 1095 1096 1097 1098 1099 1100 1101 1102 1103 1104 1105 1106 1107 1108 1109 1110 1111 1112 1113 1114 1115 1116 1117 1118 1119 1120 1121 1122 1123 1124 1125 1126 1127 1128 1129 1130 1131 1132 1133 1134 1135 1136 1137 1138 1139 1140 1141 1142 1143 1144 1145 1146 1147 1148 1149 1150 1151 1152 1153 1154 1155 1156 1157 1158 1159 1160

13

Please cite this article as: M. Masoodi, et al., Lipid signaling in adipose tissue: Connecting inflammation & metabolism, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbalip.2014.09.023

1161 1162 1163 1164 1165 1166 1167 1168 1169 1170 1171 1172 1173 1174 1175 1176 1177 1178 1179 1180 1181 1182 1183 1184 1185 1186 1187 1188 1189 1190 1191 1192 1193 1194 1195 1196 1197 1198 1199 1200 1201 1202 1203 1204 1205 1206 1207 1208 1209 1210 1211 1212 1213 1214 1215 1216 1217 1218 1219 1220 1221 1222 1223 1224 1225 1226 1227 1228 1229 1230 1231 1232 1233 1234 1235 1236 1237 1238 1239 1240 1241 1242 1243 1244 1245 1246

[147]

[148]

[149]

[150]

[151]

D

[152]

[153]

[154]

C

E

R

R

O

[155]

[156]

[157]

[158]

[159]

C

N

F

[146]

O

[145]

R O

[144]

cultured preadipocytes impairs adipogenesis program independently of endogenous prostanoids, Exp. Cell Res. 318 (2012) 408–415. S. Lu, K. Nishimura, M.A. Hossain, M. Jisaka, T. Nagaya, K. Yokota, Regulation and role of arachidonate cascade during changes in life cycle of adipocytes, Appl. Biochem. Biotechnol. 118 (2004) 133–153. Y. Mendizabal, S. Llorens, E. Nava, Vasoactive effects of prostaglandins from the perivascular fat of mesenteric resistance arteries in WKY and SHROB rats, Life Sci. 93 (2013) 1023–1032. K. Fujimori, T. Ueno, N. Nagata, K. Kashiwagi, K. Aritake, F. Amano, Y. Urade, Suppression of adipocyte differentiation by aldo-keto reductase 1B3 acting as prostaglandin F2alpha synthase, J. Biol. Chem. 285 (2010) 8880–8886. J.D. Borglum, S.B. Pedersen, G. Ailhaud, R. Negrel, B. Richelsen, Differential expression of prostaglandin receptor mRNAs during adipose cell differentiation, Prostaglandins Other Lipid Mediat. 57 (1999) 305–317. P. Strong, R.A. Coleman, P.P. Humphrey, Prostanoid-induced inhibition of lipolysis in rat isolated adipocytes: probable involvement of EP3 receptors, Prostaglandins 43 (1992) 559–566. B.K. Cole, D.C. Lieb, A.D. Dobrian, J.L. Nadler, 12- and 15-lipoxygenases in adipose tissue inflammation, Prostaglandins Other Lipid Mediat. 104–105 (2013) 84–92. D. Cota, G. Marsicano, M. Tschop, Y. Grubler, C. Flachskamm, M. Schubert, D. Auer, A. Yassouridis, C. Thone-Reineke, S. Ortmann, F. Tomassoni, C. Cervino, E. Nisoli, A.C. Linthorst, R. Pasquali, B. Lutz, G.K. Stalla, U. Pagotto, The endogenous cannabinoid system affects energy balance via central orexigenic drive and peripheral lipogenesis, J. Clin. Invest. 112 (2003) 423–431. I. Matias, M.P. Gonthier, P. Orlando, V. Martiadis, L. De Petrocellis, C. Cervino, S. Petrosino, L. Hoareau, F. Festy, R. Pasquali, R. Roche, M. Maj, U. Pagotto, P. Monteleone, V. Di Marzo, Regulation, function, and dysregulation of endocannabinoids in models of adipose and beta-pancreatic cells and in obesity and hyperglycemia, J. Clin. Endocrinol. Metab. 91 (2006) 3171–3180. G. Astarita, D. Piomelli, Lipidomic analysis of endocannabinoid metabolism in biological samples, J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 877 (2009) 2755–2767. A. Kosteli, E. Sugaru, G. Haemmerle, J.F. Martin, J. Lei, R. Zechner, A.W. Ferrante Jr., Weight loss and lipolysis promote a dynamic immune response in murine adipose tissue, J. Clin. Invest. 120 (2010) 3466–3479. D. Medrikova, J.Z. Macek, K. Bardova, P. Janovsk, M. Rossmeisl, J. Kopecky, Sex differences during the course of diet-induced obesity in mice: adipose tissue expandability and glycemic control, Int. J. Obes. 36 (2011) 262–272. G.A. Tynan, C.H. Hearnden, E. Oleszycka, C.L. Lyons, G. Coutts, J. O'Connell, M.A. Corrigan, L. Lynch, M. Campbell, J.J. Callanan, K.H. Mok, J. Geoghegan, C. O'Farrelly, S.M. Allan, H.M. Roche, D.B. O'Shea, E.C. Lavelle, Endogenous oils derived from human adipocytes are potent adjuvants that promote IL-1alpha-dependent inflammation, Diabetes 63 (2014) 2037–2050. F. Capel, E. Klimcakova, N. Viguerie, B. Roussel, M. Vitkova, M. Kovacikova, J. Polak, Z. Kovacova, J. Galitzky, J.J. Maoret, J. Hanacek, T.H. Pers, A. Bouloumie, V. Stich, D. Langin, Macrophages and adipocytes in human obesity: adipose tissue gene expression and insulin sensitivity during calorie restriction and weight stabilization, Diabetes 58 (2009) 1558–1567. T. Raclot, H. Oudart, Net release of individual fatty acids from white adipose tissue during lipolysis in vitro: evidence for selective fatty acid re-uptake, Biochem. J 348 (Pt 1) (2000) 129–136. T. Raclot, D. Langin, M. Lafontan, R. Groscolas, Selective release of human adipocyte fatty acids according to molecular structure, Biochem. J. 324 (Pt 3) (1997) 911–915. M.E. Kranendonk, F.L. Visseren, B.W. van Balkom, E.N. Nolte-'t Hoen, J.A. van Herwaarden, W. de Jager, H.S. Schipper, A.B. Brenkman, M.C. Verhaar, M.H. Wauben, E. Kalkhoven, Human adipocyte extracellular vesicles in reciprocal signaling between adipocytes and macrophages, Obesity (Silver Spring) 22 (2014) 1296–1308. P. Flachs, R. Ruhl, M. Hensler, P. Janovsk, P. Zouhar, V. Kus, J.Z. Macek, E. Papp, O. Kuda, M. Svobodova, M. Rossmeisl, G. Tsenov, V. Mohamed-Ali, J. Kopecky, Synergistic induction of lipid catabolism and anti-inflammatory lipids in white fat of dietary obese mice in response to calorie restriction and n−3 fatty acids, Diabetologia 54 (2011) 2626–2638. B. Gustafson, S. Gogg, S. Hedjazifar, L. Jenndahl, A. Hammarstedt, U. Smith, Inflammation and impaired adipogenesis in hypertrophic obesity in man, Am. J. Physiol. Endocrinol. Metab. 297 (2009) E999–E1003. P.C. Calder, N. Ahluwalia, F. Brouns, T. Buetler, K. Clement, K. Cunningham, K. Esposito, L.S. Jonsson, H. Kolb, M. Lansink, A. Marcos, A. Margioris, N. Matusheski, H. Nordmann, J. O'Brien, G. Pugliese, S. Rizkalla, C. Schalkwijk, J. Tuomilehto, J. Warnberg, B. Watzl, B.M. Winklhofer-Roob, Dietary factors and low-grade inflammation in relation to overweight and obesity, Br. J. Nutr. 106 (Suppl. 3) (2011) S5–S78. B. Vessby, M. Unsitupa, K. Hermansen, G. Riccardi, A.A. Rivellese, L.C. Tapsell, C. Nalsen, L. Berglund, A. Louheranta, B.M. Rasmussen, G.D. Calvert, A. Maffetone, E. Pedersen, I.B. Gustafsson, L.H. Storlien, Substituting dietary saturated for monounsaturated fat impairs insulin sensitivity in healthy men and women: The KANWU Study1, Diabetologia 44 (2001) 312–319. S.A. Summers, Ceramides in insulin resistance and lipotoxicity, Prog. Lipid Res. 45 (2006) 42–72. W.L. Holland, J.T. Brozinick, L.P. Wang, E.D. Hawkins, K.M. Sargent, Y. Liu, K. Narra, K.L. Hoehn, T.A. Knotts, A. Siesky, D.H. Nelson, S.K. Karathanasis, G.K. Fontenot, M.J. Birnbaum, S.A. Summers, Inhibition of ceramide synthesis ameliorates glucocorticoid-, saturated-fat-, and obesity-induced insulin resistance, Cell Metab. 5 (2007) 167–179.

P

[143]

T

[117] P. Hallenborg, C. Jorgensen, R.K. Petersen, S. Feddersen, P. Araujo, P. Markt, T. Langer, G. Furstenberger, P. Krieg, A. Koppen, E. Kalkhoven, L. Madsen, K. Kristiansen, Epidermis-type lipoxygenase 3 regulates adipocyte differentiation and PPAR{gamma} activity, Mol. Cell. Biol. (2010). [118] C. Nugent, J.B. Prins, J.P. Whitehead, J.M. Wentworth, V.K. Chatterjee, S. O'Rahilly, Arachidonic acid stimulates glucose uptake in 3T3-L1 adipocytes by increasing GLUT1 and GLUT4 levels at the plasma membrane. Evidence for involvement of lipoxygenase metabolites and peroxisome proliferator-activated receptor gamma, J. Biol. Chem. 276 (2001) 9149–9157. [119] A.D. Dobrian, D.C. Lieb, Q. Ma, J.W. Lindsay, B.K. Cole, K. Ma, S.K. Chakrabarti, N.S. Kuhn, S.D. Wohlgemuth, M. Fontana, J.L. Nadler, Differential expression and localization of 12/15 lipoxygenases in adipose tissue in human obese subjects, Biochem. Biophys. Res. Commun. 403 (2010) 485–490. [120] R. Horrillo, A. Gonzalez-Periz, M. Martinez-Clemente, M. Lopez-Parra, N. Ferre, E. Titos, E. Moran-Salvador, R. Deulofeu, V. Arroyo, J. Claria, 5-Lipoxygenase activating protein signals adipose tissue inflammation and lipid dysfunction in experimental obesity, J. Immunol. 184 (2010) 3978–3987. [121] S.K. Chakrabarti, B.K. Cole, Y. Wen, S.R. Keller, J.L. Nadler, 12/15-lipoxygenase products induce inflammation and impair insulin signaling in 3T3-L1 adipocytes, Obesity (Silver Spring) 17 (2009) 1657–1663. [122] S.K. Chakrabarti, Y. Wen, A.D. Dobrian, B.K. Cole, Q. Ma, H. Pei, M.D. Williams, M.H. Bevard, G.E. Vandenhoff, S.R. Keller, J. Gu, J.L. Nadler, Evidence for activation of inflammatory lipoxygenase pathways in visceral adipose tissue of obese Zucker rats, Am. J. Physiol. Endocrinol. Metab. 300 (2011) E175–E187. [123] E.K. Long, K. Hellberg, R. Foncea, A.V. Hertzel, J. Suttles, D.A. Bernlohr, Fatty acids induce leukotriene C4 synthesis in macrophages in a fatty acid binding proteindependent manner, Biochim. Biophys. Acta 1831 (2013) 1199–1207. [124] 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) 2597–2605. [125] A. Neuhofer, M. Zeyda, D. Mascher, B.K. Itariu, I. Murano, L. Leitner, E.E. Hochbrugger, P. Fraisl, S. Cinti, C.N. Serhan, T.M. Stulnig, Impaired local production of proresolving lipid mediators in obesity and 17-HDHA as a potential treatment for obesity-associated inflammation, Diabetes 62 (2013) 1945–1956. [126] 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) C1141–C1149. [127] P.J. White, A. Marette, Potential role of omega-3-derived resolution mediators in metabolic inflammation, Immunol. Cell Biol. 92 (2014) 324–330. [128] P.J. White, P. St-Pierre, A. Charbonneau, P.L. Mitchell, E. St-Amand, B. Marcotte, A. Marette, Protectin DX alleviates insulin resistance by activating a myokine-liver glucoregulatory axis, Nat. Med. 20 (2014) 664–669. [129] S. Banni, M.V. Di, Effect of dietary fat on endocannabinoids and related mediators: consequences on energy homeostasis, inflammation and mood, Mol. Nutr. Food Res. 54 (2010) 82–92. [130] M. Bluher, S. Engeli, N. Kloting, J. Berndt, M. Fasshauer, S. Batkai, P. Pacher, M.R. Schon, J. Jordan, M. Stumvoll, Dysregulation of the peripheral and adipose tissue endocannabinoid system in human abdominal obesity, Diabetes 55 (2006) 3053–3060. [131] M.V. Di, The endocannabinoid system in obesity and type 2 diabetes, Diabetologia 51 (2008) 1356–1367. [132] D. Osei-Hyiaman, J. Liu, L. Zhou, G. Godlewski, J. Harvey-White, W.I. Jeong, S. Batkai, G. Marsicano, B. Lutz, C. Buettner, G. Kunos, Hepatic CB1 receptor is required for development of diet-induced steatosis, dyslipidemia, and insulin and leptin resistance in mice, J. Clin. Investig. 118 (2008) 3160–3169. [133] A.R. Alvheim, M.K. Malde, D. Osei-Hyiaman, L.Y. Hong, R.J. Pawlosky, L. Madsen, K. Kristiansen, L. Froyland, J.R. Hibbeln, Dietary linoleic acid elevates endogenous 2-AG and anandamide and induces obesity, Obesity (Silver Spring) 20 (2012) 1984–1994. [134] I. Matias, S. Petrosino, A. Racioppi, R. Capasso, A.A. Izzo, M.V. Di, Dysregulation of peripheral endocannabinoid levels in hyperglycemia and obesity: effect of high fat diets, Mol. Cell. Endocrinol. 286 (2008) S66–S78. [135] T.M. D'Eon, K.A. Pierce, J.J. Roix, A. Tyler, H. Chen, S.R. Teixeira, The role of adipocyte insulin resistance in the pathogenesis of obesity-related elevations in endocannabinoids, Diabetes 57 (2008) 1262–1268. [136] R.G. Mattace, A. Santoro, R. Russo, R. Simeoli, O. Paciello, C.C. Di, S. Diano, A. Calignano, R. Meli, Palmitoylethanolamide prevents metabolic alterations and restores leptin sensitivity in ovariectomized rats, Endocrinology 155 (2014) 1291–1301. [137] C. Silvestri, A. Martella, N.J. Poloso, F. Piscitelli, R. Capasso, A. Izzo, D.F. Woodward, V. Di Marzo, Anandamide-derived prostamide F2alpha negatively regulates adipogenesis, J. Biol. Chem. 288 (2013) 23307–23321. [138] M. Rossmeisl, J.Z. Macek, 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. [139] S.E. O'Sullivan, Cannabinoids go nuclear: evidence for activation of peroxisome proliferator-activated receptors, Br. J. Pharmacol. 152 (2007) 576–582. [140] J.D. Borglum, B. Richelsen, C. Darimont, S.B. Pedersen, R. Negrel, Expression of the two isoforms of prostaglandin endoperoxide synthase (PGHS-1 and PGHS-2) during adipose cell differentiation, Mol. Cell. Endocrinol. 131 (1997) 67–77. [141] P.O. Hetu, D. Riendeau, Down-regulation of microsomal prostaglandin E2 synthase-1 in adipose tissue by high-fat feeding, Obesity (Silver Spring) 15 (2007) 60–68. [142] M.S. Hossain, A.A. Chowdhury, M.S. Rahman, K. Nishimura, M. Jisaka, T. Nagaya, F. Shono, K. Yokota, Stable expression of lipocalin-type prostaglandin D synthase in

U

1247 1248 1249 1250 Q5 1251 1252 1253 1254 1255 1256 1257 1258 1259 1260 1261 1262 1263 1264 1265 1266 1267 1268 1269 1270 1271 1272 1273 1274 1275 1276 1277 1278 1279 1280 1281 1282 1283 1284 1285 1286 1287 1288 1289 1290 1291 1292 1293 1294 1295 1296 1297 1298 1299 1300 1301 1302 1303 1304 1305 1306 1307 1308 1309 1310 1311 1312 1313 1314 1315 1316 1317 1318 1319 1320 1321 1322 1323 1324 1325 1326 1327 1328 1329 1330 1331 1332

M. Masoodi et al. / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

E

14

[160]

[161]

[162]

[163] [164]

Please cite this article as: M. Masoodi, et al., Lipid signaling in adipose tissue: Connecting inflammation & metabolism, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbalip.2014.09.023

1333 1334 1335 1336 1337 1338 1339 1340 1341 1342 1343 1344 1345 1346 1347 1348 1349 1350 1351 1352 1353 1354 1355 1356 1357 1358 1359 1360 1361 1362 1363 1364 1365 1366 1367 1368 1369 1370 1371 1372 1373 1374 1375 1376 1377 1378 1379 1380 1381 1382 1383 1384 1385 1386 1387 1388 1389 1390 1391 1392 1393 1394 1395 1396 1397 1398 1399 1400 1401 1402 1403 1404 1405 1406 1407 1408 1409 1410 1411 1412 1413 1414 1415 1416 1417 1418

M. Masoodi et al. / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

[196]

[197]

[198]

[199]

[200]

[201]

E

D

[202]

C

E

R

[203]

[204]

[205]

[206] [207]

R

N C O

F

[195]

O

[194]

R O

[193]

may mediate the ability of (n-3) fatty acids to reduce ectopic fat and inflammatory mediators in obese Zucker rats, J. Nutr. 139 (2009) 1495–1501. F. Piscitelli, G. Carta, T. Bisogno, E. Murru, L. Cordeddu, K. Berge, S. Tandy, J.S. Cohn, M. Griinari, S. Banni, M.V. Di, Effect of dietary krill oil supplementation on the endocannabinoidome of metabolically relevant tissues from high fat-fed mice, Nutr. Metab. (Lond.) 8 (2011) 51. M.G.J. Balvers, K.C.M. Verhoeckx, P. Plastina, H.M. Wortelboer, J. Meijerink, R.F. Witkamp, Docosahexaenoic acid and eicosapentaenoic acid are converted by 3T3-L1 adipocytes to N-acyl ethanolamines with anti-inflammatory properties, Biochim. Biophys. Acta Mol. Cell Biol. Lipids 1801 (2010) 1107–1114. S. Tandy, R.W. Chung, E. Wat, A. Kamili, K. Berge, M. Griinari, J.S. Cohn, Dietary krill oil supplementation reduces hepatic steatosis, glycemia, and hypercholesterolemia in high-fat-fed mice, J. Agric. Food Chem. 57 (2009) 9339–9345. A. Ferramosca, A. Conte, L. Burri, K. Berge, N.F. De, A.M. Giudetti, V. Zara, A krill oil supplemented diet suppresses hepatic steatosis in high-fat fed rats, PLoS ONE 7 (2012) e38797. K.C. Maki, M.S. Reeves, M. Farmer, M. Griinari, K. Berge, H. Vik, R. Hubacher, T.M. Rains, Krill oil supplementation increases plasma concentrations of eicosapentaenoic and docosahexaenoic acids in overweight and obese men and women, Nutr. Res. 29 (2009) 609–615. J.P. Schuchardt, I. Schneider, H. Meyer, J. Neubronner, C. von Schacky, A. Hahn, Incorporation of EPA and DHA into plasma phospholipids in response to different omega-3 fatty acid formulations — a comparative bioavailability study of fish oil vs. krill oil, Lipids Health Dis. 10 (2011) 145. S.M. Ulven, B. Kirkhus, A. Lamglait, S. Basu, E. Elind, T. Haider, K. Berge, H. Vik, J.I. Pedersen, Metabolic effects of krill oil are essentially similar to those of fish oil but at lower dose of EPA and DHA, in healthy volunteers, Lipids 46 (2011) 37–46. M.V. Chakravarthy, I.J. Lodhi, L. Yin, R.R.V. Malapaka, H.E. Xu, J. Turk, C.F. Semenkowich, Identification of a physiologically relevant endogenous ligand for PPARalfa in liver, Cell 138 (2009) 476–488. J.M. Lee, Y.K. Lee, J.L. Mamrosh, S.A. Busby, P.R. Griffin, M.C. Pathak, E.A. Ortlund, D.D. Moore, A nuclear-receptor-dependent phosphatidylcholine pathway with antidiabetic effects, Nature 474 (2011) 506–510. J.P. Schuchardt, A. Hahn, Bioavailability of long-chain omega-3 fatty acids, Prostaglandins Leukot. Essent. Fat. Acids 89 (2013) 1–8. A.C. Hoper, W. Salma, A.M. Khalid, A.D. Hafstad, S.J. Sollie, J. Raa, T.S. Larsen, E. Aasum, Oil from the marine zooplankton Calanus finmarchicus improves the cardiometabolic phenotype of diet-induced obese mice, Br. J. Nutr. 110 (2013) 2186–2193. T. Cadoudal, E. Distel, S. Durant, F. Fouque, J.M. Blouin, M. Collinet, S. Bortoli, C. Forest, C. Benelli, Pyruvate dehydrogenase kinase 4: regulation by thiazolidinediones and implication in glyceroneogenesis in adipose tissue, Diabetes 57 (2008) 2272–2279. M. Bordicchia, D. Liu, E.Z. Amri, G. Ailhaud, P. Dessi-Fulgheri, C. Zhang, N. Takahashi, R. Sarzani, S. Collins, Cardiac natriuretic peptides act via p38 MAPK to induce the brown fat thermogenic program in mouse and human adipocytes, J. Clin. Invest. 122 (2012) 1022–1036. J.N. Fain, A. Kanu, S.W. Bahouth, G.S. Cowan Jr., M.L. Hiler, C.W. Leffler, Comparison of PGE2, prostacyclin and leptin release by human adipocytes versus explants of adipose tissue in primary culture, Prostaglandins Leukot. Essent. Fat. Acids 67 (2002) 467–473. B. Richelsen, E.F. Eriksen, H. Beck-Nielsen, O. Pedersen, Prostaglandin E2 receptor binding and action in human fat cells, J. Clin. Endocrinol. Metab. 59 (1984) 7–12. A. Vegiopoulos, K. Muller-Decker, D. Strzoda, I. Schmitt, E. Chichelnitskiy, A. Ostertag, D.M. Berriel, J. Rozman, A.M. Hrabe de, R.M. Nusing, C.W. Meyer, W. Wahli, M. Klingenspor, S. Herzig, Cyclooxygenase-2 controls energy homeostasis in mice by de novo recruitment of brown adipocytes, Science 328 (2010) 1158–1161. L. Madsen, L.M. Pedersen, H.H. Lillefosse, E. Fjaere, I. Bronstad, Q. Hao, R.K. Petersen, P. Hallenborg, T. Ma, R. De Matteis, P. Araujo, J. Mercader, M.L. Bonet, J.B. Hansen, B. Cannon, J. Nedergaard, J. Wang, S. Cinti, P. Voshol, S.O. Doskeland, K. Kristiansen, UCP1 induction during recruitment of brown adipocytes in white adipose tissue is dependent on cyclooxygenase activity, PLoS ONE 5 (2010) e11391. G. Vassaux, D. Gaillard, G. Ailhaud, R. Negrel, Prostacyclin is a specific effector of adipose cell-differentiation — its dual role as a camp-elevating and Ca2+-elevating agent, J. Biol. Chem. 267 (1992) 11092–11097. S. Virtue, M. Masoodi, V. Velagapudi, C.Y. Tan, M. Dale, T. Suorti, M. Slawik, M. Blount, K. Burling, M. Campbell, N. Eguchi, G. Medina-Gomez, J.K. Sethi, M. Oresic, Y. Urade, J.L. Griffin, A. Vidal-Puig, Lipocalin prostaglandin D synthase and PPARgamma2 coordinate to regulate carbohydrate and lipid metabolism in vivo, PLoS ONE 7 (2012) e39512. M.S. Draman, F. Grennan-Jones, L. Zhang, P.N. Taylor, T.K. Tun, J. McDermott, P. Moriarty, D. Morris, C. Lane, S. Sreenan, C. Dayan, M. Ludgate, Effects of prostaglandin f2alpha on adipocyte biology relevant to graves' orbitopathy, Thyroid 23 (2013) 1600–1608. Y. Taketani, R. Yamagishi, T. Fujishiro, M. Igarashi, R. Sakata, M. Aihara, Activation of the prostanoid FP receptor inhibits adipogenesis leading to deepening of the upper eyelid sulcus in prostaglandin-associated periorbitopathy, Invest. Ophthalmol. Vis. Sci. 55 (2014) 1269–1276. R.M. Catalioto, D. Gaillard, J. Maclouf, G. Ailhaud, R. Negrel, Autocrine control of adipose cell differentiation by prostacyclin and PGF2 alpha, Biochim. Biophys. Acta 1091 (1991) 364–369. Q. Ding, T. Mracek, P. Gonzalez-Muniesa, K. Kos, J. Wilding, P. Trayhurn, C. Bing, Identification of macrophage inhibitory cytokine-1 in adipose tissue and its secretion as an adipokine by human adipocytes, Endocrinology 150 (2009) 1688–1696.

P

[192]

T

[165] P.M. Kris-Etherton, W.S. Harris, L.J. Appel, Fish consumption, fish oil, omega-3 fatty acids and cardiovascular disease, Circulation 106 (2002) 2747–2757. [166] M. Yokoyama, H. Origasa, M. Matsuzaki, Y. Matsuzawa, Y. Saito, Y. Ishikawa, S. Oikawa, J. Sasaki, H. Hishida, H. Itakura, T. Kita, A. Kitabatake, N. Nakaya, T. Sakata, K. Shimada, K. Shirato, Effects of eicosapentaenoic acid on major coronary events in hypercholesterolaemic patients (JELIS): a randomised open-label, blinded endpoint analysis, Lancet 369 (2007) 1090–1098. [167] T.A. Mori, D.Q. Bao, V. Burke, I.B. Puddey, G.F. Watts, L.J. Beilin, Dietary fish as a major component of a weight-loss diet: effect on serum lipids, glucose, and insulin metabolism in overweight hypertensive subjects, Am. J. Clin. Nutr. 70 (1999) 817–825. [168] C. Couet, J. Delarue, P. Ritz, J.M. Antoine, F. Lamisse, Effect of dietary fish oil on body fat mass and basal fat oxidation in healthy adults, Int. J. Obes. 21 (1997) 637–643. [169] N. Bender, M. Portmann, Z. Heg, K. Hofmann, M. Zwahlen, M. Egger, Fish or n3-PUFA intake and body composition: a systematic review and meta-analysis, Obes Rev. 15 (2014) 657–665. [170] L. Spadaro, O. Magliocco, D. Spampinato, S. Piro, C. Oliveri, C. Alagona, G. Papa, A.M. Rabuazzo, F. Purrello, Effects of n−3 polyunsaturated fatty acids in subjects with nonalcoholic fatty liver disease, Dig. Liver Dis. 40 (2008) 194–199. [171] P. Flachs, M. Rossmeisl, J. Kopecky, The effect of n− 3 fatty acids on glucose homeostasis and insulin sensitivity, Physiol. Res. 63 (Suppl. 1) (2014) S93–S118. [172] L.H. Storlien, E.W. Kraegen, D.J. Chisholm, G.L. Ford, D.G. Bruce, W.S. Pascoe, Fish oil prevents insulin resistance induced by high-fat feeding in rats, Science 237 (1987) 885–888. [173] J. Ruzickova, M. Rossmeisl, T. Prazak, P. Flachs, J. Sponarova, M. Vecka, E. Tvrzicka, M. Bryhn, J. Kopecky, Omega-3 PUFA of marine origin limit diet-induced obesity in mice by reducing cellularity of adipose tissue, Lipids 39 (2004) 1177–1185. [174] M. Rossmeisl, T. Jelenik, Z. Jilkova, K. Slamova, V. Kus, M. Hensler, D. Medrikova, C. Povysil, P. Flachs, V. Mohamed-Ali, M. Bryhn, K. Berge, A.K. Holmeide, J. Kopecky, Prevention and reversal of obesity and glucose intolerance in mice by DHA derivatives, Obesity 17 (2009) 1023–1031. [175] O. Kuda, T. Jelenik, Z. Jilkova, P. Flachs, M. Rossmeisl, M. Hensler, L. Kazdova, N. Ogston, M. Baranowski, J. Gorski, P. Janovska, V. Kus, J. Polak, V. Mohamed-Ali, R. Burcelin, S. Cinti, M. Bryhn, J. Kopecky, n−3 fatty acids and rosiglitazone improve insulin sensitivity through additive stimulatory effects on muscle glycogen synthesis in mice fed high-fat diet, Diabetologia 52 (2009) 941–951. [176] T. Jelenik, M. Rossmeisl, O. Kuda, Z.M. Jilkova, D. Medrikova, V. Kus, M. Hensler, P. Janovska, I. Miksik, M. Baranowski, J. Gorski, S. Hebrard, T.E. Jensen, P. Flachs, S. Hawley, B. Viollet, J. Kopecky, AMP-activated protein kinase {alpha}2 subunit is required for the preservation of hepatic insulin sensitivity by n−3 polyunsaturated fatty acids, Diabetes 59 (2010) 2737–2746. [177] J. Todoric, M. Loffler, J. Huber, M. Bilban, M. Reimers, A. Kadl, M. Zeyda, W. Waldhausl, T.M. Stulnig, Adipose tissue inflammation induced by high-fat diet in obese diabetic mice is prevented by n−3 polyunsaturated fatty acids, Diabetologia 49 (2006) 2109–2119. [178] C.E. Friedberg, M.J. Janssen, R.J. Heine, D.E. Grobbee, Fish oil and glycemic control in diabetes. A meta-analysis, Diabetes Care 21 (1998) 494–500. [179] D. Kromhout, J.M. Geleijnse, G.J. de, 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 arrhythmiarelated events, and fatal myocardial infarction in postmyocardial infarction patients with diabetes, Diabetes Care 34 (2011) 2515–2520. [180] P. Flachs, M. Rossmeisl, M. Bryhn, J. Kopecky, Cellular and molecular effects of n−3 polyunsaturated fatty acids on adipose tissue biology and metabolism, Clin. Sci. 116 (2009) 1–16. [181] G. Bannenberg, C.N. Serhan, Specialized pro-resolving lipid mediators in the inflammatory response: an update, Biochim. Biophys. Acta 1801 (2010) 1260–1273. [182] P.C. Calder, Fatty acids and inflammation: the cutting edge between food and pharma, Eur. J. Pharmacol. 668 (2011) S50–S58. [183] P. Flachs, V. Mohamed-Ali, O. Horakova, M. Rossmeisl, M.J. Hosseinzadeh-Attar, M. Hensler, J. Ruzickova, J. Kopecky, Polyunsaturated fatty acids of marine origin induce adiponectin in mice fed high-fat diet, Diabetologia 49 (2006) 394–397. [184] S. Neschen, K. Morino, J.C. Rossbacher, R.L. Pongratz, G.W. Cline, S. Sono, M. Gillum, G.I. Shulman, Fish oil regulates adiponectin secretion by a peroxisome proliferatoractivated receptor-gamma-dependent mechanism in mice, Diabetes 55 (2006) 924–928. [185] J.H. Wu, L.E. Cahill, D. Mozaffarian, Effect of fish oil on circulating adiponectin: a systematic review and meta-analysis of randomized controlled trials, J. Clin. Endocrinol. Metab. 98 (2013) 2451–2459. [186] P. Flachs, O. Matejkova, P. Pecina, N. Frassen, P. Brauner, M. Rossmeisl, J. Keijer, J. Houstek, J. Kopecky, Polyunsaturated fatty acids of marine origin induce mitochondrial biogenesis and beta-oxidation in white fat, Diabetologia 48 (2005) 2365–2375. [187] T. Yamauchi, J. Kamon, Y. Minokoshi, Y. Ito, H. Waki, S. Uchida, S. Yamashita, M. Noda, S. Kita, K. Ueki, K. Eto, Y. Akanuma, P. Froguel, F. Foufelle, P. Ferre, D. Carling, S. Kimura, R. Nagai, B.B. Kahn, T. Kadowaki, Adiponectin stimulates glucose utilization and fatty-acid oxidation by activating AMP-activated protein kinase1, Nat. Med. 8 (2002) 1288–1295. [188] C. Lipina, W. Rastedt, A.J. Irving, H.S. Hundal, New vistas for treatment of obesity and diabetes? Endocannabinoid signalling and metabolism in the modulation of energy balance, Bioassays 34 (2012) 681–691. [189] H. Butler, M. Korbonits, Cannabinoids for clinicians: the rise and fall of the cannabinoid antagonists, Eur. J. Endocrinol. 161 (2009) 655–662. [190] J. Kim, Y. Li, B.A. Watkins, Fat to treat fat: emerging relationship between dietary PUFA, endocannabinoids, and obesity, Prostaglandins Other Lipid Mediat. 104–105 (2013) 32–41. [191] B. Batetta, M. Griinari, G. Carta, E. Murru, A. Ligresti, L. Cordeddu, E. Giordano, F. Sanna, T. Bisogno, S. Uda, M. Collu, I. Bruheim, M.V. Di, S. Banni, Endocannabinoids

[208]

[209]

[210]

U

1419 1420 1421 1422 1423 1424 1425 1426 1427 1428 1429 1430 1431 1432 1433 1434 1435 1436 1437 1438 1439 1440 1441 1442 1443 1444 1445 1446 1447 1448 1449 1450 1451 1452 1453 1454 1455 1456 1457 1458 1459 1460 1461 1462 1463 1464 1465 1466 1467 1468 1469 1470 1471 1472 1473 1474 1475 1476 1477 1478 1479 1480 1481 1482 1483 1484 1485 1486 1487 1488 1489 1490 1491 1492 1493 1494 1495 1496 1497 1498 1499 1500 1501 1502 1503 1504

15

[211]

[212]

[213]

[214]

Please cite this article as: M. Masoodi, et al., Lipid signaling in adipose tissue: Connecting inflammation & metabolism, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbalip.2014.09.023

1505 1506 1507 1508 1509 1510 1511 1512 1513 1514 1515 1516 1517 1518 1519 1520 1521 1522 1523 1524 1525 1526 1527 1528 1529 1530 1531 1532 1533 1534 1535 1536 1537 1538 1539 1540 1541 1542 1543 1544 1545 1546 1547 1548 1549 1550 1551 1552 1553 1554 1555 1556 1557 1558 1559 1560 1561 1562 1563 1564 1565 1566 1567 1568 1569 1570 1571 1572 1573 1574 1575 1576 1577 1578 1579 1580 1581 1582 1583 1584 1585 1586 1587 1588 1589 1590

[223] P.J. White, M. Arita, R. Taguchi, J.X. Kang, A. Marette, 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 59 (2010) 3066–3073. [224] J.M. Schwab, N. Chiang, M. Arita, C.N. Serhan, Resolvin E1 and protectin D1 activate inflammation-resolution programmes, Nature 447 (2007) 869–874. [225] M.G. Balvers, K.C. Verhoeckx, S. Bijlsma, C.M. Rubingh, J. Meijerink, H.M. Wortelboer, R.F. Witkamp, Fish oil and inflammatory status alter the n−3 to n −6 balance of the endocannabinoid and oxylipin metabolomes in mouse plasma and tissues, Metabolomics 8 (2012) 1130–1147. [226] E. Borgeson, F.C. McGillicuddy, K.A. Harford, N. Corrigan, D.F. Higgins, P. Maderna, H.M. Roche, C. Godson, Lipoxin A4 attenuates adipose inflammation, FASEB J. 26 (2012) 4287–4294. [227] C. Gonzalez-Yanes, A. Serrano, F.J. Bermudez-Silva, M. Hernandez-Dominguez, M.A. Paez-Ochoa, F. Rodriguez de Fonseca, V. Sanchez-Margalet, Oleylethanolamide impairs glucose tolerance and inhibits insulin-stimulated glucose uptake in rat adipocytes through p38 and JNK MAPK pathways, Am. J. Physiol. Endocrinol. Metab. 289 (2005) E923–E929. [228] M. Guzman, J. Lo Verme, J. Fu, F. Oveisi, C. Blazquez, D. Piomelli, Oleoylethanolamide stimulates lipolysis by activating the nuclear receptor peroxisome proliferator-activated receptor alpha (PPAR-alpha), J. Biol. Chem. 279 (2004) 27849–27854. [229] J. Suarez, P. Rivera, S. Arrabal, A. Crespillo, A. Serrano, E. Baixeras, F.J. Pavon, M. Cifuentes, R. Nogueiras, J. Ballesteros, C. Dieguez, F. Rodriguez de Fonseca, Oleoylethanolamide enhances beta-adrenergic-mediated thermogenesis and white-to-brown adipocyte phenotype in epididymal white adipose tissue in rat, Dis. Model Mech. 7 (2014) 129–141.

F

[215] D. Sinha, S. Addya, E. Murer, G. Boden, 15-Deoxy-delta(12,14) prostaglandin J2: a putative endogenous promoter of adipogenesis suppresses the ob gene, Metabolism 48 (1999) 786–791. [216] B.M. Forman, P. Tontonoz, J. Chen, R.P. Brun, B.M. Spiegelman, R.M. Evans, 15-Deoxy-delta12,14-prostaglandin J2 is a ligand for the adipocyte determination factor PPARgama, Cell 83 (1995) 803–812. [217] A. Guilherme, G.J. Tesz, K.V. Guntur, M.P. Czech, Tumor necrosis factor-alpha induces caspase-mediated cleavage of peroxisome proliferator-activated receptor gamma in adipocytes, J. Biol. Chem. 284 (2009) 17082–17091. [218] I. Mothe-Satney, C. Filloux, H. Amghar, C. Pons, V. Bourlier, J. Galitzky, P.A. Grimaldi, C.C. Feral, A. Bouloumie, O.E. Van, J.G. Neels, Adipocytes secrete leukotrienes: contribution to obesity-associated inflammation and insulin resistance in mice, Diabetes 61 (2012) 2311–2319. [219] B.K. Cole, N.S. Kuhn, S.M. Green-Mitchell, K.A. Leone, R.M. Raab, J.L. Nadler, S.K. Chakrabarti, 12/15-Lipoxygenase signaling in the endoplasmic reticulum stress response, Am. J. Physiol. Endocrinol. Metab. 302 (2012) E654–E665. [220] A.P. Burgess, L. Vanella, L. Bellner, K. Gotlinger, J.R. Falck, N.G. Abraham, M.L. Schwartzman, A. Kappas, Heme oxygenase (HO-1) rescue of adipocyte dysfunction in HO-2 deficient mice via recruitment of epoxyeicosatrienoic acids (EETs) and adiponectin, Cell. Physiol. Biochem. 29 (2012) 99–110. [221] H. Yamada, E. Oshiro, S. Kikuchi, M. Hakozaki, H. Takahashi, K. Kimura, Hydroxyeicosapentaenoic acids from the Pacific krill show high ligand activities for PPARs, J. Lipid Res. 55 (2014) 895–904. [222] A. Gonzalez-Periz, R. Horrillo, N. Ferre, K. Gronert, B. Dong, E. Moran-Salvador, E. Titos, M. Martinez-Clemente, M. Lopez-Parra, V. Arroyo, J. Claria, Obesity-induced insulin resistance and hepatic steatosis are alleviated by omega-3 fatty acids: a role for resolvins and protectins, FASEB J. 23 (2009) 1946–1957.

O

1591 1592 1593 1594 1595 1596 1597 1598 1599 1600 1601 1602 1603 1604 1605 1606 1607 1608 1609 1610 1611 1612 1613 1614 1615 1616 1617

M. Masoodi et al. / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

R O

16

U

N

C

O

R

R

E

C

T

E

D

P

1645

Please cite this article as: M. Masoodi, et al., Lipid signaling in adipose tissue: Connecting inflammation & metabolism, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbalip.2014.09.023

1618 1619 1620 1621 1622 1623 1624 1625 1626 1627 1628 1629 1630 1631 1632 1633 1634 1635 1636 1637 1638 1639 1640 1641 1642 1643 1644

Lipid signaling in adipose tissue: Connecting inflammation & metabolism.

Obesity-associated low-grade inflammation of white adipose tissue (WAT) contributes to development of insulin resistance and other disorders. Accumula...
1MB Sizes 2 Downloads 9 Views