THE ROLE OF BOTANICAL OILS ENRICHED IN FADS2-DERIVED N-3 VS. N-6

POLYUNSATURATED FATTY ACIDS IN PREVENTION OF ATHEROSCLEROSIS

BY

SWAPNIL VIJAY SHEWALE

A Dissertation Submitted to the Graduate Faculty of

WAKE FOREST UNIVERSITY SCHOOL OF ARTS AND SCIENCES

in Partial Fulfillment of the Requirements

for the Degree of

DOCTOR OF PHILOSOPHY

Physiology and Pharmacology

AUGUST 2015

Winston Salem, North Carolina

Approved By:

John S. Parks, Ph.D., Advisor

Martha Alexander-Miller, Ph.D., Chairman

Michael C. Seeds, Ph.D.

Kylie Kavanagh, Ph.D.

Ann Tallant, Ph.D

i

DEDICATION

This dissertation is dedicated to my very loving and sweet grandmothers: Manjulabai

Kashinath Shewale (paternal) from Takali, India and Janakibai Ramchandra Hire (maternal) from Malegaon, India. My maternal and paternal grandparents were particularly keen on providing education to their children despite all odds. My largely joint family of 20 some uncles and aunts, parents: Vidhya and Vijay Shewale and in-laws: Jayashri and Sharad Sonawane have encouraged all of us, including me and my loving and beautiful wife- Poonam, my brother-

Himanshu, my sister- Swati and ~30-50 some cousins to embrace similar ideologies and to pursue our interests. My grandmothers to me are the silent sufferers and reformers of our lives.

ii

ACKNOWLEDGEMENTS

First and foremost, I would like to thank Dr. John S. Parks for his guidance, support, patience and encouragement throughout the last five years of training. I can recall many-many occasions when I may have made mistakes, asked redundant questions or spoken “too-much, too-soon”. However, Dr. Parks has always been patient with me “to get it together” and to progress as an advanced PhD candidate. It’s his originally well thought-out Botanical grant, collaborations with other established PI’s, and allowing me to navigate my way through has made this dissertation possible. I would also like to thank Dr. Rudel for his rare but extremely insightful comments, advice and for putting the original Jornal Club, Seminar and Guest-speaker series together. In the Lipid Sciences Department, there always was something to learn from every presenter every week. I am extremely fortunate and grateful to (family) members of the

Parks lab and the Lipid Sciences Department. Everyone on the “Lipid” floor welcomed, helped and taught me everything I could learn during my PhD here.

I can’t forget to thank other members from the Lipid group: Elena, Kaiser, Martha,

Matt, Xuewei, Amanda-Mark, and Xin. Elena made everything so easy like every mom does and also saved my life once. Amanda-Mark allowed me to stay in their vacant house when I was temporarily homeless. I would also like to specially thank Dr. Miller, for her guidance with design, execution and interpretation of flow-cytometry related experiments, Dr. Seeds for insights, and encouragement and Dr.’s Tallant and Kavanagh for their support, critique and evaluation of my progress as a researcher. I would also like to thank my previous mentor Dr.

Mariana Morris for “pushing” me to accept the offer from Wake. The Department of Physiology and Pharmacology not only accepted me but encouraged me to foster, explore and contribute in scientific, teaching, business development and social areas as their brand ambassador. I am eternally grateful for this experience and journey which has allowed me to look forward to my future with optimism.

iii

TABLE OF CONTENTS

LIST OF FIGURES AND TABLES………………………………...……….…….…...... V

LIST OF ABBREVATIONS…………………………………………………..…….………..... VIII

ABSTRACT...... XI

CHAPTER:

I. INTRODUCTION...... 1

II. BOTANICAL OILS ENRICHED IN N-6 AND N-3 FATTY ACID PRODUCTS OF

FADS2 ARE EQUALLY EFFECTIVE IN PREVENTING ATHEROSCLROSIS AND

HEPATOSTEATOSIS IN MICE ...... 52

III. IN VIVO ACTIVATION OF LEUKOCYTE GPR120 BY POLYUNSATURATED FATTY

ACIDS HAS MINIMAL IMPACT ON ATHEROSCLROSIS IN LDLrKO MICE...... 102

IV. DISCUSSION...... 166

CURRICULUM VITAE...... 185

iv

LIST OF FIGURES AND TABLES

CHAPTER I:

Figure 1. Outline of lipoprotein transport and metabolism ……………………………………....9

Figure 2. Outline of biosynthesis and metabolic pathways for polyunsaturated fatty acids …24

Table 1. Classification of circulating lipoproteins………………………………………………….5

Table 2. Major important to lipoprotein metabolism ………………………………….10

Table 3. Mouse models of atheroscleroses ………………………………………………………20

CHAPTER II:

Figure 1. Body weight gain and terminal liver/body weight ratios……………………………….91

Figure 2. RBC fatty acid (FA) composition. ……………………………………………………....92

Figure 3. Percentage fatty acid composition of plasma and liver lipids………………………....93

Figure 4. Plasma lipid concentrations……………………………………………………………...94

Figure 5. Plasma lipoprotein cholesterol distribution……………………………………………...95

Figure 6. Hepatic VLDL-TG secretion rate. ……………………………………………………….96

Figure 7. Hepatic response to atherogenic diets. ………………………………………………...97

Figure 8. Aortic atherosclerosis. …………………………………………………………………...98

Figure 9. Mouse atherosclerotic plaque oxidized cholesteryl ester analysis. ……………….....99

Figure 10. LPS-stimulated eicosanoid release from peritoneal macrophages. ……………….100

Figure 11. Macrophage inflammation, foam cell formation, and chemotaxis. ………………...101

Table 1. Atherogenic Diet (AD) percentage fatty acid composition (%FA) and percentage total energy equivalence (% EE) of individual fatty acid…………………………………………….….89

v

CHAPTER III:

Figure 1. L-GPR120 does not affect plasma lipids…………………………………………...131

Figure 2. L-GPR120 does not affect plasma lipoprotein cholesterol distribution…………..132

Figure 3. Effect of L-GPR120 on hepatic M1-M2 and lipogenic , neutral lipid content and inflammatory cytokines. …………………………………………………………………....133

Figure 4. Effect of L-GPR120 on neutrophilia and monocytosis. …………………………...134

Figure 5. Effect of L-GPR120 on monocyte recruitment to aortic root intima………………135

Figure 6. Histological quantification of aortic root atherosclerotic lesions…………………..136

Figure 7. Effect of L-GPR120 on aortic cholesterol content. ………………………………..137

Supplementary Table 1: Atherogenic Diet (AD) percentage fatty acid composition (%FA) and percentage total energy equivalence (% EE) of individual fatty acid……………………….145

Supplementary Figure 1. Bone marrow transplantation (BMT) efficiency. ………………...152

Supplementary Figure 2. Body weight gain and terminal body weights…………………….153

Supplementary Figure 3. Terminal organ/body weight ratios ……………………………….154

Supplementary Figure 4. RBC fatty acid (FA) composition…………………………………..155

Supplementary Figure 5. Liver Histology……………………………………………………….156

Supplementary Figure 6. Effect of L-GPR120 on circulating monocytosis. ………………...157

Supplementary Figure 7. Monocyte recruitment experiment. ………………………………. 158

Supplementary Figure 8. Representative aortic root atherosclerotic lesions stained for Oil-red-O. …………………………………………………………………………….159

Supplementary Figure 9. Representative aortic root atherosclerotic lesions stained for CD68/ macrophages. ……………………………………………………………….160

vi

Supplementary Figure 10. Representative aortic root atherosclerotic lesions stained for Sirius red/ collagen…………………………………………………………..………161

Supplementary Figure 11. Representative aortic root atherosclerotic lesions stained for CD11c/ dendritic cells……………………………………………………………….162

Supplementary Figure 12. Representative aortic root atherosclerotic lesions stained for cleaved-caspase-3/ apoptotic cells………………………………………………..163

CHAPTER IV:

Figure 1. Summary of potential pathways for decreased inflammation by n-6 and n-3 fatty acids ……………………………………………..…………….…………..175

vii

LIST OF ABBREVIATIONS

13-HODE 13-hydroxyoctadecadienoic acid

15-HETE 15-hydroxyeicosatetraenoic acid

15d-PGJ2 15-Deoxy-delta 12,14-prostaglandin J2

ABCA1 ATP-binding cassette transporter A1

AP-1 activator protein 1 apoE apolipoprotein E apoEKO apolipoprotein E knockout

Arg-1 arginase-1

CCR2 C-C chemokine receptor type 2

CCR7 C-C chemokine receptor type 7

CD cluster of differentiation

CHD coronary heart disease

COX

CVD cardiovascular disease

CX3CR1: CX3C chemokine receptor 1

DART The Diet and Reinfarction Trial

DGLA dihomo-gamma-linolenic acid

DHA docosahexaenoic acid

EPA eicosapentaenoic acid

GISSI Gruppo Italiano per lo Studio della Sopravvivenza nell’Infarto Miocardico

GLA gamma-linolenic acid

Gr-1 granulocyte differentiation antigen 1

HDL high density lipoprotein

viii

IFN-γ interferon gamma

IL-1 interleukin 1 iNOS inducible

JELIS Japan EPA Lipid Intervention Study

LP lipoprotein

LDL low density lipoprotein

LDLr low density lipoprotein receptor

LDLrKO low density lipoprotein receptor knockout

LO lipoxygenase

LP lipoprotein

LPS lipopolysaccharide

LXR liver X receptor

Ly6C lymphocyte antigen 6 C

Ly6G lymphocyte antigen 6 G mmLDL minimally modified LDL

MMP matrix metalloproteinase

MR mannose receptor

NCoR nuclear receptor corepressor

NF-κB nuclear factor kappa-light-chain-enhancer of activated B cells oxLDL oxidized LDL

PIAS1 protein inhibitor of activated STAT1

PPAR peroxisome-proliferator activated receptor

SDA stearidonic acid

SMC smooth muscle cell

STAT signal transducers and activators of transcription

Th T helper lymphocyte

ix

TNF-α tumor necrosis factor alpha

VCAM-1 vascular cell adhesion molecular 1

VLDL very low density lipoprotein

x

ABSTRACT

Swapnil V. Shewale

THE ROLE OF BOTANICAL OILS ENRICHED IN FADS2-DERIVED N-3 VS. N-6 POLYUNSATURATED FATTY ACIDS IN PREVENTION OF ATHEROSCLEROSIS

Dissertation under the direction of John S. Parks, Ph.D., Professor, Internal Medicine/Section on Molecular Medicine

Background: Dietary polyunsaturated fatty acids (PUFAs) reduce atherosclerosis in animal models and humans relative to dietary saturated and monounsaturated fatty acids. Although some of the atheroprotection of dietary PUFAs is due to plasma lipid lowering, in vivo conversion of 18 carbon PUFAs through the rate-limiting -2 (FADS2, delta-6 desaturase) step of fatty acid desaturation and elongation results in 18 and ≥ 20 carbon PUFAs that are substrates for pro-inflammatory and anti-inflammatory eicosanoid production, which affect atherosclerosis progression and inflammation. We previously showed that an atherogenic diet containing echium oil (EO), which is relatively enriched in stearidonic acid (18:4 n-3), the immediate product of FADS2-mediated desaturation of 18:3 n-3, effectively enriches plasma and tissue lipids in the anti-inflammatory PUFA 20:5 n-3 and was as atheroprotective as dietary fish oil (FO) compared to palm oil (PO), which is enriched in saturated and monounsaturated fatty acids. However, whether a similar strategy of dietary enrichment in FADS-2 n-6 products would lead to atheroprotective is unknown. To address this gap in knowledge, we tested the hypothesis that dietary borage oil (BO), enriched in the

FADS-2 product 18:3 n-6, would not be as atheroprotective as EO, due to in vivo conversion to 20:4 n-6, a pro-inflammatory eicosanoid precursor. We also investigated the role an anti- inflammatory protein G-protein coupled receptor 120(GPR120) that is activated by PUFAs in atheroprotection and hypothesized that dietary n-3 PUFAs would lead to greater activation of

GPR120 and less inflammation than n-6 PUFAs.

xi

Two studies were performed using: 1) LDL receptor knockout (LDLrKO) mice, or 2) irradiated

LDLrKO mice transplanted with wild type (WT) or GPR120 knockout (KO) bone marrow. Mice were fed one of four atherogenic diets containing 0.2% cholesterol and 10% fat as PO + an additional 10% of fat as PO, FO, EO or BO for 8-16 weeks. Measurements of lipid metabolism, atherosclerosis, and inflammation were made to test out hypotheses.

In study 1, mice fed BO, EO and FO vs. PO had significantly lower plasma total and VLDL cholesterol concentrations, hepatic neutral lipid content and inflammation, aortic CE content, aortic root intimal area and macrophage content, and peritoneal macrophage inflammation,

CE content, and ex vivo chemotaxis. We conclude that botanical oils enriched in 18:3 n-6 and

18:4 n-3 PUFAs beyond the rate-limiting FADS2 are equally effective in preventing atherosclerosis and hepatosteatosis compared to saturated/monounsaturated fat due to cellular enrichment of ≥20 PUFAs, reduced plasma VLDL, and attenuated macrophage inflammation.

In study 2, mice fed BO, EO and FO vs. PO had significantly reduced plasma cholesterol, triglycerides, VLDL cholesterol, hepatic steatosis, and atherosclerosis that were equivalent for mice transplanted with WT and GPR120 KO mouse bone marrow, demonstrating that leukocyte

GPR120 expression did not affect these outcomes. In BO, EO and FO, but not PO-fed mice, lack of leukocyte GPR120 resulted in neutrophilia, pro-inflammatory Ly6Chi monocytosis, increased monocyte recruitment into aortic roots, and increased hepatic inflammatory expression. We conclude that leukocyte GRP120 expression has minimal effect on dietary

PUFA-induced plasma lipid/lipoprotein reduction and atheroprotection, and that there is no distinction between n-3 vs. n-6 PUFAs in activating anti-inflammatory effects of leukocyte

GPR120 in vivo.

xii

CHAPTER I

INTRODUCTION

Swapnil V. Shewale

Swapnil V. Shewale prepared this chapter. Dr. John S. Parks acted in an advisory and editorial

capacity

1

1. Cardiovascular diseases and polyunsaturated fatty acids (PUFAs)

1.1 Prevalence of Cardiovascular diseases (CVD) and dietary recommendations: CVD, including coronary heart disease (CHD) and stroke, is the leading cause of mortality in North

America affecting about one third of adults and was estimated to cost $320 billion dollars in direct and indirect costs in year 20111. The hallmarks of CVD are elevated plasma cholesterol and accumulation of lipid-laden macrophage foam cells in the subendothelial space (i.e., intima) of arteries 2. Intimal monocyte recruitment followed by differentiation into macrophages that scavenge modified apolipoprotein B containing (apoB) lipoproteins (LP) and apoB LP- proteoglycan complexes leads to foam cell formation and eventually results in atherosclerosis 2

3. Progressive obstruction of the vessel lumen due to thickening of the intimal area, ulceration, thrombosis, or embolization of a plaque results in myocardial infarction or stroke2. One of the first approaches to CVD treatment is to: 1) reduce saturated fat and cholesterol intake, and 2) replace saturated fatty acids (SFA) and monounsaturated fatty acids (MUFA) with n-3 polyunsaturated fatty acids (PUFA) from fish oil (FO) and/or n-6 PUFA from vegetable oils4. As a set of goals, the AHA recommends consumption of <300 mg cholesterol/day, ~25-35% of total calories from fat, and up to 10% of total calories from PUFAs 5, 6. Understanding the mechanism by which different types of dietary fatty acids affect plasma lipid-lipoprotein metabolism, macrophage function, and atherosclerosis development is important for public health.

2. Dietary n-3 and n-6 PUFAs are atheroprotective

2.1 Epidemiological evidence for atheroprotection by n-3 PUFA: Dyerberg et al in their landmark observations in the 1970’s noted that death from CHD in Greenland Eskimos was significantly lower compared to their Danish counterparts. These effects were attributed to their diets and later on fatty acid analysis of Eskimo diets revealed that, although both Greenlanders and their Danish counterparts consumed a high fat (~40%) diet, Greenlanders consumed fewer

2

SFAs (22% vs. 55%), and increased MUFAs (~57% vs. ~35%) and PUFAs, specially n-3

PUFAs (~13.7% vs. ~2.1%). The reduced CHD events were attributed to anti-thrombotic effects of n-3 PUFAs, especially eicosapentaenoic acid (EPA, C20:5 n-3) and docosahexaenoic acid

(DHA, C22:6 n-3) obtained from fatty fish 7, 8. Other pleotropic benefits have been attributed to diets enriched in fatty fish, including plasma TG lowering and reduced inflammation to name a few 6, 9, 10. Since then, atheroprotective effects of replacing dietary SFAs and MUFAs by n-3

PUFA containing diets has been demonstrated in humans and in several animal models, including African green monkeys and transgenic mice 11-16. For example, in a primary endpoint prevention trial named Effects of eicosapentaenoic acid on major coronary events in hypercholesterolemic patients (JELIS), EPA consumption (1800 mg/day, 5 years) resulted in a

19% relative reduction in major coronary events (primary endpoint) in hypercholesterolemic subjects17. Other secondary endpoint prevention trials including, The Diet and Reinfarction Trial

(DART) 18, GISSI-Prevention Study19, Nurses’ Health Study20 and Physicians’ Health21 Study have shown a protective effects of FO consumption with regard to risk and/or outcomes for CVD including CHD and myocardial infarction.

2.2 Epidemiological evidence for atheroprotection by n-6 PUFA: Although much attention has been given to n-3 PUFAs, n-6 PUFAs are also atheroprotective in mouse models of atherosclerosis, non-human primates, and in human populations22-25. Some of the atheroprotection is likely due to plasma lipid lowering, but there may also be an anti- inflammatory role for n-6 PUFAs through the generation of prostaglandin E1 (PGE1) from dihomo-γ-linolenic acid (DGLA, 20:3 n-6), which is a potent inhibitor of thromboxane A2 (TXA2) formation and inhibits leukocyte adherence to endothelial cells26, 27. In addition, a recent meta- analysis of 13 cohort studies involving 310,602 individuals and 12,479 coronary heart disease events revealed an inverse association between dietary linoleic acid (LA, 18:2 n-6) intake and

3 coronary heart disease risk, such that, a 5% increase in energy intake from LA was associated with a 10% and 13% lower risk of coronary heart disease events and deaths, respectively 28.

3. Overview of lipid and lipoprotein metabolism

3.1. Circulating lipids and lipoproteins: Lipids in circulation are transported as hydrophilic, spherical structures called lipoproteins (LP). LP are composed of: a) amphipathic apolipoproteins on the surface that assist in providing a spherical structure to LP and can act as cofactors and ligands for lipid-processing enzymes, b) a neutral lipid core, consisting of triglycerides (TG) and cholesteryl esters (CE), and c) a monolayer of polar lipid, consisting of free cholesterol (FC) and phospholipids (PL), mainly phosphatidylcholine (PC) and sphingomyelin lipids29. Circulating lipoproteins can be categorized based on a specific apolipoprotein associated with the given LP, their electrophoretic mobility, size and/or density

(ratio of lipid to protein) as outlined in Table 1 30.

4

Table 1. General classification of circulating lipoproteins

Classification Chylomicron VLDL LDL HDL

Scheme

Density (g/ml) <0.93 0.95-1.006 1.019-1.063 1.063-1.21

Size (nm) 75-1200 30-80 18-25 7-12

Major cargo Triglyceride Triglyceride Cholesterol Cholesterol,

Phospholipid

Apolipoprotein/s ApoB48, ApoCII, ApoB100, ApoE, ApoB100, ApoA I,

ApoCIII, ApoCII, ApoCI, ApoCII, ApoAII, ApoCIII, ApoE, ApoE, ApoAIV ApoCIII ApoCII, ApoCIII, ApoAI, ApoAII Apo(a) ApoE

Electrophoretic Origin/ α2 Pre-β β α mobility

Apo, apolipoprotein.

Intermediate density lipoproteins (IDL, d=1.006-1.019 g/ml) are VLDL remnants resulting from

VLDL TG hydrolysis by lipoprotein lipase (LPL) and are usually not detectable in the circulation due to rapid removal from plasma or conversion to LDL31.

5

3.2. Overview of lipoprotein metabolism

3.2.1. Chylomicrons: Approximately 95% of dietary fat is TG and the remainder is free fatty acid (FFA), PL, CE and fat soluble vitamins. In the small intestinal lumen, several lipases, including carboxyl ester lipase / cholesterol esterase, intestinal pancreatic triglyceride lipase and group 1B phospholipase A2, hydrolyze lipids into absorbable FC, lysophospholipids, glycerol and FFAs 32, 33. Digested lipids are then solubilized by bile acids into micelles for absorption into intestinal epithelial cells (i.e., enterocytes)34. Once absorbed into enterocytes, FFAs are re- esterified into TG35, FC is esterified to CE by sterol O-acyltransferase 2 (SOAT2, also known as acyl-coenzyme A:cholesterol acyltransferase 1 (ACAT2))36, 37 and packaged with apoB-48 into chylomicrons (CM), the largest lipoprotein. ApoB-48 consists of the amino terminal 2152 amino acids of apoB-100, resulting from a unique APOBEC-1 mRNA-editing mechanism38. In humans, apoB-48 is exclusively synthesized in intestine, whereas human liver only produces only apoB-

100 39. Chylomicrons, due their large size (75-1,200 nm) (Table 1), enter the circulation through the lymphatic system40. Once chylomicrons are released in the circulation, lipoprotein lipase

(LPL) hydrolyses TG in chylomicrons, releasing glycerol and FFA41 for energy storage (adipose tissue) or utilization (muscle) (Figure 1, Table 2)).

3.2.2. Very low density lipoprotein (VLDL): VLDL is a liver derived LP enriched in TG (Table

1) and functions to distribute FFFAs to different tissues through hydrolysis of TG by LPL. In a fasting state, intracellular TG and CE is mobilized (hydrolyzed and re-esterified) from lipid droplets, assembled with apoB-100 and secreted as VLDL particles. In the ER, prior to incorporation in VLDL, FC is esterified to CE via SOAT2 36, whereas, FFA are esterified to TG and PL in the sn-glycerol-3-phosphate pathway via sequential reactions involving acyl- transferases42. Assembly of apoB-100 into VLDL involves sequential steps where an ER- localized , microsomal triglyceride transfer protein directs the cotranslational lipidation of apoB-100 in the rough ER membrane with phospholipid and TG to form precursor VLDL

6 particles with diameters of ∼10–20 nm. A second step of particle assembly involves movement of cytosolic lipid into the ER to form luminal lipid droplets, which then fuse with precursor VLDL particles, forming mature VLDL that are secreted in circulation 43, 44.

3.2.3. Low density lipoprotein (LDL): Gradual VLDL TG lipolysis gives rise to IDL and ultimately to LDL. LDL is the major plasma cholesteryl ester-transporting lipoprotein (Table 1).

LDL is catabolized via binding to the LDL receptor (LDLr) on the cell surface of the liver and extrahepatic tissues45 (Figure 1).

3.2.4. Intravascular metabolism of apoB containing LP by LPL:Several genes have been identified whose loss of function results in severe plasma hypertriglyceridemia. These include

LPL, APOC2, APOA5, lipase maturation factor 1, and glycoprotein glycosylphosphatidylinositol

(GPI)-anchored high-density lipoprotein (HDL)-binding protein 1 (GPIHBP1)46. LPL is bound to heparin sulphate proteoglycans on endothelial cell surface in peripheral tissues. Once chylomicrons and VLDLs are released in the circulation, LPL hydrolyse TG, releasing glycerol and FFA41 for energy storage (adipose tissue) or utilization (muscle) (Figure 1). APOC2 and

APOA5, are known to stimulate LPL activity, whereas APOC1 and APOC3 inhibit LPL activity.

GPIHBP-1 is known to bind both chylomicrons and LPL and is thought to act as an anchoring protein that stabilizes LPL. Additionally, GPIHBP-1 can interact with angiopoietin-like protein 4

(Angptl4), which is known to inhibit LPL activity47. After chylomicron TG lipolysis, the residual

CE-rich, TG-depleted lipoprotein particles, referred to as chylomicron remnants, are released into the plasma compartment and rapidly cleared by the liver through chylomicron remnant receptors, LDLr, and LDL-receptor related protein 48, 49.

3.2.6. High density lipoprotein (HDL): HDL biogenesis involves the synthesis and secretion of the major HDL-associated apolipoproteins, apoA-I, synthesized in liver and intestine, and apoA-

II, synthesize in liver. Following secretion, apoA-I and apoA-II are lipidated via ATP binding

7 cassette transporter A1 (ABCA1) at the cell membrane surface, resulting in nascent HDL

(nHDL) particle formation. Subsequent maturation of nHDL to spherical plasma HDL occurs in the circulation50.

3.2.7. Intravascular HDL maturation and reverse cholesterol transport: Maturation of nHDL involves acquisition of additional PL from redundant VLDL surface generated during LPL lipolysis of TG, and PL transfer from other plasma apoB LP via phospholipid transfer protein

(PLTP) (Figure 1)51. Maturation of HDL in the circulation also involves esterification of FC to CE via lecithin: cholesterol acyltransferase (LCAT), generating a neutral lipid core (Figure 1)52.

Liver is the largest contributor to the plasma HDL pool 53. CE-enriched HDL returns to the liver in a process known as reverse cholesterol transport (RCT), which is the only quantitative way of removing extra-hepatic cholesterol from body54, 55. Liver is the primary site of HDL cholesterol uptake from plasma via scavenger receptor class BI (SR-BI)-mediated selective HDL CE uptake without HDL apolipoprotein degradation56, 57. Another circulating protein, CE transfer protein

(CETP), plays a critical role in human HDL cholesterol catabolism. CETP exchanges TG from apoB LP for CE in HDL, leading to TG enrichment and CE depletion in HDL particles.

Individuals with loss of function mutations in the CETP gene have high HDL cholesterol levels, whereas in rodents, which lack CETP, overexpression of CETP results in significant reductions in HDL-C58-60 (also reviewed in Figure 1, Table 2).

8

Figure 1. Outline of lipoprotein transport and metabolism

(adapted from Michael R. Flock et al. Adv Nutr 2011;2:261-274)

9

Table 2. Major enzymes important in lipoprotein metabolism

Enzyme / Location Function

transporter

ABCA1 Cell membrane Free cholesterol and phospholipid efflux

CETP Circulation/HDL Cholesteryl ester (CE) transfer from VLDL to HDL and

TG transfer from HDL to VLDL

LPL Endothelium Catalyzes chylomicron and VLDL TG hydrolysis to

release FFA. Activated by apoC-II and inhibited by

apoC-III

LCAT Circulation/HDL Esterification of free cholesterol to CE to generate

mature HDL particles. Activated by apoA-I and other

apolipoproteins

PLTP Circulation Phospholipid transfer from VLDL/LDL to HDL

SOAT2 Liver and Generation of intracellular CE from esterification of FC.

Intestine CE can be stored in lipid droplets or secreted in VLDL

particles

ABCA1: ATP binding cassette transporter 1, CETP: cholesteryl ester transfer protein, LPL: lipoprotein lipase, LCAT: lecithin cholesteryl acetyl , PLTP: phospholipid transfer protein, SOAT2: sterol O-acyltransferase 2. (Note: Above enzymes are also depicted in

Figure. 1).

10

4. PUFA-induced transcriptional regulation of lipoprotein metabolism and inflammation

PUFAs, in addition to providing energy through β-oxidation, modulate cellular functions in multiple ways including regulation of membrane composition and fluidity, cell signaling (i.e. phosphorylation, ubiquitination or proteolytic cleavage), and binding to transcription factors and/or their co-activators61. Nuclear receptors (NR) consist of a DNA-binding domain (DBD) and a ligand-binding domain (LBD). Ligand binding induces a conformational change that results in

NR binding to DNA (as homo/heterodimers) or to DNA-bound transcription factors. NR themselves can act as co-activators or co-repressors by binding to other transcription factors present in regulatory regions of target genes (reviewed in 62).

4.1. Nuclear receptors that directly bind to unsaturated fatty acids:

Fatty acids share the property of hydrophobicity with hormones (e.g. steroid, thyroid hormones) that activate NR. Several NR including peroxisome proliferator receptor (PPAR: α, β, γ1, γ2 isoforms), hepatocyte nuclear factor-5 (HNF4), retinoic acid receptor (RXR: α) and liver X receptor (LXR: α and β isoforms) contain fatty acid binding domains. Although, fatty acid

(especially linoleic acid) binding (reversible as well as irreversible) to HNF4 has been documented, its effects on transcriptional activity and physiological implications are relatively underexplored.

4.1.1: PPARs: Via their DBD, PPARs bind to peroxisome proliferator response elements

(PPREs) in enhancer region of target genes as heterodimers of RXR 63. Among three known human PPAR isotypes, α, γ and δ, PPARγ and PPAR δ are abundantly expressed in the macrophage64. PPARα is predominantly expressed in (rodent and human) liver and brown adipose tissue, and increases expression of genes involved in fatty acid elongation,

11 desaturation, oxidation and transport to accommodate increased requirement for fatty acid oxidation upon fasting 65. Macrophage PPARα also has atheroprotective potential 66.

PPARγ is unique in having a large T-shaped cavity of the apparent volume (~1,300 A˚) as a ligand for rosiglitazone and naturally occuring hydrophobic carboxylic acids, including essential fatty acids and their oxidized derivatives (such as linoleic acid derived 13-

67-69 HODE, 15-HETE), and prostaglandin J2 metabolites, such as (15d-PGJ2) . Longer chain

PUFAs typically bind to PPARγ in the micromolar concentration range 68 and induce

“transrepression” of NF-κB activated inflammatory genes70, 71.

Macrophage PPARγ is implicated in differentiation, ox-LDL uptake, and foam cell formation67, 72.

PPARγ is also expressed in human atherosclerotic plaque macrophages 73 and its deletion in mice exacerbates atherosclerosis compared to their WT counterparts given synthetic PPARγ agonist 74. Additionally, n-3 PUFAs share anti-inflammatory, insulin-sensitizing, and triglyceride- lowering effects with PPAR agonists, thiazolidinediones (TZDs) and fibrates, which target

PPARγ and PPARα, respectively (reviewed in 75).

4.1.2. LXRs: LXRs (α and β isotypes) are central regulators of lipid homeostasis. LXRα is expressed primarily in liver, intestine, adipose tissue, and macrophages, whereas LXRβ is expressed in many cell types76. In peripheral cells, such as macrophages, LXRs directly control transcription of genes involved in cholesterol efflux pathways, including ABCA1, ABCG1 and apoE77-79. In the intestine, ligand activation of LXR/RXR heterodimers dramatically reduces dietary cholesterol absorption, an effect postulated to be mediated by upregulation of LXR target genes ABCA1 and ATP binding casette sterol transporters, ABCG5/G8 80, 81. In the liver, LXRs regulate cholesterol and fatty acid metabolism and bile acid synthesis82. Additional work has

12 focused on identifying unique functions for LXRα and β isoforms and on synthetic antagonists that could be used to treat atherosclerosis and metabolic disorders (reviewed in83-85).

In vitro, unsaturated fatty acids prevent LXRs binding to its coactivators86, resulting in lower transcription of LXR target genes, including SREBP1-c, even in presence of synthetic or endogeneous LXR ligands, such as 24(S),25-epoxycholesterol and 22(R)-hydroxycholesterol87.

Additionally, PUFAs induce PPARγ activation, which inhibits LXR-RXR heterodimer formation and reduces LXR activation88.

4.2. Regulation of transcription factors by unsaturated fatty acids without direct binding

4.2.1. Sterol Regulatory Element-Binding Proteins (SREBPs): SREBPs belong to the basic helix-loop-helix-leucine zipper (bHLH-Zip) family of transcription factors and include SREBP-1a,

SREPB-1c and SREBP2. SREBP-1a and -1c are “splice variants” produced from a common gene. SREBP-1c is predominantly expressed in liver and is selective for the fatty acid pathway, whereas SREBP-1a is expressed in all other cells and is a strong activator of cholesterol and fatty acid biosynthesis. SREBP-2 is encoded by a separate gene and is selective for cholesterol biosynthesis in liver 89.

4.2.2. Negative feedback regulation of de novo lipogenesis via SREBPs: SREBPs are synthesized as inactive ER membrane protein precursors (consisting of 1150 amino acids).

SREBP cleavage-activating protein (Scap) is a tetrameric ER membrane protein that binds both sterol and SREBPs. ER cholesterol content is estimated to be ~1 molar percent of total cellular cholesterol and when ER membrane cholesterol content falls below ~5 mole percent with respect to phospholipid, Scap acts as an escort protein by binding to COPII vesicular proteins and mediating ER to Golgi translocation of SREBP2 90. In the Golgi, two proteases (site 1 and

13 site 2 protease, S1P and S2P, respectively) cleave off the N-terminus of SREBPs to release a transcriptionally active basic helix-loop-helix leucine zipper fragment, which in the nucleus activates lipogenic genes. When ER cholesterol content rises above ~5 mole percent, SCAP undergoes a conformational change that results in its binding to Insulin-induced gene protein

(INSIG). Upon Scap-Insig binding, COPII complex can no longer bind to Scap, hence retaining it in the ER. Inhibition of ER-Golgi translocation of the SREBP/Scap complex results in SREBP feedback inhibition (reviewed in90, 91).

4.2.3. Carbohydrate/glucose responsive transcription factor (ChREBP): In liver, SREBP-1c overexpression induces hepatosteatosis 92, whereas its deletion does not completely shut down fatty acid synthesis (~50% reduction) 93. This was postulated to be likely due to the presence of

Glucose-or-carbohydrate-response elements (GhoREs) in promoters of many lipogenic genes94.

Since, GhoREs can be recognized by bHLH family members (including SREBPs) 95, a large protein (864 amino acids) containing a bHLH domain was discovered and identified as

ChREBP96. ChREBP and SREPB-1c activation induces glycolytic and lipogenic gene expression in a synergistic fashion and in a gluckokinase-dependent manner97, whereas

ChREBP deletion reduces glycolysis and lipolysis, resulting in decreased hepatic steatosis, insulin resistance, and increased β-oxidation in mice98, 99. PUFAs (in vitro as well as in vivo) inhibit ChREBP activity by decreasing its mRNA stability and its translocation from the cytosol to the nucleus100.

4.2.4. Nuclear factor kappa B (NF-κB): NF-kB regulates expression of the genes involved in innate and adaptive immune responses. Activation of canonical NF-kB pathway up-regulates genes encoding cytokines, adhesion molecules anti-apoptotic and proliferative pathways. NF-kB exists in the form of a cytosolic trimer consisting of RelA or RelB, p50 or p52, and an inhibitory

14 protein, IκB. NF-kB activation by extracellular signals, such as endotoxins, cytokines, free radicals, results in phosphorylation, dissociation, and degradation of IkB, releasing the NF-kB dimer to translocate to the nucleus where it upregulates pro-inflammatory target gene expression 101.

In vitro, EPA and DHA inhibit of inflammatory proteins in endothelial cells, monocytes, macrophages and dendritic cells. The inhibitory effects of n-3 PUFA involve decreased IkB phosphorylation resulting in inactivation of NF-kB 102, and inactivation of key early signaling proteins, such as mitogen-activated protein kinases103. Dietary supplementation of n-3 PUFA in animals decreases NF-kB activation104, whereas in humans, 10 weeks of FO feeding

(18gm/day) reduced IL-1β and TNF-α secretion from endotoxin-stimulated mononuclear cells105.

Additional evidence suggests that ligand bound (i.e., activated) PPARs physically interact with

NF-kB, preventing its translocation to the nucleus106.

5. Immunomodulation by n-3 and n-6 PUFAs and their oxidized derivatives via G-protein coupled receptors (GPCR)

5.1. Eicosanoid generation via n-3 and n-6 PUFA: Arachidonic acid (AA, 20:4 n-6) esterified at the sn-2 position in cellular phospholipids (PLs) is a precursor for generation of oxidized metabolites, collectively called eicosanoids, via the cyclooxygenase (COX) and lipoxygenase (LO) pathways, resulting in families of inflammatory mediators, such as leukotrienes (LTs), prostaglandins (PGs), and thromboxanes (TXs)107. These mediators are rapidly synthesized and released from cells in response to extracellular stimuli. They act on cells in an autocrine or paracrine fashion through GPCRs to initiate downstream second messengers that modulate inflammation, chemotaxis, and vascular function12. Conversely, EPA, an n-3 counterpart of AA,

15 can generate oxidized metabolites that include series 3 PGs and TXs, series 5 LTs, and hydroxy- eicosatetraenoic acids (HETEs) by the same pathways, resulting in n-3 eicosanoid products that are typically less active than their n-6 counterpart in eliciting an inflammatory response. For instance, leukotriene B5 (LTB5), a major end product of EPA catalysis by 5-LO, has 10% of the

12, 108 biological activity of LTB4, the analogous end product of 5-LO from AA . Additionally, EPA can replace AA in the PL fraction of the cell and compete with AA for the pathways involved in eicosanoid production, leading to reduced series 2 PGs and TXs, and series 4 LTs and HETEs108.

5.2. Free fatty acid receptors (FFARs): Recent deorphanization of FFARs has allowed further understanding of FFAs as signaling molecules. FFA receptors are G-protein coupled receptors

(GPRs) that are activated by short, medium or long chain FFAs109-114 and include GPR40

(FFAR1), GPR43 (FFAR2), GPR41 (FFAR3), GPR84 and GPR120 (FFAR4). GPR120 is highly expressed in intestine, adrenals, lung, adipose tissue, and macrophages and is described as an n-3 PUFA receptor115. Upon activation by n-3 PUFA, GPR120 inhibits transforming growth factor beta-activated kinase 1 (TAK1) activation, resulting in attenuation of NF-κB and Jun N-terminal kinases (JNK) signaling115. GPR120 regulates glucose metabolism and obesity in mice and humans; a non-synonymous mutation (p.R270H) inhibited GPR120 signaling activity, resulting in increased risk of obesity in European populations116.

Although macrophages, adipose tissue, and gut predominantly express GPR120, the relative contribution of its expression in various cell types is just beginning to be determined. Bone marrow transplantation studies by Oh et al demonstrate that the insulin-resistance and pro- inflammatory macrophage phenotypes observed in whole body GPR120 KO mice can largely be attributed to hematopoietic cell/ macrophage GPR120 deletion. In LPS primed macrophages,

GPR120 activation by DHA or a synthetic agonist results in its internalization via βarr2 and binding to TAB1, preventing TAB1 and TAK1 kinases interaction required to activate pro- inflammatory NF-κB and JNK pathways 115, 117. Downstream of GPR120 and GPR40, βarr2-

16

TAB-1 interaction also prevents βarr2 and NLRP3/NLRP1c binding required for inflammasome activation, reducing caspase-1 activation and IL-1β maturation 118. GPR120 activation by DHA attenuates LPS-induced macrophage COX2 induction, PGE2 production, and IL-6 gene expression119. In mouse liver, GPR120 expression is localized to Kupffer cells and its activation protects mouse livers from ischemia reperfusion injury.

6. Macrophages and monocytes in atherosclerosis

6.1. Macrophage activation and polarization: Macrophages are professional phagocytotic cells of the innate immune system120 that recognize pathogen-associated molecular patterns

(PAMPs) via expression of the toll like receptors (TLRs)121, 122. Downstream signaling events of

TLR activation are mediated by association of intracellular adaptor proteins with TLR signaling domains. TLR-4 is activated by lipopolysaccharide (LPS) (a cell wall component of gram negative bacteria) and signals through its co-receptor, CD14. The TLR-4-mediated signaling pathway is well characterized and known to activate NF-κB and MAP kinases.

Tissue resident macrophages can be “polarized” toward a pro-inflammatory M1 or anti- inflammatory M2 phenotype123, 124. Activation by Th1 stimuli (e.g. IFN-γ, LPS) induces iNOS synthase required for intracellular pathogen killing and results in secretion of proinflammatory cytokines, such as TNF-α, IL-1β, IL-6, IL-18, and IL-12p40. Activation by Th2 stimuli (e.g. IL-4,

IL-13) derived from various cells of the innate immune system125, as well as adipocytes 126, induces expression of the mannose receptor (i.e. CD206)127 and arginase-1. M2 macrophages secrete anti-inflammatory cytokines, such as IL-10, and TGF-β, and are implicated in resolution of inflammation and wound healing 124, 128.

TLRs can also be aseptically activated via SFAs and endotoxins. FetA, a liver-derived circulating glycoprotein, is an adaptor protein that can act as a carrier of SFAs resulting in TLR4

17 activation22, 23. Additionally, n-3 PUFAs can limit TLR4 dimerization to membrane lipid rafts and inactivate the NF-κB pathway129. Deletion of MyD88, an adaptor protein downstream of TLR4-

CD14 complex, results in reduced atherosclerotic plaque and circulating levels of monocyte chemoattractant protein-1 (MCP-1)130, whereas deletion of TLR4, but not CD14, also reduces atherosclerosis and is accompanied by reduced circulating levels of MCP-1131.

6.2. Monocyte Subsets and monocyte recruitment to atherosclerotic lesions: Like macrophages, monocytes differ in their effector function and can be classified so. Circulating, undifferentiated monocytes respond to their environment and chemotactic cues by recruitment and differentiation into macrophages and/or dendritic cells. In mice, two major monocyte populations have been described based on cell surface C-C chemokine receptor (CCR) and

CX3C chemokine receptor (CX3CR) expression as: 1) CCR2+, CX3CR1+, Ly-6Chi (i.e., Ly-6Chi) or pro-inflammatory monocytes and 2) CCR2-, CX3CR1++,Ly-6Clo (i.e. Ly-6Clo) or patrolling monocytes 132, 133. CCR2 is the receptor for MCP1, which has been shown to be important in attracting blood monocytes to developing atherosclerotic lesions25. Ly6Chi monocytes are phenotypic correspondents to human CD16-CD14+ monocytes134. Nonclassical, Ly6Clo monocytes have been termed resident/reparative in mice because of their longer half-life in vivo; they are thought to replenish tissue-resident dendritic cells and macrophages. Ly6Clo monocytes are phenotypic correspondents to CD16+ CD14dim human monocytes134, 135.

6.3. Hypercholesterolemia-induced monocytosis: Hypercholesterolemia is associated with increased monocytes in the circulation that display a more inflammatory phenotype associated with greater surface expression of Ly6Chi and CCR2, the receptor for the chemokine CCL2 (i.e.,

MCP1) that is involved in monocyte recruitment to sites of inflammation 25, 33. A greater proportion of these monocytes are found in atherosclerotic lesions as macrophages, suggesting that these monocytes are preferentially recruited to developing atherosclerotic lesions34, 35. The accumulation/influx of blood monocytes in the artery wall correlates with lesion development and

18 occurs throughout the progression of atherosclerosis136. Deficiency CCR2 or CX3CR1 or their ligands inhibits recruitment of specific monocyte subsets to the artery wall137, 138. Furthermore, in models of atherosclerotic regression, particularly after reversal of hyperlipidemia, monocyte- derived cells emigrate out of plaques139-141, and monocyte recruitment is also attenuated142.

During regression, monocyte-derived cells take on dendritic cell-like characteristics and employ

CCR7 to exit lesions143.

7. Mouse models of atherosclerosis: C57BL/6 mice do not develop atherosclerosis when fed a chow diet; however feeding a diet containing 15% fat, 1.25% cholesterol, and 0.5% cholic acid

(Paigen diet)144 results in small lesions in the aortic root region of the heart, with minimal cholesterol accumulation in the aorta145. Since the advent of gene targeting, two genetically- modified mouse models have been widely used for atherosclerosis research: 1) mice lacking apolipoprotein E (ApoEKO)146 and 2) mice lacking low density lipoprotein receptor (LDLrKO)147, both characterized by elevated plasma cholesterol concentrations (Table 3).

19

Table: 3. Mouse models of Athero-progression

Baseline Mouse Features / Characteristics Causative factor plasma Model cholesterol

Spontaneous development of complex Accumulation of

plaques when mice are fed a chow diet; cholesterol rich ApoEKO mice ~500mg/dl acceleration of plaque formation when remnant lipoprotein

mice are fed a Western/high fat diet particles

Development of plaques following

feeding mice a cholesterol and fat- Delayed clearance of

LDLrKO mice ~250mg/dl enriched diet; lipoprotein profile is similar VLDL and LDL

to that of humans

Adapted from: Nature Reviews Immunology 13,709–721(2013) doi:10.1038/nri3520

8. Pathophysiology of atherosclerotic lesion development: Atherosclerosis is a characterized by dyslipidemia and inflammation. Multiple cell types including endothelial, smooth muscle, and immune cells play critical roles in atherosclerosis progression10.

Macrophages are involved in lipid metabolism as well as innate immunity 11. Cholesteryl ester

(CE)-loaded macrophages (i.e., foam cells) are a hallmark of atherosclerotic lesions2. Since LDL is a major transporters of CE in plasma and elevated plasma apoB-100/apoA-1 ratio is an independent risk factor for CVD148, atherosclerosis is initiated when plasma LDL is elevated.

20

LDL crosses the endothelial monolayer by transcytosis and gets trapped and retained in the arterial intima through binding to electronegative proteoglycans149. LDL retention is followed by lipoprotein aggregation and modification via enzymatic (e.g. 12/15 lipooxygenase (12-15 LO), secretory phospholipase 2 (sPLA2), myeloperoxidases (MPO) etc.) and non-enzymatic (e.g. free radicals, reactive oxygen species etc.) oxidation, primarily in the phospholipid (PL) fraction, resulting in minimally-modified or oxidized LDL (mmLDL/oxLDL)150-152.

The mmLDL/oxLDLs stimulate endothelial cells to produce chemokines (MCP-1, interferon- γ

(IFN-γ)), adhesion molecules (e.g. intracellular adhesion molecule (ICAM/ CD54), vascular cell adhesion protein-1 (VCAM-1/CD106), P and E selectins) and growth factors, such as macrophage colony-stimulating factor (M-CSF) and granulocyte macrophage colony stimulating factor (GM-CSF), resulting in chemo-attraction, rolling, and infiltration of monocytes into the lesional intima, where recruited monocytes differentiate into macrophages2, 153-157. MCP-1/CCR2 interaction plays a critical role in monocyte recruitment in atherosclerosis158. mmLDL/oxLDL is rapidly taken up via scavenger receptors, SR-A and CD36, expressed on activated macrophages, leading to foam cell formation159-162. The expression of scavenger receptors is further regulated by PPARγ and pro-inflammatory cytokines, such as TNF-α and IFN-γ72.

Accumulation of cholesterol and oxygenated derivatives of cholesterol results in downregulation of LDL receptor45, and upregulation of LXR, resulting in the transcriptional activation of a program of genes that function to efflux the excess cholesterol from the cell78, 79, 83, 163.

Cholesterol efflux to lipid-free apolipoproteins (i.e., apoA-I, apoE) occurs via ABCA1, whereas efflux to HDL particles occurs via ABCG1164, 165. Unabated sterol accumulation can ultimately result in apoptotic and/or necrotic cell death and fibrous plaque formation.

In addition to macrophages, smooth muscle cells (SMCs), T cells, and B cells play a role in atherosclerotic lesions. SMCs proliferation results in fibrous cap formation. In fibrous plaques, interaction of CD40 with its ligand CD40L (CD154) (expressed on T cells, B cells, macrophages,

21

ECs, and SMCs) results in production of inflammatory cytokines, matrix-degrading proteases, and adhesion molecules. Activated T cells produce proatherogenic cytokines, such as IFN-γ, which inhibit extracellular matrix production by SMCs, whereas macrophage secreted interstitial collagenases, gelatinases, and stromolysin that degrade extracellular matrix 166-168 169-171. In advanced lesions, calcification and neovascularization (growth of small vessels from media) both influence stability of atherosclerotic lesions2. Intimal calcification occurs due to secretion of matrix scaffold by specialized cells of osteoblastic potential in response to cytokines and oxysterols172. The thrombogenicity of lesion depends on tissue factor, a key protein in coagulation cascade initiation. Production of tissue factor by ECs and macrophages is enhanced by oxLDL, infection or CD40-CD40L interaction2, 173, 174. Ultimately, if this pathologic process is not halted or reversed, atherosclerotic plaque rupture occurs, resulting in myocardial infarction or stroke.

9. Potential of botanically-derived FADS2 n-6 and n-3 PUFAs as a FO substitute for CVD prevention: PUFAs have pleotropic functions; n-3 PUFAs are cardioprotective and anti- inflammatory molecules, whereas n-6 PUFAs have cardioprotective and immunomodulatory properties.

Reduction of dietary cholesterol and total fat intake and a replacement of dietary saturated fatty acids by PUFAs is still a primary intervention for CVD4. Despite known cardioprotective and anti- inflammatory properties of FO, consumption of fatty fish or FO supplements is low in the USA175.

Both n-3 and n-6 PUFAs are atheroprotective in mice, non-human primates, and humans11, 12, 15,

16, 20, 22, 23, 28, 176, 177. Despite high consumption of LA and alpha linoleic acid (ALA, 18:3 n-3)

(mostly consumed in vegetable oils), tissue enrichment of their longer chain bioactive products,

AA and EPA, respectively, is limited due to inefficiency of the rate-limiting FADS2 enzyme in the fatty acid desaturation and elongation pathway(depicted in Figure 2)178. Botanical oils that are

22 enriched in ALA do not offer the plasma lipid lowering and atheroprotective properties that are observed with FO179. Similarly, most sources of botanical n-6 PUFAs are highly enriched in LA, constituting ~85-90% of the North American dietary PUFA intake 22, 28, which is inefficiently converted by FADS2 to longer chain PUFAs (i.e, AA) that are precursors of bioactive eicosanoid species.

To circumvent the inefficient conversion of LA and ALA to longer chain PUFAs, we have identified botanical oils enriched in n-3 (echium oil, EO; enriched in 18:4 n-3) or n-6 (borage oil,

BO; enriched in 18:3 n-6 (GLA)) PUFAs that are the immediate products of FADS2 desaturation of ALA and LA, respectively (Figure 2).. We previously showed that EO effectively enriches plasma and tissue lipids in EPA 180. In LDLrKO mice, isocaloric replacement of palm oil (PO) with EO attenuated atherosclerosis severity, splenic monocytosis, monocyte influx into aortic intima, and aortic root intimal macrophage content to an equivalent extent as FO, lending proof of principle for this strategy16. However, whether a similar strategy will be atheroprotective for the n-6 PUFA pathway is unknown. GLA can be elongated to DGLA, a precursor of the anti- inflammatory PGE1, or AA, a substrate for proinflammatory eicosanoid species. No atherosclerosis studies have been performed with BO and little is known little about its effect on plasma lipids and lipoproteins.

23

EO BO

Figure 2. Outline of metabolic pathways for PUFA adapted from Calder and Grimble, 2002.

24

10. Statement of Research Intent

The purpose of this project was to determine the predominant mechanism (i.e. effect on lipid- lipoproteins vs. immunomodulatory/ anti-inflammatory) by which FADS2 derived n-6 PUFAs vs. n-3 PUFAs affect atherosclerosis in LDLrKO mice. We hypothesized that a botanical oil enriched in the FADS2 product 18:3 n-6 (i.e. BO) will not be as atheroprotective as one enriched in the FADS2 product 18:4 n-3 (i.e. EO) (Specific Aim 1). Contrary to our hypothesis, we conclude that BO is as atheroprotective and hepatoprotective as EO and FO relative to Palm Oil

(PO) via VLDL-cholesterol lowering and attenuation of macrophage inflammation and chemotaxis in vitro.

We then determined the relative contribution of G-coupled protein receptor (GPR) 120, to

FADS2-derived n-3 and n-6 PUFA-induced atheroprotection in LDLrKO mice. GPR120 acts as a negative regulator of macrophage inflammation and chemotaxis upon activation by n-3 PUFA.

Based on available information, we hypothesized that atheroprotection in mice fed n-3 PUFA- enriched EO or FO is due to a GPR120 (n-3 fatty acid receptor)-mediated switch of pro- inflammatory M1 macrophages to anti-inflammatory M2 macrophages, whereas atheroprotection associated with n-6 PUFA-enriched BO is independent of GPR120 (Specific

Aim 2). We transplanted bone marrow (BM) from wild type (WT) and GPR120 knockout

(GPR120KO) mice into lethally-irradiated LDLrKO recipient mice and investigated lipid metabolism, inflammatory response and atherosclerosis. We conclude that, although in vivo activation of L-GPR120 attenuates Kupffer cell inflammation, neutrophilia, splenic monocytosis, and monocyte recruitment into aortic intima, BO, EO and FO-induced atheroprotective is independent of leukocyte GPR120 activation.

In summary, atheroprotection in n-6 as well as n-3 PUFA (i.e. BO vs. EO and FO) fed LDLrKO mice is primarily driven by VLDL-cholesterol lowering.

25

This thesis addressed the considerable gaps in knowledge regarding atheroprotective and anti- inflammatory potential of BO supplementation. This work helped elucidate the role of FADS2 derived n-6 fatty acids in atheroprotection. Furthermore, this work elucidated the relative contribution of a macrophage anti-inflammatory protein GPR120 on atherosclerosis in the context of n-6 as well as n-3 PUFA feeding.

26

Reference List

1. Mozaffarian D, Benjamin EJ, Go AS, Arnett DK, Blaha MJ, Cushman M, de Ferranti S,

Despres JP, Fullerton HJ, Howard VJ, Huffman MD, Judd SE, Kissela BM, Lackland DT,

Lichtman JH, Lisabeth LD, Liu S, Mackey RH, Matchar DB, McGuire DK, Mohler ER, 3rd, Moy

CS, Muntner P, Mussolino ME, Nasir K, Neumar RW, Nichol G, Palaniappan L, Pandey DK,

Reeves MJ, Rodriguez CJ, Sorlie PD, Stein J, Towfighi A, Turan TN, Virani SS, Willey JZ, Woo

D, Yeh RW, Turner MB, American Heart Association Statistics C and Stroke Statistics S. Heart disease and stroke statistics--2015 update: a report from the American Heart Association.

Circulation. 2015;131:e29-322.

2. Lusis AJ. Atherosclerosis. Nature. 2000;407:233-41.

3. Moore KJ and Tabas I. Macrophages in the pathogenesis of atherosclerosis. Cell.

2011;145:341-55.

4. Mozaffarian D, Micha R and Wallace S. Effects on coronary heart disease of increasing polyunsaturated fat in place of saturated fat: a systematic review and meta-analysis of randomized controlled trials. PLoS medicine. 2010;7:e1000252.

5. American Heart Association Nutrition C, Lichtenstein AH, Appel LJ, Brands M,

Carnethon M, Daniels S, Franch HA, Franklin B, Kris-Etherton P, Harris WS, Howard B, Karanja

N, Lefevre M, Rudel L, Sacks F, Van Horn L, Winston M and Wylie-Rosett J. Diet and lifestyle recommendations revision 2006: a scientific statement from the American Heart Association

Nutrition Committee. Circulation. 2006;114:82-96.

6. Kris-Etherton PM, Harris WS, Appel LJ and American Heart Association. Nutrition C.

Fish consumption, fish oil, omega-3 fatty acids, and cardiovascular disease. Circulation.

2002;106:2747-57.

27

7. Dyerberg J and Bang HO. Lipid metabolism, atherogenesis, and haemostasis in

Eskimos: the role of the prostaglandin-3 family. Haemostasis. 1979;8:227-33.

8. Dyerberg J and Bang HO. A hypothesis on the development of acute myocardial infarction in Greenlanders. Scandinavian journal of clinical and laboratory investigation

Supplementum. 1982;161:7-13.

9. Brown AA and Hu FB. Dietary modulation of endothelial function: implications for cardiovascular disease. The American journal of clinical nutrition. 2001;73:673-86.

10. Abeywardena MY and Head RJ. Longchain n-3 polyunsaturated fatty acids and blood vessel function. Cardiovascular research. 2001;52:361-71.

11. Bigger JT, Jr. and El-Sherif T. Polyunsaturated fatty acids and cardiovascular events: a fish tale. Circulation. 2001;103:623-5.

12. Calder PC and Grimble RF. Polyunsaturated fatty acids, inflammation and immunity. Eur

J Clin Nutr. 2002;56 Suppl 3:S14-9.

13. Calder PC and Yaqoob P. Omega-3 polyunsaturated fatty acids and human health outcomes. Biofactors. 2009;35:266-72.

14. Parks JS and Bullock BC. Effect of fish oil versus lard diets on the chemical and physical properties of low density lipoproteins of nonhuman primates. Journal of lipid research.

1987;28:173-82.

15. Parks JS, Kaduck-Sawyer J, Bullock BC and Rudel LL. Effect of dietary fish oil on coronary artery and aortic atherosclerosis in African green monkeys. Arteriosclerosis.

1990;10:1102-12.

28

16. Brown AL, Zhu X, Rong S, Shewale S, Seo J, Boudyguina E, Gebre AK, Alexander-

Miller MA and Parks JS. Omega-3 fatty acids ameliorate atherosclerosis by favorably altering monocyte subsets and limiting monocyte recruitment to aortic lesions. Arteriosclerosis, thrombosis, and vascular biology. 2012;32:2122-30.

17. Yokoyama M, Origasa H, Matsuzaki M, Matsuzawa Y, Saito Y, Ishikawa Y, Oikawa S,

Sasaki J, Hishida H, Itakura H, Kita T, Kitabatake A, Nakaya N, Sakata T, Shimada K, Shirato K and Japan EPAlisI. Effects of eicosapentaenoic acid on major coronary events in hypercholesterolaemic patients (JELIS): a randomised open-label, blinded endpoint analysis.

Lancet. 2007;369:1090-8.

18. Burr ML, Fehily AM, Gilbert JF, Rogers S, Holliday RM, Sweetnam PM, Elwood PC and

Deadman NM. Effects of changes in fat, fish, and fibre intakes on death and myocardial reinfarction: diet and reinfarction trial (DART). Lancet. 1989;2:757-61.

19. Dietary supplementation with n-3 polyunsaturated fatty acids and vitamin E after myocardial infarction: results of the GISSI-Prevenzione trial. Gruppo Italiano per lo Studio della

Sopravvivenza nell'Infarto miocardico. Lancet. 1999;354:447-55.

20. Hu FB, Bronner L, Willett WC, Stampfer MJ, Rexrode KM, Albert CM, Hunter D and

Manson JE. Fish and omega-3 fatty acid intake and risk of coronary heart disease in women.

JAMA : the journal of the American Medical Association. 2002;287:1815-21.

21. Sun Q, Ma J, Campos H, Rexrode KM, Albert CM, Mozaffarian D and Hu FB. Blood concentrations of individual long-chain n-3 fatty acids and risk of nonfatal myocardial infarction.

The American journal of clinical nutrition. 2008;88:216-23.

22. Harris WS, Mozaffarian D, Rimm E, Kris-Etherton P, Rudel LL, Appel LJ, Engler MM,

Engler MB and Sacks F. Omega-6 fatty acids and risk for cardiovascular disease: a science

29 advisory from the American Heart Association Nutrition Subcommittee of the Council on

Nutrition, Physical Activity, and Metabolism; Council on Cardiovascular Nursing; and Council on

Epidemiology and Prevention. Circulation. 2009;119:902-7.

23. Harris WS and Shearer GC. Omega-6 Fatty Acids and Cardiovascular Disease: Friend or Foe? Circulation. 2014;130:1562-1564.

24. Rudel LL, Johnson FL, Sawyer JK, Wilson MS and Parks JS. Dietary polyunsaturated fat modifies low-density lipoproteins and reduces atherosclerosis of nonhuman primates with high and low diet responsiveness. The American journal of clinical nutrition. 1995;62:463S-470S.

25. Degirolamo C, Shelness GS and Rudel LL. LDL cholesteryl oleate as a predictor for atherosclerosis: evidence from human and animal studies on dietary fat. Journal of lipid research. 2009;50 Suppl:S434-9.

26. Chopra J and Webster RO. PGE1 inhibits neutrophil adherence and neutrophil-mediated injury to cultured endothelial cells. Am Rev Respir Dis. 1988;138:915-20.

27. Jones DA and Fitzpatrick FA. "Suicide" inactivation of thromboxane A2 synthase.

Characteristics of mechanism-based inactivation with isolated enzyme and intact platelets. The

Journal of biological chemistry. 1990;265:20166-71.

28. Farvid MS, Ding M, Pan A, Sun Q, Chiuve SE, Steffen LM, Willett WC and Hu FB.

Dietary Linoleic Acid and Risk of Coronary Heart Disease: A Systematic Review and Meta-

Analysis of Prospective Cohort Studies. Circulation. 2014;130:1568-1578.

29. Mahley RW, Innerarity TL, Rall SC, Jr. and Weisgraber KH. Plasma lipoproteins: apolipoprotein structure and function. Journal of lipid research. 1984;25:1277-94.

30

30. Cox RA and Garcia-Palmieri MR. Cholesterol, Triglycerides, and Associated

Lipoproteins. In: H. K. Walker, W. D. Hall and J. W. Hurst, eds. Clinical Methods: The History,

Physical, and Laboratory Examinations. 3rd ed. Boston; 1990.

31. Mahley RW and Innerarity TL. Lipoprotein receptors and cholesterol homeostasis.

Biochimica et biophysica acta. 1983;737:197-222.

32. Whitcomb DC and Lowe ME. Human pancreatic digestive enzymes. Digestive diseases and sciences. 2007;52:1-17.

33. Iqbal J and Hussain MM. Intestinal lipid absorption. American journal of physiology

Endocrinology and metabolism. 2009;296:E1183-94.

34. Chiang JY. Bile acids: regulation of synthesis. Journal of lipid research. 2009;50:1955-

66.

35. Coleman RA and Lee DP. Enzymes of triacylglycerol synthesis and their regulation.

Progress in lipid research. 2004;43:134-76.

36. Lee RG, Shah R, Sawyer JK, Hamilton RL, Parks JS and Rudel LL. ACAT2 contributes cholesteryl esters to newly secreted VLDL, whereas LCAT adds cholesteryl ester to LDL in mice. Journal of lipid research. 2005;46:1205-12.

37. Nguyen TM, Sawyer JK, Kelley KL, Davis MA, Kent CR and Rudel LL. ACAT2 and

ABCG5/G8 are both required for efficient cholesterol absorption in mice: evidence from thoracic lymph duct cannulation. Journal of lipid research. 2012;53:1598-609.

38. Chen SH, Habib G, Yang CY, Gu ZW, Lee BR, Weng SA, Silberman SR, Cai SJ,

Deslypere JP, Rosseneu M and et al. Apolipoprotein B-48 is the product of a messenger RNA with an organ-specific in-frame stop codon. Science. 1987;238:363-6.

31

39. Young SG. Recent progress in understanding apolipoprotein B. Circulation.

1990;82:1574-94.

40. Dixon JB. Mechanisms of chylomicron uptake into lacteals. Annals of the New York

Academy of Sciences. 2010;1207 Suppl 1:E52-7.

41. Cheng CF, Oosta GM, Bensadoun A and Rosenberg RD. Binding of lipoprotein lipase to endothelial cells in culture. The Journal of biological chemistry. 1981;256:12893-8.

42. Buhman KK, Chen HC and Farese RV, Jr. The enzymes of neutral lipid synthesis. The

Journal of biological chemistry. 2001;276:40369-72.

43. Shelness GS, Ingram MF, Huang XF and DeLozier JA. Apolipoprotein B in the rough endoplasmic reticulum: translation, translocation and the initiation of lipoprotein assembly. The

Journal of nutrition. 1999;129:456S-462S.

44. Shelness GS and Ledford AS. Evolution and mechanism of apolipoprotein B-containing lipoprotein assembly. Current opinion in lipidology. 2005;16:325-32.

45. Brown MS and Goldstein JL. A receptor-mediated pathway for cholesterol homeostasis.

Science. 1986;232:34-47.

46. Surendran RP, Visser ME, Heemelaar S, Wang J, Peter J, Defesche JC, Kuivenhoven

JA, Hosseini M, Peterfy M, Kastelein JJ, Johansen CT, Hegele RA, Stroes ES and Dallinga-Thie

GM. Mutations in LPL, APOC2, APOA5, GPIHBP1 and LMF1 in patients with severe hypertriglyceridaemia. Journal of internal medicine. 2012;272:185-96.

47. Kersten S and Bensadoun A. Stabilizing lipoprotein lipase. Journal of lipid research.

2009;50:2335-6.

32

48. Brown MS, Kovanen PT and Goldstein JL. Regulation of plasma cholesterol by lipoprotein receptors. Science. 1981;212:628-35.

49. Rohlmann A, Gotthardt M, Hammer RE and Herz J. Inducible inactivation of hepatic LRP gene by cre-mediated recombination confirms role of LRP in clearance of chylomicron remnants. The Journal of clinical investigation. 1998;101:689-95.

50. Rader DJ. Molecular regulation of HDL metabolism and function: implications for novel therapies. The Journal of clinical investigation. 2006;116:3090-100.

51. Huuskonen J, Olkkonen VM, Jauhiainen M and Ehnholm C. The impact of phospholipid transfer protein (PLTP) on HDL metabolism. Atherosclerosis. 2001;155:269-81.

52. Kuivenhoven JA, Pritchard H, Hill J, Frohlich J, Assmann G and Kastelein J. The molecular pathology of lecithin:cholesterol acyltransferase (LCAT) deficiency syndromes.

Journal of lipid research. 1997;38:191-205.

53. Timmins JM, Lee JY, Boudyguina E, Kluckman KD, Brunham LR, Mulya A, Gebre AK,

Coutinho JM, Colvin PL, Smith TL, Hayden MR, Maeda N and Parks JS. Targeted inactivation of hepatic Abca1 causes profound hypoalphalipoproteinemia and kidney hypercatabolism of apoA-I. The Journal of clinical investigation. 2005;115:1333-42.

54. Fielding CJ and Fielding PE. Molecular physiology of reverse cholesterol transport.

Journal of lipid research. 1995;36:211-28.

55. Brewer HB, Jr. and Rader DJ. HDL: structure, function and metabolism. Progress in lipid research. 1991;30:139-44.

33

56. Kozarsky KF, Donahee MH, Rigotti A, Iqbal SN, Edelman ER and Krieger M.

Overexpression of the HDL receptor SR-BI alters plasma HDL and bile cholesterol levels.

Nature. 1997;387:414-7.

57. Silver DL, Wang N, Xiao X and Tall AR. High density lipoprotein (HDL) particle uptake mediated by scavenger receptor class B type 1 results in selective sorting of HDL cholesterol from protein and polarized cholesterol secretion. The Journal of biological chemistry.

2001;276:25287-93.

58. Glass C, Pittman RC, Weinstein DB and Steinberg D. Dissociation of tissue uptake of cholesterol ester from that of apoprotein A-I of rat plasma high density lipoprotein: selective delivery of cholesterol ester to liver, adrenal, and gonad. Proceedings of the National Academy of Sciences of the United States of America. 1983;80:5435-9.

59. Ikewaki K, Rader DJ, Sakamoto T, Nishiwaki M, Wakimoto N, Schaefer JR, Ishikawa T,

Fairwell T, Zech LA, Nakamura H and et al. Delayed catabolism of high density lipoprotein apolipoproteins A-I and A-II in human cholesteryl ester transfer protein deficiency. The Journal of clinical investigation. 1993;92:1650-8.

60. Agellon LB, Walsh A, Hayek T, Moulin P, Jiang XC, Shelanski SA, Breslow JL and Tall

AR. Reduced high density lipoprotein cholesterol in human cholesteryl ester transfer protein transgenic mice. The Journal of biological chemistry. 1991;266:10796-801.

61. Jump DB, Tripathy S and Depner CM. Fatty Acid–Regulated Transcription Factors in the

Liver. Annual review of nutrition. 2013;33:249-269.

62. Glass CK and Rosenfeld MG. The coregulator exchange in transcriptional functions of nuclear receptors. Genes & development. 2000;14:121-41.

34

63. Miyata KS, McCaw SE, Marcus SL, Rachubinski RA and Capone JP. The peroxisome proliferator-activated receptor interacts with the retinoid X receptor in vivo. Gene. 1994;148:327-

30.

64. Li AC, Binder CJ, Gutierrez A, Brown KK, Plotkin CR, Pattison JW, Valledor AF, Davis

RA, Willson TM, Witztum JL, Palinski W and Glass CK. Differential inhibition of macrophage foam-cell formation and atherosclerosis in mice by PPARalpha, beta/delta, and gamma. The

Journal of clinical investigation. 2004;114:1564-76.

65. Kersten S, Seydoux J, Peters JM, Gonzalez FJ, Desvergne B and Wahli W. Peroxisome proliferator-activated receptor alpha mediates the adaptive response to fasting. The Journal of clinical investigation. 1999;103:1489-98.

66. Babaev VR, Ishiguro H, Ding L, Yancey PG, Dove DE, Kovacs WJ, Semenkovich CF,

Fazio S and Linton MF. Macrophage expression of peroxisome proliferator-activated receptor- alpha reduces atherosclerosis in low-density lipoprotein receptor-deficient mice. Circulation.

2007;116:1404-12.

67. Nagy L, Tontonoz P, Alvarez JG, Chen H and Evans RM. Oxidized LDL regulates macrophage gene expression through ligand activation of PPARgamma. Cell. 1998;93:229-40.

68. Xu HE, Lambert MH, Montana VG, Parks DJ, Blanchard SG, Brown PJ, Sternbach DD,

Lehmann JM, Wisely GB, Willson TM, Kliewer SA and Milburn MV. Molecular recognition of fatty acids by peroxisome proliferator-activated receptors. Molecular cell. 1999;3:397-403.

69. Nolte RT, Wisely GB, Westin S, Cobb JE, Lambert MH, Kurokawa R, Rosenfeld MG,

Willson TM, Glass CK and Milburn MV. Ligand binding and co-activator assembly of the peroxisome proliferator-activated receptor-gamma. Nature. 1998;395:137-43.

35

70. Ghisletti S, Huang W, Ogawa S, Pascual G, Lin ME, Willson TM, Rosenfeld MG and

Glass CK. Parallel SUMOylation-dependent pathways mediate gene- and signal-specific transrepression by LXRs and PPARgamma. Molecular cell. 2007;25:57-70.

71. Pascual G, Fong AL, Ogawa S, Gamliel A, Li AC, Perissi V, Rose DW, Willson TM,

Rosenfeld MG and Glass CK. A SUMOylation-dependent pathway mediates transrepression of inflammatory response genes by PPAR-gamma. Nature. 2005;437:759-63.

72. Tontonoz P, Nagy L, Alvarez JG, Thomazy VA and Evans RM. PPARgamma promotes monocyte/macrophage differentiation and uptake of oxidized LDL. Cell. 1998;93:241-52.

73. Marx N, Sukhova G, Murphy C, Libby P and Plutzky J. Macrophages in human atheroma contain PPARgamma: differentiation-dependent peroxisomal proliferator-activated receptor gamma(PPARgamma) expression and reduction of MMP-9 activity through PPARgamma activation in mononuclear phagocytes in vitro. The American journal of pathology. 1998;153:17-

23.

74. Babaev VR, Yancey PG, Ryzhov SV, Kon V, Breyer MD, Magnuson MA, Fazio S and

Linton MF. Conditional knockout of macrophage PPARgamma increases atherosclerosis in

C57BL/6 and low-density lipoprotein receptor-deficient mice. Arteriosclerosis, thrombosis, and vascular biology. 2005;25:1647-53.

75. Berger J and Moller DE. The mechanisms of action of PPARs. Annual review of medicine. 2002;53:409-35.

76. Repa JJ and Mangelsdorf DJ. Nuclear receptor regulation of cholesterol and bile acid metabolism. Current opinion in biotechnology. 1999;10:557-63.

36

77. Costet P, Luo Y, Wang N and Tall AR. Sterol-dependent transactivation of the ABC1 promoter by the liver X receptor/retinoid X receptor. The Journal of biological chemistry.

2000;275:28240-5.

78. Laffitte BA, Repa JJ, Joseph SB, Wilpitz DC, Kast HR, Mangelsdorf DJ and Tontonoz P.

LXRs control lipid-inducible expression of the apolipoprotein E gene in macrophages and adipocytes. Proceedings of the National Academy of Sciences of the United States of America.

2001;98:507-12.

79. Venkateswaran A, Laffitte BA, Joseph SB, Mak PA, Wilpitz DC, Edwards PA and

Tontonoz P. Control of cellular cholesterol efflux by the nuclear oxysterol receptor LXR alpha.

Proceedings of the National Academy of Sciences of the United States of America.

2000;97:12097-102.

80. Repa JJ, Turley SD, Lobaccaro JA, Medina J, Li L, Lustig K, Shan B, Heyman RA,

Dietschy JM and Mangelsdorf DJ. Regulation of absorption and ABC1-mediated efflux of cholesterol by RXR heterodimers. Science. 2000;289:1524-9.

81. Repa JJ, Berge KE, Pomajzl C, Richardson JA, Hobbs H and Mangelsdorf DJ.

Regulation of ATP-binding cassette sterol transporters ABCG5 and ABCG8 by the liver X receptors alpha and beta. The Journal of biological chemistry. 2002;277:18793-800.

82. Peet DJ, Turley SD, Ma W, Janowski BA, Lobaccaro JM, Hammer RE and Mangelsdorf

DJ. Cholesterol and bile acid metabolism are impaired in mice lacking the nuclear oxysterol receptor LXR alpha. Cell. 1998;93:693-704.

83. Hong C and Tontonoz P. Liver X receptors in lipid metabolism: opportunities for drug discovery. Nat Rev Drug Discov. 2014;13:433-44.

37

84. Calkin AC and Tontonoz P. Transcriptional integration of metabolism by the nuclear sterol-activated receptors LXR and FXR. Nature reviews Molecular cell biology. 2012;13:213-

24.

85. Tontonoz P. Transcriptional and posttranscriptional control of cholesterol homeostasis by liver X receptors. Cold Spring Harb Symp Quant Biol. 2011;76:129-37.

86. Heery DM, Kalkhoven E, Hoare S and Parker MG. A signature motif in transcriptional co- activators mediates binding to nuclear receptors. Nature. 1997;387:733-6.

87. Janowski BA, Grogan MJ, Jones SA, Wisely GB, Kliewer SA, Corey EJ and Mangelsdorf

DJ. Structural requirements of ligands for the oxysterol liver X receptors LXRalpha and

LXRbeta. Proceedings of the National Academy of Sciences of the United States of America.

1999;96:266-71.

88. Yoshikawa T, Ide T, Shimano H, Yahagi N, Amemiya-Kudo M, Matsuzaka T, Yatoh S,

Kitamine T, Okazaki H, Tamura Y, Sekiya M, Takahashi A, Hasty AH, Sato R, Sone H, Osuga J,

Ishibashi S and Yamada N. Cross-talk between peroxisome proliferator-activated receptor

(PPAR) alpha and liver X receptor (LXR) in nutritional regulation of fatty acid metabolism. I.

PPARs suppress sterol regulatory element binding protein-1c promoter through inhibition of

LXR signaling. Molecular endocrinology. 2003;17:1240-54.

89. Horton JD, Goldstein JL and Brown MS. SREBPs: activators of the complete program of cholesterol and fatty acid synthesis in the liver. Journal of Clinical Investigation. 2002;109:1125-

31.

90. Radhakrishnan A, Goldstein JL, McDonald JG and Brown MS. Switch-like control of

SREBP-2 transport triggered by small changes in ER cholesterol: a delicate balance. Cell metabolism. 2008;8:512-21.

38

91. Goldstein JL, DeBose-Boyd RA and Brown MS. Protein sensors for membrane sterols.

Cell. 2006;124:35-46.

92. Shimano H, Horton JD, Shimomura I, Hammer RE, Brown MS and Goldstein JL. Isoform

1c of sterol regulatory element binding protein is less active than isoform 1a in livers of transgenic mice and in cultured cells. The Journal of clinical investigation. 1997;99:846-54.

93. Liang G, Yang J, Horton JD, Hammer RE, Goldstein JL and Brown MS. Diminished hepatic response to fasting/refeeding and liver X receptor agonists in mice with selective deficiency of sterol regulatory element-binding protein-1c. The Journal of biological chemistry.

2002;277:9520-8.

94. Kaytor EN, Shih H and Towle HC. Carbohydrate regulation of hepatic gene expression.

Evidence against a role for the upstream stimulatory factor. The Journal of biological chemistry.

1997;272:7525-31.

95. Shih HM, Liu Z and Towle HC. Two CACGTG motifs with proper spacing dictate the carbohydrate regulation of hepatic gene transcription. The Journal of biological chemistry.

1995;270:21991-7.

96. Yamashita H, Takenoshita M, Sakurai M, Bruick RK, Henzel WJ, Shillinglaw W, Arnot D and Uyeda K. A glucose-responsive transcription factor that regulates carbohydrate metabolism in the liver. Proceedings of the National Academy of Sciences of the United States of America.

2001;98:9116-21.

97. Dentin R, Pegorier JP, Benhamed F, Foufelle F, Ferre P, Fauveau V, Magnuson MA,

Girard J and Postic C. Hepatic glucokinase is required for the synergistic action of ChREBP and

SREBP-1c on glycolytic and lipogenic gene expression. The Journal of biological chemistry.

2004;279:20314-26.

39

98. Iizuka K, Bruick RK, Liang G, Horton JD and Uyeda K. Deficiency of carbohydrate response element-binding protein (ChREBP) reduces lipogenesis as well as glycolysis.

Proceedings of the National Academy of Sciences of the United States of America.

2004;101:7281-6.

99. Iizuka K, Miller B and Uyeda K. Deficiency of carbohydrate-activated transcription factor

ChREBP prevents obesity and improves plasma glucose control in leptin-deficient (ob/ob) mice.

American journal of physiology Endocrinology and metabolism. 2006;291:E358-64.

100. Dentin R, Benhamed F, Pegorier JP, Foufelle F, Viollet B, Vaulont S, Girard J and Postic

C. Polyunsaturated fatty acids suppress glycolytic and lipogenic genes through the inhibition of

ChREBP nuclear protein translocation. The Journal of clinical investigation. 2005;115:2843-54.

101. Perkins ND. Integrating cell-signalling pathways with NF-kappaB and IKK function.

Nature reviews Molecular cell biology. 2007;8:49-62.

102. Novak TE, Babcock TA, Jho DH, Helton WS and Espat NJ. NF-kappa B inhibition by omega -3 fatty acids modulates LPS-stimulated macrophage TNF-alpha transcription. American journal of physiology Lung cellular and molecular physiology. 2003;284:L84-9.

103. Lo CJ, Chiu KC, Fu M, Chu A and Helton S. Fish oil modulates macrophage P44/P42 mitogen-activated protein kinase activity induced by lipopolysaccharide. JPEN Journal of parenteral and enteral nutrition. 2000;24:159-63.

104. Xi S, Cohen D and Chen LH. Effects of fish oil on cytokines and immune functions of mice with murine AIDS. Journal of lipid research. 1998;39:1677-87.

105. Endres S, Ghorbani R, Kelley VE, Georgilis K, Lonnemann G, van der Meer JW, Cannon

JG, Rogers TS, Klempner MS, Weber PC and et al. The effect of dietary supplementation with

40 n-3 polyunsaturated fatty acids on the synthesis of interleukin-1 and tumor necrosis factor by mononuclear cells. New England Journal of Medicine. 1989;320:265-71.

106. Vanden Berghe W, Vermeulen L, Delerive P, De Bosscher K, Staels B and Haegeman

G. A paradigm for gene regulation: inflammation, NF-kappaB and PPAR. Advances in experimental medicine and biology. 2003;544:181-96.

107. Funk CD. Prostaglandins and leukotrienes: advances in eicosanoid biology. Science.

2001;294:1871-5.

108. Jump DB. The biochemistry of n-3 polyunsaturated fatty acids. The Journal of biological chemistry. 2002;277:8755-8.

109. Wellendorph P, Johansen LD and Brauner-Osborne H. Molecular pharmacology of promiscuous seven transmembrane receptors sensing organic nutrients. Molecular pharmacology. 2009;76:453-65.

110. Briscoe CP, Tadayyon M, Andrews JL, Benson WG, Chambers JK, Eilert MM, Ellis C,

Elshourbagy NA, Goetz AS, Minnick DT, Murdock PR, Sauls HR, Jr., Shabon U, Spinage LD,

Strum JC, Szekeres PG, Tan KB, Way JM, Ignar DM, Wilson S and Muir AI. The orphan G protein-coupled receptor GPR40 is activated by medium and long chain fatty acids. The Journal of biological chemistry. 2003;278:11303-11.

111. Brown AJ, Goldsworthy SM, Barnes AA, Eilert MM, Tcheang L, Daniels D, Muir AI,

Wigglesworth MJ, Kinghorn I, Fraser NJ, Pike NB, Strum JC, Steplewski KM, Murdock PR,

Holder JC, Marshall FH, Szekeres PG, Wilson S, Ignar DM, Foord SM, Wise A and Dowell SJ.

The Orphan G protein-coupled receptors GPR41 and GPR43 are activated by propionate and other short chain carboxylic acids. The Journal of biological chemistry. 2003;278:11312-9.

41

112. Hirasawa A, Tsumaya K, Awaji T, Katsuma S, Adachi T, Yamada M, Sugimoto Y,

Miyazaki S and Tsujimoto G. Free fatty acids regulate gut incretin glucagon-like peptide-1 secretion through GPR120. Nature medicine. 2005;11:90-4.

113. Oh Da Y and Olefsky Jerrold M. Omega 3 Fatty Acids and GPR120. Cell metabolism.

2012;15:564-565.

114. Wang J, Wu X, Simonavicius N, Tian H and Ling L. Medium-chain fatty acids as ligands for orphan G protein-coupled receptor GPR84. The Journal of biological chemistry.

2006;281:34457-64.

115. Oh DY, Talukdar S, Bae EJ, Imamura T, Morinaga H, Fan W, Li P, Lu WJ, Watkins SM and Olefsky JM. GPR120 is an omega-3 fatty acid receptor mediating potent anti-inflammatory and insulin-sensitizing effects. Cell. 2010;142:687-98.

116. Ichimura A, Hirasawa A, Poulain-Godefroy O, Bonnefond A, Hara T, Yengo L, Kimura I,

Leloire A, Liu N, Iida K, Choquet H, Besnard P, Lecoeur C, Vivequin S, Ayukawa K, Takeuchi

M, Ozawa K, Tauber M, Maffeis C, Morandi A, Buzzetti R, Elliott P, Pouta A, Jarvelin MR,

Korner A, Kiess W, Pigeyre M, Caiazzo R, Van Hul W, Van Gaal L, Horber F, Balkau B, Levy-

Marchal C, Rouskas K, Kouvatsi A, Hebebrand J, Hinney A, Scherag A, Pattou F, Meyre D,

Koshimizu TA, Wolowczuk I, Tsujimoto G and Froguel P. Dysfunction of lipid sensor GPR120 leads to obesity in both mouse and human. Nature. 2012;483:350-4.

117. Oh da Y, Walenta E, Akiyama TE, Lagakos WS, Lackey D, Pessentheiner AR, Sasik R,

Hah N, Chi TJ, Cox JM, Powels MA, Di Salvo J, Sinz C, Watkins SM, Armando AM, Chung H,

Evans RM, Quehenberger O, McNelis J, Bogner-Strauss JG and Olefsky JM. A Gpr120- selective agonist improves insulin resistance and chronic inflammation in obese mice. Nature medicine. 2014;20:942-7.

42

118. Yan Y, Jiang W, Spinetti T, Tardivel A, Castillo R, Bourquin C, Guarda G, Tian Z,

Tschopp J and Zhou R. Omega-3 Fatty Acids Prevent Inflammation and Metabolic Disorder through Inhibition of NLRP3 Inflammasome Activation. Immunity. 2013;38:1154-1163.

119. Li X, Yu Y and Funk CD. Cyclooxygenase-2 induction in macrophages is modulated by docosahexaenoic acid via interactions with free fatty acid receptor 4 (FFA4). FASEB journal : official publication of the Federation of American Societies for Experimental Biology.

2013;27:4987-97.

120. Tobias P and Curtiss LK. Thematic review series: The immune system and atherogenesis. Paying the price for pathogen protection: toll receptors in atherogenesis. Journal of lipid research. 2005;46:404-11.

121. Triantafilou M, Miyake K, Golenbock DT and Triantafilou K. Mediators of innate immune recognition of bacteria concentrate in lipid rafts and facilitate lipopolysaccharide-induced cell activation. Journal of cell science. 2002;115:2603-11.

122. Triantafilou M and Triantafilou K. Lipopolysaccharide recognition: CD14, TLRs and the

LPS-activation cluster. Trends Immunol. 2002;23:301-4.

123. Mantovani A, Sica A, Sozzani S, Allavena P, Vecchi A and Locati M. The chemokine system in diverse forms of macrophage activation and polarization. Trends Immunol.

2004;25:677-86.

124. Gordon S and Martinez FO. Alternative activation of macrophages: mechanism and functions. Immunity. 2010;32:593-604.

125. Gordon S. Alternative activation of macrophages. Nature reviews Immunology.

2003;3:23-35.

43

126. Kang K, Reilly SM, Karabacak V, Gangl MR, Fitzgerald K, Hatano B and Lee C-H.

Adipocyte-Derived Th2 Cytokines and Myeloid PPARδ Regulate Macrophage Polarization and

Insulin Sensitivity. Cell metabolism. 2008;7:485-495.

127. Stein M, Keshav S, Harris N and Gordon S. Interleukin 4 potently enhances murine macrophage mannose receptor activity: a marker of alternative immunologic macrophage activation. The Journal of experimental medicine. 1992;176:287-92.

128. Varin A and Gordon S. Alternative activation of macrophages: immune function and cellular biology. Immunobiology. 2009;214:630-41.

129. Wong SW, Kwon MJ, Choi AM, Kim HP, Nakahira K and Hwang DH. Fatty acids modulate Toll-like receptor 4 activation through regulation of receptor dimerization and recruitment into lipid rafts in a reactive oxygen species-dependent manner. The Journal of biological chemistry. 2009;284:27384-92.

130. Bjorkbacka H, Kunjathoor VV, Moore KJ, Koehn S, Ordija CM, Lee MA, Means T,

Halmen K, Luster AD, Golenbock DT and Freeman MW. Reduced atherosclerosis in MyD88- null mice links elevated serum cholesterol levels to activation of innate immunity signaling pathways. Nature medicine. 2004;10:416-21.

131. Ding Y, Subramanian S, Montes VN, Goodspeed L, Wang S, Han C, Teresa AS, 3rd,

Kim J, O'Brien KD and Chait A. Toll-like receptor 4 deficiency decreases atherosclerosis but does not protect against inflammation in obese low-density lipoprotein receptor-deficient mice.

Arteriosclerosis, thrombosis, and vascular biology. 2012;32:1596-604.

132. Tacke F, Alvarez D, Kaplan TJ, Jakubzick C, Spanbroek R, Llodra J, Garin A, Liu J,

Mack M, van Rooijen N, Lira SA, Habenicht AJ and Randolph GJ. Monocyte subsets

44 differentially employ CCR2, CCR5, and CX3CR1 to accumulate within atherosclerotic plaques.

The Journal of clinical investigation. 2007;117:185-194.

133. Taylor PR and Gordon S. Monocyte heterogeneity and innate immunity. Immunity.

2003;19:2-4.

134. Geissmann F, Jung S and Littman DR. Blood monocytes consist of two principal subsets with distinct migratory properties. Immunity. 2003;19:71-82.

135. Geissmann F, Manz MG, Jung S, Sieweke MH, Merad M and Ley K. Development of monocytes, macrophages, and dendritic cells. Science. 2010;327:656-61.

136. Swirski FK, Pittet MJ, Kircher MF, Aikawa E, Jaffer FA, Libby P and Weissleder R.

Monocyte accumulation in mouse atherogenesis is progressive and proportional to extent of disease. Proceedings of the National Academy of Sciences of the United States of America.

2006;103:10340-5.

137. Combadiere C, Potteaux S, Rodero M, Simon T, Pezard A, Esposito B, Merval R,

Proudfoot A, Tedgui A and Mallat Z. Combined inhibition of CCL2, CX3CR1, and CCR5 abrogates Ly6C(hi) and Ly6C(lo) monocytosis and almost abolishes atherosclerosis in hypercholesterolemic mice. Circulation. 2008;117:1649-57.

138. Landsman L, Bar-On L, Zernecke A, Kim KW, Krauthgamer R, Shagdarsuren E, Lira SA,

Weissman IL, Weber C and Jung S. CX3CR1 is required for monocyte homeostasis and atherogenesis by promoting cell survival. Blood. 2009;113:963-72.

139. Randolph GJ. The fate of monocytes in atherosclerosis. Journal of thrombosis and haemostasis : JTH. 2009;7 Suppl 1:28-30.

45

140. Randolph GJ. Emigration of monocyte-derived cells to lymph nodes during resolution of inflammation and its failure in atherosclerosis. Current opinion in lipidology. 2008;19:462-8.

141. Llodra J, Angeli V, Liu J, Trogan E, Fisher EA and Randolph GJ. Emigration of monocyte-derived cells from atherosclerotic lesions characterizes regressive, but not progressive, plaques. Proceedings of the National Academy of Sciences of the United States of

America. 2004;101:11779-84.

142. Potteaux S, Gautier EL, Hutchison SB, van Rooijen N, Rader DJ, Thomas MJ, Sorci-

Thomas MG and Randolph GJ. Suppressed monocyte recruitment drives macrophage removal from atherosclerotic plaques of Apoe-/- mice during disease regression. The Journal of clinical investigation. 2011;121:2025-36.

143. Trogan E, Feig JE, Dogan S, Rothblat GH, Angeli V, Tacke F, Randolph GJ and Fisher

EA. Gene expression changes in foam cells and the role of chemokine receptor CCR7 during atherosclerosis regression in ApoE-deficient mice. Proceedings of the National Academy of

Sciences of the United States of America. 2006;103:3781-6.

144. Nishina PM, Verstuyft J and Paigen B. Synthetic low and high fat diets for the study of atherosclerosis in the mouse. Journal of lipid research. 1990;31:859-69.

145. Furbee JW, Jr. and Parks JS. Transgenic overexpression of human lecithin: cholesterol acyltransferase (LCAT) in mice does not increase aortic cholesterol deposition. Atherosclerosis.

2002;165:89-100.

146. Zhang SH, Reddick RL, Piedrahita JA and Maeda N. Spontaneous hypercholesterolemia and arterial lesions in mice lacking apolipoprotein E. Science. 1992;258:468-71.

46

147. Ishibashi S, Brown MS, Goldstein JL, Gerard RD, Hammer RE and Herz J.

Hypercholesterolemia in low density lipoprotein receptor knockout mice and its reversal by adenovirus-mediated gene delivery. The Journal of clinical investigation. 1993;92:883-93.

148. Walldius G and Jungner I. The apoB/apoA-I ratio: a strong, new risk factor for cardiovascular disease and a target for lipid-lowering therapy--a review of the evidence. Journal of internal medicine. 2006;259:493-519.

149. Skalen K, Gustafsson M, Rydberg EK, Hulten LM, Wiklund O, Innerarity TL and Boren J.

Subendothelial retention of atherogenic lipoproteins in early atherosclerosis. Nature.

2002;417:750-4.

150. Cyrus T, Witztum JL, Rader DJ, Tangirala R, Fazio S, Linton MF and Funk CD.

Disruption of the 12/15-lipoxygenase gene diminishes atherosclerosis in apo E-deficient mice.

The Journal of clinical investigation. 1999;103:1597-604.

151. Goldstein JL, Ho YK, Basu SK and Brown MS. Binding site on macrophages that mediates uptake and degradation of acetylated low density lipoprotein, producing massive cholesterol deposition. Proceedings of the National Academy of Sciences of the United States of America. 1979;76:333-7.

152. Watson AD, Leitinger N, Navab M, Faull KF, Horkko S, Witztum JL, Palinski W,

Schwenke D, Salomon RG, Sha W, Subbanagounder G, Fogelman AM and Berliner JA.

Structural identification by mass spectrometry of oxidized phospholipids in minimally oxidized low density lipoprotein that induce monocyte/endothelial interactions and evidence for their presence in vivo. The Journal of biological chemistry. 1997;272:13597-607.

47

153. Dong ZM, Chapman SM, Brown AA, Frenette PS, Hynes RO and Wagner DD. The combined role of P- and E-selectins in atherosclerosis. The Journal of clinical investigation.

1998;102:145-52.

154. Collins RG, Velji R, Guevara NV, Hicks MJ, Chan L and Beaudet AL. P-Selectin or intercellular adhesion molecule (ICAM)-1 deficiency substantially protects against atherosclerosis in apolipoprotein E-deficient mice. The Journal of experimental medicine.

2000;191:189-94.

155. Moore KJ, Sheedy FJ and Fisher EA. Macrophages in atherosclerosis: a dynamic balance. Nature reviews Immunology. 2013;13:709-21.

156. Gupta S, Pablo AM, Jiang X, Wang N, Tall AR and Schindler C. IFN-gamma potentiates atherosclerosis in ApoE knock-out mice. The Journal of clinical investigation. 1997;99:2752-61.

157. de Villiers WJ, Smith JD, Miyata M, Dansky HM, Darley E and Gordon S. Macrophage phenotype in mice deficient in both macrophage-colony-stimulating factor (op) and apolipoprotein E. Arteriosclerosis, thrombosis, and vascular biology. 1998;18:631-40.

158. Gu L, Okada Y, Clinton SK, Gerard C, Sukhova GK, Libby P and Rollins BJ. Absence of monocyte chemoattractant protein-1 reduces atherosclerosis in low density lipoprotein receptor- deficient mice. Molecular cell. 1998;2:275-81.

159. Podrez EA, Febbraio M, Sheibani N, Schmitt D, Silverstein RL, Hajjar DP, Cohen PA,

Frazier WA, Hoff HF and Hazen SL. Macrophage scavenger receptor CD36 is the major receptor for LDL modified by monocyte-generated reactive nitrogen species. The Journal of clinical investigation. 2000;105:1095-108.

160. Febbraio M, Podrez EA, Smith JD, Hajjar DP, Hazen SL, Hoff HF, Sharma K and

Silverstein RL. Targeted disruption of the class B scavenger receptor CD36 protects against

48 atherosclerotic lesion development in mice. The Journal of clinical investigation. 2000;105:1049-

56.

161. Gough PJ, Greaves DR, Suzuki H, Hakkinen T, Hiltunen MO, Turunen M, Herttuala SY,

Kodama T and Gordon S. Analysis of macrophage scavenger receptor (SR-A) expression in human aortic atherosclerotic lesions. Arteriosclerosis, thrombosis, and vascular biology.

1999;19:461-71.

162. Suzuki H, Kurihara Y, Takeya M, Kamada N, Kataoka M, Jishage K, Ueda O, Sakaguchi

H, Higashi T, Suzuki T, Takashima Y, Kawabe Y, Cynshi O, Wada Y, Honda M, Kurihara H,

Aburatani H, Doi T, Matsumoto A, Azuma S, Noda T, Toyoda Y, Itakura H, Yazaki Y, Kodama T and et al. A role for macrophage scavenger receptors in atherosclerosis and susceptibility to infection. Nature. 1997;386:292-6.

163. Chawla A, Boisvert WA, Lee CH, Laffitte BA, Barak Y, Joseph SB, Liao D, Nagy L,

Edwards PA, Curtiss LK, Evans RM and Tontonoz P. A PPAR gamma-LXR-ABCA1 pathway in macrophages is involved in cholesterol efflux and atherogenesis. Molecular cell. 2001;7:161-71.

164. Rosenson RS, Brewer HB, Davidson WS, Fayad ZA, Fuster V, Goldstein J, Hellerstein

M, Jiang X-C, Phillips MC, Rader DJ, Remaley AT, Rothblat GH, Tall AR and Yvan-Charvet L.

Cholesterol Efflux and Atheroprotection. Circulation. 2012;125:1905-1919.

165. Wang X, Collins HL, Ranalletta M, Fuki IV, Billheimer JT, Rothblat GH, Tall AR and

Rader DJ. Macrophage ABCA1 and ABCG1, but not SR-BI, promote macrophage reverse cholesterol transport in vivo. The Journal of clinical investigation. 2007;117:2216-24.

166. Mach F, Schonbeck U and Libby P. CD40 signaling in vascular cells: a key role in atherosclerosis? Atherosclerosis. 1998;137 Suppl:S89-95.

49

167. Swanson SJ, Rosenzweig A, Seidman JG and Libby P. Diversity of T-cell antigen receptor V beta gene utilization in advanced human atheroma. Arteriosclerosis and thrombosis : a journal of vascular biology / American Heart Association. 1994;14:1210-4.

168. Amento EP, Ehsani N, Palmer H and Libby P. Cytokines and growth factors positively and negatively regulate interstitial collagen gene expression in human vascular smooth muscle cells. Arteriosclerosis and thrombosis : a journal of vascular biology / American Heart

Association. 1991;11:1223-30.

169. Sukhova GK, Schonbeck U, Rabkin E, Schoen FJ, Poole AR, Billinghurst RC and Libby

P. Evidence for increased collagenolysis by interstitial collagenases-1 and -3 in vulnerable human atheromatous plaques. Circulation. 1999;99:2503-9.

170. Newby AC, Libby P and van der Wal AC. Plaque instability--the real challenge for atherosclerosis research in the next decade? Cardiovascular research. 1999;41:321-2.

171. Libby P, Schoenbeck U, Mach F, Selwyn AP and Ganz P. Current concepts in cardiovascular pathology: the role of LDL cholesterol in plaque rupture and stabilization. The

American journal of medicine. 1998;104:14S-18S.

172. Watson KE, Bostrom K, Ravindranath R, Lam T, Norton B and Demer LL. TGF-beta 1 and 25-hydroxycholesterol stimulate osteoblast-like vascular cells to calcify. The Journal of clinical investigation. 1994;93:2106-13.

173. Schonbeck U, Sukhova GK, Shimizu K, Mach F and Libby P. Inhibition of CD40 signaling limits evolution of established atherosclerosis in mice. Proceedings of the National

Academy of Sciences of the United States of America. 2000;97:7458-63.

50

174. Schonbeck U, Mach F, Sukhova GK, Herman M, Graber P, Kehry MR and Libby P.

CD40 ligation induces tissue factor expression in human vascular smooth muscle cells. The

American journal of pathology. 2000;156:7-14.

175. Kris-Etherton PM, Taylor DS, Yu-Poth S, Huth P, Moriarty K, Fishell V, Hargrove RL,

Zhao G and Etherton TD. Polyunsaturated fatty acids in the food chain in the United States. The

American journal of clinical nutrition. 2000;71:179S-88S.

176. Kang JX and Leaf A. Antiarrhythmic effects of polyunsaturated fatty acids. Recent studies. Circulation. 1996;94:1774-80.

177. Zampolli A, Bysted A, Leth T, Mortensen A, De Caterina R and Falk E. Contrasting effect of fish oil supplementation on the development of atherosclerosis in murine models.

Atherosclerosis. 2006;184:78-85.

178. Horrobin DF. Fatty acid metabolism in health and disease: the role of delta-6- desaturase. American Journal of Clinical Nutrition. 1993;57:732S-736S; discussion 736S-737S.

179. Degirolamo C, Kelley KL, Wilson MD and Rudel LL. Dietary n-3 LCPUFA from fish oil but not alpha-linolenic acid-derived LCPUFA confers atheroprotection in mice. Journal of lipid research. 2010;51:1897-905.

180. Zhang P, Boudyguina E, Wilson MD, Gebre AK and Parks JS. Echium oil reduces plasma lipids and hepatic lipogenic gene expression in apoB100-only LDL receptor knockout mice. Journal of Nutritional Biochemistry. 2008;19:655-63.

51

CHAPTER II

BOTANICAL OILS ENRICHED IN N-6 AND N-3 FATTY ACID PRODUCTS OF FADS2 ARE

EQUALLY EFFECTIVE IN PREVENTING ATHEROSCLEROSIS AND HEPATOSTEATOSIS

IN MICE

Swapnil V. Shewale, Elena Boudyguina, Xuewei Zhu , Lulu Shen, Patrick M. Hutchins, Robert

M. Barkley, Robert C. Murphy, John S. Parks.

This chapter is published in J. Lipid Res. 2015 Apr 28. pii: jlr.M059170 (PMID 25921305). Stylistic variations are due to the requirements of the journal.

Swapnil V. Shewale performed the experiments and prepared the manuscript. Dr. John S. Parks acted in an advisory and editorial capacity.

52

Botanical oils enriched in n-6 and n-3 fatty acid products of FADS2 are equally effective

in preventing atherosclerosis and hepatosteatosis in mice

Swapnil V. Shewale 1, 3, Elena Boudyguina1, Xuewei Zhu 1, Lulu Shen1,3, Patrick M. Hutchins4,

Robert M. Barkley4, Robert C. Murphy4, John S. Parks1,2.

1Departments of Internal Medicine-Section on Molecular Medicine, 2Biochemistry, and

3Physiology/Pharmacology, Wake Forest School of Medicine, Winston-Salem, NC 27157 and

4Department of Pharmacology, University of Colorado Denver, Aurora, CO 80045

Address correspondence to: Dr. John S. Parks, Department of Internal Medicine-Section on

Molecular Medicine, Wake Forest School of Medicine, Medical Center Blvd, Winston-Salem, NC

27157, USA.

Phone: 336-716-2145; Fax: 336-716-6279; Email: [email protected]

Running title: PUFA-mediated atheroprotection and hepatoprotection in mice

Abbreviations: AD, Atherogenic Diet; AUC, area under the curve; CE, cholesteryl ester;

FADS2, fatty acid desaturase 2 (FADS2) / delta six fatty acid desaturase enzyme; HODES,

Hydroxyoctadecadienoic acids; LDLrKO, LDL receptor KO; LPS, Lipopolysaccharide; Ox-CE,

Oxidized cholesteryl ester-fatty acyl species

3- current address: Lulu Shen, School of Life Science, Jiangsu Normal University. Jiangsu,

China, 221116

53

Abstract

Echium oil (EO), which is enriched in 18:4 n-3, the immediate product of FADS2 desaturation of

18:3 n-3, is as atheroprotective as fish oil (FO). The objective of this study was to determine whether botanical oils enriched in the FADS2 products 18:3 n-6 vs. 18:4 n-3 are equally atheroprotective. LDL receptor KO mice were fed one of four atherogenic diets containing 0.2% cholesterol and 10% calories as palm oil (PO) plus 10% calories as: 1) PO, 2) borage oil (BO;

18:3 n-6 enriched), 3) EO; 18:4 n-3 enriched, or 4) FO for 16 weeks. Mice fed BO, EO and FO vs. PO had significantly lower plasma total and VLDL cholesterol concentrations; hepatic neutral lipid content and inflammation, aortic CE content, aortic root intimal area and macrophage content; and peritoneal macrophage inflammation, CE content, and ex vivo chemotaxis.

Atheromas lacked oxidized CEs despite abundant generation of macrophage 12/15 lipooxygenase-derived metabolites. We conclude that botanical oils enriched in 18:3 n-6 and

18:4 n-3 PUFAs beyond the rate-limiting FADS2 enzyme are equally effective in preventing atherosclerosis and hepatosteatosis compared to saturated/monounsaturated fat due to cellular enrichment of ≥20 PUFAs, reduced plasma VLDL, and attenuated macrophage inflammation.

Keywords: atherosclerosis, fatty acid/metabolism, inflammation, lipoproteins/metabolism, macrophages/monocytes, mass spectrometry, VLDL

54

Introduction

Despite widespread use of lipid-lowering drugs, cardiovascular disease is still the leading cause of death in the United States 1. One of the first lines of cardiovascular disease treatment is reducing total fat intake and replacing saturated fatty acids with polyunsaturated fatty acids

(PUFAs) 2. In particular, n-3 PUFAs from fatty fish or fish oil (FO) reduce the extent of cardiovascular and other chronic diseases in humans and experimental animals 3-6. The atheroprotective benefits of fatty fish and FO are attributed to two n-3 PUFAs, eicosapentaenoic acid (EPA, 20:5 n-3) and docosahexaenoic acid (DHA, 22:6 n-3). However, despite the well- documented health benefits of n-3 PUFAs, consumption is low in the US 7. Furthermore, most botanical sources of n-3 fatty acids (i.e., flaxseed oil) contain alpha linolenic acid (ALA, 18:3 n-

3), which is inefficiently converted to EPA (4-15%) because of the rate-limiting nature of fatty acid desaturase 2 (FADS2, i.e., delta-6 desaturase) in the fatty acid elongation and desaturation pathway 8. Thus, botanical oils enriched in ALA lack the plasma lipid-lowering and atheroprotective properties observed with FO 9.

Dietary n-6 PUFAs are also atheroprotective in mouse and nonhuman primate models of atherosclerosis, and in human populations 10-13. Some of the atheroprotection is likely due to plasma lipid lowering, but there may also be an anti-inflammatory role for n-6 PUFAs through the generation of prostaglandin E1 (PGE1) from dihomo-gamma-linolenic acid (DGLA, 20:3 n-6), a potent inhibitor of thromboxane A2 (TXA2) formation that inhibits leukocyte adherence to endothelial cells 14, 15. However, most sources of botanical n-6 PUFAs are highly enriched in linoleic acid (LA; 18:2 n-6, ~85-90% of North American dietary PUFA intake) 10, 12. LA also must be converted by FADS2 to longer-chain bioactive PUFAs, although some bioactive derivatives of LA are described as both pro- and anti-inflammatory 13.

55

One strategy to circumvent the poor conversion of dietary 18-carbon PUFAs to their ≥20 carbon bioactive products is to identify botanical oils enriched in PUFAs beyond the FADS2 rate-limiting step of desaturation and elongation. We previously showed that echium oil (EO), which is relatively enriched in stearidonic acid (SDA, 18:4 n-3), the immediate product of FADS2- mediated desaturation of ALA, effectively enriches plasma and tissue lipids in EPA 16. In LDL receptor KO (LDLrKO) mice, isocaloric replacement of palm oil (PO) with EO attenuated atherosclerosis severity, splenic monocytosis, monocyte influx into aortic intima, and aortic root intimal macrophage content to an equivalent extent as FO, lending proof of principle for this strategy 17. However, whether a similar strategy will be atheroprotective for the n-6 PUFA pathway is unknown.

To address this gap in knowledge, we identified borage oil (BO) as a potential botanical oil to test this strategy in the n-6 PUFA conversion pathway. BO is enriched (~20%) in gamma linolenic acid (18:3 n-6; GLA), the immediate product of FADS2-mediated desaturation of LA.

GLA can be elongated to DGLA, a precursor of the anti-inflammatory PGE1, or arachidonic acid

(AA, 20:4 n-6), a precursor of proinflammatory eicosanoid species. The atheroprotective effects of BO have not been explored, and little is known about its effects on plasma lipids and lipoproteins. The purpose of this study was to directly compare the atheroprotective potential of

BO with EO and elucidate potential mechanisms for atheroprotection.

56

Materials and methods

Dietary oils

The seed oil of Borago officinalis L. (a member of the Boraginaceae family) and the seed oil of

Echium plantagineum L. (a member of the Boraginaceae family) were generous gifts from

Croda Europe Ltd. (Leek, Staffordshire, UK) and authenticated by the Wake Forest University

Center for Botanical Lipids and Inflammatory Disease Prevention. The seed oil of palm, Elaeis guineensis Jacq (a member of the Arecaceae family) was purchased from Shay and Company

(Portland, OR, USA). For these oils, a certificate of analysis is on file and retention samples are deposited at the Wake Forest School of Medicine. The fish oil source, Brevoortia tyrannis

Latrobe (a member of the Clupeidae family), was manufactured and generously provided by

Omega Protein (Reidsville, VA, USA) with a report of analysis on file for reference.

Animals and atherogenic diets (AD)

Female LDLrKO (C57BL/6 background) mice (5-6 weeks of age) were purchased from The

Jackson Laboratory (Bar Harbor, Maine, USA). Mice were housed in a specific pathogen-free facility on a 12h light/dark cycle. All protocols and procedures were approved by the Institutional

Animal Care and Use Committee. Mice were allowed to acclimate for 1-2 weeks during which time they ate a chow diet. At 7-8 weeks of age mice were randomly assigned to one of four AD groups (n=15/ diet group) containing 10% calories as PO and 0.2% cholesterol, supplemented with an additional 10% of calories as 1) PO, 2) BO (18:3 n-6 enriched), 3) EO (18:4 n-3 enriched) or 4) FO (20:5 n-3 and 22:6 n-3 enriched) for 16 weeks. The synthetic ADs were prepared by the diet kitchen in the Department of Pathology at Wake Forest School of Medicine as previously described 18. Detailed composition and quality control data for similar ADs have been published 16.

57

Body and organ weights

Mice were weighed every 2-4 weeks. After 16 weeks of AD feeding, body weights were recorded after a 4-hour fast. Blood was collected via the tail vein at baseline and after 2, 4, 8, and 16 weeks of AD feeding. Mice were then anesthetized using ketamine-xylazine and perfused via the left ventricle using cold PBS at the rate of ~3 ml/minute for 3-4 minutes before organs were harvested. After perfusion, liver wet weights were measured and normalized to terminal body weight.

Fatty acid analysis

Lipid extraction of lyophilized diets, RBCs, plasma, and liver was performed using the Bligh-

Dyer method 19. The total lipid extracts from plasma and liver were separated into cholesteryl ester (CE), triglyceride (TG), and phospholipid (PL) fractions by thin-layer chromatography.

Lipids were re-extracted from CE, TG, PL fractions and then trans-methylated using boron trifluoride 20, and percentage fatty acid composition was quantified by gas-liquid chromatography (GLC) as described previously 18. To estimate recovery of fatty acids from TLC and the transmethylation procedure, a known amount of tripentadecanoin (C15:0) and cholesteryl nonadecanoate (C19:0) were used as internal standards prior to TLC and tri- heptadecanoin (C17:0) was used as an internal standard prior to transmethylation. All internal standards were purchased from Nu-Chek-Prep, Inc, (T-145, Ch-818 and T-155). Percentage loss during TLC and transmethylation was <10% and <15%, respectively. Diet and RBC lipid extracts were directly trans-methylated for fatty acid analysis.

Plasma lipid and lipoprotein analysis

Plasma was isolated by low-speed centrifugation. Plasma total and free cholesterol (Wako), and triglyceride (Roche) concentrations were determined using enzymatic assays as described earlier 21 at baseline and after 2, 4, 8, and 16 weeks of AD feeding. Cholesteryl ester content

58 was calculated as (total-free) × 1.67; this calculation corrects for loss of fatty acid during the cholesterol esterase step of the assay. Data were expressed as area under the curve (AUC), which integrates plasma lipid concentrations over the 16 weeks, representing a time average estimate of hypercholesterolemia during the 16 week experiment. Plasma lipoprotein cholesterol mass distribution was determined using fast protein liquid chromatography (FPLC) fractionation on a Superose 6 10/300 column (GE Healthcare). Three equal volumes of plasma samples from each time point (5 mice/group) were pooled and subjected to FPLC fractionation and cholesterol quantification. VLDL, LDL and HDL cholesterol concentrations were determined and then expressed as AUC.

In vivo quantification of VLDL TG secretion rate

VLDL TG secretion rate was determined using detergent inhibition of intravascular TG lipolysis

20, 22. Tyloxapol (500 mg/kg body weight) was intravenously injected into anesthetized mice

(n=5/group) fed the AD for 16 weeks and fasted for 4 hours. Plasma TG concentrations before

(0 minutes), and 60, 120 and 180 minutes after injection were determined by enzymatic assay.

TG secretion rate was calculated using linear regression analysis to determine the slope of the time vs. TG concentration plot for each animal 20.

Liver lipid analysis

Livers were harvested at necropsy, flash frozen in liquid N2, and stored at -80°C. Liver lipids were quantified using a detergent-based enzymatic assay 23.

Western blot analysis

Nuclear proteins were prepared from frozen livers using ultracentrifugation as described 24, 25.

Equal aliquots (20 μg) of nuclear protein from individual livers were subjected to SDS-PAGE on

4-12% gels and transferred to a PVDF membrane. Immunoblot analyses were performed using

59 monoclonal anti-mouse SREBP-1 and rabbit polyclonal SREBP-2 antibodies as described 26.

Anti-SREBP1 and 2 antibodies were generously donated by Dr. Timothy Osborne (Sanford

Burnham Medical Research Institute). Anti-YY1 antibody (Abcam # 43058) was used as a nuclear loading control. Whole cell lysates were prepared using frozen livers 27 and equal aliquots of (50 g) of protein from individual livers were subjected to SDS-PAGE on 4-12% gels and transferred to PVDF membranes. Immunoblot analyses were performed using rabbit monoclonal anti-FAS antibody (Cell Signalling # 3180) and goat polyclonal anti SCD-1 antibody

(Santa Cruz Biotechnology # 14719).

Atherosclerotic lesion analysis

A subset of mice (n=3/group) was sacrificed after 8 weeks of AD feeding to measure aortic cholesteryl ester content. In the remaining mice, aortic cholesteryl ester content, aortic root intimal area, and intimal macrophage content (CD68+) were measured after 16 weeks of AD feeding 17. Aortic root intimal area was measured with Oil Red O staining.

Aortic oxidized-CE and aortic cholesterol content analysis

At necropsy, aortas were cleaned of visible adventitial fat and placed into a 15 ml etched glass tissue grinder containing 1 ml of 1:1 MeOH: H2O (v/v) and a known amount of internal standard

5-α cholestane. Aortas were homogenized on ice and neutral lipids were extracted twice using

75:25 isooctane: ethyl acetate (v/v). The combined organic layers were brought to a volume of 2 ml with 75:25 isooctane:ethyl acetate. One ml of the extract was added to an argon-purged ampule, heat-sealed, and shipped on dry ice to the University of Colorado Denver for LC-

MS/MS analysis. Before analysis, the organic layer was dried under N2. The residue was then weighed and diluted to 1 g/l with isooctane, and stored at -20°C until analysis of oxidized CE

(ox-CE) using LC/MS (/MS) 28. The remaining 1 ml of the original aortic extract in 75:25

60 isooctane: ethyl acetate was used to quantify aortic TC and FC content by GLC 17. Results were normalized to aortic wet weight.

Peritoneal macrophage studies

Thioglycollate-elicited peritoneal macrophages were isolated and cholesterol content was measured as described previously 21. Macrophage gene expression was measured after 2 hours of PBS or LPS (200 ng/ml) stimulation 21 and eicosanoids were identified and quantified using LC/MS/MS in macrophage- conditioned media 29, 30. In vitro macrophage chemotaxis in response to MCP-1 and MIP-1α was performed in a 48-well microtaxis chamber 31.

Real-time PCR analysis

Total RNA was isolated from mouse macrophages and liver using TRIzol (Invitrogen) and quantitative real-time RCR was performed to determine gene expression 21. Primer sequences have been reported previously 20, 27, 32. GAPDH was used as the control for normalization of results.

Statistics

Data are reported as mean ± SEM. Statistical analyses were performed using one-way ANOVA and individual paired diet comparisons were made using Tukey's post-hoc analysis. All statistical analyses were performed using GraphPad Prism software.

61

Results

Systemic response to atherogenic diets enriched in n-3 vs. n-6 fatty acid products of

FADS2

Dietary fatty acid compositions are given in Table 1 and show relative enrichment of 18-carbon fatty acids beyond FADS2 in the BO (19.7% 18:3 n-6) and EO (6% 18:4 n-3) diets. AD feeding over 16 weeks resulted in uniform food consumption (~3-4 g/day/mouse), body weight gain, and terminal liver/body weight ratios among all groups (Figure 1). Upon initiation of AD diet feeding, significant RBC fatty acid enrichment over baseline (chow diet) was evident within 1 week and appeared to reach equilibrium by 4 weeks (Figure 2). RBC fatty acid compositions reflected efficient in vivo elongation-desaturation of dietary 18:3 n-6 derived from BO to its longer chain

(20-carbon) counterparts, 20:3 n-6 and 20:4 n-6. Similarly, 18:4 n-3 derived from EO was elongated-desaturated to 20:5 n-3. These data indicate that dietary enrichment of 18-carbon fatty acid beyond FADS2 is sufficient to result in membrane enrichment of their respective 20- carbon chain counterparts. After 4 weeks of diet feeding, compared to chow (baseline), BO-fed mice had significant RBC fatty acid enrichment in GLA (~5.0% vs. 0.1%), DGLA (~4.0% vs. 1.1

%) and AA (~20.0% vs. 12.0%); whereas EO-fed mice had significant RBC fatty acid enrichment in ALA (~2.0% vs. 0.4%) and EPA (~3% vs. 2%). FO-fed mice had the greatest enrichment in EPA (~10% vs. 2%) and DHA (~12% vs. ~10.0 %), whereas PO-fed mice had the greatest enrichment in OA (~30% vs. 11%) and lowest enrichment in SA (below 10% vs. ~12-

13%).

Fatty acid composition was also measured in plasma and liver CE, TG, and PL fractions (Figure

3). In plasma from BO-fed mice, equivalent (~20%) AA enrichment was observed in CE, TG, and PL fractions relative to plasma from PO-fed mice (Figure 3 A-C). Plasma neutral lipids from mice fed BO were also relatively enriched in GLA (CE (~6%) and TG (~15%)), unlike plasma

62

PL. In BO-fed livers, AA was modestly, but significantly enriched in PL (~20%), but not in neutral lipids (below 5%) relative to the other diet groups (Figure 3 D-F). In general, PUFA enrichment in liver CE and TG was low for all diet groups relative to liver PL.

Plasma lipid and lipoprotein response to dietary BO vs. EO

Chow-fed mice had similar baseline measurements, but after 2 weeks of AD diet feeding, plasma TC, CE, and TG concentrations increased significantly, peaked by week 4, and remained elevated over the 16 weeks of diet feeding for PO-fed mice (Figure 4 A-C). This pattern was significantly attenuated in the other diet groups. BO and EO induced equivalent cholesterol lowering, whereas FO induced even further TC and CE lowering, compared to PO.

Plasma TG concentrations increased equivalently in PO and BO fed mice after 2 weeks of AD diet feeding, but stayed at baseline levels in the EO- and FO-fed mice.

A sharp increase in VLDL cholesterol (VLDL-c) concentrations occurred within 2 weeks of AD feeding that peaked and equilibrated by 4 weeks for PO-fed mice (Figure 5A). VLDL-c concentrations were significantly and equivalently attenuated in the other diet groups (Figure 5

A, D). LDL-c levels showed a similar pattern; only FO-fed mice had significantly lower LDL-c concentrations versus PO-fed mice (Figure 5B, E). HDL-c concentrations also increased sharply within 4 weeks among BO-, EO-, and FO-fed mice, and were relatively stable thereafter.

HDL-c concentrations were significantly lower in PO-fed mice compared to the other groups

(Figure 5C, F); however, HDL-c was <10% of the total pool of plasma cholesterol. Hence, most of the total cholesterol lowering came from decreases in VLDL-c. BO-induced reduction in

VLDL-c was equivalent to that of EO and FO, but BO did not reduce plasma TG concentrations.

EO, but not BO, diet reduces hepatic VLDL triglyceride secretion rate

We investigated whether the difference in plasma TG concentrations between EO vs. BO-fed mice could be explained by hepatic TG secretion. Using detergent inhibition of plasma TG

63 lipolysis, hepatic TG secretion rates were significantly lower for EO and FO-fed groups compared to BO and PO-fed mice (Figure 6). These data suggest that the higher plasma TG concentrations in BO vs. EO fed mice are likely due, in part, to a higher hepatic VLDL TG secretion rate.

BO- and EO-containing ADs are equally effective in reducing hepatosteatosis

We and others have shown that ADs enriched in FO reduce hepatic lipid content 16, 33-35. In the present study, ADs containing BO, EO or FO for 16 weeks reduced hepatic total neutral lipid content (i.e., TG and CE) relative to PO (Figure 7A), but FC and PL contents were similar among all diet groups. Furthermore, BO and EO were equally effective in reducing hepatic neutral lipid content. To determine the reason for the decreased hepatic neutral lipid content with BO and EO feeding, we examined hepatic lipogenic gene expression using quantitative real-time PCR. Compared to PO, all three ADs significantly lowered SREBP2 mRNA and its target gene HMG-CoA reductase, but not HMG-CoA synthase (Figure 7B). Although SREBP1-c mRNA expression levels were similar among all diet groups, expression of its target genes fatty acid synthase (FAS) and stearoyl-CoA desaturase-1 (SCD1) was significantly reduced in all three ADs relative to PO (Figure 7B). There were no significant differences among diet groups in LXRα and ABCA1 expression, whereas PPARα expression was significantly induced in FO- fed mice (Figure 7B). Since mice consuming the BO diet had reduced hepatosteatosis, similar to that of mice fed the EO or FO diets, we determined whether BO feeding attenuated hepatic inflammation as well. mRNA expression of CD68, TNF-α, and MCP-1 were significantly and equivalently reduced in BO, EO and FO-fed mice compared to those fed PO (Figure 7B).

Expression of other pro-inflammatory cytokines, such as IL-1β, IL-6, and IL-18, and the alternatively activated macrophage markers arginase-1 and anti-inflammatory cytokine IL-10 were similar among all four diet groups (data not shown). To determine whether FADS2

64 products reduce hepatic lipogenic gene expression via attenuation of the SREBP pathway, we measured accumulation of proteolytically cleaved mature/nuclear SREBP 1 and 2 isoforms in liver nuclear preparations. Content of nuclear SREBP 1 (but not 2) was significantly reduced in

BO-, EO-, and FO-fed mice relative to PO (Figure 7C). Hepatic protein content of the SREBP-

1c target genes FAS and SCD-1 was also significantly lower in those groups relative to PO

(Figure 7D).

BO and EO are equally atheroprotective

Aortic FC content was comparable among the groups after 8 weeks; however, CE content was significantly lower (~2 fold lower) in all groups vs. PO-fed mice (Figure 8A). After 16 weeks of

AD feeding, aortic FC and CE content was significantly lower in all groups (~ 2 fold reduction in

FC; ~3-4 fold reduction in CE) vs. PO-fed mice (Figure 8B), indicating that BO, EO and FO induced equivalent atheroprotection compared to PO. Significant aortic FC accumulation occurred between 8 and 16 weeks of AD feeding among all groups (~3-4 g/mg at 8 weeks vs.

6-14 g/mg at 16 weeks). CE accumulation increased modestly among all groups (~1-3 g/mg at 8 weeks vs. ~2-4 g/mg at 16 weeks) relative to the PO group (~4 g/mg at 8 weeks to ~16

g/mg at 16 weeks), indicating a disproportionate increase of CE deposition in PO-fed mice.

Aortic root intimal area was significantly lower (300-400 vs. 700 mm2) in the three PUFA- containing ADs compared to PO (Figure 8C), as was aortic root intimal macrophage content

(CD68+) (Figure 8D). Collectively, these results show that BO, EO, and FO were equally effective in reducing aortic atherosclerosis and macrophage content.

Mouse aortas lack of oxidized cholesteryl ester species

Human atheromas contain several distinct families of ox-CE species derived from PUFAs (e.g.

18:2 n-6, 20:4 n-6 and 22:6 n-3) that may play a role in atherogenesis 28. Since three of the four

65

ADs showed substantial PUFA enrichment, we examined ox-CE content in mouse aortas after

16 weeks of AD feeding. Aortic cholesteryl ester fatty acyl molecular species reflected the fatty acid enrichment of the diet; however, aortas had undetectable levels of ox-CE in all diet groups

(Figure 9). The relative abundances of the non-oxidized CE molecular species presented in

Figure 9 insets do not precisely reflect absolute differences in molecular species due to different electrospray ionization response factors (mass spectrometric parameters) for the polyunsaturated CEs as previously noted (29). Nonetheless, this raw data reveals dietary modification of PUFA-containing CE molecular species in aorta tissues reflecting the dietary fats.

Thioglycollate-elicited peritoneal macrophage eicosanoid release is similar for BO- and

EO-fed mice

We studied the effects of AD feeding on thioglycollate-elicited peritoneal macrophage eicosanoid biosynthesis. After 16 weeks of AD feeding, thioglycollate-elicited peritoneal macrophages were incubated ± LPS for 2 hours before media eicosanoids were quantified. This model of stimulated eicosanoid biosynthesis has been recently reported, including details of gene expressions related to PUFA metabolism stimulated by LPS (30). Among the eicosanoids measured, 12/15 lipoxygenase-derived LA metabolites 9 and 13 HODE were most abundant

(Figure 10). Furthermore, the type of dietary fat had little impact on production of thioglycollate- elicited peritoneal macrophage eicosanoid species in the basal state or after LPS stimulation.

Only in the FO group were significant reductions observed for generation of TXB2, PGE2, and 12 hydroxyeicosatetraenoic acids after LPS stimulation (~2-4 fold lowering vs. PO, EO and BO).

These data indicate that LPS stimulation revealed little differences in thioglycollate-elicited peritoneal macrophage eicosanoid biosynthesis among BO-, EO-, and PO-fed mice.

66

BO and EO attenuate macrophage inflammatory response to LPS

HODEs are natural ligands for macrophage PPARγ, which, when activated, can result in down- regulation of the canonical NF-κB pathway 36, 37. Thus, we measured expression of NF-κB target genes after LPS-induced activation in thioglycollate-elicited PMs isolated from mice after 16 weeks of AD feeding. Basal (-LPS) gene expression of the pro-inflammatory cytokines IL-6, IL-

1β, TNF-α, and chemokine MCP-1 was comparable among all groups (Figure 11A). LPS- treated macrophages from mice fed BO, EO, and FO vs. PO had decreased mRNA expression of IL-6, IL-1β and MCP-1, whereas TNF-α expression was significantly higher (Figure 11A).

However, 6 hours after LPS stimulation, TNF-α mRNA expression was significantly lower in macrophages from BO, EO and FO vs. PO-fed mice (data not shown), similar to the results for

IL-6, IL-1β and MCP-1 at 2 hours. mRNA abundance for the alternatively activated macrophage markers arginase 1 and CD206 was not different among groups (data not shown). Collectively, these data indicate that BO attenuates macrophage LPS-induced pro-inflammatory gene expression to the same extent as EO and FO, without affecting the gene expression of alternative-activated macrophage markers.

BO and EO equally attenuate macrophage CE accumulation and chemotaxis

Given the equivalent effectiveness of BO in attenuating aortic CE and macrophage content, we sought to determine whether macrophages from BO-fed mice contributed to reduced aortic atherosclerosis via reduced macrophage foam cell formation. We used thioglycollate-elicited

PMs (as a surrogate for aortic macrophages) from mice after 8 and 16 weeks of AD exposure to estimate cholesteryl ester content. At 8 weeks, macrophage CE content ranged from 5 to 20

g/mg among the groups, but only the FO-fed mice had significantly lower macrophage CE content compared to PO-fed mice (Figure 11B). After 16 weeks of AD feeding, macrophage CE content increased considerably (~2-8 fold) compared to 8 weeks and was significantly

67 attenuated in BO, EO and FO vs. PO-fed mice (Figure 11B). We next investigated whether reduced macrophage chemotaxis might partially explain the reduction in aortic root macrophage content in BO-, EO-, and FO-fed mice. All 3 PUFA containing ADs were equally effective in reducing chemotaxis to MIP-1α and MCP-1 compared to PO-fed mice (Figure 11C).

68

Discussion

In the current study, we tested the hypothesis that botanical oils enriched in 18:3 n-6 (BO) and

18:4 n-3 (EO) PUFAs beyond the rate-limiting FADS2 enzyme are equally atheroprotective compared to saturated/monounsaturated fat (PO) in LDLrKO mice. Although BO differs from EO and FO in its inability to lower plasma triglycerides, BO and EO were comparable in their ability to lower plasma cholesterol concentrations, especially VLDL-c. Additionally, BO was as effective as EO and FO in alleviating hepatic steatosis, and in regulating hepatic lipogenic and inflammatory gene expression compared to PO. As a result, BO and EO resulted in significant atheroprotection relative to PO at early (8 weeks) and advanced (16 weeks) atherosclerotic stages. We also report that ox-CEs were undetectable in early and advanced atherosclerotic arteries, suggesting that ox-CEs play a minimal role in murine atherosclerosis progression.

Additionally, BO and EO significantly and equivalently attenuated macrophage inflammatory gene expression, CE content, and chemotaxis. BO had these atheroprotective outcomes despite significant enrichment of RBC membranes, and plasma and liver lipids with 20:4 n-6, a precursor for several proinflammatory eicosanoids. Our results support the conclusion that botanical oils enriched in 18:3 n-6 and 18:4 n-3 PUFAs beyond the rate-limiting FADS2 enzyme are equally atheroprotective and hepatoprotective compared to saturated/monounsaturated fat.

Although previous studies in nonhuman primates and mice have demonstrated n-6 PUFA- mediated atheroprotection, these studies used dietary fats primarily enriched in LA 38-40. Our study focused on the hypothesis that dietary enrichment with n-6 PUFAs beyond FADS2 (i.e.,

GLA) would be as atheroprotective as EO, which we have previously demonstrated equal to FO in preventing atherosclerosis progression 17, 41. On the other hand, flaxseed oil, which is enriched in ALA, a substrate for FADS2, was not as atheroprotective as FO; this outcome occurred despite significant enrichment of liver phospholipids with EPA and was likely due to the lower plasma LDL-c concentrations in the FO vs. flaxseed oil group 9. The combined results of

69 the flaxseed oil and EO studies support our hypothesis that dietary enrichment in n-3 PUFAs beyond FADS2 is atheroprotective.

Our current results show that this is also true for the n-6 pathway of fatty acid elongation- desaturation. BO, which is enriched in GLA, was as atheroprotective as EO and FO despite its significant enrichment of RBCs, and plasma and liver lipid fractions with AA, a fatty acid precursor to pro-inflammatory leukotrienes and prostaglandins 42. Concerns that elevated membrane AA may result in increased cellular inflammation that exacerbates atherosclerosis lack support in human studies 42. Moreover, LA-enriched diets have not enriched AA in plasma and tissue lipid fractions 40, 43, 44, likely due to inefficient FADS2 conversion of LA to AA 45, 46. In addition, a recent meta-analysis of 13 cohort studies (involving 310,602 individuals and 12,479 coronary heart disease events) revealed an inverse association between dietary LA intake and coronary heart disease risk, such that, a 5% increase in energy intake from LA was associated with a 10% and 13% lower risk of coronary heart disease events and deaths, respectively 12.

Collectively, these results suggest that increased consumption of dietary LA and GLA is not harmful and is potentially atheroprotective in the general population. Our results also suggest that assessing cardiovascular risk and inflammatory potential by dietary n-3/n-6 PUFA ratio may be an oversimplification that does not take into account differences between 18 vs. ≥20 carbon

PUFAs.

A recent study reported a single nucleotide polymorphism (rs174537) in the FADS1/2 gene cluster that affects plasma AA levels 47. The GG allele, which is nearly twice as frequent in

African Americans as in European Americans, was associated with a small (~3%) but statistically significant enrichment in plasma AA. A subsequent study demonstrated that GG homozygotes for the rs174537 allele had increased plasma AA enrichment and production of

LTB4 and 5-hydroxyeicosatetraenoic acids in zymosan- stimulated blood, suggesting that genetic polymorphisms may influence tissue AA content and inflammatory response to external

70 pathogens 48. Thus, individuals harboring rs174537 homozygous GG alleles may be hyperresponsive to diets enriched in LA and GLA. Further studies are required to determine whether this potential hyperresponsiveness affects coronary heart disease risk.

Our results also suggest that BO and EO slow atherosclerosis progression in multiple ways – including reduced plasma VLDL-c concentrations, macrophage cholesterol content, inflammatory gene expression, and decreased macrophage migration towards a chemokine gradient. Plasma VLDL-c reduction likely had the greatest influence on atherosclerosis outcome in this study, since VLDL-c concentrations are the best lipoprotein/lipid predictor of aortic root atherosclerosis in LDLrKO mice 49 and we observed a strong positive association between plasma VLDL-c and aortic CE content (r2=0.88; p<0.0001, data not shown). LDL-c concentrations were similar among PO-, BO-, and EO-fed mice, suggesting they had a minimal effect on atheroprotection in BO and EO-fed mice. Plasma HDL-c was significantly elevated in all three diet groups and may have contributed to atheroprotection relative to the PO group; however, only a small fraction (~10%) of plasma cholesterol was distributed in HDL particles, making this a less likely possibility. In other studies investigating the influence of dietary n-3 and n-6 PUFA on atherogenesis, LDL was the predominant plasma atherogenic lipoprotein 9, 38.

Differences in plasma apoB lipoprotein response (VLDL vs. LDL) among these studies are likely due to genetic background (LDLrKO vs. apoB100 only-LDLrKO) and diet composition (0.02 vs.

0.2% cholesterol; 10% vs. 20% calories as fat).

A significant driver of atherogenesis in nonhuman primates and LDLrKO mice is hepatic production of saturated and monounsaturated CEs that are core constituents of secreted VLDL particles 11. Plasma VLDL particles undergo intravascular metabolism to become LDL particles, a major atherogenic lipoprotein particle in plasma. Deletion of the cholesterol esterification enzyme steroyl O-acyltransferase 2 (SOAT2) strikingly reduces atherosclerosis regardless of dietary fat saturation 38, supporting a critical role for SOAT2 in the generation of atherogenic

71

CEs in apoB-containing lipoproteins, such as VLDL and LDL. Hepatic CE content was significantly and equivalently reduced in mice fed BO and EO, suggesting the production of atherogenic CEs was blunted in these mice compared to those fed PO. This likely contributed to lower VLDL-c concentrations for the EO and BO groups, and reduced monounsaturated 18:1 n-

9 CE species in plasma at the expense of FADS2-derived CEs.

One of the earliest events in atherogenesis is arterial retention of apoB lipoproteins by proteoglycans 50. A recent study showed that SOAT2 KO vs. WT mice or mice fed FO vs. monounsaturated fat had plasma LDL enriched in PUFA CE species and bound with less affinity to proteoglycans, suggesting a mechanism by which apoB lipoproteins from mice fed dietary

PUFA may be less atherogenic 51. Although PUFA lipid species are more easily oxidized than their saturated and monounsaturated counterparts 52, 53 and ox-CE species have been identified in human tissue samples from endarterectomies 28, no arterial ox-CEs were detected in our study, despite significant enrichment of circulating and tissue lipids with AA in the BO group

(Figures 2, 3, and 9). Thus, ox-CEs did not appear contribute significantly to atherogenesis in our study.

CE-loaded macrophages are a hallmark of atherosclerosis, which result from chemokine- induced chemotaxis of monocytes to atherosclerotic lesions, differentiation of monocytes into macrophages expressing scavenger receptors, and unregulated uptake of modified apoB lipoproteins and apoB lipoprotein-proteoglycan complexes 54. In addition to reduced aortic cholesterol and aortic root intimal area in mice fed BO and EO vs. PO, we also observed decrease aortic root macrophage (CD68+ cells) content and decreased macrophage chemotaxis in vitro. Thioglycollate-elicited peritoneal macrophages isolated from BO and EO fed mice also had reduced sterol content and inflammatory response to LPS compared to their PO- fed counterparts. Macrophage FC accumulation resulting from genetic deletion of macrophage efflux genes ABCA1 and ABCG1 results in hyper-responsiveness to proinflammatory stimuli and

72 increased chemotaxis in vitro and in vivo 21, 55. However, in this study, the cholesterol elevation was due to CE, not FC, and there was no increase in macrophage ABCA1 and ABCG1 gene expression among diet groups (data not shown). These results suggest that decreased uptake of cholesterol from apoB lipoproteins likely explained the decreased macrophage atherogenic phenotype for mice fed EO and BO.

Diets enriched in n-3 PUFAs reduce hepatosteatosis 9, 16, 33, 56, unlike those containing n-6

PUFAs 27, 56. The reduced hepatic neutral lipid content in animals fed n-3 PUFAs is primarily mediated through decreased hepatic lipogenesis and is usually accompanied by reduced hepatic TG secretion 34, 57. In vitro, n-3 and n-6 PUFAs suppress SREBP-1c gene transcription

58, 59, decreasing hepatic lipogenesis, by competing with activators of LXR, a potent inducer of hepatic TG synthesis 58, 60. n-3 and n-6 PUFAs also increase mRNA degradation 61, inhibit the proteolytic processing of SREBP-1c in vitro 59, and accelerate the degradation of nuclear

SREBP-1c 62, reducing lipogenesis. Thus, we hypothesize that BO, but not the other n-6 PUFA- enriched diets (27,56), reduces hepatosteatosis relative to diets containing saturated/monounsaturated fatty acids via its ability to enrich liver lipids in AA through elongation and desaturation of GLA. Many n-6 PUFA-enriched botanical oils contain LA as the predominant fatty acyl species, accounting for 85-90% of n-6 PUFA consumption in the US 10.

However, as discussed above, diets enriched in LA do not result in plasma and tissue AA enrichment 44, 63. Genetic deletion of Elovl5, the gene encoding the enzyme that elongates 18:3 n-6 to 20:3 n-6 and 18:4 n-3 to 20:4 n-3, results in diminished hepatic lipid AA and DHA and increased neutral lipid storage 64. Feeding Elovl5 knockout mice AA or DHA rescued the hepatosteatosis phenotype by decreasing nuclear SREBP-1c and de novo lipogenesis with no changes in mRNA or membrane-bound SREBP-1c, supporting a role for AA in blocking SREBP-

1c cleavage and activation. In our study, liver membrane (i.e., phospholipid) AA content was elevated, nuclear SREBP-1 content was reduced, and SREBP-1c targeted genes (FAS, SCD-1)

73 were reduced in BO vs. PO fed mice, supporting a role for elevated AA in reducing hepatosteatosis in BO-fed mice. In another study, BO reduced ethanol-induced hepatosteatosis

65, suggesting that BO protects against hepatosteatosis regardless of the method of induction.

How elevated liver AA inhibits the proteolytic processing of membrane SREBP-1c is unclear, but may be related to fluidity of the endoplasmic reticulum membrane. Regardless of the detailed mechanism for decreasing hepatic lipogenesis, our study reveals a distinct advantage of BO in reducing hepatosteatosis and atherosclerosis in contrast to other n-6 PUFA-enriched botanical oils that are atheroprotective, but do not prevent hepatosteatosis.

Although BO, EO, and FO all reduced hepatosteatosis comparably relative to PO, BO did not reduce hepatic TG secretion nor plasma TG concentrations, whereas EO and FO did (Figure 6).

FO consistently reduces hepatic TG secretion, likely due to reduced hepatic lipogenesis 34, 41.

This paradox is likely due to the fact that only a small fraction of hepatic TG is mobilized for secretion; therefore, a large decrease in hepatic TG content does not necessarily result in decreased TG secretion. For example, knockdown of SCD-1 with a targeting anti-sense oligonucleotide resulted in a 90% reduction in hepatic TG content, but did not affect hepatic TG secretion compared to mice treated with a non-targeting ASO 66. We also observed a 50% decrease in newly synthesized TG secreted from livers of monkeys fed FO vs. lard diets, although liver TG synthesis was similar between diet groups 34. These combined results suggest a unique secretory pool of TG that may be regulated differently than the bulk TG storage pool in hepatocytes.

Collectively, our results support the hypothesis that dietary enrichment with FADS2 fatty acid products, such as SDA and GLA, results in membrane and plasma lipid enrichment in EPA and

AA, which in turn, is associated with reduced plasma lipids, atherosclerosis, and hepatosteatosis. This hypothesis was based on data in humans and animal models showing that conversion of 18-carbon PUFAs to ≥20 carbon PUFAs is inefficient, and that bypassing the

74 rate-limiting FADS2 step is possible by feeding botanical oils enriched in FADS2 products.

Furthermore, allowing for body surface area differences between mice and humans 67, achieving this dose of botanical oils in the diet is feasible. For example, our ADs contained BO and EO as 10% energy, which corresponds to a human equivalent dose of 0.81% energy (10%

× 3/37; 67), well within the range of n-3 PUFA doses administered in many randomized clinical trials 68 and the American Heart Association’s recommended n-3 PUFA intake for individuals with documented coronary heart disease 69. While replacing dietary saturated fat with PUFA is viewed as atheroprotective in general, our study suggests a more targeted approach of dietary

PUFA replacement using known biochemical pathways may enhance the beneficial outcomes of increased dietary PUFA consumption.

Acknowledgments

Anti-SREBP1 and 2 antibodies were generously donated by Dr. Timothy Osborne (Sanford

Burnham Medical Research Institute). We gratefully acknowledge Karen Klein (Biomedical

Research Services and Administration, Wake Forest School of Medicine) for editing the manuscript. This work was supported by grants from National Institute of Health P50 AT002782 and R01HL119962 (to JSP).

75

Reference List

1. Lloyd-Jones D, Adams R, Carnethon M, De Simone G, Ferguson TB, Flegal K, Ford E,

Furie K, Go A, Greenlund K, Haase N, Hailpern S, Ho M, Howard V, Kissela B, Kittner S,

Lackland D, Lisabeth L, Marelli A, McDermott M, Meigs J, Mozaffarian D, Nichol G, O'Donnell C,

Roger V, Rosamond W, Sacco R, Sorlie P, Stafford R, Steinberger J, Thom T, Wasserthiel-

Smoller S, Wong N, Wylie-Rosett J and Hong Y. Heart disease and stroke statistics--2009 update: a report from the American Heart Association Statistics Committee and Stroke Statistics

Subcommittee. Circulation. 2009;119:480-6.

2. Mozaffarian D, Micha R and Wallace S. Effects on coronary heart disease of increasing polyunsaturated fat in place of saturated fat: a systematic review and meta-analysis of randomized controlled trials. PLoS medicine. 2010;7:e1000252.

3. Bigger JT, Jr. and El-Sherif T. Polyunsaturated fatty acids and cardiovascular events: a fish tale. Circulation. 2001;103:623-5.

4. Calder PC and Grimble RF. Polyunsaturated fatty acids, inflammation and immunity.

European Journal of Clinical Nutrition. 2002;56 Suppl 3:S14-9.

5. Kang JX and Leaf A. Antiarrhythmic effects of polyunsaturated fatty acids. Recent studies. Circulation. 1996;94:1774-80.

6. Parks JS, Kaduck-Sawyer J, Bullock BC and Rudel LL. Effect of dietary fish oil on coronary artery and aortic atherosclerosis in African green monkeys. Arteriosclerosis.

1990;10:1102-12.

7. Hu FB, Bronner L, Willett WC, Stampfer MJ, Rexrode KM, Albert CM, Hunter D and

Manson JE. Fish and omega-3 fatty acid intake and risk of coronary heart disease in women.

JAMA : the journal of the American Medical Association. 2002;287:1815-21.

8. Horrobin DF. Fatty acid metabolism in health and disease: the role of delta-6- desaturase. American Journal of Clinical Nutrition. 1993;57:732S-736S; discussion 736S-737S.

76

9. Degirolamo C, Kelley KL, Wilson MD and Rudel LL. Dietary n-3 LCPUFA from fish oil but not alpha-linolenic acid-derived LCPUFA confers atheroprotection in mice. Journal of lipid research. 2010;51:1897-905.

10. Harris WS, Mozaffarian D, Rimm E, Kris-Etherton P, Rudel LL, Appel LJ, Engler MM,

Engler MB and Sacks F. Omega-6 fatty acids and risk for cardiovascular disease: a science advisory from the American Heart Association Nutrition Subcommittee of the Council on

Nutrition, Physical Activity, and Metabolism; Council on Cardiovascular Nursing; and Council on

Epidemiology and Prevention. Circulation. 2009;119:902-7.

11. Degirolamo C, Shelness GS and Rudel LL. LDL cholesteryl oleate as a predictor for atherosclerosis: evidence from human and animal studies on dietary fat. Journal of lipid research. 2009;50 Suppl:S434-9.

12. Farvid MS, Ding M, Pan A, Sun Q, Chiuve SE, Steffen LM, Willett WC and Hu FB.

Dietary Linoleic Acid and Risk of Coronary Heart Disease: A Systematic Review and Meta-

Analysis of Prospective Cohort Studies. Circulation. 2014;130:1568-1578.

13. Harris WS and Shearer GC. Omega-6 Fatty Acids and Cardiovascular Disease: Friend or Foe? Circulation. 2014;130:1562-1564.

14. Chopra J and Webster RO. PGE1 inhibits neutrophil adherence and neutrophil-mediated injury to cultured endothelial cells. Am Rev Respir Dis. 1988;138:915-20.

15. Jones DA and Fitzpatrick FA. "Suicide" inactivation of thromboxane A2 synthase.

Characteristics of mechanism-based inactivation with isolated enzyme and intact platelets. The

Journal of biological chemistry. 1990;265:20166-71.

16. Zhang P, Boudyguina E, Wilson MD, Gebre AK and Parks JS. Echium oil reduces plasma lipids and hepatic lipogenic gene expression in apoB100-only LDL receptor knockout mice. J Nutr Biochem. 2008;19:655-63.

17. Brown AL, Zhu X, Rong S, Shewale S, Seo J, Boudyguina E, Gebre AK, Alexander-

Miller MA and Parks JS. Omega-3 fatty acids ameliorate atherosclerosis by favorably altering

77 monocyte subsets and limiting monocyte recruitment to aortic lesions. Arteriosclerosis, thrombosis, and vascular biology. 2012;32:2122-30.

18. Rudel LL, Kelley K, Sawyer JK, Shah R and Wilson MD. Dietary monounsaturated fatty acids promote aortic atherosclerosis in LDL receptor-null, human ApoB100-overexpressing transgenic mice. Arteriosclerosis, thrombosis, and vascular biology. 1998;18:1818-27.

19. Bligh EG and Dyer WJ. A rapid method of total lipid extraction and purification. Canadian journal of biochemistry and physiology. 1959;37:911-7.

20. Bi X, Zhu X, Gao C, Shewale S, Cao Q, Liu M, Boudyguina E, Gebre AK, Wilson MD,

Brown AL and Parks JS. Myeloid cell-specific ATP-binding cassette transporter A1 deletion has minimal impact on atherogenesis in atherogenic diet-fed low-density lipoprotein receptor knockout mice. Arteriosclerosis, thrombosis, and vascular biology. 2014;34:1888-99.

21. Zhu X, Lee JY, Timmins JM, Brown JM, Boudyguina E, Mulya A, Gebre AK, Willingham

MC, Hiltbold EM, Mishra N, Maeda N and Parks JS. Increased cellular free cholesterol in macrophage-specific Abca1 knock-out mice enhances pro-inflammatory response of macrophages. The Journal of biological chemistry. 2008;283:22930-41.

22. Millar JS, Cromley DA, McCoy MG, Rader DJ and Billheimer JT. Determining hepatic triglyceride production in mice: comparison of poloxamer 407 with Triton WR-1339. Journal of lipid research. 2005;46:2023-8.

23. Carr TP, Andresen CJ and Rudel LL. Enzymatic determination of triglyceride, free cholesterol, and total cholesterol in tissue lipid extracts. Clin Biochem. 1993;26:39-42.

24. Hattori M, Tugores A, Veloz L, Karin M and Brenner DA. A simplified method for the preparation of transcriptionally active liver nuclear extracts. DNA and cell biology. 1990;9:777-

81.

25. Lee JH, Giannikopoulos P, Duncan SA, Wang J, Johansen CT, Brown JD, Plutzky J,

Hegele RA, Glimcher LH and Lee AH. The transcription factor cyclic AMP-responsive element- binding protein H regulates triglyceride metabolism. Nature medicine. 2011;17:812-5.

78

26. Jeon TI, Esquejo RM, Roqueta-Rivera M, Phelan PE, Moon YA, Govindarajan SS, Esau

CC and Osborne TF. An SREBP-responsive microRNA operon contributes to a regulatory loop for intracellular lipid homeostasis. Cell metabolism. 2013;18:51-61.

27. Rong S, Cao Q, Liu M, Seo J, Jia L, Boudyguina E, Gebre AK, Colvin PL, Smith TL,

Murphy RC, Mishra N and Parks JS. Macrophage 12/15 lipoxygenase expression increases plasma and hepatic lipid levels and exacerbates atherosclerosis. Journal of lipid research.

2012;53:686-95.

28. Hutchins PM, Moore EE and Murphy RC. Electrospray MS/MS reveals extensive and nonspecific oxidation of cholesterol esters in human peripheral vascular lesions. Journal of lipid research. 2011;52:2070-83.

29. Murphy RC, Barkley RM, Zemski Berry K, Hankin J, Harrison K, Johnson C, Krank J,

McAnoy A, Uhlson C and Zarini S. Electrospray ionization and tandem mass spectrometry of eicosanoids. Analytical biochemistry. 2005;346:1-42.

30. Maurya MR, Gupta S, Li X, Fahy E, Dinasarapu AR, Sud M, Brown HA, Glass CK,

Murphy RC, Russell DW, Dennis EA and Subramaniam S. Analysis of inflammatory and lipid metabolic networks across RAW264.7 and thioglycolate-elicited macrophages. Journal of lipid research. 2013;54:2525-42.

31. Zhu X, Westcott MM, Bi X, Liu M, Gowdy KM, Seo J, Cao Q, Gebre AK, Fessler MB,

Hiltbold EM and Parks JS. Myeloid cell-specific ABCA1 deletion protects mice from bacterial infection. Circulation research. 2012;111:1398-409.

32. Bi X, Zhu X, Gao C, Shewale S, Cao Q, Liu M, Boudyguina E, Gebre AK, Wilson MD,

Brown AL and Parks JS. Myeloid Cell–Specific ATP-Binding Cassette Transporter A1 Deletion

Has Minimal Impact on Atherogenesis in Atherogenic Diet–Fed Low-Density Lipoprotein

Receptor Knockout Mice. Arteriosclerosis, thrombosis, and vascular biology. 2014;34:1888-

1899.

79

33. Brown JM, Chung S, Sawyer JK, Degirolamo C, Alger HM, Nguyen TM, Zhu X, Duong

MN, Brown AL, Lord C, Shah R, Davis MA, Kelley K, Wilson MD, Madenspacher J, Fessler MB,

Parks JS and Rudel LL. Combined therapy of dietary fish oil and stearoyl-CoA desaturase 1 inhibition prevents the metabolic syndrome and atherosclerosis. Arteriosclerosis, thrombosis, and vascular biology. 2010;30:24-30.

34. Parks JS, Johnson FL, Wilson MD and Rudel LL. Effect of fish oil diet on hepatic lipid metabolism in nonhuman primates: lowering of secretion of hepatic triglyceride but not apoB.

Journal of lipid research. 1990;31:455-66.

35. Parks JS, Wilson MD, Johnson FL and Rudel LL. Fish oil decreases hepatic cholesteryl ester secretion but not apoB secretion in African green monkeys. Journal of lipid research.

1989;30:1535-44.

36. Nagy L, Tontonoz P, Alvarez JG, Chen H and Evans RM. Oxidized LDL regulates macrophage gene expression through ligand activation of PPARgamma. Cell. 1998;93:229-40.

37. Ricote M, Li AC, Willson TM, Kelly CJ and Glass CK. The peroxisome proliferator- activated receptor-gamma is a negative regulator of macrophage activation. Nature.

1998;391:79-82.

38. Bell TA, 3rd, Kelley K, Wilson MD, Sawyer JK and Rudel LL. Dietary fat-induced alterations in atherosclerosis are abolished by ACAT2-deficiency in ApoB100 only, LDLr-/- mice.

Arteriosclerosis, thrombosis, and vascular biology. 2007;27:1396-402.

39. Rudel LL, Johnson FL, Sawyer JK, Wilson MS and Parks JS. Dietary polyunsaturated fat modifies low-density lipoproteins and reduces atherosclerosis of nonhuman primates with high and low diet responsiveness. The American journal of clinical nutrition. 1995;62:463S-470S.

40. Rudel LL, Parks JS and Sawyer JK. Compared with dietary monounsaturated and saturated fat, polyunsaturated fat protects African green monkeys from coronary artery atherosclerosis. Arteriosclerosis, thrombosis, and vascular biology. 1995;15:2101-10.

80

41. Forrest LM, Boudyguina E, Wilson MD and Parks JS. Echium oil reduces atherosclerosis in apoB100-only LDLrKO mice. Atherosclerosis. 2012;220:118-21.

42. Harris WS and Shearer GC. Omega-6 Fatty Acids and Cardiovascular Disease: Friend,

Not Foe? Circulation. 2014;130:1562-1564.

43. Thornburg JT, Parks JS and Rudel LL. Dietary fatty acid modification of HDL phospholipid molecular species alters lecithin: cholesterol acyltransferase reactivity in cynomolgus monkeys. Journal of lipid research. 1995;36:277-89.

44. Rett BS and Whelan J. Increasing dietary linoleic acid does not increase tissue arachidonic acid content in adults consuming Western-type diets: a systematic review. Nutrition

& metabolism. 2011;8:36.

45. Demmelmair H, Iser B, Rauh-Pfeiffer A and Koletzko B. Comparison of bolus versus fractionated oral applications of [13C]-linoleic acid in humans. European journal of clinical investigation. 1999;29:603-9.

46. Hussein N, Ah-Sing E, Wilkinson P, Leach C, Griffin BA and Millward DJ. Long-chain conversion of [13C]linoleic acid and alpha-linolenic acid in response to marked changes in their dietary intake in men. Journal of lipid research. 2005;46:269-80.

47. Mathias RA, Sergeant S, Ruczinski I, Torgerson DG, Hugenschmidt CE, Kubala M,

Vaidya D, Suktitipat B, Ziegler JT, Ivester P, Case D, Yanek LR, Freedman BI, Rudock ME,

Barnes KC, Langefeld CD, Becker LC, Bowden DW, Becker DM and Chilton FH. The impact of

FADS genetic variants on omega6 polyunsaturated fatty acid metabolism in African Americans.

BMC genetics. 2011;12:50.

48. Hester AG, Murphy RC, Uhlson CJ, Ivester P, Lee TC, Sergeant S, Miller LR, Howard

TD, Mathias RA and Chilton FH. Relationship between a Common Variant in the Fatty Acid

Desaturase (FADS) Cluster and Eicosanoid Generation in Humans. Journal of Biological

Chemistry. 2014;289:22482-9.

81

49. VanderLaan PA, Reardon CA, Thisted RA and Getz GS. VLDL best predicts aortic root atherosclerosis in LDL receptor deficient mice. Journal of lipid research. 2009;50:376-85.

50. Skalen K, Gustafsson M, Rydberg EK, Hulten LM, Wiklund O, Innerarity TL and Boren J.

Subendothelial retention of atherogenic lipoproteins in early atherosclerosis. Nature.

2002;417:750-4.

51. Melchior JT, Sawyer JK, Kelley KL, Shah R, Wilson MD, Hantgan RR and Rudel LL. LDL particle core enrichment in cholesteryl oleate increases proteoglycan binding and promotes atherosclerosis. Journal of lipid research. 2013;54:2495-503.

52. Thomas MJ, Thornburg T, Manning J, Hooper K and Rudel LL. Fatty acid composition of low-density lipoprotein influences its susceptibility to autoxidation. Biochemistry. 1994;33:1828-

34.

53. Whitman SC, Fish JR, Rand ML and Rogers KA. n-3 fatty acid incorporation into LDL particles renders them more susceptible to oxidation in vitro but not necessarily more atherogenic in vivo. Arteriosclerosis, thrombosis, and vascular biology. 1994;14:1170-6.

54. Moore KJ and Tabas I. Macrophages in the pathogenesis of atherosclerosis. Cell.

2011;145:341-55.

55. Yvan-Charvet L, Welch C, Pagler TA, Ranalletta M, Lamkanfi M, Han S, Ishibashi M, Li

R, Wang N and Tall AR. Increased inflammatory gene expression in ABC transporter-deficient macrophages: free cholesterol accumulation, increased signaling via toll-like receptors, and neutrophil infiltration of atherosclerotic lesions. Circulation. 2008;118:1837-47.

56. Bell TA, 3rd, Wilson MD, Kelley K, Sawyer JK and Rudel LL. Monounsaturated fatty acyl-coenzyme A is predictive of atherosclerosis in human apoB-100 transgenic, LDLr-/- mice.

Journal of lipid research. 2007;48:1122-31.

57. Jump DB and Clarke SD. Regulation of gene expression by dietary fat. Annual review of nutrition. 1999;19:63-90.

82

58. Ou J, Tu H, Shan B, Luk A, DeBose-Boyd RA, Bashmakov Y, Goldstein JL and Brown

MS. Unsaturated fatty acids inhibit transcription of the sterol regulatory element-binding protein-

1c (SREBP-1c) gene by antagonizing ligand-dependent activation of the LXR. Proceedings of the National Academy of Sciences of the United States of America. 2001;98:6027-32.

59. Hannah VC, Ou J, Luong A, Goldstein JL and Brown MS. Unsaturated fatty acids down- regulate srebp isoforms 1a and 1c by two mechanisms in HEK-293 cells. The Journal of biological chemistry. 2001;276:4365-72.

60. Chen G, Liang G, Ou J, Goldstein JL and Brown MS. Central role for liver X receptor in insulin-mediated activation of Srebp-1c transcription and stimulation of fatty acid synthesis in liver. Proceedings of the National Academy of Sciences of the United States of America.

2004;101:11245-50.

61. Xu J, Teran-Garcia M, Park JH, Nakamura MT and Clarke SD. Polyunsaturated fatty acids suppress hepatic sterol regulatory element-binding protein-1 expression by accelerating transcript decay. The Journal of biological chemistry. 2001;276:9800-7.

62. Botolin D, Wang Y, Christian B and Jump DB. Docosahexaneoic acid (22:6,n-3) regulates rat hepatocyte SREBP-1 nuclear abundance by Erk- and 26S proteasome-dependent pathways. Journal of lipid research. 2006;47:181-92.

63. Rioux FM and Innis SM. Arachidonic acid concentrations in plasma and liver phospholipid and cholesterol esters of piglets raised on formulas with different linoleic and linolenic acid contents. The American journal of clinical nutrition. 1992;56:106-12.

64. Moon YA, Hammer RE and Horton JD. Deletion of ELOVL5 leads to fatty liver through activation of SREBP-1c in mice. Journal of lipid research. 2009;50:412-23.

65. Lukivskaya OY, Naruta E, Sadovnichy V, Kirko S and Buko VU. Reversal of experimental ethanol-induced liver steatosis by borage oil. Phytotherapy research : PTR.

2012;26:1626-31.

83

66. Brown JM, Chung S, Sawyer JK, Degirolamo C, Alger HM, Nguyen T, Zhu X, Duong

MN, Wibley AL, Shah R, Davis MA, Kelley K, Wilson MD, Kent C, Parks JS and Rudel LL.

Inhibition of stearoyl-coenzyme A desaturase 1 dissociates insulin resistance and obesity from atherosclerosis. Circulation. 2008;118:1467-75.

67. Reagan-Shaw S, Nihal M and Ahmad N. Dose translation from animal to human studies revisited. FASEB J. 2008;22:659-61.

68. Rizos EC, Ntzani EE, Bika E, Kostapanos MS and Elisaf MS. Association between omega-3 fatty acid supplementation and risk of major cardiovascular disease events: a systematic review and meta-analysis. JAMA : the journal of the American Medical Association.

2012;308:1024-33.

69. Kris-Etherton PM, Harris WS, Appel LJ and American Heart Association. Nutrition C.

Fish consumption, fish oil, omega-3 fatty acids, and cardiovascular disease. Circulation.

2002;106:2747-57.

84

Figures

Fig. 1. Body weight gain and terminal liver/body weight ratios. (A) Body weight gain was monitored periodically from baseline (chow diet) to 16 weeks of feeding the indicated atherogenic diets. (B) Mice were then euthanized and liver wet weights were measured and normalized to terminal body weight. Data are expressed as mean ± SEM; n=15/diet group. No significant differences were found by one-way ANOVA.

Fig. 2. RBC fatty acid (FA) composition. RBCs were harvested from mice consuming the indicated atherogenic diets from baseline (chow diet) to 10 weeks; percentage of fatty acid distribution was measured as described in the methods. Data for individual fatty acids are expressed as percentage composition of total fatty acids. An equal volume pool of RBCs from 8 mice/diet group/week was used for analysis. Horizontal line denotes extrapolation of chow data

(0 weeks) across the 10-week period.

Fig. 3. Percentage fatty acid composition of plasma and liver lipids. LDLrKO mice were fed the indicated atherogenic diets for 16 weeks before harvesting plasma and liver for fatty acid analysis. Data for individual fatty acids are expressed as percentage (mean ± SEM) composition of total fatty acids; n=3/diet group. Bars with different letters denote significant (p<0.05) differences among diet groups by one-way ANOVA and Tukey’s post-hoc analysis.

Fig. 4. Plasma lipid concentrations. Fasting (4h) plasma (A) total cholesterol (TC), (B) cholesteryl ester (CE), and (C) triglyceride (TG) concentrations in LDLrKO mice during a 16 week AD feeding were measured by enzymatic assays (n=15). CE= (TC-FC)*1.67. Plasma TC,

CE and TG concentrations were integrated over the 16-week study and expressed as area under the curve (AUC) (D, E, F). Data are expressed as mean ± SEM, n=15/diet. Groups with

85 different letters are significantly different (p<0.05) by one-way ANOVA and Tukey’s post-hoc analysis.

Fig. 5. Plasma lipoprotein cholesterol distribution. LDLrKO mice were fed the indicated atherogenic diets for 16 weeks. Fasted (4h) plasma was harvested and fractionated by FPLC and cholesterol distributions among VLDL (A), LDL (B) and HDL (C) fractions was measured.

Plasma VLDL-c (D), LDL-c (E) and HDL-c (F) concentrations were integrated over the 16-week study and expressed as area under the curve (AUC). Data are expressed as mean ± SEM, n=3 equal volume pooled samples from 5 mice/group. Groups with different letters are significantly different (p<0.05) by one-way ANOVA and Tukey’s post-hoc analysis.

Fig. 6. Hepatic VLDL-TG secretion rate. (A) Plasma TG levels were measured by enzymatic assay before (0 min) and after (60, 120, 180 min) intravenous Triton® injection. (B) Hepatic TG secretion rate during the 3h experiment was calculated for each animal as the slope of the regression line. Data are plotted as mean ± SEM, n=4-5. Groups with different letters are significantly different (p<0.05) by one-way ANOVA and Tukey’s post-hoc analysis.

Fig. 7. Hepatic response to atherogenic diets. LDLrKO mice were fed the indicated atherogenic diets for 16 weeks before liver was harvested to measure lipid content and gene expression. (A) Hepatic lipid content. Neutral lipid= CE+TG. (B) Hepatic mRNA expression of genes involved in cholesterol biosynthesis, lipogenesis, and inflammation. (C) Nuclear accumulation of mature SREBP-1c and 2 isoforms in liver. (D) Hepatic FAS and SCD-1 protein content. Data are expressed as mean ± SEM, n=5. Groups with different letters are significantly different (p<0.05) by one-way ANOVA and Tukey’s post-hoc analysis.

86

Fig. 8. Aortic atherosclerosis. LDLrKO mice were fed the indicated atherogenic diets for 16 weeks before aortas and hearts were harvested for atherosclerosis quantification. Aortas were lipid extracted for quantification of total cholesterol and free cholesterol (FC) content using gas- liquid chromatography. Cholesterol ester (CE) content was calculated as (TC-FC) x1.67. (A)

Aortic FC and CE content of mice fed PO, BO, EO and FO for 8 wks or (B) 16 wks. Each data point represents an individual mouse aorta. (C) Quantification of aortic root Oil Red O positive intimal area (lesion area) and representative Oil Red O-stained aortic root sections. (D)

Quantification of percentage aortic root lesional area occupied by CD68+ cells and representative CD68+ immunohistochemically stained aortic root sections. Each point represents the average lesion area of 6-8 sections per mouse. Horizontal lines denote the mean for each diet group. Groups with different letters are significantly different (p<0.05) by one-way

ANOVA and Tukey’s post-hoc analysis.

Fig. 9. Mouse atherosclerotic plaque oxidized cholesteryl ester analysis. LDLrKO mice were fed the indicated atherogenic diets for 16 weeks before aortas were harvested for oxidized

CE analysis using normal phase-HPLC-MS/MS. Ion chromatograms of cholesteryl ester (2-4 min elution) and oxidized cholesteryl esters (6-36 min elution) for one aorta from each diet group. Inset presents the raw mass spectrometric data corresponding to CE molecular species labeled with m/z values and their acyl components, denoted by total acyl carbons and double bonds in this qualitative study. No oxCE species were detectable for any of the four diet groups.

Fig. 10. LPS-stimulated eicosanoid release from peritoneal macrophages. LDLrKO mice were fed the indicated atherogenic diets for 16 weeks. Mice were then injected with thioglycollate in the peritoneal cavity and macrophages were isolated 4 days later as described in the methods. The macrophages were cultured for 3 hours in serum-free medium and then

87 treated with LPS (200 ng/ml) or PBS (-LPS) for 2 hours before 1 ml media were collected for eicosanoid quantification by LC-MS/MS analysis. Eicosanoids below the minimum detection level are not shown. Data are expressed as mean ± SEM, n=5/group. Groups with different letters are significantly different (p<0.05) by one-way ANOVA and Tukey’s post-hoc analysis.

Fig. 11. Macrophage inflammation, foam cell formation, and chemotaxis. LDLrKO mice were fed the indicated atherogenic diets for 8 or 16 weeks before thioglycollate-elicited macrophages were isolated. (A) Inflammatory gene expression was measured using RT-PCR after 2 hour treatment with LPS (200 ng/ml) or PBS (-LPS) in thioglycollate-elicited peritoneal macrophages. (B) Peritoneal macrophage cholesteryl ester content was measured using gas- liquid chromatography after 8 and 16 weeks of atherogenic diet feeding. (C) Ex vivo chemotaxis of thioglycollate-elicited peritoneal macrophages towards MCP-1 and MIP-1α was measured after 16 weeks of atherogenic diet feeding. Data are expressed as mean ± SEM, n=5. In panels

B and C, data points for individual mice are also shown. Groups with different letters are significantly different (p<0.05) by one-way ANOVA and Tukey’s post-hoc analysis.

88

TABLE 1. Fatty acid composition (% FA) and total energy equivalence (% EE) of individual fatty acids in each atherogenic diet

Palm oil (PO) Borage oil (BO) Echium oil (EO) Fish oil (FO)

Fatty Acid % FA % EE % FA % EE % FA % EE % FA % EE

Palmitic Acid 43.2 8.64 24.5 4.9 25.6 5.12 30.6 6.12

(C16:0)

Palmitoleic Acid 0.37 0.074 0.3 0.06 0.4 0.08 4.2 0.84

(C16:1)

Stearic Acid 4.5 0.90 4.4 0.88 4.1 0.82 4.5 0.9

(C18:0)

Oleic Acid 37.3 7.46 24.1 4.82 26.1 5.22 24.9 4.98

(C18:1 n-9)

Linoleic Acid 11.1 2.22 17.4 3.48 15.4 3.08 8.2 1.64

(C18:2 n-6), LA

α-Linolenic Acid 0.37 0.074 0.5 0.11 13.8 2.76 1 0.2

(C18:3 n-3), ALA

γ-Linolenic Acid 19.7 3.94 5.0 1.0

(C18:3 n-6), GLA

Stearidonic Acid 0.2 0.04 6.0 1.2 1.5 0.3

(C18:4 n-3), SDA

89

Euric acid 2.6 0.52 0.2 0.04

(C22:1 n-9)

Eicosapentaenoic 0.1 0.02 0.3 0.06 0.3 0.06 6.9 1.38

Acid (C20:5 n-3),

EPA

Docasahexaenoic 0.2 0.04 0.3 0.06 0.3 0.06 7 1.40

Acid (C22:6 n-3),

DHA

Diets contained 0.2% cholesterol + 10% calories as Palm Oil (PO) + 10% calories as: PO,

Borage oil (BO), Echium oil (EO) or Fish oil (FO). Percent (%) fatty acid composition of PO, BO,

EO and FO diets determined using gas-liquid chromatography. % EE for individual fatty acid was calculated using total energy derived from fatty acids (i.e. 20%) / diet and % fatty acid composition of respective diet.

90

Fig. 1.

91

Fig. 2.

92

Fig. 3.

93

Fig. 4.

94

Fig. 5.

95

Fig. 6.

96

Fig. 7.

97

Fig. 8.

98

Fig. 9.

99

Fig. 10.

100

Fig. 11.

101

CHAPTER III

IN VIVO ACTIVATION OF LEUKOCYTE GPR120 BY POLYUNSATURATED FATTY ACIDS HAS MINIMAL IMPACT ON ATHEROSCLEROSIS IN LDLrKO MICE

Swapnil V. Shewale, Amanda L. Brown, Xin Bi, Elena Boudyguina, Martha Alexander-Miller, Da

Young Oh, Jerrold M. Olefsky, John S. Parks.

This manuscript will be submitted to Circulation Research.

Stylistic variations are due to the requirements of the journal.

Swapnil V. Shewale performed the experiments and prepared the manuscript.

Dr. John S. Parks acted in an advisory and editorial capacity.

102

In vivo activation of leukocyte GPR120 by polyunsaturated fatty acids has minimal impact on atherosclerosis in LDLrKO mice

Shewale: Leukocyte GPR120 and PUFA-induced atheroprotection

Swapnil V. Shewale 1, 2, Amanda L. Brown1, Xin Bi1, Elena Boudyguina1, Martha Alexander- Miller3, Da Young Oh4, Jerrold M. Olefsky4, John S. Parks1.

1Departments of Internal Medicine/Section on Molecular Medicine, 2Physiology/Pharmacology, and 3Microbiology and Immunology, Wake Forest School of Medicine, Winston-Salem, NC 27157 and 4Department of Medicine, University of San Diego, La Jolla, CA 92093

Address correspondence to:

Dr. John S. Parks,

Department of Internal Medicine/Section on Molecular Medicine,

Wake Forest School of Medicine,

Medical Center Blvd, Winston-Salem, NC 27157, USA.

Phone: 336-716-2145;

Fax: 336-716-6279;

Email: [email protected]

Total word count: XXXX/7000 (including Title Page, Abstract, Text, References, Tables and Figures Legends)

Subject Codes:

[90] Lipid and lipoprotein metabolism

[130] Animal models of human disease

103

ABSTRACT

Rationale: G protein-coupled receptor 120 (GPR120) activation by n-3 polyunsaturated fatty acids (PUFAs) attenuates NF-κB inflammatory signaling. However, the impact of GPR120 expression on atherosclerosis is unknown.

Objective: To determine whether in vivo activation of leukocyte GPR120 by n-3 vs. n-6 PUFA is atheroprotective.

Methods and Results: Leukocyte GPR120 wildtype (WT) or knockout (KO) mice in the LDL receptor knockout background were generated by bone marrow transplantion. Mice were fed one of the four atherogenic diets containing 0.2% cholesterol and 10% calories as palm oil (PO)

+ 10% calories as: 1) PO, 2) fish oil (FO; 20:5 n-3 and 22:6 n-3 enriched), 3) echium oil (EO;

18:4 n-3 enriched), or 4) borage oil (BO; 18:3 n-6 enriched) for 16 weeks. Plasma lipids and lipoproteins, hepatic lipid content, neutrophilia, monocytosis, aortic root monocyte recruitment, and aortic atherosclerosis were analyzed. Compared to PO, mice fed BO, EO and FO had significantly reduced plasma cholesterol, triglycerides, VLDL cholesterol, hepatic neutral lipid, and atherosclerosis that were equivalent for WT and KO mice, demonstrating that leukocyte

GPR120 expression did not affect these outcomes. In BO, EO and FO, but not PO-fed mice, lack of leukocyte GPR120 resulted in neutrophilia, pro-inflammatory Ly6Chi monocytosis, increased monocyte recruitment into aortic roots, and increased hepatic inflammatory gene expression.

Conclusions: We conclude that leukocyte GRP120 expression has minimal effects on dietary

PUFA-induced plasma lipid/lipoprotein reduction and atheroprotection, and that there is no distinction between n-3 vs. n-6 PUFAs in activating anti-inflammatory effects of leukocyte

GPR120 in vivo.

104

Keywords:

GPR120, atherosclerosis, n-3 PUFA, n-6 PUFA, hepatosteatosis, monocytes

Non-standard Abbreviations and Acronyms:

AD: Atherogenic diet

AP-1: activator protein-1

BMT: Bone marrow transplantation

CMPs: common myeloid progenitors

CVD: cardiovascular disease

CE: Cholesteryl ester

ECM: Extracellular matrix

FADS2: fatty acid desaturase 2 (FADS2) / delta six fatty acid desaturase (D6D) enzyme

FFAR: free fatty acid receptor

GPR120: G protein coupled receptor 120

HSCPs: hematopoeitic stem cell precursor (HSCPs)

IKK: I kappaB kinase

JNK: c-Jun N-terminal kinase

PUFA: polyunsaturated fatty acids

PGE1/2: prostaglandin E1/2

105

LDLrKO: LDL receptor knockout mice

LP: Lipoproteins

LPS: Lipopolysaccharide

MMP’s: matrix metalloproteases n-3/n-6 PUFA: omega-3/omega-6 polyunsaturated fatty acids

NF-κB: nuclear factor kappa-light-chain-enhancer of activated B cells

SOAT2: cholesterol esterification enzyme steroyl O-acyltransferase 2

SREBP: Sterol responsive element binding protein

106

INTRODUCTION

Deorphanization of free fatty acid receptors (FFARs) has allowed further understanding of free fatty acids (FA) as signaling molecules. These class A, G-protein coupled receptors

(GPRs) include GPR40 (FFAR1), GPR43 (FFAR2), GPR41 (FFAR3), GPR84 and GPR120

(FFAR4) that are activated by short, medium or long chain FA1-6. GPR120 is highly expressed in intestine, adrenals, lung, adipose tissue, and macrophages, and is described as the n-3 polyunsaturated fatty acid (PUFA) receptor. Upon activation by n-3 PUFA, GPR120 inhibits transforming growth factor beta-activated kinase 1 (TAK1) activation, resulting in attenuation of

IKKB/NF-κB and JNK/AP1 signaling7. GPR120 expression regulates obesity in mice and humans; a non-synonymous mutation (p.R270H) inhibited GPR120 signaling activity, resulting in increased risk of obesity in European populations8. In vivo, selective activation of GPR120 by n-3 PUFAs is anti-inflammatory and insulin sensitizing7, 8. High fat diet-fed GPR120 knockout

(KO) mice vs. WT counterparts supplemented with n-3 PUFA or a selective GPR120 agonist

(cpdA)9 have: 1) increased adipose tissue F4/80+ macrophage infiltration and pro- inflammatory/M1 gene expression, 2) increased M1 gene expression in LPS-stimulated peritoneal macrophages, and 3) increased insulin resistance. Collectively, these findings highlight the anti-inflamatory potential of macrophage GPR120 activation by n-3 PUFA.

However, the impact of GPR120 expression on atherosclerosis progression, particularly in the context of dietary fatty acid composition, is unknown.

One of the first lines of cardiovascular disease (CVD) treatment is reducing total fat intake and replacing saturated fatty acids with PUFAs10. n-3 as well as n-6 PUFAs are atheroprotective in mice, non-human primates, and humans11-20. In humans, dietary n-3 PUFAs, eicosapentaenoic acid (EPA; 20:5 n-3) and docasahexaenoic acid (DHA; 22:6 n-3) found in fish

107 oil (FO), are anti-inflammatory and lower plasma triglycerides (TG), but not plasma LDL cholesterol, a primary risk factor for atherosclerosis in humans 21, 22. However, FO consumption is low in the USA 15. Dietary linoleic acid (LA; 18:2 n-6) is cardioprotective in humans, such that, a 5% increase in energy intake from LA is associated with a 10% and 13% lower risk of coronary heart disease events and deaths, respectively 18. LA-enriched diets in nonhuman primates and mice also are atheroprotective compared to saturated/monounsaturated diets 23-25.

Concerns that increased n-6 PUFA consumption may result in elevated membrane arachidonic acid (AA; 20:4 n-6), increased cellular inflammation, and atherosclerosis exacerbation lack support in humans 26. Moreover, LA-enriched diets do not enrich AA in plasma and tissue lipid fractions 25, 27, 28, likely due to inefficient fatty acid desaturase-2 (FADS2) conversion of LA to AA

29, 30. Although some of the atheroprotection by n-6 PUFAs is likely due to plasma lipid lowering, diets enriched in desaturation (via FADS2)-elongation products of LA, such as dihomo gamma linolenic acid (DGLA; 20:3 n-6), have anti-inflammatory potential through generation of prostaglandin E1 (PGE1), a potent thromboxane A2 inhibitor that reduces leukocyte endothelial cells adherence31, 32. Currently, in North American diets, plant-derived LA and alpha linolenic acid (ALA; 18:3 n-3) constitute the majority of PUFA intake 16, 18. Despite high consumption of

LA and ALA, tissue enrichment of their longer chain bioactive products, AA and EPA, respectively, is limited due to inefficiency of the rate limiting FADS2 enzyme (16,17). To circumvent this, we have identified botanical oils enriched in n-3 or n-6 PUFAs beyond FADS2,

18:4 n-3 (stearidonic acid; SDA)-enriched echium oil (EO) and 18:3 n-6 (gamma linolenic acid;

GLA)-enriched borage oil (BO). We previously demonstrated that EO and BO are as atheroprotective and hepatoprotective as FO, compared to a diet enriched in saturated/monounsaturated fat (PO)33.

108

Based on available information, we hypothesized that leukocyte GPR120 (L-GPR120) activation contributes to the atheroprotective effects of n-3, but not n-6, PUFAs. Selective activation of L-GPR120 by n-3 PUFAs should result in decreased leukocyte inflammation and atherosclerosis. We tested this hypothesis by feeding PO, FO, EO or BO diets to leukocyte

GPR120 WT and KO mice in the C57BL/6-LDLrKO background generated by bone marrow transplantation (BMT). To our knowledge, this is the first study to determine the in vivo role of leukocyte GPR120 in the context of n-3 vs. n-6 PUFA-induced atheroprotection.

METHODS

Detailed experimental procedures and methods are reported in the online data supplement. Methods include generation and validation of LDLrKO/GPR120WT (WT) and

LDLrKO/GPR120KO (KO) mice using BMT, measurements of plasma lipids and lipoproteins, hepatic lipid content, circulating and splenic monocytosis and neutrophilia, monocyte labeling and recruitment into aortic intima, aortic and macrophage cholesterol content, and aortic root atherosclerosis and immunohistochemistry/histology.

STATISTICS

Data are presented as mean ± SEM. All data were tested for significant differences using 2-way ANOVA (diet vs. genotype) with post-hoc Tukey’s multiple comparison test. All analyses were performed with Statistica.

109

RESULTS

Systemic response to atherogenic diets (AD) enriched in n-3 and n-6 PUFAs

Eight week (wk) old female LDLrKO mice were irradiated and then received bone marrow from male GPR120 WT or KO mice. LDLrKO recipients recovered from radiation/BMT induced body weight loss within one week and showed similar and positive weight gain during 6 wks of recovery (Supplemental Fig. I A). Transplantation efficiency was ~95% based on quantification of the ratio of Syr/SOAT2 intensity of the agarose gel scans (Supplemental Fig. I

B) and verification of leukocyte GPR120 alleles using PCR analysis (Supplemental Fig. I C).

Eight wks after BMT, WT and KO mice were switched from chow to one of the four AD containing PO, BO, EO, or FO for additional 16 wks. Dietary FA compositions showed relative enrichment of 18 carbon FAs beyond FADS2 in the BO (11.7% 18:3 n-6) and EO (6% 18:4 n-3)

ADs. AD FA compositions are given in Supplemental table I and are similar to those published previously33 , except that GLA content (2.3% energy) was approximately half that of the previous study (3.9% energy) because a different source of BO was used. AD feeding over 16 wks resulted in uniform food consumption (~3-4 gm/day/mouse), body weight gain, and terminal liver, spleen and gonadal adipose/body weight ratios among all groups irrespective of L-

GPR120 expression (Supplemental Figs. II & III). Additionally, L-GPR120 did not affect percentage (%) red blood cell (RBC) fatty acid (FA) enrichment. Because WT and KO mice fed the same AD had similar percentage RBC FA enrichment, data from both genotypes were pooled and are plotted together (Supplemental Fig. IV). We previously demonstrated that percentage FA enrichment in plasma and liver phospholipids was similar to that in RBCs33.

110

L-GPR120 does not affect plasma lipids and cholesterol lipoprotein distribution

Consistent with previous reports, FO, EO and BO reduced total plasma cholesterol (Figure 1 A and B), plasma triglycerides (Figure 1 C and D), VLDL cholesterol (Figure 2 A and B), LDL cholesterol (Figure 2 C and D) and raised HDL cholesterol (Figure 2 E and F) relative to PO.

The absence of L-GPR120 had no effect on any of the plasma lipid and lipoprotein concentrations.

L-GPR120 does not affect hepatosteatosis

Previous reports have shown a phenotype switching (M2 to M1-like) of macrophages in

GPR120 KO vs. WT mice fed high fat diet containing n-3 PUFA or GPR120 selective agonist7, 9.

We determined the extent to which L-GPR120 affects hepatic inflammatory and lipogenic gene expression, and hepatic lipid content. PO-fed mice had similar hepatic inflammatory gene expression regardless of L-GPR120 expression (Figure 3A). However, FO, EO and BO-fed

GPR120KO mice had increased M1 (CD11c, MCP-1, TNF-α, IL-1β, and IL-6) and reduced M2

(IL-10, Arginase-1, and iNOS) gene expression compared to WT mice, and increased hepatic

TNF-α and IL-1β protein content (Figure 3 B). FO KO vs. WT mice also had reduced expression of macrophage marker genes, CD68 and F4/80 (Figure 3A), whereas other diet groups did not. In general, L-GPR120 expression had no effect on hepatic lipogenic gene expression (Figure 3C) and hepatic lipid content (Figure 3D), except for FO-fed mice, in which expression of several lipogenic genes (SREBP-1c, FAS, SCD-1, SREBP-2, HMGCoA synthase,

HMGCoA reductase, and PPARα) was significantly reduced in L-GPR120 KO mice.

Additionally, liver sections of WT and KO mice fed the same AD had similar morphological

111 appearance by H&E and CD68 IHC staining (Supplemental Fig. V). Thus, reduced hepatosteatosis in FO, EO and BO vs. PO-fed mice is independent of L-GPR120 expression and likely occurs via reduced SREBP1 activation and lipogenic gene expression.

PUFAs attenuate neutrophilia and monocytosis via leukocyte GPR120

Hypercholesterolemia induces leukocytosis, primarily by elevating circulating monocytes

(monocytosis) and neutrophils (neutrophilia), both of which are positively associated with atherosclerosis in mice and humans 34, 35. The percentage of circulating neutrophils

(CD11b+,CD115-, Ly6G+) was decreased with FO, EO, and BO feeding relative to PO, which was reversed in the absence of L-GPR120 (Figure 4 A). The spleen is an extramedullary reservoir for circulating, undifferentiated monocytes, which upon recruitment to a site of injury/ inflammation may differentiate into macrophages, dendritic cells, or other tissue descendants36,

37. Splenic neutrophilia (Figure 4 B) and monocytosis (%CD11b+,CD115+, Ly6G-) (Figure 4C) were significantly attenuated by dietary FO, EO and BO in mice expressing L-GPR210, but not those lacking L-GPR120, relative to PO-fed mice. Splenic Ly6Chi monocytosis was also significantly attenuated in FO, EO and BO-fed WT mice, relative to PO, but not KO mice, suggesting that L-GPR120 plays a significant role in maintaining a favorable Ly6Clo monocyte profile in FO, EO and BO-fed mice (Figure 4 D). Despite L-GPR120 expression and dietary

PUFA effects on splenic monocytosis, neither affected the percentage of circulating monocytosis or Ly6Chi monocytes (Supplemental Fig. VI).

112

L-GPR120 attenuates monocyte infiltration into atherosclerotic lesions

GPR120 negatively regulates macrophage chemotaxis ex vivo7. We previously showed that

ADs containing FO, EO and BO, compared to PO, reduce macrophage chemotaxis ex vivo and result in reduced aortic root CD68+ intimal area in LDLrKO mice33. We hypothesized that activated L-GPR120 may limit monocyte trafficking into aortic root intima and performed monocyte-labeling experiments to test our hypothesis. Blood monocytes were phagocytically labeled with 1µm fluorescent beads using the procedure developed by Randolph and collegues38, 39. Using this procedure, ~1% of blood leukocytes were labelled with beads (bead+) and of these, 60-80% were monocytes, resulting in a range of 6-8 % monocyte labeling among the groups (Supplemental figure VII A). Monocyte recruitment into aortic root intima was significantly reduced for mice consuming FO, EO and BO, relative to PO, as anticipated; however, recruitment for all three PUFA groups was equivalent to PO in mice lacking L-GPR120

(Figure 5 A and Supplemental figure VII B). Additionally, splenic Ly6Chi monocytosis positively correlated with monocyte infiltration into aortic root intima (Figure 5 B).

L-GPR120 expression does not affect atherosclerosis in LDLrKO mice

Aortic root intimal area was significantly lower (500-700 vs. 950-1050 mm2) in FO, EO and BO fed vs. PO-fed mice (Figure 6 A), as was aortic root intimal macrophage content (CD68+)

(Figure 6 B). L-GPR20 expression did not affect aortic root intimal area; however, unexpectedly, we observed further reduction in intimal CD68+ content in KO vs. WT mice fed

FO. Intimal collagen content was similar among all groups of mice except for higher values for

WT mice fed FO (Figure 6 C). We hypothesized that paradoxical lowering of CD68+

113 macrophage content despite increased monocyte trafficking in KO FO (vs. WT FO) is plausible if recruited monocytes egress out of the intima and into the lymphatic system undifferentiated or that the recruited monocytes differentiate into CD11c+ dendritic cells within intima. We quantified CD11c+ intimal area and found that KO FO mice had increased CD11c+ intimal area relative to WT FO mice (Figure 6 D). Representative aortic root images of WT and KO mice corresponding to Fig. 6 A, B, C and D can be found in Supplemental figures VIII, IX, X and XI, respectively. We also found the number of cleaved caspase-3+ apoptotic cells was similar among all groups of mice (Supplemental Figure XII). To examine atherosclerosis at another site, we measured whole aorta FC (Figure 7 A) and CE (Figure 7 B) content, which was significantly lower in FO, EO and BO fed (~2 fold reduction in FC; ~3-4 fold reduction in CE) vs.

PO-fed mice. In agreement with aortic root intimal area results, L-GPR120 expression had no impact on this measurement of atherosclerosis, indicating that PUFA-induced atheroprotection is not significantly affected by L-GPR120 expression.

114

DISCUSSION

Whether in vivo L-GPR120 activation by PUFAs is atheroprotective is unknown. In this study, we determined the extent to which L-GPR120 expression was atheroprotective in the context of n-3 vs. n-6 PUFA feeding. We hypothesized that L-GPR120 expression contributes to atheroprotective effects of n-3, but not n-6, PUFAs, since in vitro studies suggest that n-3

PUFAs preferentially activate GPR1207. Our study led to several novel observations. First, we show that L-GPR120 deletion had minimal effects on plasma lipid concentrations, lipoprotein cholesterol distribution, hepatic neutral lipid content, aortic root intimal area, and aortic cholesterol content regardless of dietary fat type fed to the mice. Second, in vivo activation of L-

GPR120 by PUFA-enriched diets limits: a) hepatic macrophage (i.e., Kupffer cell) pro- inflammatory gene expression, b) splenic and blood neutrophilia, c) splenic Ly6Chi monocytosis, and d) monocyte recruitment into aortic root atherosclerotic lesions. Third, there was no distinction between n-3 vs. n-6 PUFAs with regard to in vivo activation of GPR120. These data suggest that dietary n-3 and n-6 PUFAs, though atheroprotective in LDLrKO mice relative to

PO, have GRP120-independent (i.e., attenuation of hypercholesterolemia, hepatosteatosis, and aortic atherosclerosis) and -dependent (i.e., attenuation of pro-inflammatory gene expression, neutrophilia, splenic monocytosis and monocyte trafficking) atheroprotective roles.

We investigated whether GPR120 activation by n-3 PUFAs would reduce atherogenesis based on its role in attenuating inflammation and insulin resistance in diet-induced obesity7, 9.

GPR120 is highly expressed in macrophages, which are important inflammatory cells in the pathogenesis of obesity and atherosclerosis7. Mice fed a high fat diet have increased monocyte recruitment to adipose tissue, adipose tissue macrophage inflammation, and insulin resistance7.

This phenotype was reversed in mice fed a high fat diet in which FO was isocalorically (27% of

115 calories) substituted for saturated fat; however, this was not the case for GPR120 KO mice or mice transplanted with GPR120 KO bone marrow, suggesting that most of the beneficial effect of FO was due to macrophage GPR120 activation7. Oh et al have shown that activation of

GPR120 results in a beta arrestin 2-mediated internalization of GPR120 and binding to TAB1, preventing its activation of TAK1 and blunting a common node of inflammatory signaling for Toll- like receptors, TNFα receptor, and inflammasome activation7. However, another study using a different GPR120 gene targeting construct showed that reversal of insulin resistance by feeding a high n-3 PUFA-containing diet was independent of GPR120 expression, suggesting the in vivo metabolic impact of GPR120 expression is not fully elucidated40.

Central to the pathogenesis of atherosclerosis is lipid-laden macrophages that initiate a chronic inflammatory state41. Dietary PUFAs are atheroprotective in mice, non-human primates, and humans11-20; however, the extent to which this atheroprotection is related to GPR120 expression, in general, and macrophage GPR120 expression, in particular, is unknown. Our results show that dietary PUFA-induced atheroprotection was independent of BM GPR120 expression by several measurements, including aortic root intimal area and whole aorta cholesterol content. There may be several explanations for this unexpected outcome. First, inflammation likely plays a minor role in the pathogenesis of atherosclerosis in LDLrKO mice, in which apoB lipoproteins are elevated in plasma. Past studies from our lab and others have shown that atherosclerosis is highly associated with plasma VLDL cholesterol concentrations in

LDLrKO mice33, 42. Second, cells outside the BM compartment (i.e., endothelial and smooth muscle cells) that express GPR120 may play important atheroprotective roles in the context of dietary PUFA feeding and these would have been missed in our study. Finally, the intake of dietary PUFAs may not have been sufficient to activate GPR120 in vivo, as our study used much lower n-3 PUFA enriched diets compared to previous studies7, 40.

116

To determine whether diet enrichment of PUFAs was sufficient for in vivo activation of macrophage GPR120 in our study, we made several measurements of macrophage function.

Mice fed the PUFA-containing diets had decreased percentage of splenic Ly6Chi monocytes and decreased trafficking of blood monocytes into aortic root atherosclerotic lesions relative to PO- fed mice and these trends were reversed in mice transplanted with GPR120 KO vs. WT BM.

Blood and splenic neutrophilia was also suppressed in mice transplanted with WT BM and fed

PUFAs, but not PUFA-fed mice transplanted with GPR120 KO BM. In addition, PUFA-fed mice lacking BM GPR120 had higher hepatic pro-inflammatory gene expression than their WT counterparts. Since liver has low GPR120 expression4, 8, whereas macrophages have relatively high expression7, these results suggest that hepatic Kupffer cells, as well as circulating monocytes and neutrophils, display dietary PUFA-mediated GPR120 activation in vivo. Although hypercholesterolemia is associated with monocytosis19 , our outcomes were unrelated to plasma cholesterol concentrations, which were equivalent among all PUFA fed mice regardless of BM GPR120 expression. Past studies showed in vivo GPR120 activation (i.e., decreased inflammation, increased insulin sensitivity) at much higher, super physiological intakes of FO7.

As discussed in our previous publication33, the levels of dietary PUFA intake in our study were in a physiological range that would be achievable for human consumption. Whether diminished monocytosis and inflammation will occur in humans consuming diets with a similar PUFA enrichment is unknown. However, adipose tissue GPR120 expression is increased in obese humans and a GPR120 coding variant that inhibits its activation increases obesity risk in

Europeans. Collectively, our results support the conclusion that sufficient PUFA enrichment occurred in vivo to activate GPR120, but the anticipated GPR120-mediated reduction in atherosclerosis was not observed because the plasma lipid lowering effect of the PUFA diets was overwhelming.

117

Our study also demonstrated that the in vivo activation of GPR120 was not specific for n-

3 PUFAs, but was equally effective with dietary n-6 PUFAs. The fatty acid specificity of GPR120 activation has been investigated primarily using in vitro techniques and results are contradictory.

Some report a preference of n-3 PUFAs for GPR120 activation, whereas other studies report broader fatty acyl specificity. In vivo studies also have yielded conflicting results. Oh et al showed that a high fat diet enriched in FO reduced glucose intolerance in a GPR120 dependent manner, whereas Bjursell et al demonstrated that FO containing high fat diet reversed glucose intolerance in both WT and GPR120 KO mice. The explanation for these disparate outcomes has been discussed at length by Bjursell et al; possibilities include different gene targeting constructs, mouse backgrounds, and dietary n-3 PUFA content and length of diet feeding.

Although our primary outcome, which was atherosclerosis progression, was not affected by L-

GPR120 expression, similar to results published by Bjursell et al on glucose intolerance, our results clearly showed an impact of L-GPR120 on monocyte/macrophage function with both n-3 and n-6 PUFA diets, suggesting the in vivo specificity for L-GPR120 activation is equal for both classes of PUFAs.

Whole body deletion of GPR120 has led to increased hepatosteatosis in some7, 8, but not all studies40. In our study, PUFA-containing diets uniformly reduced hepatic neutral lipid content relative to PO regardless L-GPR120 expression. However, LDLrKO mice transplanted with

GPR120 KO vs. WT BM had increased liver inflammatory gene expression, which we attribute to Kupffer cells (see above), but only in the PUFA-fed groups. The reason for a PUFA-induced increase in inflammatory gene expression only in mice lacking L-GPR120 and its relationship to hepatic lipid content is unknown. We speculate that hepatic Kupffer cell GPR120 expression

118 normally suppresses inflammation, perhaps initiated by increased PUFA oxidation, in the context of PUFA-enriched diets. The increase in hepatic inflammation in PUFA-fed L-GPR120

KO mice may suppress hepatic lipogenesis (ref for increased inflammation decreases hepatic lipogenesis), resulting in similar hepatic lipid content for L-GPR120 WT and KO mice. The difference in hepatic lipid content between our study and that of Oh et al may relate to global vs.

BM KO of GPR120 or differences in dietary content of n-3 PUFAs.

119

ACKNOWLEDGEMENTS

We gratefully acknowledge Karen Klein (Biomedical Research Services and Administration,

Wake Forest School of Medicine) for editing the manuscript; Xuewei Zhu for help with bone marrow isolation and retro-orbital injections during monocyte recruitment experiment and Sandy

Chan for help in organ harvesting during terminal necropsies and quantification of aortic root intimal areas to determine intersubject variability.

SOURCES OF FUNDING

This study was supported by grants from National Institute of Health P50 AT002782 and

R01HL119962 (to JSP).

DISCLOSURES

None.

120

Reference List

1. Wellendorph P, Johansen LD and Brauner-Osborne H. Molecular pharmacology of promiscuous seven transmembrane receptors sensing organic nutrients. Molecular pharmacology. 2009;76:453-65.

2. Briscoe CP, Tadayyon M, Andrews JL, Benson WG, Chambers JK, Eilert MM, Ellis C,

Elshourbagy NA, Goetz AS, Minnick DT, Murdock PR, Sauls HR, Jr., Shabon U, Spinage LD,

Strum JC, Szekeres PG, Tan KB, Way JM, Ignar DM, Wilson S and Muir AI. The orphan G protein-coupled receptor GPR40 is activated by medium and long chain fatty acids. The Journal of biological chemistry. 2003;278:11303-11.

3. Brown AJ, Goldsworthy SM, Barnes AA, Eilert MM, Tcheang L, Daniels D, Muir AI,

Wigglesworth MJ, Kinghorn I, Fraser NJ, Pike NB, Strum JC, Steplewski KM, Murdock PR,

Holder JC, Marshall FH, Szekeres PG, Wilson S, Ignar DM, Foord SM, Wise A and Dowell SJ.

The Orphan G protein-coupled receptors GPR41 and GPR43 are activated by propionate and other short chain carboxylic acids. The Journal of biological chemistry. 2003;278:11312-9.

4. Hirasawa A, Tsumaya K, Awaji T, Katsuma S, Adachi T, Yamada M, Sugimoto Y,

Miyazaki S and Tsujimoto G. Free fatty acids regulate gut incretin glucagon-like peptide-1 secretion through GPR120. Nature medicine. 2005;11:90-4.

5. Oh Da Y and Olefsky Jerrold M. Omega 3 Fatty Acids and GPR120. Cell metabolism.

2012;15:564-565.

121

6. Wang J, Wu X, Simonavicius N, Tian H and Ling L. Medium-chain fatty acids as ligands for orphan G protein-coupled receptor GPR84. The Journal of biological chemistry.

2006;281:34457-64.

7. Oh DY, Talukdar S, Bae EJ, Imamura T, Morinaga H, Fan W, Li P, Lu WJ, Watkins SM and Olefsky JM. GPR120 is an omega-3 fatty acid receptor mediating potent anti-inflammatory and insulin-sensitizing effects. Cell. 2010;142:687-98.

8. Ichimura A, Hirasawa A, Poulain-Godefroy O, Bonnefond A, Hara T, Yengo L, Kimura I,

Leloire A, Liu N, Iida K, Choquet H, Besnard P, Lecoeur C, Vivequin S, Ayukawa K, Takeuchi

M, Ozawa K, Tauber M, Maffeis C, Morandi A, Buzzetti R, Elliott P, Pouta A, Jarvelin MR,

Korner A, Kiess W, Pigeyre M, Caiazzo R, Van Hul W, Van Gaal L, Horber F, Balkau B, Levy-

Marchal C, Rouskas K, Kouvatsi A, Hebebrand J, Hinney A, Scherag A, Pattou F, Meyre D,

Koshimizu TA, Wolowczuk I, Tsujimoto G and Froguel P. Dysfunction of lipid sensor GPR120 leads to obesity in both mouse and human. Nature. 2012;483:350-4.

9. Oh da Y, Walenta E, Akiyama TE, Lagakos WS, Lackey D, Pessentheiner AR, Sasik R,

Hah N, Chi TJ, Cox JM, Powels MA, Di Salvo J, Sinz C, Watkins SM, Armando AM, Chung H,

Evans RM, Quehenberger O, McNelis J, Bogner-Strauss JG and Olefsky JM. A Gpr120- selective agonist improves insulin resistance and chronic inflammation in obese mice. Nature medicine. 2014;20:942-7.

10. Mozaffarian D, Micha R and Wallace S. Effects on coronary heart disease of increasing polyunsaturated fat in place of saturated fat: a systematic review and meta-analysis of randomized controlled trials. PLoS medicine. 2010;7:e1000252.

11. Bigger JT, Jr. and El-Sherif T. Polyunsaturated fatty acids and cardiovascular events: a fish tale. Circulation. 2001;103:623-5.

122

12. Calder PC and Grimble RF. Polyunsaturated fatty acids, inflammation and immunity. Eur

J Clin Nutr. 2002;56 Suppl 3:S14-9.

13. Kang JX and Leaf A. Antiarrhythmic effects of polyunsaturated fatty acids. Recent studies. Circulation. 1996;94:1774-80.

14. Parks JS, Kaduck-Sawyer J, Bullock BC and Rudel LL. Effect of dietary fish oil on coronary artery and aortic atherosclerosis in African green monkeys. Arteriosclerosis.

1990;10:1102-12.

15. Hu FB, Bronner L, Willett WC, Stampfer MJ, Rexrode KM, Albert CM, Hunter D and

Manson JE. Fish and omega-3 fatty acid intake and risk of coronary heart disease in women.

JAMA : the journal of the American Medical Association. 2002;287:1815-21.

16. Harris WS, Mozaffarian D, Rimm E, Kris-Etherton P, Rudel LL, Appel LJ, Engler MM,

Engler MB and Sacks F. Omega-6 fatty acids and risk for cardiovascular disease: a science advisory from the American Heart Association Nutrition Subcommittee of the Council on

Nutrition, Physical Activity, and Metabolism; Council on Cardiovascular Nursing; and Council on

Epidemiology and Prevention. Circulation. 2009;119:902-7.

17. Harris WS and Shearer GC. Omega-6 Fatty Acids and Cardiovascular Disease: Friend or Foe? Circulation. 2014;130:1562-1564.

18. Farvid MS, Ding M, Pan A, Sun Q, Chiuve SE, Steffen LM, Willett WC and Hu FB.

Dietary Linoleic Acid and Risk of Coronary Heart Disease: A Systematic Review and Meta-

Analysis of Prospective Cohort Studies. Circulation. 2014;130:1568-1578.

19. Brown AL, Zhu X, Rong S, Shewale S, Seo J, Boudyguina E, Gebre AK, Alexander-

Miller MA and Parks JS. Omega-3 fatty acids ameliorate atherosclerosis by favorably altering

123 monocyte subsets and limiting monocyte recruitment to aortic lesions. Arteriosclerosis, thrombosis, and vascular biology. 2012;32:2122-30.

20. Zampolli A, Bysted A, Leth T, Mortensen A, De Caterina R and Falk E. Contrasting effect of fish oil supplementation on the development of atherosclerosis in murine models.

Atherosclerosis. 2006;184:78-85.

21. Parks JS and Rudel LL. Effect of fish oil on atherosclerosis and lipoprotein metabolism.

Atherosclerosis. 1990;84:83-94.

22. Harris WS. Fish oils and plasma lipid and lipoprotein metabolism in humans: a critical review. Journal of lipid research. 1989;30:785-807.

23. Bell TA, 3rd, Kelley K, Wilson MD, Sawyer JK and Rudel LL. Dietary fat-induced alterations in atherosclerosis are abolished by ACAT2-deficiency in ApoB100 only, LDLr-/- mice.

Arteriosclerosis, thrombosis, and vascular biology. 2007;27:1396-402.

24. Rudel LL, Johnson FL, Sawyer JK, Wilson MS and Parks JS. Dietary polyunsaturated fat modifies low-density lipoproteins and reduces atherosclerosis of nonhuman primates with high and low diet responsiveness. The American journal of clinical nutrition. 1995;62:463S-470S.

25. Rudel LL, Parks JS and Sawyer JK. Compared with dietary monounsaturated and saturated fat, polyunsaturated fat protects African green monkeys from coronary artery atherosclerosis. Arteriosclerosis, thrombosis, and vascular biology. 1995;15:2101-10.

26. Harris WS and Shearer GC. Omega-6 Fatty Acids and Cardiovascular Disease: Friend,

Not Foe? Circulation. 2014;130:1562-1564.

124

27. Thornburg JT, Parks JS and Rudel LL. Dietary fatty acid modification of HDL phospholipid molecular species alters lecithin: cholesterol acyltransferase reactivity in cynomolgus monkeys. Journal of lipid research. 1995;36:277-89.

28. Rett BS and Whelan J. Increasing dietary linoleic acid does not increase tissue arachidonic acid content in adults consuming Western-type diets: a systematic review. Nutrition

& metabolism. 2011;8:36.

29. Demmelmair H, Iser B, Rauh-Pfeiffer A and Koletzko B. Comparison of bolus versus fractionated oral applications of [13C]-linoleic acid in humans. European journal of clinical investigation. 1999;29:603-9.

30. Hussein N, Ah-Sing E, Wilkinson P, Leach C, Griffin BA and Millward DJ. Long-chain conversion of [13C]linoleic acid and alpha-linolenic acid in response to marked changes in their dietary intake in men. Journal of lipid research. 2005;46:269-80.

31. Chopra J and Webster RO. PGE1 inhibits neutrophil adherence and neutrophil-mediated injury to cultured endothelial cells. Am Rev Respir Dis. 1988;138:915-20.

32. Jones DA and Fitzpatrick FA. "Suicide" inactivation of thromboxane A2 synthase.

Characteristics of mechanism-based inactivation with isolated enzyme and intact platelets. The

Journal of biological chemistry. 1990;265:20166-71.

33. Shewale SV, Boudyguina E, Zhu X, Shen L, Hutchins PM, Barkley RM, Murphy RC and

Parks JS. Botanical oils enriched in n-6 and n-3 fatty acid products of FADS2 are equally effective in preventing atherosclerosis and hepatosteatosis in mice. Journal of lipid research.

2015.

34. Drechsler M, Megens RT, van Zandvoort M, Weber C and Soehnlein O. Hyperlipidemia- triggered neutrophilia promotes early atherosclerosis. Circulation. 2010;122:1837-45.

125

35. Swirski FK, Libby P, Aikawa E, Alcaide P, Luscinskas FW, Weissleder R and Pittet MJ.

Ly-6Chi monocytes dominate hypercholesterolemia-associated monocytosis and give rise to macrophages in atheromata. The Journal of clinical investigation. 2007;117:195-205.

36. Swirski FK, Nahrendorf M, Etzrodt M, Wildgruber M, Cortez-Retamozo V, Panizzi P,

Figueiredo JL, Kohler RH, Chudnovskiy A, Waterman P, Aikawa E, Mempel TR, Libby P,

Weissleder R and Pittet MJ. Identification of splenic reservoir monocytes and their deployment to inflammatory sites. Science. 2009;325:612-6.

37. Geissmann F, Manz MG, Jung S, Sieweke MH, Merad M and Ley K. Development of monocytes, macrophages, and dendritic cells. Science. 2010;327:656-61.

38. Tacke F, Alvarez D, Kaplan TJ, Jakubzick C, Spanbroek R, Llodra J, Garin A, Liu J,

Mack M, van Rooijen N, Lira SA, Habenicht AJ and Randolph GJ. Monocyte subsets differentially employ CCR2, CCR5, and CX3CR1 to accumulate within atherosclerotic plaques. J

Clin Invest. 2007;117:185-94.

39. Tacke F, Ginhoux F, Jakubzick C, van Rooijen N, Merad M and Randolph GJ. Immature monocytes acquire antigens from other cells in the bone marrow and present them to T cells after maturing in the periphery. J Exp Med. 2006;203:583-97.

40. Bjursell M, Xu X, Admyre T, Bottcher G, Lundin S, Nilsson R, Stone VM, Morgan NG,

Lam YY, Storlien LH, Linden D, Smith DM, Bohlooly YM and Oscarsson J. The beneficial effects of n-3 polyunsaturated fatty acids on diet induced obesity and impaired glucose control do not require Gpr120. PloS one. 2014;9:e114942.

41. Lusis AJ. Atherosclerosis. Nature. 2000;407:233-41.

42. VanderLaan PA, Reardon CA, Thisted RA and Getz GS. VLDL best predicts aortic root atherosclerosis in LDL receptor deficient mice. Journal of lipid research. 2009;50:376-85.

126

FIGURE LEGENDS

Figure 1. L-GPR120 expression does not affect plasma lipids. Irradiated LDLrKO mice received bone marrow from WT or GPR120 KO donors and were fed ADs (PO, FO, EO or BO) for 16 wks. Fasting (4h) plasma total cholesterol (TC) (A) and triglyceride (TG) (B) concentrations were measured by enzymatic assays. Plasma TC (C) and TG (D) concentrations were integrated over time (0-14 weeks) and expressed as area under the curve (AUC). Mean ±

SEM, n=13-15/diet. Groups with different letters are significantly different (p<0.05) by two-way

ANOVA and Tukey’s post-hoc analysis.

Figure 2. L-GPR120 does not affect plasma lipoprotein cholesterol distribution. Fasted

(4h) plasma was harvested from AD fed WT and KO mice and fractionated by FPLC to determine cholesterol distribution among VLDL (A), LDL (B) and HDL (C) fractions. Plasma

VLDL-C (D), LDL-C (E) and HDL-C (F) concentrations were integrated over time (0-14 weeks) and expressed as area under the curve (AUC). Mean ± SEM, n=3 equal volume pooled samples from 4 mice/group. Groups with different letters are significantly different (p<0.05) by two-way

ANOVA and Tukey’s post-hoc analysis.

Figure 3. Effect of L-GPR120 expression on hepatic gene expression and lipid content.

Livers were harvested from 16 week AD fed WT and KO mice to measure lipid content and mRNA abundance. Macrophage gene expression (A), lipogenic gene expression (B), and hepatic neutral lipid content (C) Neutral lipid= CE+TG. Liver was lipid extracted and TC, FC, and

TG were quantified by enzymatic assay. CE mass was calculated as (TC-FC)*1.67 and

127 normalized to protein content. Mean ± SEM, n=6. Groups with different letters are significantly different (p<0.05) by two-way ANOVA and Tukey’s post-hoc analysis.

Figure 4. Effect of L-GPR120 on neutrophilia and monocytosis. Blood and spleens were harvested from 16 week AD fed WT and KO mice to measure the distribution of neutrophils and monocytes, including Ly6Chi monocytes, using flow cytometry. Percentage blood neutrophils

(A), percentage splenic neutrophils (B), percentage splenic monocytes (C) and percentage splenic Ly6Chi monocytes (D). Mean ± SEM, n=12-15/group. Groups with different letters are significantly different (p<0.05) by two-way ANOVA and Tukey’s post-hoc analysis.

Figure 5. Effect of L-GPR120 on monocyte recruitment to aortic root intima. Blood monocytes were labeled using 1-µm Fluoresbrite microparticles in WT and KO mice fed AD for

15 weeks. Percentage of FITC bead-labeled blood monocytes was measured by flow cytometry

24 hr after injection. Five days later, the number of FITC+ beads was quantified in the intimal area of 8 µm thick aortic root sections and normalized by the percentage of FITC+ blood monocytes to obtain normalized frequency of monocytes recruited into aortic intima (A). Mean ±

SEM, n=6-7/group. Groups with different letters are significantly different (p<0.05) by two-way

ANOVA and Tukey’s post-hoc analysis. Percentage of splenic Ly6Chi monocytes was plotted vs. monocyte recruitment into aortic intima (B). Each data point represents individual animal and symbols represent their respective groups, n=6-7/group. The line of best fit, determined by linear regression analysis, is also plotted.

128

Figure 6. Histological quantification of aortic root atherosclerotic lesions. WT and KO mice were fed AD for 16 weeks. At necropsy, hearts were frozen in OCT, and aortic roots were serially sectioned. Quantification of aortic root Oil Red O positive intimal area (lesion area) (A).

Quantification of percentage aortic root lesional area occupied by CD68+ cells (macrophages)

(B). Quantification of percentage aortic root lesional area occupied by Sirius red positive collagen fibers under plane polarized light (C). Quantification of percentage aortic root lesional area occupied by CD11c+ cells (D). Each point represents the average area from 8-10 sections for an individual mouse. Horizontal lines denote the mean for each diet group. Representative images for each parameter are depicted in Supplemental Figures VIII, IX, X and XI. n=6-

8/group. Groups with different letters are significantly different (p<0.05) by two-way ANOVA and

Tukey’s post-hoc analysis.

Figure 7. Effect of L-GPR120 expression on aortic cholesterol content. Aortas were harvested from 16 week AD fed WT and KO mice for atherosclerosis quantification. Aortas were lipid extracted for measurement of total cholesterol and free cholesterol (FC) content using gas- liquid chromatography. Cholesterol ester (CE) content was calculated as (TC-FC) x1.67. Aortic

FC content (A). Aortic CE content (B). n=12-14/group, each data point represents an individual mouse aorta. Horizontal lines denote the mean for each diet group. Groups with different letters are significantly different (p<0.05) by two-way ANOVA and Tukey’s post-hoc analysis.

129

Novelty and Significance

What Is Known?

 GPR120 is an n-3 fatty acid receptor mediating potent anti-inflammatory and insulin-

sensitizing effects

 Botanical oils enriched in n-6 and n-3 fatty acid products of FADS2 are equally effective

in preventing atherosclerosis and hepatosteatosis in mice

What New Information Does This Article Contribute?

 L-GPR120 is neither necessary nor sufficient for atheroprotection in LDLrKO mice fed n-

6 or n-3 PUFA-enriched diets

 Physiological levels of dietary PUFA intake in LDLrKO were sufficient for in vivo

activation of L-GPR120, leading to attenuated neutrophilia, splenic monocytosis, and

monocyte recruitment into aortic intima

 GPR120 activation in vivo was not n-3 PUFA selective; n-6 PUFAs also activate

GPR120

 Deletion of GPR120 expression in hepatic Kupffer cells results in increased inflammation

in the context of increased dietary consumption of PUFAs independent of hepatic neutral

lipid content

130

Figures

131

132

133

134

135

136

137

CHAPTER III: Supplemental Material

Supplemental Methods and Results

Dietary oils

The seed oil of Borago officinalis L., a member of the Boraginaceae family, was generously donated by Nordic Naturals (Watsonville, CA, USA). The seed oil of Echium plantagineum L., a member of the Boraginaceae family was a generous gift from Croda Europe Ltd. (Leek,

Staffordshire, UK). All oils were authenticated by the Wake Forest University Center for

Botanical Lipids and Inflammatory Disease Prevention. A certificate of analysis is on file for reference along with retention samples deposited at the Wake Forest School of Medicine. The seed oil of palm, Elaeis guineensis Jacq., a member of the Arecaceae family, was purchased from Shay and Company (Portland, OR, USA). A certificate of analysis is on file for reference along with retention samples deposited at the Wake Forest School of Medicine. The fish oil source was Brevoortia tyrannis Latrobe, a member of the Clupeidae family, was manufactured and generously provided by Omega Protein (Houston,TX, USA) with a report of analysis on file for reference.

Animals and Atherogenic Diets

Female LDLrKO (C57BL/6 background) mice (5-6 weeks of age) were purchased from The

Jackson Laboratory (Bar Harbor, Maine, USA). Mice were housed in a specific pathogen-free facility on a 12h light/dark cycle. Mice were allowed to acclimate to in-house animal resource facilities for 1-2 weeks. At 8 weeks of age, mice received bone marrow from either WT or KO

GPR120 male mice (see below). After ~6 weeks of recovery from bone marrow transplantation

(BMT) and at 13-14 weeks age, mice were randomly assigned to one of four atherogenic diet

138

(AD) groups (n=15/ diet group) containing 10% calories as PO and 0.2% cholesterol, supplemented with an additional 10% of calories as: 1) PO, 2) Borage oil (BO; 18:3 n-6 enriched), 3) Echium Oil (18:4 n-3 enriched), or 4) Fish oil (FO; 20:5 n-3 and 22:6 n-3 enriched) for additional 16 weeks. All protocols and procedures were approved by the Institutional Animal

Care and Use Committee. ADs were prepared by the diet kitchen in the Department of

Pathology at Wake Forest School of Medicine as previously described 1. Detailed composition and quality control data for similar ADs has been published2, 3

Bone marrow transplantation

BM cells were harvested from cleaned femurs and tibias of male GPR120 WT and GR120 KO mice and re-suspended in serum-free RPMI 1640 medium. Female LDLr-/- recipient mice

(Jackson Laboratories) were fasted overnight and received a sublethal dose of radiation (900 rads) 4h prior to BM injection. BM cells (~7 × 10 6/ mouse) were injected into the retro-orbital venous plexus of anesthetized recipient mice. Recipient mice received autoclaved, acidified (pH

2.7) water supplemented with 100 mg/l neomycin and 10 mg/l polymyxin B sulfate 3 days before and 2 weeks after the transplantation. Mice were then given autoclaved acidified water until the end of the study as described previously 4.

Repopulation of blood leukocytes by transplanted GPR120 WT or KO hematopoietic stem cells in female LDLrKO recipients was evaluated after 6 weeks of recovery by determining percentage expression of the Y--associated sex-determining region Y gene (Sry) in genomic DNA obtained from white blood cells. Additionally, white blood cell GPR120 WT and

KO genotype was also confirmed using PCR. Briefly, genomic DNA from whole blood was extracted with a Wizard Genomic DNA Purification kit (Promega). Sry was amplified by PCR to

139 a linear amplification phase under the following conditions: 95°C for 3 min followed by 30 cycles of 94°C for 30 s, 61°C for 1 min, and 72°C for 1 min. Acyl-CoA: cholesterol acetyltransferase 2

(ACAT2) was amplified by PCR as an internal control. The PCR products were separated on

0.8% agarose gels and visualized with ethidium bromide. A series of male and female genomic

DNA mixtures (100, 50, 25, and 0% male DNA) were made and used to construct a standard curve of male:female DNA ratio vs. Sry:ACAT2 ratio, determined by quantification of density of

PCR bands. Male genomic DNA in the whole blood samples of transplanted mice was estimated using the standard curve.

Body weight gain and organ weights

Mice were weighed one day prior to and daily for 3 consecutive days after BMT and then on a weekly basis for 6 weeks until they were started on the ADs. During 16 weeks of AD feeding, mice were weighed every 2-4 weeks. At necropsy, terminal body weights were recorded after 4 hours of fasting. Mice were then anesthetized using ketamine-xylazine (intramuscularly) and perfused through the left ventricle using cold PBS at the rate of ~3ml/minute for 3-4 minutes prior to organ harvesting. After perfusion, liver, spleen, and adipose wet weight was measured and normalized to terminal body weight.

Fatty acid analysis

Lipid extraction of lyophilized diets and RBCs was performed using the Bligh-Dyer method 5.

The lipid extracts were trans-methylated using boron trifluoride (BF3), and percentage fatty acid composition was quantified by gas–liquid chromatography as described previously 1.

140

Plasma lipid and lipoprotein analysis

Blood was collected from 4h-fasted mice by tail bleeding and plasma was isolated by low speed centrifugation. Plasma total and free cholesterol (Wako), and triglyceride (Roche) concentrations were determined using enzymatic assays as described earlier 6 at baseline and after 2, 4, 8, and 16 weeks of AD feeding. Cholesteryl ester content was calculated as (TC-FC) x 1.67; the multiplication by 1.67 corrects for loss of fatty acid during the cholesterol esterase step of the assay. Data were expressed as area under the curve (AUC), which integrates plasma lipid concentrations over the 16 weeks of AD feeding, representing a time average estimate of hypercholesterolemia throughout atherosclerosis progression. Plasma lipoprotein cholesterol mass distribution was determined using fast protein liquid chromatography (FPLC) fractionation on a Superose 6 10/300 column (GE Healthcare). Three equal volume plasma samples from 5 mice/group were pooled and subjected to FPLC fractionation and cholesterol quantification at baseline (chow diet) and at 2, 4, 8, 12 and 16 weeks of AD feeding. VLDL, LDL and HDL cholesterol concentrations were determined and then expressed as AUC.

Liver Lipids

Livers were harvested at necropsy, flash frozen in liquid N2, and stored at -80 C. Liver lipids were quantified using a detergent-based enzymatic assay as described earlier 7.

Flow cytometry

Peripheral blood was obtained for circulating leukocyte analysis. For splenic cells, tissue was digested with an enzyme cocktail as published8. The cell suspension was subsequently passed through a 70-µm nylon cell strainer (BD Falcon). Red blood cells were removed from flow

141 cytometry preparations by treatment with ACK lysing buffer (Gibco). The remaining white blood cells were incubated with the following mAbs: CD11b-APC-Cy7, clone:M1/70 (BioLegend);

CD115-APC, clone:AFS98; CD62L-PerCP-Cy5.5, clone:MEL-14 (eBiosciences); Ly6G-PE, clone:1A8; CD11c-PE-Cy7, clone:HL3; and Ly6C-FITC or PE clone:AL-21 (BD Pharmingen).

Data were acquired on a BD FACSCanto II instrument (BD Biosciences) and analyzed using

FacsDiva software (BD Biosciences).

Monocyte labeling

Monocytes were labeled following the Gr-1lo method as described by Tacke et al. 9, 10. One-µm

Fluoresbrite Yellow Green microspheres (2.5% solids [wt/vol], Polysciences, Inc., Warrington,

PA) were diluted in PBS (1:4) and retroorbitally injected into anesthetized mice. Blood was harvest from mice 24 hours after labeling for evaluation of bead-containing monocytes. The method labels predominantly Ly6Clo monocytes 9.

Quantification of monocyte recruitment

Monocyte recruitment was evaluated as reported previously11. Briefly, aortic root sections were taken as described below and the number of beads per section in atherosclerotic lesions was counted manually at 20x magnification. A total of 8 sections representing the length of the aortic root were analyzed per mouse. Because of slight variations in monocyte labeling (6-8%) among animals, data were normalized for individual mice based on the percentage of blood monocytes labeled. This adjustment gave a normalized value, which represented the actual number of monocytes entering the lesion (i.e. normalized frequency).

142

Quantification of aortic root lesion area

Aortic root atherosclerosis was assessed according to the method of Daugherty et. al12. Aortic roots were embedded in Optimal Cutting Temperature (Tissuetek) media in a plastic mold, frozen, and cut at 8 μm intervals. Sections were collected from the aortic root moving toward the apex of the heart and sequentially placed on 6 slides, such that each slide had sections 48 μm apart. The sections were fixed in 10% buffered formalin, stained in 0.5% Oil Red O for 25 minutes and counterstained with hematoxylin. Stained sections were photographed with a Nikon

DigitalSight DSFi1 camera and quantified using NIS Elements (Nikon) software. Intimal area measurements were obtained with ImagePro 6.2 software (Media Cybernetics Inc., Rockville,

Md.). We defined intimal area as the lesion area between the internal elastic lamina and the luminal surface of the aorta. Eight-10 aortic root sections were quantified from 6-8 mice/group.

The mean coefficient of variation for these measurements from individual mice on atherogenic diets was 15.0%.

Sirius red staining and quantification

Fresh frozen aortic root sections were fixed (10% buffered formalin, 10 minutes), washed (DI water, 5 minutes) and dehydrated in 50, 75 and 80 % ethanol, sequentially, for 1 minute each.

Sections were stained using hematoxylin, and then stained in 1% Sirius red dye (Sigma Aldrich

#365548) in saturated solution of picric acid (1.3 % picric acid in H2O Sigma Aldrich # P6744) for 60 minutes. After staining with Sirius red, sections were washed in acidified water (5% glacial acetic acid) for 5 minutes and dehydrated in increasing ethanol concentrations, cleaned in xylene and cover slipped using xylene based mounting media. Images were acquired under plane polarized light using a Nikon DigitalSight DSFi1 camera fitted with T-A2-DIC analyzer

(Nikon # MEN51921) and T-P2 DIC polarizer (Nikon # MEN51951).

143

Immunohistochemistry

Aortic root cross-sections and liver sections were fixed and incubated with anti-CD68 antibody

(AbDSerotec, clone FA11), anti-CD11c antibody (BD Pharminogen, clone HL3) or anti-cleaved caspase -3 antibody (Abcam, polyclonal # ab2302) (1:200, overnight) as described previously11.

Analysis of aortic cholesterol content

Aortic total, free and cholesteryl ester content was measured as described earlier 11.

Statistical analyses

Data are presented as mean ± SEM. All data were tested for significant differences using 2-way

ANOVA with post-hoc Tukey’s multiple comparison test. All the analyses were performed with

Statistica.

144

Supplementary Table 1

Table 1: Atherogenic Diet (AD) percentage fatty acid composition (%FA) and percentage total energy equivalence (% EE) of individual fatty acid

Palm oil (PO) Fish oil (FO) Echium oil (EO Borage oil (BO)

Fatty Acid % FA % FA % EE % EE % FA % EE % FA % EE

Palmitic Acid 43.2 8.89 30.6 6.89 25.6 5.27 26.0 5.20 (C16:0)

Palmitoleic Acid 0.37 0.08 4.2 0.95 0.4 0.02 0.4 0.08 (C16:1)

Stearic Acid 4.5 0.93 4.5 1.01 4.1 0.84 3.5 0.7 (C18:0)

Oleic Acid 37.3 7.68 24.9 5.61 26.1 5.37 25.1 5.02 (C18:1 n-9)

Linoleic Acid 11.1 2.29 8.2 1.85 15.4 3.17 25.0 5.0 (C18:2 n-6), LA

α-Linolenic Acid 0.37 0.08 1.0 0.23 13.8 2.84 0.4 0.08 (C18:3 n-3), ALA

γ-Linolenic Acid 5.0 1.03 11.40 2.28 (C18:3 n-6), GLA

Stearidonic Acid 1.5 0.34 6.0 1.23 0.2 0.04 (C18:4 n-3), SDA

Euric acid 0.2 0.04 0.9 0.18 (C22:1 n-9)

Eicosapentaenoic 0.1 0.02 6.9 1.55 0.3 0.06 0.3 0.06 Acid (C20:5 n-3), EPA

Docasahexaenoic 0.2 0.04 7 1.58 0.3 0.06 0.3 0.06 Acid (C22:6 n-3), DHA

145

ADs contained 0.2% cholesterol + 10% calories as Palm Oil (PO) + 10% calories as: PO,

Borage oil (BO), Echium oil (EO) or Fish oil (FO). Percent (%) fatty acid composition was measured using gas-liquid chromatography. Percent total energy equivalence (% EE) for individual fatty acid is calculated using total energy derived from fatty acids (i.e. 20%) / diet and

% fatty acid composition of respective diet.

146

Supplementary Figure Legends

Supplementary Fig. I. Bone marrow transplantation (BMT) efficiency. Irradiated LDLrKO mice received bone marrow from WT or GPR120 KO donors. After 6 weeks of recovery, mice were bled to isolate genomic DNA from circulating leukocytes and were subjected to genotyping for male gene “SRY” to verify BMT efficiency and GPR120 deletion. Body weight prior to and after BMT (A), expression of SRY and SOAT22 (positive control) in leukocyte genomic DNA obtained from female LDLrKO recipients (B), and GPR120 genotyping for WT and KO alleles in leukocyte genomic DNA isolated from WT or KO BM recipients (C).

Supplementary Fig. II. Body weight gain and terminal body weights. Irradiated LDLrKO mice received bone marrow from WT or GPR120 KO donors, were fed the indicated ADs for 16 weeks, and body weight gain was measured periodically (mean ± SEM; n=12-15/diet group) (A).

At 16 weeks, mice were euthanized and terminal body weights were noted. Each point represents terminal body weight for an individual mouse. Horizontal lines denote the mean for each diet group, n=12-15/diet group (B). No significant differences were found by two-way

ANOVA.

Supplementary Fig. III. Terminal organ/body weight ratios Irradiated LDLrKO mice received bone marrow from WT or GPR120 KO donors and were fed ADs (PO, FO, EO or BO) for 16 weeks. Mice were then euthanized and liver (A), spleen (B), and gonadal adipose tissue (C) wet weights were measured and normalized to terminal body weight., Each point represents terminal body weight/ organ weigh ratio for an individual mouse,. Horizontal lines denote the

147 mean for each diet group, n=12-15/diet group. No significant differences were found by two-way

ANOVA.

Supplementary Fig. IV. RBC fatty acid (FA) composition. RBCs were harvested from WT and KO mice consuming the indicated AD at baseline (chow diet) and after 6 weeks of AD feeding; fatty acid percentage distribution was measured as described in the methods. Data for individual fatty acids are expressed as percentage composition of total fatty acids. An equal volume pool of RBCs from 4 mice/diet group/genotype was used for analysis. Since no significant differences were detected between WT and KO mice fed the same AD, WT and KO data were pooled and plotted for mice fed the same AD. Mean ± SEM, n=8. Groups with different letters are significantly different (p<0.05) by one-way ANOVA and Tukey’s post-hoc analysis.

Supplementary Fig. V. Liver Histology. Livers were harvested from 16 week AD-fed WT and

KO mice, formalin fixed, sectioned (8um thick), and stained for H&E and CD68 using immunohistochemistry (IHC). Representative H&E stained liver sections (A) and CD68 IHC images (B) from WT and KO mice fed each AD. Qualitative microscopic observation did not reveal any histological differences between WT and KO mice fed the same AD.

Supplementary Fig. VI. Effect of L-GPR120 on circulating monocytosis. Irradiated LDLrKO mice received bone marrow from WT or GPR120 KO donors and were fed ADs (PO, FO, EO or

BO) for 16 weeks before blood was harvested to measure the distribution of monocytes by flow cytometry. Monocytes were defined as CD11b+, CD115+, Ly6G- cells. Percentage blood

148 monocytes (A) and % blood Ly6Chi monocytes (B). n=12-14/group. Groups with different letters are significantly different (p<0.05) by two-way ANOVA and Tukey’s post-hoc analysis.

Supplementary Fig. VII. Monocyte recruitment experiment. Irradiated LDLrKO mice received bone marrow from WT or GPR120 KO donors and were fed ADs (PO, FO, EO or BO) for 16 weeks. Percentage blood monocyte labeling efficiency was measured as % FITC+, CD115+,

Ly6G- cells 24 hours after Fluoresbrite microparticle injection. n=12-14/group. No significant differences were found by two-way ANOVA (A). 5 days later, the number of FITC+ beads was quantified in intimal area of serially sectioned 8 µm thick aortic roots using fluorescent microscope. A minimum of 8-10 sections were analyzed and then averaged for individual aortic root/ animal. n=6-8 animals/group. Bars with different letters denote statistically significant difference, p<0.05 (B).

Supplementary Fig. VIII. Representative aortic root atherosclerotic lesions stained for Oil- red-O. Irradiated LDLrKO mice received bone marrow from WT or GPR120 KO donors and were fed ADs (PO, FO, EO or BO) for 16 weeks. At necropsy, hearts were frozen in OCT, and aortic roots (8 µm thick) were serially sectioned and stained using Oil Red O. Representative images of aortic root Oil Red O positive intimal area (lesion area) of individual WT and KO mice for each diet group.

Supplementary Fig. IX. Representative aortic root atherosclerotic lesions stained for

CD68/ macrophages. Irradiated LDLrKO mice received bone marrow from WT or GPR120 KO donors and were fed ADs (PO, FO, EO or BO) for 16 weeks. At necropsy, hearts were frozen in

149

OCT, and aortic roots (8 µm thick) were serially sectioned and stained using anti-CD68 antibody. Representative images of aortic root CD68+ intimal area (lesion area) of individual WT and KO mice for each diet group.

Supplementary Fig. X. Representative aortic root atherosclerotic lesions stained for

Sirius red/ collagen. Irradiated LDLrKO mice received bone marrow from WT or GPR120 KO donors and were fed ADs (PO, FO, EO or BO) for 16 weeks. At necropsy, hearts were frozen in

OCT, and aortic roots (8 µm thick) were serially sectioned and stained using picrosirius red dye.

Representative images of aortic roots obtained under plane polarized light of individual WT and

KO mice fed respective AD.

Supplementary Fig. XI. Representative aortic root atherosclerotic lesions stained for

CD11c/ dendritic cells. Irradiated LDLrKO mice received bone marrow from WT or GPR120

KO donors and were fed ADs (PO, FO, EO or BO) for 16 weeks. At necropsy, hearts were frozen in OCT, and aortic roots (8 µm thick) were serially sectioned and stained using anti-CD68 antibody. Representative images of aortic root CD68+ intimal area of individual WT and KO mice fed FO. Mice fed other (than FO) atherogenic diets did not differ from one another (not depicted).

Supplementary Fig. XII. Representative aortic root atherosclerotic lesions stained for cleaved-caspase-3/ apoptotic cells. Irradiated LDLrKO mice received bone marrow from WT or GPR120 KO donors and were fed ADs (PO, FO, EO or BO) for 16 weeks. At necropsy, hearts were frozen in OCT, and aortic roots (8 µm thick) were serially sectioned and stained

150 using anti-cleaved caspase-3 (CC-3) antibody. Representative images of aortic root CC-3+ intimal area of individual WT and KO mice fed FO. Mice fed other (than FO) atherogenic diets did not differ from one another (not depicted).

151

Supplementary Figures

152

153

154

155

156

157

158

159

160

161

162

163

Reference List

1. Rudel LL, Kelley K, Sawyer JK, Shah R and Wilson MD. Dietary monounsaturated fatty acids promote aortic atherosclerosis in LDL receptor-null, human ApoB100-overexpressing transgenic mice. Arteriosclerosis, thrombosis, and vascular biology. 1998;18:1818-27.

2. Shewale SV, Boudyguina E, Zhu X, Shen L, Hutchins PM, Barkley RM, Murphy RC and

Parks JS. Botanical oils enriched in n-6 and n-3 fatty acid products of FADS2 are equally effective in preventing atherosclerosis and hepatosteatosis in mice. Journal of lipid research.

2015.

3. Zhang P, Boudyguina E, Wilson MD, Gebre AK and Parks JS. Echium oil reduces plasma lipids and hepatic lipogenic gene expression in apoB100-only LDL receptor knockout mice. J Nutr Biochem. 2008;19:655-63.

4. Rong S, Cao Q, Liu M, Seo J, Jia L, Boudyguina E, Gebre AK, Colvin PL, Smith TL,

Murphy RC, Mishra N and Parks JS. Macrophage 12/15 lipoxygenase expression increases plasma and hepatic lipid levels and exacerbates atherosclerosis. Journal of lipid research.

2012;53:686-95.

5. Bligh EG and Dyer WJ. A rapid method of total lipid extraction and purification. Canadian journal of biochemistry and physiology. 1959;37:911-7.

6. Zhu X, Lee JY, Timmins JM, Brown JM, Boudyguina E, Mulya A, Gebre AK, Willingham

MC, Hiltbold EM, Mishra N, Maeda N and Parks JS. Increased cellular free cholesterol in macrophage-specific Abca1 knock-out mice enhances pro-inflammatory response of macrophages. The Journal of biological chemistry. 2008;283:22930-41.

7. Carr TP, Andresen CJ and Rudel LL. Enzymatic determination of triglyceride, free cholesterol, and total cholesterol in tissue lipid extracts. Clin Biochem. 1993;26:39-42.

164

8. Galkina E, Kadl A, Sanders J, Varughese D, Sarembock IJ and Ley K. Lymphocyte recruitment into the aortic wall before and during development of atherosclerosis is partially L- selectin dependent. J Exp Med. 2006;203:1273-82.

9. Tacke F, Ginhoux F, Jakubzick C, van Rooijen N, Merad M and Randolph GJ. Immature monocytes acquire antigens from other cells in the bone marrow and present them to T cells after maturing in the periphery. J Exp Med. 2006;203:583-97.

10. Tacke F, Alvarez D, Kaplan TJ, Jakubzick C, Spanbroek R, Llodra J, Garin A, Liu J,

Mack M, van Rooijen N, Lira SA, Habenicht AJ and Randolph GJ. Monocyte subsets differentially employ CCR2, CCR5, and CX3CR1 to accumulate within atherosclerotic plaques. J

Clin Invest. 2007;117:185-94.

11. Brown AL, Zhu X, Rong S, Shewale S, Seo J, Boudyguina E, Gebre AK, Alexander-

Miller MA and Parks JS. Omega-3 fatty acids ameliorate atherosclerosis by favorably altering monocyte subsets and limiting monocyte recruitment to aortic lesions. Arteriosclerosis, thrombosis, and vascular biology. 2012;32:2122-30.

12. Daugherty A and Whitman SC. Quantification of atherosclerosis in mice. Methods in molecular biology. 2003;209:293-309.

165

CHAPTER IV

DISCUSSION

Swapnil V. Shewale

166

DISCUSSION

Chronic inflammation, primarily mediated via intimal macrophage accumulation and activation, plays a critical role in atherosclerotic lesion progression13. Dietary PUFAs impart atheroprotection in mice, non-human primates and humans11, 14-22. However, the extent to which this atheroprotection is imparted via attenuation of macrophage inflammation, particularly via

GPR120, and whether activation of leukocyte/ macrophage GPR120 affects lipid-lipoprotein metabolism was unknown.

We used a dietary approach to achieve efficient in vivo enrichment of cell membranes and plasma lipoproteins in >18 carbon n-3 or n-6 PUFAs by identifying botanical oils relatively enriched in n-3 or n-6 PUFAs beyond rate limiting FADS2 enzyme. Using this approach, we previously showed that an atherogenic diet containing EO, which is relatively enriched in stearidonic acid (18:4 n-3), the immediate product of FADS2-mediated desaturation of 18:3 n-3, effectively enriches plasma and tissue lipids in the anti-inflammatory PUFA 20:5 n-3 and was as atheroprotective as dietary fish oil (FO) compared to the saturated/monounsaturated fatty acid enriched palm oil (PO)11, 23. However, whether a similar strategy of dietary enrichment in FADS-

2 n-6 products would lead to atheroprotection is unknown. To address this gap in knowledge, we tested the hypothesis that dietary borage oil (BO), enriched in the FADS-2 product 18:3 n-6, would not be as atheroprotective as EO, due to in vivo conversion of 18:3 n-6 to 20:4 n-6, a pro- inflammatory eicosanoid precursor.

Contrary to our hypothesis, we found that BO and EO equally attenuated atherosclerosis and hepatic neutral lipid accumulation relative to PO. The reduction in atherosclerosis resulting from BO feeding appeared due to plasma VLDL-c reduction as well as attenuation of macrophage inflammation and chemotaxis despite significant enrichment of plasma and liver lipids in 20:4 n-6, a pro-inflammatory eicosanoid precursor (Chapter II).

167

However, BO vs. EO feeding did not result in increased 20:4 n-6 derived eicosanoid release in

LPS-stimulated peritoneal macrophage cultured media. In fact, we found that 12/15 lipooxygenase-derived oxidized 18:2 n-6 metabolites, 9 and 13 HODE, were predominantly secreted over 20:4 n-6-derived eicosanoids. Since oxidized cholesteryl ester (ox-CE) species derived from 18:2 n-6 have been detected in advanced human atheromas24, we determined whether ox-CE species may contribute to athero-progression in mouse aortas as well. We found that mouse aortas were devoid of ox-CE’s, implying a minimal role of ox-CE species in mouse atherosclerosis.

BO reduced hepatic neutral lipid accumulation in similar fashion to that of EO and FO.

This is a novel observation since diets enriched in 18:2 n-6 do not reduce hepatic neutral lipid content25, 26. Reduced nuclear SREBP-1 content is likely the key determinant of reduced hepatic neutral lipid content in BO, EO and FO-fed mice. Multiple mechanisms have been proposed that may explain reduced nuclear SREBP-1 content including reduced mRNA expression27, mRNA decay28, proteosomal degradation and inactivation by LXR agonists29.

However, understanding of detailed mechanisms regarding how PUFAs inhibit proteolytic cleavage of SREBP-1 is limited. Additionally, BO, EO and FO induced hepato-protection, in part, is mediated via reduction in monounsaturated cholesteryl ester formation via SOAT230.

SOAT2 deletion, similar to BO, EO and FO diet feeding, is atheroprotective and attenuates hepatic neutral lipid content 31 32. The reduction in hepatic neutral lipid accumulation in BO-fed mice is despite elevated hepatic TG secretion and plasma TG levels relative to EO and FO. This paradox is likely due to the fact that quantitatively, only a fraction of hepatic TG is mobilized for secretion in VLDL; therefore, a large decrease in hepatic TG content does not necessarily result in decreased hepatic VLDL TG secretion. For example, SCD-1 silencing using anti-sense oligonucleotide results in a 90% reduction in hepatic TG content, but does not affect hepatic

TG secretion compared to mice treated with a non-targeting anti-sense oligonucleotide33,

168 suggesting that a secretory pool of TG may be regulated differently than the bulk TG storage pool in hepatocytes. Hence, collectively our results show that BO and EO-enriched FADS2 fatty acid products 18:4 n-3 and 18:3 n-6 PUFAs are efficiently converted to longer chain ≥ 20 carbon

PUFAs, resulting in membrane and plasma lipid enrichment in 20:5 n-3 and 20:4 n-6, respectively, which in turn, reduces hepatic steatosis and aortic atherosclerosis.

In addition to reduced hepatic neural lipid content, BO, EO and FO fed mice also had reduced expression of the macrophage marker, CD68, and a chemokine, MCP-1, in liver. We also observed decreased aortic root macrophage (CD68+ cells) content, decreased macrophage chemotaxis in vitro, and decreased peritoneal macrophage inflammatory response to LPS in vitro. These observations prompted us to consider the atheroprotective potential of

GPR120, a negative regulator of macrophage inflammation and chemotaxis.

We determined the atheroprotective potential of GPR120, which is an n-3 fatty acid receptor mediating potent anti-inflammatory and insulin-sensitizing effects34, 35. Since its deorphanization, GPR120 has been described as a physiological mediator of long chain

PUFAs36 and as the physiological mediatory of n-3 PUFAs37. GPR120 activation has been implicated in energy homeostasis, regulation of obesity, insulin sensitivity and chronic inflammation34-36, 38 . However, whether leukocyte GPR120 (L-GPR120) activation by n-3 vs. n-6

PUFA is atheroprotective in vivo was unknown. We hypothesized that dietary n-3 PUFAs would lead to greater activation of GPR120 and less inflammation than n-6 PUFAs, since in vivo activation of GPR120 by n-3 PUFAs has been shown to reverse chronic inflammation and glucose intolerance resulting from high fat diet feeding (ref). Additionally, GPR120 is highly expressed in macrophages and bone marrow transplantation studies indicate that most of the beneficial effects of n-3 PUFA are mediated via macrophage GPR12037. GPR120 activation blunts the activation of TAK137, a common node of Toll-like receptor mediated pro-inflammatory signaling and NLRP3 inflammasome activation 39.

169

Our study on whether L-GPR120 activation is atheroprotective led to several novel observations. We report that GPR120 activation in vivo was not n-3 PUFA selective; n-6 PUFAs also activated GPR120. In our studies, physiological levels of dietary PUFA intake in LDLrKO were sufficient for in vivo activation of L-GPR120, leading to attenuated neutrophilia, splenic monocytosis, and monocyte recruitment into aortic intima, whereas deletion of L-GPR120 expression in hepatic Kupffer cells resulted in increased inflammation in the context of increased dietary consumption of PUFAs, independent of hepatic neutral lipid content (Chapter

III). Although, hypercholesterolemia is known to induce monocytosis, in our studies, L-GPR120 activation attenuated monocytosis and neutrophilia independent of plasma cholesterol content which was equivalent among all PUFA fed groups. Additionally, L-GPR120 deletion had minimal impact on plasma lipid concentration, lipoprotein cholesterol distribution, and aortic root intimal area and aortic cholesterol content. Hence, we conclude that L-GPR120 is neither sufficient nor required to impart significant atheroprotection or hepatoprotection in LDLrKO mice fed n-3 or n-

6 PUFA-enriched diets.

Contradictory results exist regarding specificity of GPR120 activation by n-3 PUFAs and its influence on hepatic steatosis34, 35 40. Our study, indicates that either class of PUFAs may activate GPR120 in vivo. To our knowledge, other studies have not investigated the possibility of GPR120 activation by FADS2 derived PUFAs although Initial studies had identified: α-linoleic acid and γ-linoleic acid as high potency fatty acid activators of GPR12035. In a recent study,

DHA, oleic, palmitoleic, palmitic acid were shown to activate GPR120; however n-6 PUFA including linoleic, γ-linoleic acid and AA were not reported/tested 41.

Additionally, in our study L-GPR120 deletion did not significantly affect hepatic neutral lipid content, in contrast to other studies using whole body GPR120 KO mice 34, 35, 37. In studies using whole body GPR120 KO (GPR120 KO) mice, a super-physiological fish oil (27% w/w) supplementation was fed35, 37. Additionally, only GPR120 KO mice bred on a mixed

170

129Sv/C57BL/6 background, in contrast to GPR120 KO mice bred on a pure C57BL/6N background show increased hepatic neutral lipid content40. In summary, possible explanations for above discrepancies include different gene targeting constructs, mouse backgrounds, and dietary n-3 PUFA content and length of diet feeding.

Our results regarding effects of L-GPR120 deletion on monocytosis and neutrophilia and hepatic Kupffer cell inflammation support the conclusion that sufficient PUFA enrichment occurred in vivo to activate GPR120, but the anticipated GPR120-mediated reduction in atherosclerosis was not observed. Factors that may have contributed to above results regarding L-GPR120 deletion include activation of L-GPR120 in non-bone marrow derived cells that express GPR120 and may impart an atheroprotective effect. For example, a study showed

GPR120 activation in perivascular adipose tissue (PVAT) by endogenously synthesized n-3

PUFAs in fat-1 transgenic mice protected from femoral arterial thrombosis, neointimal hyperplasia and vascular inflammation 42. Secondly, in LDLrKO mice, elevated cholesterol content in ApoB containing lipoproteins, particularly VLDL, rather than increased macrophage inflammation, is likely the determining factor in pathogenesis of aortic root atherosclerosis43.

As described in Chapter II, the levels of dietary PUFA intake in our study were in a physiological range that would be achievable for human consumption. Whether similar effects regarding macrophage inflammation will occur in humans consuming diets with a similar PUFA enrichment is unknown.

171

Implications for human health

A meta-analysis of 13 cohort studies (involving 310,602 individuals and 12,479 coronary heart disease events) indicate that increased consumption of dietary n-6 PUFAs, particularly LA is potentially atheroprotective in the general population. A 5% increase in energy intake from LA is associated with a 10% and 13% lower risk of coronary heart disease events and deaths, respectively44. However, populations harbouring a single nucleotide polymorphism (rs174537) in the FADS1/2 gene cluster have a small, but significant, enrichment in plasma AA levels and show increased production of LTB4 and 5-hydroxyeicosatetraenoic acids in zymosan- stimulated blood45, 46. Thus, individuals harboring genetic FADS polymorphisms (rs174537) may be hypersensitive to LA and GLA-enriched diets. However, whether this potential hyperresponsiveness due to rs174537 affects coronary heart disease risk is unknown.

Additionally, a non-synonymous mutation (p.R270H) inhibits GPR120 signaling activity, resulting in increased risk of obesity in European populations35, indicating that GPR120 activation by

PUFAs has implications for metabolic disorders. Whether the p.R270H mutation affects coronary heart disease risk is currently unknown. However, since PUFAs are potential GPR120 ligands and known to reduce coronary heart disease risk, it may be worth determining coronary heart disease risk in individuals with the p.R270H mutation.

Overall Conclusion

Dietary n-3 and n-6 PUFAs regulate multiple pathways involved in lipid metabolism and macrophage inflammation. Notably FADS2 derived n-6 PUFAs are as atheroprotective and hepatoprotective as their n-3 counterparts. Although, some of the effects of dietary PUFAs with regard to attenuation of macrophage inflammation are primarily mediated via activation of

GPR120, dietary PUFA-induced atheroprotection in LDLrKO mice was independent of leukocyte

172

GPR120 expression by several measurements, including aortic root intimal area and whole aorta cholesterol content.

Future Studies

PUFA-induced attenuation of inflammation involves multiple pathways. Based on our results and other animal studies, attenuation of macrophage/Kupffer cell inflammation, monocytosis, and neutrophilia via activation of GPR120 may have implications for improving chronic inflammation associated with hight fat diet feeding 34, 47. Although lipopolysaccharide

(LPS) is known to activate Toll-like receptor 4 (TLR4), during chronic inflammation, aseptic activation of TLR4 by saturated fatty acids can also activate the canonical NF-κB proinflammatory pathway (Figure 1, pathway 1). GPR120 activation by n-3 and n-6 PUFAs attenuates NF-κB pathway due to GPR120 internalization via binding with β-arrestin2. The internalized complex then binds TAB1 (TAK1 binding protein 1), effectively blocking TAB1 binding to TAK1 (transforming growth factor-β activated kinase) required for activation of the

NF-κB pathway (Figure 1, pathway 2). Recent evidence suggests that inactivation of NLRP3 inflammasome (a large multimeric protein complex that is activated by pathogen-associated molecular patterns) occurs due to GPR120 (Figure 1, pathway 2) activation and may mediate part of the anti-inflamatory effects of PUFAs39. Whether n-3 and n-6 PUFAs equally attenuate inflammasome activation is unknown. Additionally, since GPR120 is highly expressed in adipose tissue and macrophages, whether adipose tissue macrophage inflammasome inactivation by PUFAs protects from obesity and insulin resistance needs to be determined.

Autophagy is a homeostatic process that degrades cytosolic macromolecules, damaged organelles and pathogens, whereas lipophagy is a lysosomal degradative pathway for

TG and cholesteryl esters in lipid droplets. Both autophagy and lipophagy can regulate intracellular lipid trafficking and storage (reviewed in 48, 49). Inhibition of autophagy results in

173 increased intracellular lipid storage in the form of lipid droplets and attenuation of fatty acid β- oxidation50, whereas n-3 PUFA can induce activation of autophagy via AMPK pathway51 (Figure

1, pathway 3). Hence, autophagy likely mediates part of the lipid lowering effects of n-3 PUFAs.

Activation of autophagy can also limit inflammasome activation by degrading the inflammasome and pro-interleukin 1-β (IL-1β) (Figure 1, pathway 4) 52, 53. Secondly, since n-3 PUFAs inactivate NLRP3 inflammasome, activation of autophagy may be a pathway by which n-3

PUFAs regulate both inflammation and lipid metabolism and this pathway needs to be further studied in context of n-3 and n-6 PUFAs.

174

Figure 1.

.

Figure 1. Summary of potential macrophage anti-inflammatory pathways for botanical

PUFAs. (1) TLR4 activation by LPS or saturated fatty acids such as palmitate activates canonical NFκB pathway via interaction of TAB1 and Tak1 adapter proteins, resulting in increased pro-inflammatory gene transcription. (2) GPR120 activation by n-3 and n-6 PUFA results in internalization via in β-arrestin2. The GPR120-β arrestin complex binds TAB1, and inhibits NFκB activation by preventing TAB1 -TAK1 interaction. A similar pathway inhibits

NLRP3 inflammasome activation with β arrestin binding to NLRP3, resulting in decreased processing of pro-caspase 1 and pro-IL-1β, and decreased secretion of mature IL-1β. 3) n-3 and n-6 PUFAs or their oxidized derivatives, increase AMPK activation resulting in activation of autophagy. LPS binding to TLR4 also activates autophagy which is inhibited by palmitate via

175

AMPK. 4) Activation of autophagy can inhibit inflammation by inducing NLRP3 inflammasome degradation and pro-IL-1β degradation. Red arrows denote proinflammatory pathways; green lines denote various pathways activated by PUFAs, black lines denote the pro-inflammatory pathways blocked by PUFAs and blue lines denote potential pathway by which PUFA-induced autophagy may inhibit inflammation.

176

Reference List

1. Rudel LL, Kelley K, Sawyer JK, Shah R and Wilson MD. Dietary monounsaturated fatty acids promote aortic atherosclerosis in LDL receptor-null, human ApoB100-overexpressing transgenic mice. Arteriosclerosis, thrombosis, and vascular biology. 1998;18:1818-27.

2. Shewale SV, Boudyguina E, Zhu X, Shen L, Hutchins PM, Barkley RM, Murphy RC and

Parks JS. Botanical oils enriched in n-6 and n-3 fatty acid products of FADS2 are equally effective in preventing atherosclerosis and hepatosteatosis in mice. Journal of lipid research.

2015.

3. Zhang P, Boudyguina E, Wilson MD, Gebre AK and Parks JS. Echium oil reduces plasma lipids and hepatic lipogenic gene expression in apoB100-only LDL receptor knockout mice. J Nutr Biochem. 2008;19:655-63.

4. Rong S, Cao Q, Liu M, Seo J, Jia L, Boudyguina E, Gebre AK, Colvin PL, Smith TL,

Murphy RC, Mishra N and Parks JS. Macrophage 12/15 lipoxygenase expression increases plasma and hepatic lipid levels and exacerbates atherosclerosis. Journal of lipid research.

2012;53:686-95.

5. Bligh EG and Dyer WJ. A rapid method of total lipid extraction and purification. Canadian journal of biochemistry and physiology. 1959;37:911-7.

6. Zhu X, Lee JY, Timmins JM, Brown JM, Boudyguina E, Mulya A, Gebre AK, Willingham

MC, Hiltbold EM, Mishra N, Maeda N and Parks JS. Increased cellular free cholesterol in macrophage-specific Abca1 knock-out mice enhances pro-inflammatory response of macrophages. The Journal of biological chemistry. 2008;283:22930-41.

7. Carr TP, Andresen CJ and Rudel LL. Enzymatic determination of triglyceride, free cholesterol, and total cholesterol in tissue lipid extracts. Clin Biochem. 1993;26:39-42.

177

8. Galkina E, Kadl A, Sanders J, Varughese D, Sarembock IJ and Ley K. Lymphocyte recruitment into the aortic wall before and during development of atherosclerosis is partially L- selectin dependent. J Exp Med. 2006;203:1273-82.

9. Tacke F, Ginhoux F, Jakubzick C, van Rooijen N, Merad M and Randolph GJ. Immature monocytes acquire antigens from other cells in the bone marrow and present them to T cells after maturing in the periphery. J Exp Med. 2006;203:583-97.

10. Tacke F, Alvarez D, Kaplan TJ, Jakubzick C, Spanbroek R, Llodra J, Garin A, Liu J,

Mack M, van Rooijen N, Lira SA, Habenicht AJ and Randolph GJ. Monocyte subsets differentially employ CCR2, CCR5, and CX3CR1 to accumulate within atherosclerotic plaques. J

Clin Invest. 2007;117:185-94.

11. Brown AL, Zhu X, Rong S, Shewale S, Seo J, Boudyguina E, Gebre AK, Alexander-

Miller MA and Parks JS. Omega-3 fatty acids ameliorate atherosclerosis by favorably altering monocyte subsets and limiting monocyte recruitment to aortic lesions. Arteriosclerosis, thrombosis, and vascular biology. 2012;32:2122-30.

12. Daugherty A and Whitman SC. Quantification of atherosclerosis in mice. Methods in molecular biology. 2003;209:293-309.

13. Lusis AJ. Atherosclerosis. Nature. 2000;407:233-41.

14. Bigger JT, Jr. and El-Sherif T. Polyunsaturated fatty acids and cardiovascular events: a fish tale. Circulation. 2001;103:623-5.

15. Calder PC and Grimble RF. Polyunsaturated fatty acids, inflammation and immunity. Eur

J Clin Nutr. 2002;56 Suppl 3:S14-9.

178

16. Kang JX and Leaf A. Antiarrhythmic effects of polyunsaturated fatty acids. Recent studies. Circulation. 1996;94:1774-80.

17. Parks JS, Kaduck-Sawyer J, Bullock BC and Rudel LL. Effect of dietary fish oil on coronary artery and aortic atherosclerosis in African green monkeys. Arteriosclerosis.

1990;10:1102-12.

18. Hu FB, Bronner L, Willett WC, Stampfer MJ, Rexrode KM, Albert CM, Hunter D and

Manson JE. Fish and omega-3 fatty acid intake and risk of coronary heart disease in women.

JAMA : the journal of the American Medical Association. 2002;287:1815-21.

19. Harris WS, Mozaffarian D, Rimm E, Kris-Etherton P, Rudel LL, Appel LJ, Engler MM,

Engler MB and Sacks F. Omega-6 fatty acids and risk for cardiovascular disease: a science advisory from the American Heart Association Nutrition Subcommittee of the Council on

Nutrition, Physical Activity, and Metabolism; Council on Cardiovascular Nursing; and Council on

Epidemiology and Prevention. Circulation. 2009;119:902-7.

20. Harris WS and Shearer GC. Omega-6 Fatty Acids and Cardiovascular Disease: Friend or Foe? Circulation. 2014;130:1562-1564.

21. Farvid MS, Ding M, Pan A, Sun Q, Chiuve SE, Steffen LM, Willett WC and Hu FB.

Dietary Linoleic Acid and Risk of Coronary Heart Disease: A Systematic Review and Meta-

Analysis of Prospective Cohort Studies. Circulation. 2014;130:1568-1578.

22. Zampolli A, Bysted A, Leth T, Mortensen A, De Caterina R and Falk E. Contrasting effect of fish oil supplementation on the development of atherosclerosis in murine models.

Atherosclerosis. 2006;184:78-85.

23. Forrest LM, Boudyguina E, Wilson MD and Parks JS. Echium oil reduces atherosclerosis in apoB100-only LDLrKO mice. Atherosclerosis. 2012;220:118-21.

179

24. Hutchins PM, Moore EE and Murphy RC. Electrospray MS/MS reveals extensive and nonspecific oxidation of cholesterol esters in human peripheral vascular lesions. Journal of lipid research. 2011;52:2070-83.

25. Rudel LL, Parks JS and Sawyer JK. Compared with dietary monounsaturated and saturated fat, polyunsaturated fat protects African green monkeys from coronary artery atherosclerosis. Arteriosclerosis, thrombosis, and vascular biology. 1995;15:2101-10.

26. Thornburg JT, Parks JS and Rudel LL. Dietary fatty acid modification of HDL phospholipid molecular species alters lecithin: cholesterol acyltransferase reactivity in cynomolgus monkeys. Journal of lipid research. 1995;36:277-89.

27. Hannah VC, Ou J, Luong A, Goldstein JL and Brown MS. Unsaturated fatty acids down- regulate srebp isoforms 1a and 1c by two mechanisms in HEK-293 cells. The Journal of biological chemistry. 2001;276:4365-72.

28. Xu J, Teran-Garcia M, Park JH, Nakamura MT and Clarke SD. Polyunsaturated fatty acids suppress hepatic sterol regulatory element-binding protein-1 expression by accelerating transcript decay. The Journal of biological chemistry. 2001;276:9800-7.

29. Ou J, Tu H, Shan B, Luk A, DeBose-Boyd RA, Bashmakov Y, Goldstein JL and Brown

MS. Unsaturated fatty acids inhibit transcription of the sterol regulatory element-binding protein-

1c (SREBP-1c) gene by antagonizing ligand-dependent activation of the LXR. Proceedings of the National Academy of Sciences of the United States of America. 2001;98:6027-32.

30. Temel RE, Hou L, Rudel LL and Shelness GS. ACAT2 stimulates cholesteryl ester secretion in apoB-containing lipoproteins. Journal of lipid research. 2007;48:1618-27.

31. Alger HM, Brown JM, Sawyer JK, Kelley KL, Shah R, Wilson MD, Willingham MC and

Rudel LL. Inhibition of acyl-coenzyme A:cholesterol acyltransferase 2 (ACAT2) prevents dietary

180 cholesterol-associated steatosis by enhancing hepatic triglyceride mobilization. The Journal of biological chemistry. 2010;285:14267-74.

32. Melchior JT, Sawyer JK, Kelley KL, Shah R, Wilson MD, Hantgan RR and Rudel LL. LDL particle core enrichment in cholesteryl oleate increases proteoglycan binding and promotes atherosclerosis. Journal of lipid research. 2013;54:2495-503.

33. Brown JM, Chung S, Sawyer JK, Degirolamo C, Alger HM, Nguyen TM, Zhu X, Duong

MN, Brown AL, Lord C, Shah R, Davis MA, Kelley K, Wilson MD, Madenspacher J, Fessler MB,

Parks JS and Rudel LL. Combined therapy of dietary fish oil and stearoyl-CoA desaturase 1 inhibition prevents the metabolic syndrome and atherosclerosis. Arteriosclerosis, thrombosis, and vascular biology. 2010;30:24-30.

34. Oh da Y, Walenta E, Akiyama TE, Lagakos WS, Lackey D, Pessentheiner AR, Sasik R,

Hah N, Chi TJ, Cox JM, Powels MA, Di Salvo J, Sinz C, Watkins SM, Armando AM, Chung H,

Evans RM, Quehenberger O, McNelis J, Bogner-Strauss JG and Olefsky JM. A Gpr120- selective agonist improves insulin resistance and chronic inflammation in obese mice. Nature medicine. 2014;20:942-7.

35. Ichimura A, Hirasawa A, Poulain-Godefroy O, Bonnefond A, Hara T, Yengo L, Kimura I,

Leloire A, Liu N, Iida K, Choquet H, Besnard P, Lecoeur C, Vivequin S, Ayukawa K, Takeuchi

M, Ozawa K, Tauber M, Maffeis C, Morandi A, Buzzetti R, Elliott P, Pouta A, Jarvelin MR,

Korner A, Kiess W, Pigeyre M, Caiazzo R, Van Hul W, Van Gaal L, Horber F, Balkau B, Levy-

Marchal C, Rouskas K, Kouvatsi A, Hebebrand J, Hinney A, Scherag A, Pattou F, Meyre D,

Koshimizu TA, Wolowczuk I, Tsujimoto G and Froguel P. Dysfunction of lipid sensor GPR120 leads to obesity in both mouse and human. Nature. 2012;483:350-4.

181

36. Hirasawa A, Tsumaya K, Awaji T, Katsuma S, Adachi T, Yamada M, Sugimoto Y,

Miyazaki S and Tsujimoto G. Free fatty acids regulate gut incretin glucagon-like peptide-1 secretion through GPR120. Nature medicine. 2005;11:90-4.

37. Oh DY, Talukdar S, Bae EJ, Imamura T, Morinaga H, Fan W, Li P, Lu WJ, Watkins SM and Olefsky JM. GPR120 is an omega-3 fatty acid receptor mediating potent anti-inflammatory and insulin-sensitizing effects. Cell. 2010;142:687-98.

38. Gotoh C, Hong YH, Iga T, Hishikawa D, Suzuki Y, Song SH, Choi KC, Adachi T,

Hirasawa A, Tsujimoto G, Sasaki S and Roh SG. The regulation of adipogenesis through

GPR120. Biochemical and biophysical research communications. 2007;354:591-7.

39. Yan Y, Jiang W, Spinetti T, Tardivel A, Castillo R, Bourquin C, Guarda G, Tian Z,

Tschopp J and Zhou R. Omega-3 Fatty Acids Prevent Inflammation and Metabolic Disorder through Inhibition of NLRP3 Inflammasome Activation. Immunity. 2013;38:1154-1163.

40. Bjursell M, Xu X, Admyre T, Bottcher G, Lundin S, Nilsson R, Stone VM, Morgan NG,

Lam YY, Storlien LH, Linden D, Smith DM, Bohlooly YM and Oscarsson J. The beneficial effects of n-3 polyunsaturated fatty acids on diet induced obesity and impaired glucose control do not require Gpr120. PloS one. 2014;9:e114942.

41. Suckow AT, Polidori D, Yan W, Chon S, Ma JY, Leonard J and Briscoe CP. Alteration of the glucagon axis in GPR120 (FFAR4) knockout mice: a role for GPR120 in glucagon secretion.

The Journal of biological chemistry. 2014;289:15751-63.

42. Li X, Ballantyne LL, Che X, Mewburn JD, Kang JX, Barkley RM, Murphy RC, Yu Y and

Funk CD. Endogenously generated omega-3 Fatty acids attenuate vascular inflammation and neointimal hyperplasia by interaction with free Fatty Acid receptor 4 in mice. Journal of the

American Heart Association. 2015;4.

182

43. VanderLaan PA, Reardon CA, Thisted RA and Getz GS. VLDL best predicts aortic root atherosclerosis in LDL receptor deficient mice. Journal of lipid research. 2009;50:376-85.

44. Harris WS and Shearer GC. Omega-6 Fatty Acids and Cardiovascular Disease: Friend,

Not Foe? Circulation. 2014;130:1562-1564.

45. Hester AG, Murphy RC, Uhlson CJ, Ivester P, Lee TC, Sergeant S, Miller LR, Howard

TD, Mathias RA and Chilton FH. Relationship between a Common Variant in the Fatty Acid

Desaturase (FADS) Cluster and Eicosanoid Generation in Humans. Journal of Biological

Chemistry. 2014;289:22482-9.

46. Mathias RA, Sergeant S, Ruczinski I, Torgerson DG, Hugenschmidt CE, Kubala M,

Vaidya D, Suktitipat B, Ziegler JT, Ivester P, Case D, Yanek LR, Freedman BI, Rudock ME,

Barnes KC, Langefeld CD, Becker LC, Bowden DW, Becker DM and Chilton FH. The impact of

FADS genetic variants on omega6 polyunsaturated fatty acid metabolism in African Americans.

BMC genetics. 2011;12:50.

47. Raptis DA, Limani P, Jang JH, Ungethum U, Tschuor C, Graf R, Humar B and Clavien

PA. GPR120 on Kupffer cells mediates hepatoprotective effects of omega3-fatty acids. Journal of hepatology. 2014;60:625-32.

48. Singh R and Cuervo AM. Lipophagy: connecting autophagy and lipid metabolism.

International journal of cell biology. 2012;2012:282041.

49. Singh R. Autophagy and regulation of lipid metabolism. Results and problems in cell differentiation. 2010;52:35-46.

50. Singh R, Kaushik S, Wang Y, Xiang Y, Novak I, Komatsu M, Tanaka K, Cuervo AM and

Czaja MJ. Autophagy regulates lipid metabolism. Nature. 2009;458:1131-5.

183

51. Xue B, Yang Z, Wang X and Shi H. Omega-3 polyunsaturated fatty acids antagonize macrophage inflammation via activation of AMPK/SIRT1 pathway. PloS one. 2012;7:e45990.

52. Shi CS, Shenderov K, Huang NN, Kabat J, Abu-Asab M, Fitzgerald KA, Sher A and

Kehrl JH. Activation of autophagy by inflammatory signals limits IL-1beta production by targeting ubiquitinated inflammasomes for destruction. Nature immunology. 2012;13:255-63.

53. Harris J, Hartman M, Roche C, Zeng SG, O'Shea A, Sharp FA, Lambe EM, Creagh EM,

Golenbock DT, Tschopp J, Kornfeld H, Fitzgerald KA and Lavelle EC. Autophagy controls IL-

1beta secretion by targeting pro-IL-1beta for degradation. The Journal of biological chemistry.

2011;286:9587-97.

184

CURRICULUM VITAE

SWAPNIL VIJAY SHEWALE

230 Ridge Forest Ct, Winston-Salem, NC 27104 978-729-0185

[email protected]

HTTPS://WWW.LINKEDIN.COM/PUB/SWAPNIL-SHEWALE/19/B5B/938

Education

2007 Bachelor of Pharmacy/ Pharmaceutical Sciences, Pune University, India

2010 Master of Sciences, Boonshoft School of Medicine, Dayton, OH

2015 PhD, Wake Forest University, Winston Salem, NC

Honors and Awards

2008-2010 Fellowship: Graduate Research Assistant, Department of Pharmacology

and Toxicology, Boonshaft School of Medicine, Dayton, OH

2010 Graduate Student Excellence Award, Wright State University, Dayton, OH

2011 Best Poster, Integrative Category, Annual Research Retreat,

Wake Forest University

2012 Treasurer and Dept. Representative,

Graduate Student’s Association,Wake Forest University

2013-2014 Chair, Graduate Student’s Association, Wake Forest University

185

2013 Member: Wake Forest Team, 4th Annual Biotech Case Competition

(runner up team)

2014 Team Leader: Wake Forest Team, 5th Annual Biotech Case Competition

Positions and Titles

2005 Undergraduate Research Fellow, VP School of Biotechnology, Pune,

India

2006 Undergraduate Research Fellow, National Chemical Laboratory, Pune,

India

2007 Summer Pharmacy Intern, Blue Cross laboratories, Nashik, India

2007 Key Accounts Manager, Baxter Transfusion Therapeutics, Delhi, India

2008 Project Manager, Criterium Clinical Research, Pune, India

2013-2015 Senior Commercialization Intern, Wake Forest Innovations

2011-2015 Teaching Intern, Physical Therapy program, Winston Salem State

University, NC

Travel Awards

2012 Travel Award, Annual Botanical Centers Meeting, NIH, Bethesda, MD

2014 Travel Award: Gordon Research Seminars: Lipoprotein Metabolism, NH

2014 Travel Award: Kern Lipid Conference, CO

186

Oral Talks

2013 South East Lipid Research Conference, Pine Mountain, GA

2014 MHI/IVB 2014 Symposium, UNC McAllister Heart Institute, NC

2014 Columbia University, Department of Molecular Medicine, NY

2014 Lipoprotein Metabolism Section, National Heart, Lung and Blood Institute,

NIH, DC

2014 Perelman School of Medicine, University of Pennsylvania, PA

Professional Societies

2009-Current Member, American Heart Association

2013-Current Member, American Society for Pharmacology and Experimental

Therapeutics (ASPET)

2013-Current Member, American Physiological Society (APS)

2015 Member, Association of University Technology Transfer Managers (AUTM)

Publications

1. Swapnil V. Shewale, Elena Boudyguina, Xuewei Zhu, Lulu Shen, Patrick M. Hutchins,

Robert M. Barkley, Robert C. Murphy, John S. Parks. Botanical oils enriched in n-6 and n-

3 fatty acid products of FADS2 are equally effective in preventing atherosclerosis and

hepatosteatosis in mice. J Lipid Res. 2015 Apr 28. pii: jlr.M059170

187

2. Xin Bi, Xuewei Zhu, Chuan Gao, Shewale SV, Qiang Cao, Mingxia Liu, Elena

Boudyguina, Hermina Borgerink, Martha D. Wilson, Amanda L. Brown, and John S. Parks.

Myeloid Cell-Specific ABCA1 Deletion Has Minimal Impact on Atherogenesis in

Atherogenic Diet-Fed LDL Receptor Knockout Mice. Arterioscler Thromb Vasc Biol.

2014;34:1888-1899

3. Brown AL, Zhu X, Rong S, Shewale S, Seo J, Boudyguina E, Gebre AK, Alexander-Miller

MA, Parks JS. Omega-3 fatty acids ameliorate atherosclerosis by favorably altering

monocyte subsets and limiting monocyte recruitment to aortic lesions. Arterioscler Thromb

Vasc Biol. 2012 Sep; 32(9):2122-30.

4. Shewale S, Anstadt MP, Horenziak M, Izu B, Morgan EE, Lucot JB, Morris M. Sarin

causes autonomic imbalance and cardiomyopathy: an important issue for military and

civilian health.

J Cardiovasc Pharmacol. 2012 Jul; 60(1):76-87.

5. Senador D, Shewale S, Irigoyen MC, Elased KM, Morris M. Effects of restricted fructose

access on body weight and blood pressure circadian rhythms. Exp Diabetes Res. 2012;

2012:459087.

188

Published First Author Abstracts

6. Swapnil V Shewale; Elena Boudyguina; Xin Bi; Xuewei Zhu; Nilamadhab Mishra; Da Young

Oh; Jerrold Olefsky; John S. Parks. Role of Leukocyte GPR120 in Atherosclerosis

Development in Polyunsaturated Fatty Acid Diet Induced Atheroprotection in LDLrKO Mice.

Arterioscler Thromb Vasc Biol. 33: A121, 2013

7. Swapnil V Shewale, Elena Boudyguina, Patrick M Hutchins, Robert C Murphy, John S

Parks. Echium and Borage Oil Are as Atheroprotective as Fish Oil in Mice. Arterioscler

Thromb Vasc Biol. 32: A190, 2012

8. Swapnil V Shewale, Michael Horenziak, Brent Izu, James Lucot, Mark Anstadt, Mariana

Morris. Dilated Cardiomyopathy in Nerve Agent Sarin Treated Mice. Hypertension, 56: e66,

2010

9. Swapnil V Shewale, Felipe Lino, Danielle Senador, Khalid M Elased, Mariana Morris,

Restricted Fructose Intake Accentuates Body Weight Gain and Corticosterone without

Changing Blood Pressure. Hypertension , 54: e78, 2009

189