regulation of inflammatory macrophage activation and

160
Regulation of Inflammatory Macrophage Activation and Tolerance by Modulation of Glucose Metabolism The Harvard community has made this article openly available. Please share how this access benefits you. Your story matters Citation Langston, Paul Kent. 2019. Regulation of Inflammatory Macrophage Activation and Tolerance by Modulation of Glucose Metabolism. Doctoral dissertation, Harvard University, Graduate School of Arts & Sciences. Citable link http://nrs.harvard.edu/urn-3:HUL.InstRepos:42029764 Terms of Use This article was downloaded from Harvard University’s DASH repository, and is made available under the terms and conditions applicable to Other Posted Material, as set forth at http:// nrs.harvard.edu/urn-3:HUL.InstRepos:dash.current.terms-of- use#LAA

Upload: others

Post on 19-Dec-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Regulation of Inflammatory Macrophage Activation and

Regulation of InflammatoryMacrophage Activation and Toleranceby Modulation of Glucose Metabolism

The Harvard community has made thisarticle openly available. Please share howthis access benefits you. Your story matters

Citation Langston, Paul Kent. 2019. Regulation of Inflammatory MacrophageActivation and Tolerance by Modulation of Glucose Metabolism.Doctoral dissertation, Harvard University, Graduate School of Arts &Sciences.

Citable link http://nrs.harvard.edu/urn-3:HUL.InstRepos:42029764

Terms of Use This article was downloaded from Harvard University’s DASHrepository, and is made available under the terms and conditionsapplicable to Other Posted Material, as set forth at http://nrs.harvard.edu/urn-3:HUL.InstRepos:dash.current.terms-of-use#LAA

Page 2: Regulation of Inflammatory Macrophage Activation and

REGULATION OF INFLAMMATORY MACROPHAGE ACTIVATION AND TOLERANCE BY

MODULATION OF GLUCOSE METABOLISM

A DISSERTATION PRESENTED

BY

P. KENT LANGSTON

TO

THE COMMITTEE ON HIGHER DEGREES IN BIOLOGICAL SCIENCES IN PUBLIC HEALTH

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

IN THE SUBJECT OF

BIOLOGICAL SCIENCES IN PUBLIC HEALTH

HARVARD UNIVERSITY

CAMBRIDGE, MASSACHUSETTS

APRIL 2019

Page 3: Regulation of Inflammatory Macrophage Activation and

© 2019 PAUL KENT LANGSTON

ALL RIGHTS RESERVED

Page 4: Regulation of Inflammatory Macrophage Activation and

iii

ADVISORS: TIFFANY HORNG AND JAMES MITCHELL P. KENT LANGSTON

REGULATION OF INFLAMMATORY MACROPHAGE ACTIVATION AND TOLERANCE BY

MODULATION OF GLUCOSE METABOLISM

ABSTRACT

The reactions in an organism’s metabolic network take on dynamic changes in fluxes in response

to imposed stress to support calibrated functional responses. Macrophages, phagocytic cells of the

myeloid lineage, are activated during microbial infection to coordinate inflammatory responses

and host defense. Here we show that in macrophages activated by bacterial lipopolysaccharide

(LPS), mitochondrial glycerol 3-phosphate dehydrogenase (GPD2) regulates glucose oxidation to

drive inflammatory responses. GPD2, a component of the glycerol phosphate shuttle (GPS), boosts

glucose oxidation to fuel acetyl-CoA (Ac-CoA) production, histone acetylation and inflammatory

gene induction. While acute LPS exposure drives macrophage activation, prolonged exposure

triggers entry into LPS tolerance, where macrophages orchestrate immunosuppression to limit the

detrimental effects of sustained inflammation. We find that the shift from activation to tolerance

is modulated by GPD2, which coordinates a shutdown of oxidative metabolism that limits Ac-

CoA availability for histone acetylation at inflammatory genes, thus contributing to suppression

of inflammatory responses. Therefore, GPD2 and the GPS integrate the extent of microbial

stimulation with glucose oxidation to balance the beneficial and detrimental effects of the

inflammatory response. In addition, we show that LPS-inducible glucose oxidation is downstream

of the serine/threonine kinase Akt. We find that Akt signaling controls glucose oxidation to couple

the strength and duration of LPS exposure to the GPD2-dependent increase and decrease in

Page 5: Regulation of Inflammatory Macrophage Activation and

iv

mitochondrial respiration, promoter region histone acetylation, and inflammatory gene induction

in macrophages over the transition from activation to tolerance. Furthermore, transcriptional

profiling of tolerant macrophages re-exposed to LPS revealed a subset of highly-responsive,

primed (P) genes, important for antimicrobial activity and mitochondrial respiration. Induction of

P genes is Akt-dependent and fueled by acetate-derived Ac-CoA production catalyzed by

nucleocytosolic Ac-CoA synthetase (ACSS2) rather than by citrate-derived Ac-CoA from glucose

oxidation. Therefore, Akt and ACSS2 act as additional regulatory modules integrating TLR

signaling with shifts in macrophage metabolism to induce gene-specific responses to microbial

encounter. Taken together, our findings support a paradigm shift in immunometabolism,

demonstrating that glucose oxidation regulates induction and suppression of macrophage

inflammatory responses while also identifying several novel metabolic pathways controlling gene-

specific transcriptional responses in tolerant macrophages.

Page 6: Regulation of Inflammatory Macrophage Activation and

v

TABLE OF CONTENTS

Abstract.........................................................................................................................................iii

Table of Contents...........................................................................................................................v

Acknowledgements......................................................................................................................vii

List of Figures..............................................................................................................................xii

Chapter 1: Introduction

Sepsis: a tale of two syndromes...........................................................................................2

Innate response to microbial encounter: activation.............................................................3

Innate response to microbial encounter: suppression and tolerance...................................6

Immunometabolism: a new perspective on the regulation of immunity.............................9

Metabolic regulation of inflammatory macrophage polarization......................................14

Mitochondria as hubs in inflammatory activation of macrophages..................................18

Integration of metabolic signals by chromatin..................................................................20

Thesis summary.................................................................................................................21

References.........................................................................................................................23

Chapter 2: Glycerol phosphate shuttle enzyme GPD2 drives glucose oxidation to regulate

macrophage inflammatory responses

Abstract.............................................................................................................................35

Introduction.......................................................................................................................36

Results...............................................................................................................................39

Page 7: Regulation of Inflammatory Macrophage Activation and

vi

Discussion..........................................................................................................................61

Materials and Methods.......................................................................................................67

Acknowledgements............................................................................................................73

References..........................................................................................................................83

Chapter 3: Additional metabolic modules regulating macrophage responses to microbial

encounter: roles for Akt and nucleocytosolic acetyl-CoA synthetase ACSS2

Abstract..............................................................................................................................88

Introduction........................................................................................................................89

Results................................................................................................................................92

Discussion........................................................................................................................103

Materials and Methods.....................................................................................................109

Acknowledgements..........................................................................................................112

References........................................................................................................................116

Chapter 4: Discussion and Future Directions

Summary and Conclusions..............................................................................................121

Lessons from a worked example in immunometabolism................................................124

Concluding Thoughts and Future Directions...................................................................131

References........................................................................................................................140

Page 8: Regulation of Inflammatory Macrophage Activation and

vii

ACKNOWLEDGEMENTS

I would like to express my most sincere gratitude for those that have supported my scientific career

thus far. First, I must thank my advisor, Tiffany Horng for giving me the freedom to formulate and

test my own thesis while also providing indispensable insight throughout its execution. Tiffany

exhibits a tenacity that is truly inspiring and stems from the strength of her character and passion

for science. Although I likely could have learned more from her had I listened to her instructions

or taken her advice more frequently, our discourse over ideas and constant discussions about data

benefitted me tremendously. I will miss such conversations dearly. I must also thank my first

scientific advisor, Jason Grayson, who gave me my start in basic biological science while I was

studying Chemistry as an undergraduate at Wake Forest University in North Carolina. Jason is a

fantastic mentor, role model, and friend. I attribute many of the scientific leanings and

idiosyncrasies I have at the present to the formative years I spent under his tutelage.

An immense thank you also goes to the many lab mates and students I have had until now. In the

Horng Laboratory: Ibrahim Aksoylar, Munehiko Shibata, Anthony Covarrubias, Jiujiu Yu, Aya

Nambu, Jiahui Lei, Noriaki Maeshige, and Jon Jung. Ibrahim and Anthony taught me many of the

fundamental macrophage techniques I use today. They were also good friends who loved to talk

about science. Noriaki was only a temporary member of the lab and came from a very different

background as a physical therapist, yet he possessed levels of enthusiasm and kindness that made

teaching him a joy and a privilege. Our conversations about his work gave me an outlet for my

side passion for skeletal muscle biology. Jon Jung was my partner in crime for the summers of

2016 and 2018. His love for science and bench work almost made me change my opinions about

medical students. In 2016, after my PQE, Jon provided the additional manpower needed to get the

Page 9: Regulation of Inflammatory Macrophage Activation and

viii

“metabolic basis of tolerance” project going at full speed. In 2018, he came back and assisted

greatly in advancing the “metabolic basis of priming” project. Jon has been an outstanding friend

and colleague over the years and has set my standards very high for what a good student can

accomplish.

In the Mitchell Laboratory, I would like to thank: Jay Mitchell, Humberto Treviño, Lear Brace,

Pedro Meija, Chris Hine, Alban Longchamp, Sarah Mitchell, Luis Hernandez, Peter Kipp, Stuart

Adamson, and Mike MacArthur. Jay allowed me to rotate in his lab in my first summer as a

graduate student, which was a very memorable experience. He has also shown me immense

generosity and kindness by serving on my PQE, DAC, and Thesis Defense committees as well as

adopting me into his lab after Tiffany started her new lab at ShanghaiTech University at the end

of 2018. Enough positive things cannot be said about Jay, but above all, he is an outstanding mentor.

It has been a pleasure to watch his research program thrive and grow over the years. The members

of the Mitchell Lab, especially Humberto, Lear, Chris, Luis, Stu, and Mike, have played a major

role in my life in and outside of the lab at different times. I will sincerely miss the energy and

laughter they have brought.

In the Grayson Laboratory, I owe thanks to Katie Crump and Maddie Price. I am immensely

grateful to have had such a positive research experience as an undergraduate, and working with

Katie and Maddie was a significant source of that positivity. Katie was an outstanding example of

a bright and hard-working graduate student and mentor and was the one who showed me a 2011

article in Nature Reviews in Immunology by Diane Mathis that introduced me to the field of

immunometabolism. Katie taught me many things, but most important were how to be a careful

Page 10: Regulation of Inflammatory Macrophage Activation and

ix

scientist and how to eat lunch during a 12h-long intracellular cytokine staining protocol. Maddie

was the first student I ever trained and was a fellow Wake Forester that has gone on to do well in

the Immunology PhD Program at Emory University. She was a very level-headed and teachable

student and showed me how fulfilling it can be to train someone and watch their spark for science

grow into a bright flame.

I would also like to thank Renata Goncalves, Benny Garfinkel, Issam Ben-Sahra, Gerta Hoxhaj,

Ryan Alexander, and Vanessa Byles for their friendship and for all of the stimulating scientific

discussions and support they provided over the years I have been at HSPH. They are all outstanding

people and scientists, and I am fortunate to have shared so much time with them.

Many thanks to the other members of my PQE, DAC, and Thesis Defense Committees: Marianne

Wessling-Resnick, Jon Kagan, Edward Chouchani, Chih-Hao Lee, and Zhi-Min Yuan. Each of

these faculty has been a source of valuable advice, inspiration, and support. Special thanks to

Marianne for our many hallway conversations and for being such a great example to all of us BPH

students. And additional thanks to Ed for taking the time to chat about mitochondrial biology and

exercise physiology.

I would also like to thank Brendan Manning for his direction for the BPH Program and for

supporting me as I struggled to find footing at the start of my graduate training. His conduct and

contributions as a scientist are an inspiration to me and to others in BPH. Gratitude also goes out

to Deirdre Duckett and Tom Brazda for helping the Program run properly and for answering all of

my questions over the years.

Page 11: Regulation of Inflammatory Macrophage Activation and

x

It would be negligent to fail to mention some of the other people with whom I have worked and

had helpful discussions over the years at HSPH, so thanks also to Molly McNamara, Nelson

Knudson, and Urmila Powale and to my BPH cohort and members of the cohort above me, who

welcomed me to the Program: Joe Beyene, Maji Nwanaji-Enwerem, Cory Gerlach, Nate Hicks,

Kristine Werling, Allison Andraski, Allyson Morton, Andie Smidler, and Gabe Rangel.

Most of my life over the years I have lived in Boston have been spent in the laboratory (which I

would not trade for any other use); similarly, my undergraduate years were a mix of time spent in

the lab and in the ZSR Library and the Chemistry Reading Room. However, the friends I have had

outside the lab, although receiving a lesser portion of my time, have had an equivalent or greater

influence on my life thus far. First, I would like to thank my best friends of nearly a decade:

Matthew Gray, John Walsh, Jay Butterfield, and Alec Birmingham. You are my brothers, and there

are not enough words in my lexicon to tell you how much I appreciate you. Thank you for your

unconditional love and support and for taking time to have reunions. I hope we will have many

more years of experiences together. I would also like to thank Jerry Martin, Greg Jessup, and the

other members of Jack King’s Gym who overlapped with me from 2011-2014. These men showed

me that exhaustion from a long day of work is only in your mind. Thank you for being such positive

role models over the years and for your perpetual positivity and brotherly support.

Last but certainly not least on the list of friends to whom I am grateful is my buddy Brendon Smith,

who is the embodiment of a disciplined and driven mindset. I am thankful for the positive influence

Brendon has had on my life in the few years that I have known him. Our counter-culture caffeinated

conversations called “Muscle Meals,” which cover everything from how to cook sweet potatoes

Page 12: Regulation of Inflammatory Macrophage Activation and

xi

for optimal nutritional value to muscle repair in lobsters, have kept my passion for science alive

through times when my work has not been particularly exciting.

Finally, the biggest thanks goes to my family. To my dad: thank you for being an example of what

it looks like to live a purpose-driven life and for pushing me when I needed it early on in life. To

my mom: thank you for unconditional love and support and for passing down your creativity and

love for music, which played a vital role in teaching me dedication and discipline. To my sister:

thank you for still loving me, despite my deficiencies as a sibling stemming from my work-focused

way of living. To my extended family, especially Uncle Ken and Aunt Susan and my grandparents

on both sides, I am also forever thankful for your love and support and for the many ways that you

have aided in my development to this point. To all: thank you.

Page 13: Regulation of Inflammatory Macrophage Activation and

xii

LIST OF FIGURES Figure 2.1 ACLY activity supports inflammatory gene induction in LPS-activated macrophages..................................................................................................................................40 Figure 2.2 LPS activation increases glucose oxidation in macrophages......................................41 Figure 2.3 Glucose oxidation supports inflammatory gene induction in LPS-activated macrophages by providing carbon substrate for histone acetylation.............................................42 Figure 2.4 LPS activation enhances glycerol phosphate shuttle activity in macrophages............44 Figure 2.5 Generation of GPD2 KO mice for studying the role of glycerol phosphate shuttle activity in macrophage inflammatory responses...........................................................................46 Figure 2.6 The GPS enzyme GPD2 regulates glucose oxidation in LPS-activated macrophages..................................................................................................................................47 Figure 2.7 GPD2 activity controls inflammatory gene induction in LPS-activated macrophages by regulating Ac-CoA production and histone acetylation................................................................48 Figure 2.8 Naïve and LPS tolerant macrophages exhibit profound differences in responsiveness to LPS stimulation.........................................................................................................................50 Figure 2.9 Glucose utilization impairs respiration in tolerant macrophages................................50 Figure 2.10 Glucose utilization limits inflammatory gene induction in tolerant macrophages....51 Figure 2.11 Glucose utilization impairs respiration and limits inflammatory gene induction in tolerant macrophages without affecting TLR signaling................................................................52 Figure 2.12 GPD2 activity drives suppression of inflammatory responses in tolerant macrophages..................................................................................................................................54 Figure 2.13 Schematic depicting FET and RET during LPS activation and tolerance................56 Figure 2.14 GPD2 activity drives induction of reverse electron transport (RET) in tolerant macrophages..................................................................................................................................57 Figure 2.15 GPD2 activity supports LPS tolerance in vivo..........................................................60 Supplementary Figure 2.1 ATP-citrate lyase (ACLY) and p300 activity are necessary for supporting pro-inflammatory gene induction during LPS activation in BMDMs.........................74 Supplementary Figure 2.2 LPS exposure modulates mitochondrial respiration and glucose uptake as a function of time...........................................................................................................75 Supplementary Figure 2.3 Glucose starvation attenuates pro-inflammatory gene induction and promoter region histone acetylation in LPS-activated BMDMs....................................................75 Supplementary Figure 2.4 Depiction of the spatial and biochemical position of GPD2 at the nexus between glycolysis and electron transport...........................................................................76 Supplementary Figure 2.5 LPS activation increases expression and activity of GPD2 in BMDMs.........................................................................................................................................77 Supplementary Figure 2.6 GPD2 regulates glucose flux through glycolysis and the TCA cycle. .............................................................................................................................................78 Supplementary Figure 2.7 Model integrating GPD2-regulated mitochondrial oxidative metabolism with LPS activation....................................................................................................79 Supplementary Figure 2.8 Prolonged LPS stimulation disrupts forward electron transport, leading to a shift in mitochondrial redox balance..........................................................................79 Supplementary Figure 2.9 Nicotinamide (NAM) reverses RET-redox imbalance to rescue pro-inflammatory gene induction and promoter region histone acetylation in tolerant BMDMs........80 Supplementary Figure 2.10 GPD2 deficiency does not affect some clinical parameters altered during sepsis...................................................................................................................................80

Page 14: Regulation of Inflammatory Macrophage Activation and

xiii

Supplementary Figure 2.11 Pharmacological inhibition of GPD2 modulates LPS activation and tolerance.........................................................................................................................................81 Supplementary Figure 2.12 Model of the metabolic adaptations underpinning the biphasic functional response to LPS exposure.............................................................................................82 Figure 3.1 Dose-dependent metabolic shifts couple strength of LPS stimulation to amplitude and duration of inflammatory gene induction in macrophages............................................................93 Figure 3.2 Akt signaling integrates LPS dose with glucose oxidation to regulate LPS activation in BMDMs.........................................................................................................................................94 Figure 3.3 Akt signaling leads to impaired glucose oxidation and inflammatory gene suppression during prolonged LPS exposure.....................................................................................................95 Figure 3.4 Inhibition of Akt protects against induction of LPS tolerance in macrophages..........96 Figure 3.5 The LPS-inducible transcriptome of macrophages falls into three classes based on expression after repeated LPS stimulation.....................................................................................97 Figure 3.6 LPS-inducible Akt signaling regulates induction of P genes in tolerant BMDMs......99 Figure 3.7 ACSS2 is expressed in tolerant macrophages and regulates induction of P genes...100 Figure 3.8 Histone deacetylation supports induction of P genes in tolerant BMDMs................101 Figure 3.9 Sirt1 does not regulate the induction of P genes in tolerant macrophages................102 Supplementary Figure 3.1 Attenuating glucose oxidation impairs suppression of inflammatory gene expression and cytokine production in macrophages and mice..........................................113 Supplementary Figure 3.2 LPS-induced expression of Gpd2 is regulated by Akt...................113 Supplementary Figure 3.3 P genes belong to an array of gene ontologies...............................114 Supplementary Figure 3.4 Inhibiting the glucose-ACLY-Ac-CoA axis in tolerant macrophages does not attenuate induction of P genes.......................................................................................114 Supplementary Figure 3.5 Integration of Akt and ACSS2 into previous model of metabolic regulation of gene induction in LPS stimulated macrophages.....................................................115

Page 15: Regulation of Inflammatory Macrophage Activation and

CHAPTER 1

Introduction

Page 16: Regulation of Inflammatory Macrophage Activation and

2

Sepsis: a tale of two syndromes Systemic microbial infection in vertebrate animals elicits an acute systemic inflammatory response

syndrome1, characterized by activation of the innate and adaptive arms of the immune cells and

elevated local and circulating pro-inflammatory cytokines (e.g. IL-6, IL-1β, and TNFα) and

enhanced antimicrobial activity, including phagocytosis and production of reactive oxygen species

(ROS), nitric oxide (NO), and hypochlorous acid2. The clinical presentation of SIRS in patients

with a documented pathogen is referred to as sepsis, or septic shock, and is diagnosed by elevated

temperature, heart rate, respiratory rate, and/or blood leukocyte count2. Hospital mortality statistics

for sepsis give pause, as sepsis-related illness contributes to nearly half of all hospital deaths

despite presentation in only 10% of patients3. Although SIRS may progress to multiorgan

dysfunction syndrome resulting from tissue damage in the wake of the primary pro-inflammatory

response, standard care practices and advances in medical technology in intensive care units have

minimized acute risk of mortality from sepsis2,4. However, patients who survive SIRS are at an

elevated risk of succumbing to secondary nosocomial infections or reactivated latent infections5,6,

as the initial robust pro-inflammatory response is followed by prolonged immunoparalysis,

characterized by general immunosuppression due to decreased pro-inflammatory cytokine

production, impaired antigen presentation and costimulation, and marked T cell anergy7–10.

Immunosuppression in this context, referred to as compensatory anti-inflammatory response

syndrome (CARS)11, is believed to be a compensatory mechanism to limit damage to the host;

however, protracted immunosuppression contributes a high percentage of sepsis-related hospital

deaths12.

Page 17: Regulation of Inflammatory Macrophage Activation and

3

There have been many attempts at suppressing inflammation during sepsis, resulting in numerous

phase-two clinical trials, including anti-TNFα antibodies (Abs)13–15, IL-1 receptor antagonists16,17,

and endotoxin-neutralizing Abs18. Such attempts have so far been unsusccessful, turning attention

toward preventing immunoparalysis and restoring immunocompetence during CARS12,19,20.

Providing clarity to known mechanisms of immunosuppression, as well as uncovering novel

modules of regulation, may inform the design of targeted therapies to improve patient outcome.

Innate response to microbial encounter: activation

Macrophages, phagocytic cells of the innate immune system, are a critical component of the

biphasic host response during sepsis, mounting an initial pro-inflammatory response to microbial

encounter followed by a persistent state of tolerance to subsequent challenges. First described in

the late 19th century by Metchnikoff, macrophages discriminate between invading microbes and

the host’s own tissues through detection of various structures common to pathogens, including

lipopolysaccharide (LPS), peptidoglycans, and mannose-rich oligosaccharides21. Collectively,

these structures are referred to as pathogen-associated molecular patterns (PAMPs), and the

receptors expressed on macrophages that identify PAMPs are called pattern recognition receptors

(PRRs). There are many types of PRRs in different cell compartments developed to generate

specific responses to a wide array of PAMPs from viruses, bacteria, fungi, and parasites21–23.

Toll-like receptors (TLRs), originally identified in Drosophila24, are the best-characterized of the

many PRRs. The TLR family consists of 10 members in humans and 12 members in mice25. This

repertoire allows cells of the innate immune system to sense a wide array of PAMPs to drive

inflammatory responses and enhance adaptive immunity. Signaling through TLRs is initiated by

Page 18: Regulation of Inflammatory Macrophage Activation and

4

homo- or heterodimer formation, followed by transduction through a cytoplasmic Toll/IL-1

receptor (TIR) domain26 to adaptor proteins capable of sorting signals from multiple TLRs into a

broad range of transcriptional responses22,27. All TLRs except for TLR3, which senses specific

viral components, signal through the adaptor protein MyD88 (myeloid differentiation primary

response 88), which interacts through its TIR domain with the TIR domain of TLRs. This MyD88-

depenent pathway signals through activation of downstream IL-1 receptor-associated kinases

(IRAK), of which there are 4 members (1, 2, M, and 4)28–31. IRAK4 is first recruited to MyD88,

followed by activation and phosphorylation of IRAK1. Both IRAKs interact with MyD88 through

their death domains and then associate with TRAF6 (tumor-necrosis factor-receptor-associated

factor 6), leading to activation of the transcription factors NF-κB and AP-1 through activation of

IKKs α/β/γ (inhibitor of NF-κB (IκB) kinases) and JNK, respectively23. TLRs also signal through

a MyD88-independent pathway, which ultimately activates IRF3 (IFN-regulatory factor 3) to

induce IFN-β and, consequently, IFN-inducible genes. IRF3 activation occurs as a result of

phosphorylation by the non-canonical IKKs TANK-binding kinase 1 (TBK1) and IKKε/IKKi32,33.

As mentioned above, transcriptional responses induced downstream of TLR signaling are further

calibrated to specific PAMPs through the incorporation of additional adaptor molecules. In the

MyD88-dependent pathway, MyD88 functions with TIRAP (TIR domain-containing adaptor

protein)34–37 to constitute an adapter set, while the proteins TRIF (TIR domain-containing adaptor

inducing IFN-β)38 and TRAM (TRIF-related adapter molecule)39–41 form a second adapter set.

Importantly, TLR4, which senses lipopolysaccharide (LPS) from gram-negative bacteria, has the

unique ability to signal through both of these adapter sets, first through TIRAP/MyD88 at the

plasma membrane and then through TRIF/TRAM from endosomes following CD14-dependent

Page 19: Regulation of Inflammatory Macrophage Activation and

5

endocytosis of TLR442,43. Notably, CD14 is also involved in signaling from the cell surface,

transferring LPS from LPS-binding protein (LBP) to TLR4-bound MD-2, which promotes

dimerization and activation of TLR444–47. This spatiotemporal regulation of signaling from TLR4

results in early and late phases of inflammatory gene induction through NF-κB (late phase being

TRIF/TRAM-dependent) and induction of antiviral immunity through IRF3.

The enhanced production of pro-inflammatory cytokines, antimicrobial species, and expression of

costimulatory molecules by innate immune cells as a result of TLR-dependent signaling is critical

for mounting an appropriate and effective defense against microbial invasion. Indeed, disruption

of pathogen sensing and signaling by mutations in TLRs or deficiencies in adaptor molecules and

associated kinases results in incomplete protection48. However, TLR signaling is also involved in

the pathogenesis of autoimmune, chronic inflammatory, and infectious diseases49. As described at

the beginning of this chapter, systemic bacterial infection induces a robust pro-inflammatory

response, often referred to as a “cytokine storm,” which is necessary to control and eliminate the

microbe by enhancing innate and adaptive immunity at sites of infection. However, if the

magnitude and duration of the inflammatory response are left unchecked, then serious damage to

the host’s own tissues will ensue, resulting in fatal multi-organ dysfunction4,12. Therefore, many

mechanisms of attenuating the pro-inflammatory response have evolved to balance the double-

edged sword of TLR activation.

Page 20: Regulation of Inflammatory Macrophage Activation and

6

Innate response to microbial encounter: suppression and tolerance

The myriad mechanisms of negative regulation of early TLR-driven inflammatory responses can

be conveniently grouped into three classifications based on spatial distribution: extracellular,

transmembrane, and intracellular50. Extracellular mechanisms of negative regulation of TLR

signaling involve competition for binding PAMPs and co-receptor complexes by soluble decoy

TLRs (sTLRs). sTLR2 and sTLR4 have been identified as a first layer of negative regulation of

inflammatory responses induced by TLR2 and TLR4 agonists51,52. sTLR2 exists in six isoforms

that each result from post-translational modification of transmembrane TLR2. In contrast, sTLR4

is encoded by several mRNAs. Expression of the orphan receptor SIGIRR (single immunoglobulin

IL-1-related receptor) constitutes a layer of negative regulation at the transmembrane level, as the

level of surface residency of SIGIRR is negatively correlated with NF-kB activation53,54.

Supportive of a suppressive role in inflammatory responses, SIGIRR-deficient mice are more

susceptible to septic shock, and DCs from the bone marrow of these mice are more sensitive to

TLR4 and TLR9 stimulation. Finally, intracellular mechanisms represent the largest class of

negative regulators, including truncated MyD88, catalytically-inactive IRAKs, SOCS1, TOLLIP,

and A20. The first in this list, truncated MyD88, or short form MyD88 (MyD88s), is a splice

variant of MyD88 that lacks the normal interdomain. When MyD88s is expressed at high levels

(as is true in the spleen), it forms a heterodimer with MyD88 and prevents the association with

IRAK4 that is required for IRAK1 phosphorylation and ultimately for activation of NF-kB55,56.

Second on the list, are IRAK2 and IRAKM. Expression of splice variants of IRAK2 (Irak2a, b, c,

and d) and IRAKM attenuates LPS-induced inflammatory gene induction31,57. Perhaps most

striking of the intracellular negative regulators is SOCS1, as deficiency is lethal within 3 weeks of

life, and SOCS1-deficient mice are also profoundly susceptible to sepsis-induced multi-organ

Page 21: Regulation of Inflammatory Macrophage Activation and

7

dysfunction and mortality58,59. The remaining two intracellular negative regulators, TOLLIP (Toll-

interacting protein) and A20 act at IRAK1 and TRAF6, respectively. TOLLIP is phosphorylated

by IRAK1 upon LPS stimulation but impinges on IRAK1 phosphorylation in a negative feedback

loop, while TLR-induced expression of A20 leads to enhanced deubiquitination of TRAF6,

attenuating MyD88-dependent and -indepedenent signaling60–63.

While modulation of the negative regulators of TLR signaling discussed above leads to important

changes in inflammation, both in normal animals (i.e. SOCS1) and in animals subjected to septic

shock induced by TLR agonists (i.e. A20), these mechanisms do not provide protection against

subsequent exposures to PAMPs50. However, in addition to fine tuning the magnitude and duration

of inflammation, macrophages have also evolved mechanisms that drive entry into a state of

profound refractoriness to subsequent microbial encounters following prolonged engagement of

TLRs. This phenomenon, termed endotoxin tolerance (not to be confused with central and

peripheral tolerance, which pertain to immune regulation in the adaptive arm of the immune system)

is a protective mechanism, evolved to protect the host’s tissues from sustained inflammation64.

One of the first observations of this form of tolerance was reported by Paul Beeson in 1946 after

he found that administering serial injections of typhoid vaccine to rabbits resulted in progressive

protection against the vaccine’s pyrogenicity65. Consistent with this observation, it is now well

known that injection of mice with a sublethal dose of LPS protects against a subsequent lethal dose

of LPS. Such protection can be largely attributed to cell autonomous endotoxin tolerance in

macrophage66. Indeed, suppression of inflammatory activity and decreased responsiveness to

repeated encounters with the same PAMP, or in some cases a different PAMP, is one of the primary

adaptations underpinning the profound immunoparalysis that ensues following septic shock11.

Page 22: Regulation of Inflammatory Macrophage Activation and

8

Similar to mechanisms of negative regulation of early TLR-driven inflammation, tolerance

mechanisms can also be classified as transmembrane and intracellular. Transmembrane tolerance

mechanisms include expression of the surface receptors ST2 and TRAILR (TNF-related apoptosis-

inducing ligand receptor), and the expression of TLRs themselves. The transmembrane form of

ST2, ST2L, is known to suppress TLR signaling by sequestering MyD88 and TIRAP through their

TIR domains; however, translocation of ST2L occurs after the early events of TLR activation, such

that ST2-deficiency does not affect the response to LPS-induced septic shock but does prevent

development of LPS tolerance in vivo67. Similarly, TRAILR-deficient cells do not have attenuated

TLR signaling acutely during activation by agonists but do exhibit differences at late timepoints68.

Expression of some TLRs is also downregulated in response to specific PAMPs, contributing to

tolerance. Ubiquitylation of TLR4 and TLR9 by the RING-finger E3 ubiquitin ligase TRIAD3

promotes downregulation of these TLRs by enhanced degradation following LPS and CpG DNA

exposure69. Anti-inflammatory cytokines such as TGF-β and IL-10 can also impinge on TLR4

expression70–72.

Intracellular mechanisms contributing to acquisition of a tolerant phenotype can be attributed to

decreased signaling as a result of changes in surface expression of the receptors described above,

as well as to impaired formation and activity of the myddosome (MyD88, TIRAP, and IRAKs)

and decreased NF-kB and JNK activation due to negative feedback loops73,74. Strikingly, changes

in responsiveness of genes encoding inflammatory cytokines and antimicrobial effector functions

in tolerant cells is also underpinned by gene-specific chromatin modifications induced by TLR

signaling64,66. Specifically, genes encoding pro-inflammatory cytokines and chemokines lose

transcription-activating histone lysine acetylation during prolonged exposure to the TLR4 ligand

Page 23: Regulation of Inflammatory Macrophage Activation and

9

LPS, while genes responsible for antimicrobial effector functions are not silenced by decreased

histone acetylation and are in some instances hyperacetylated to promote enhanced expression in

re-stimulated tolerant cells66. Since this finding, additional work has strengthened the epigenetic

basis of LPS tolerance in macrophages by characterizing the regulation and function of histone

acetylation and methylation by chromatin modifying enzymes in the context of inflammation75.

However, a complete understanding of the mechanisms underpinning these chromatin

modifications and the overall phenotype of endotoxin tolerance remains elusive. Thus, providing

clarity to known mechanisms of tolerance, as well as uncovering novel modules of regulation, may

inform the design of targeted therapies to relieve immunosuppression in sepsis and improve patient

outcome.

Immunometabolism: a new perspective on the regulation of immunity

Just as macrophage responsiveness changes over the course of extended stimulation, the field of

immunology itself has changed over the span of time humans have been studying host defense.

Indeed, from Edward Jenner’s discovery of vaccination at the end of the 18th century to the “Toll

Rush” at the turn of the 21st century to the transcriptional profiling of tissue-resident leukocyte

subsets at the present, immunology has evolved. Each step forward in the pursuit of knowledge on

the subject of host defense leads to an increasing number of questions to fuel further exploration.

One such step forward in the last decade was towards understanding the role of metabolism in

immune cells. Continued exploration in this area has led to the genesis of a new subfield in

immunology, referred to as “immunometabolism.” This relatively new discipline is focused on

investigating how shifts in intermediary metabolism support the four tasks of the immune system:

recognition, effector functions, regulation, and memory. Although it is still common to find

Page 24: Regulation of Inflammatory Macrophage Activation and

10

colleagues, at all levels of experience, who believe that metabolism is a passive network of

chemical reactions that simply supplies energy and does not have any direct bearing on specific

cellular functions, this belief has been thoroughly debunked by the flood of evidence

demonstrating that innate and adaptive immunity depends on dynamic changes in metabolism76–

82. Furthermore, an understanding and appreciation of this new underpinning to immunity does not

require translation and memorization of the impenetrable charts of metabolic pathways that are

commonly presented in textbooks and seminars – although such knowledge is necessary to

participate in productive research in immunometabolism – but rather depends only on a modest

understanding of the fundamental nature of biochemistry and the overall functions of some major

metabolic pathways. Before summarizing metabolic regulation of macrophage activation and

function, an overview of these topics is provided.

The opening lines of Josiah Willard Gibbs’ masterwork On the Equilibrium of Heterogeneous

Substances is a quote from Rudolf Clausius, summarizing the first and second laws of

thermodynamics:

“Die Energie der Welt ist konstant

Die Entropie der Welt strebt einem Maximum zu.”

- Clausius, 1865

These laws, stating that the energy in the universe is constant and that entropy in the universe tends

to a maximum, were the basis for the Gibbs-Helmholtz equation relating the free-energy change

in a system (ΔG) to quantities of heat change (ΔH) and entropy change (ΔS) in the system under

constant temperature and pressure (Eq. 1.1).

Page 25: Regulation of Inflammatory Macrophage Activation and

11

∆" = ∆$ − &∆' (Equation 1.1)

This equation, in addition to Walther Nernst’s equation describing the electrochemical behavior

of chemical substances participating in oxidation-reduction reactions, paved the way for the

systematic discovery and mapping of biochemical reactions in the 20th century. As stated

previously, knowledge of individual biochemical reactions is not necessary here, rather these

equations provide a conceptual framework within which to think about metabolism. Specifically,

one must understand that because the existence of animate objects (organisms) depends on the sum

of these biochemical reactions to resist crumbling into complete randomness (S at a maximum),

organisms are bound by the same thermodynamic laws that govern the reactions of inanimate

chemicals.

The behavior of heterogeneous substances such as those in a biochemical network depends on the

adherence to and utilization of chemical and electrochemical gradients. For example, energy-

producing (exergonic; ΔG<0) reactions involved in the systematic degradation of nutrients, such

as the oxidation of carbohydrates, are coupled to energy-consuming (endergonic; ΔG>0) reactions

used to build and replenish highly-ordered structures of the body, such as cholesterol and proteins83.

Based on the Gibbs equation, exergonic reactions increase the entropy of a system, while

endergonic reactions have the opposite effect. Coupling these two types of reactions keeps the net

change in entropy in the system close to zero. The change in free energy in a system is such a

powerful way of analyzing the behavior of heterogeneous substances that it can be used as an

organizing principle for biochemical networks. Indeed, it is possible to reconstruct experimentally-

validated metabolic pathways simply from knowledge of metabolites and enzymes in the system

and the Gibbs free energy of the reactions in which they can participate84,85. In lower organisms

Page 26: Regulation of Inflammatory Macrophage Activation and

12

and in some mammalian cells, thermodynamic analysis has also been used to predict metabolic

fluxes, optimized to specific cellular functions86–89. Thus, when thinking about central metabolic

pathways, the heat and entropy changes occurring over a series of chemical transformations is

often enough logic to understand the direction of flux.

In addition to gradients in chemical energy, electrochemical gradients also provide a useful lens

through which to view metabolism90. Indeed, electrochemical gradients generated from the

separation of charged species across biological surfaces or from the specific spatial orientation or

stoichiometry of redox-active substances are used as a motive force to drive myriad processes vital

for life. For example, the mitochondrial respiratory chain, also called the electron transport chain

(ETC), is responsible for making the majority of ATP in many types of cells, and its function

depends on multiple electrochemical gradients91–93. The ETC is comprised of protein complexes

situated in the tightly-knit inner mitochondrial membrane and is organized such that electrons

stripped from nutrients during oxidative metabolism enter at complexes of the lowest reduction

potential (most negative) and flow from complex-to-complex until ultimately reducing oxygen,

which has the highest reduction potential (most positive). The potential difference from “negative”

to “positive” ends of the ETC is about 1.5 V94, which is equivalent to the nominal charge of a

standard alkaline battery! This electrical circuit within the inner mitochondrial membrane is

coupled to displacement of protons across the membrane by pumping action of some ETC

complexes, establishing a parallel chemical circuit with an electrochemical potential equivalent to

250,000 V/cm. This gradient is sufficient to drive production of ATP through the activity of the

F0/F1 ATPase at the end of the respiratory chain92, fueling innumerable energy-consuming

biochemical reactions necessary for life.

Page 27: Regulation of Inflammatory Macrophage Activation and

13

Beyond these thermodynamic constraints, metabolic networks are also subject to other layers of

regulation, including concentrations of metabolites, concentrations of enzymes, and modification

status and thus intrinsic activities of enzymes. Although metabolite concentrations can be factored

into thermodynamic equations, in most cases application of Le Chatelier’s Principle or the Law of

Mass Action is sufficient to address concerns with how levels of reactants and products affect

flux94. These additional layers of regulation are also important for calculating the velocity and

directionality (if bidirectional) of reactions through application of Michaelis-Menton kinetic

equations; however, this will not be discussed here. Instead, it will be noted that a final

fundamental principle of metabolism that is useful for general understanding is that, because the

study of living organisms typically involves application of controlled stress(es) in an effort to

induce and dissect a specific response, the metabolism of the organism is rarely at true steady-state.

As a result, metabolite concentrations are displaced, inducing measurable changes in fluxes

through the individual reactions of the organism’s metabolic network. Thus, focusing on

metabolites with large displacement from control conditions and study of the reactions in which

they participate is a powerful way to gain insight into the specific metabolic requirements for an

organism’s response to a particular stress.

Finally, there are numerous mechanisms through which metabolic adaptations underpin specific

functions of immune cells. These include meeting bioenergetics demands (supply of ATP for

energy-consuming processes), supplying carbon substrates for growth and proliferation or for

expanding ER to enhance secretory capacity, producing reactive species like ROS or metabolites

that affect signaling, and increasing substrate flux through pathways producing metabolite-derived

effector molecules77–80. Although significant progress has been made towards uncovering and

Page 28: Regulation of Inflammatory Macrophage Activation and

14

understanding metabolic instruction of many aspects of immunity, the following section is devoted

to summarizing the body of knowledge regarding how metabolic shifts support LPS-induced

macrophage activation (classical, or M1 polarization) and effector functions.

Metabolic regulation of inflammatory macrophage polarization

In addition to the canonical signal transduction pathways discussed in the previous section of this

Introduction, pathogen recognition by TLRs is also coupled to activation of metabolic signaling

pathways, enabling coordinate induction of effector functions and the metabolic processes needed

to sustain them. Downstream of LPS-TLR4, Akt is activated by TBK/IKKε and PI3K95. Akt

kinases regulate cell survival, metabolism, and cytoskeleton, in part through activation of another

metabolic signaling pathway, the mammalian target of rapamycin (mTOR)96,97. The mTOR kinase

is found in two complexes in mammals, mTORC1 and mTORC2. Other subunits of these

complexes are unique and define the complexes, such as Raptor in mTORC1 and Rictor in

mTORC2. mTORC1 links availability of nutrients (especially amino acids) and growth factor

signaling to anabolic processes such as macromolecule synthesis and nutrient storage in

proliferating cells and tumor cells. In LPS-activated DCs, mTORC1 activity inhibits oxidative

metabolism, apparently by stimulating production of NO98, which is a long-appreciated toxin to

the ETC99. Importantly, although Akt and mTORC1 are both activated downstream of TLR

signaling, Akt activity can be dissociated from mTORC1 activity in this setting95. Relevant

metabolic targets of Akt and mTORC1 in control of macrophage activation remain incompletely

characterized and represent an important avenue of future investigations. The Akt target HIF-1α

has been demonstrated to increase in macrophages following LPS stimulation100, supporting an

upregulation of glycolysis, the process by which glucose is incompletely oxidized in the cytosol,

Page 29: Regulation of Inflammatory Macrophage Activation and

15

yielding lactate as its final product. HIF-1α is normally an unstable protein but becomes stabilized

by LPS-triggered alterations to ETC activity that increases mitochondrial ROS (mtROS)

production101. Moreover, LPS signaling increases HIF-1α activity by upregulating the PKM2

isoform of the metabolic enzyme pyruvate kinase102. In addition to its role in glycolysis, PKM2 is

a HIF-1α coactivator, which binds to HIF-1α to enhance its transcriptional activity. Finally, LPS

acutely inhibits activation of AMP protein kinase (AMPK), a key regulator of cellular energy

homeostasis that stimulates ATP-generating metabolic processes (i.e. oxidation of fatty acids) in

response to increasing ADP/ATP ratio, but prolonged LPS exposure leads to enhanced

phosphorylation and activation of AMPK due to autocrine/paracrine IL-10 signaling through the

IL-10 receptor103,104. Thus, LPS signaling co-opts metabolic machinery to coordinate metabolic

shifts, which are described in more detail below.

Early studies demonstrating increased expression of the glucose transporter GLUT1 on M1

macrophages suggested that LPS activation preferentially boosts glucose utilization105–107.

Experimental attenuation of glycolysis reduces inflammatory cytokine production and bacterial

killing100,107,108, indicating that glucose utilization in glycolysis is necessary for optimal induction

of a pro-inflammatory phenotype. This metabolic switch is driven by LPS-induced activation of

Akt to enhance glucose uptake95, concomitant with HIF-1α-dependent upregulation of multiple

genes encoding glycolytic enzymes. For example, upregulation of PFKFB3, a tissue-specific

isoform of PFKFB, drives overall flux through glycolysis, while induction of PDK1, a negative

regulator of pyruvate oxidation in the mitochondria, and LDHA, the enzyme that converts pyruvate

into lactate, increases aerobic glycolysis109,110. This upregulation of glycolysis at the expense of

mitochondrial respiration is referred to as the “Warburg effect” after the German scientist Otto

Page 30: Regulation of Inflammatory Macrophage Activation and

16

Warburg, who discovered that cancer cells prefer to generate ATP from glycolysis rather than

respiration even in the presence of sufficient oxygen111,112. Glycolytic production of ATP may be

important to fuel optimal induction of effector activities and maintain viability in macrophages,

given that oxidative metabolism is actually impaired. Glucose oxidation in the mitochondria at

very early timepoints before impairment of respiration appears to be important for Ac-CoA and

phospholipid synthesis and inflammatory cytokine production in LPS-stimulated DCs and

macrophages95. How glucose oxidation and glycolysis are coordinated is incompletely understood

but seems to be a pivotal node in control of classical macrophage polarization, as various

manipulations that alter this balance perturb the activation phenotype108,113.

In addition to the above changes in glycolysis, TLR signaling also results in reprogramming of the

tricarboxylic acid cycle (TCA cycle, or Krebs cycle), the series of oxidative processes in the

mitochondrial matrix formulated in the theory of the “citric acid cycle” proposed in 1937 by Hans

Krebs and William Johnson. This central pathway in carbon substrate metabolism is necessary for

oxidation of glucose-derived pyruvate, fatty acids, and glutamine to generate ATP in the ETC and

provide biosynthetic precursors for numerous cellular processes. A series of important experiments

in which substrates containing heavy isotopes of carbon were fed to macrophages during LPS

stimulation to trace their flux through intermediary metabolism revealed two “breaks” in the TCA

cycle of LPS-stimulated macrophages114. The first break regulates the fate of the TCA cycle

intermediate citrate, which is normally converted to α-ketoglutarate by IDH1. In these

macrophages, downregulation of IDH1, but upregulation of immunoresponsive gene 1 (IRG1), an

enzyme with decarboxylase activity for citrate-derived aconitase, drives the production of

itaconate. By inhibiting the glyoxylate cycle, a variation of the TCA cycle discovered by Krebs’

Page 31: Regulation of Inflammatory Macrophage Activation and

17

student Hans Kornberg115 and found in microorganisms, itaconate has direct microbicidal activity

against several pathogens116. In addition, itaconate is believed to competitively inhibit succinate

dehydrogenase (SDH) in the TCA cycle leading to the second “break”117. Because SDH is a

subunit of complex II (CII) of the ETC, itaconate production perturbs ETC activity and oxidative

metabolism, and has been linked to regulation of mtROS production and inflammatory gene

induction. Of note, itaconate has also been demonstrated to activate the anti-inflammatory and

antioxidant transcription factor NRF2 via alkylation of the redox-sensing protein KEAP1 (kelch-

like ECH-associated protein 1)118 and through electrophilic stress119, both of which may underpin

itaconate’s inhibitory effects on inflammatory cytokine production.

As described above, altered carbon flux through the TCA cycle impacts LPS activation. As an

additional adaptation to the second break in the TCA cycle, LPS-stimulated macrophages utilize

the aspartate-arginosuccinate shunt to re-supply fumarate and also produce arginine (a substrate of

iNOS) through a series of reactions involving transformation of aspartate and citrulline via the

activity of argininosuccinate synthase (Ass1) and argininosuccinate lyase (Asl)114. In the context

of Mycobacterium infection, macrophages initially import extracellular arginine and export

citrulline (a by-product of arginine metabolism by iNOS) but switch to citrulline import for

arginine regeneration through this shunt only after extracellular arginine becomes depleted. Such

import of extracellular arginine may allow the macrophages to limit T cell responses120,121.

Another key node in glucose utilization, the pentose phosphate pathway (PPP), has also been

shown to control classical macrophage activation. The PPP is a glycolytic shunt that produces

NADPH for maintaining cellular redox balance and nucleotide synthesis. LPS increases PPP

Page 32: Regulation of Inflammatory Macrophage Activation and

18

activity in macrophages by downregulating the sedoheptulose kinase carbohydrate kinase-like

protein CARKL, and knockdown or overexpression of CARKL perturbed induction of LPS-

inducible genes122. NADPH production is also likely to be important for fueling the activities of

NADPH oxidases (NOX) and iNOS, which produce ROS and NO, respectively, and are important

components of antimicrobial defense.

More recent studies indicate that glutamine consumption is also enhanced following LPS

stimulation and serves to compensate for the first break in the TCA cycle below citrate by

supplying α-ketoglutarate in a process called anaplerosis114,123, in which carbons are funneled into

the TCA cycle at points other than pyruvate (of note, anaplerosis is another contribution from

Kornberg124).

Mitochondria as hubs in inflammatory activation of macrophages

It has long been appreciated that stimulation of macrophages and dendritic cells with LPS or

LPS+IFN-γ shuts down oxidative metabolism. An underpinning mechanism could be interactions

of NO with iron–sulfur clusters in the ETC, since iNOS inhibition rescues oxidative

metabolism125,126. In addition, M1 macrophages produce ROS from ETC complex I (CI), which

has been linked to enhanced bactericidal activity127. The underlying mechanism involved

mitochondria-localized interactions between TRAF6 and ECSIT, a regulator of CI assembly. More

recently, a flurry of papers has further underscored the connection between oxidative metabolism,

ETC activity, and inflammatory gene induction, thus positioning the mitochondria at the center of

LPS activation79,82,128.

Page 33: Regulation of Inflammatory Macrophage Activation and

19

The IRG1/itaconate axis appears to play a key role in regulating oxidative metabolism and

inflammatory gene induction in M1 macrophages. As discussed above, induction of IRG1 allows

for itaconate production and inhibition of SDH117 Addition of exogenous itaconate inhibits

oxidative metabolism, while IRG1 deficiency enhances oxidative metabolism129. Importantly,

itaconate addition inhibits production of IL-6 and IL-12, while IRG1 deficiency augments their

production129, suggesting that the IRG1/itaconate axis serves as a built-in negative feedback that

impinges on the ETC to limit excessive inflammatory responses. Another recent study also

described ETC alterations during E. coli infection, specifically a transient increase in CII activity

but a decrease in CI activity130. Increased CII activity was triggered by the activities of NADPH

oxidase and the Src kinase Fgr, which phosphorylates the SDHA subunit of CII. Consistent with

the other studies referenced in this section, ETC remodeling was linked to inflammatory gene

induction130.

How exactly these ETC alterations regulate inflammatory gene induction remains unclear. It has

been suggested that CII activity drives reverse electron transport (RET) to potentiate mtROS

production and HIF-1α stabilization101. High CII activity in a setting of impaired respiration (i.e.

due to elevated NO) leads to over-reduction of the mobile electron carrier ubiquinone (Q), which

under normal conditions accepts electrons from CI, II, and glycerol 3-phosphate dehydrogenase

(GPD2) and donates to CIII. This over-reduction changes the thermodynamic configuration of the

ETC such that the reduction potential of the oxidized/reduced Q pair favors RET from reduced Q

(ubiquinol, QH2) to CI131–133. Such RET is associated with mtROS production from the Q binding

site in CI, which may stabilize HIF-1α to allow for transcriptional induction of its target Il1b123,134.

In support of this model, mtROS scavengers and pharmacological inhibitors of CII reduce

Page 34: Regulation of Inflammatory Macrophage Activation and

20

inflammatory gene induction101,129. Ectopic expression of the mitochondrial alternative oxidase

AOX, which should alleviate RET by accepting electrons from over-reduced Q/QH2, also

recapitulates the phenotype101. However, ETC activities similarly potentiate induction of

inflammatory genes not thought to be regulated by HIF-1α (e.g., Il6), so clarification of additional

underpinning mechanisms remains an outstanding question.

Integration of metabolic signals by chromatin

A final mode of metabolic regulation of inflammatory macrophage polarization is through gene-

specific chromatin modifications. Recall that LPS induces gene-specific changes in promoter

region histone acetylation to regulate activation and tolerization66,75. More recent studies have

indicated that metabolism impinges on chromatin for transcriptional control of macrophage

activation. Cheng et al. showed that priming with the Candida albicans product β-glucan enhances

the ability of a subsequent LPS challenge to induce genes encoding inflammatory cytokines. This

process is dependent on Akt and mTORC1 as well as glycolytic activity and correlates with

chromatin changes at the promoters of these genes135. There is also evidence macrophages

polarized to an anti-inflammatory (M2) phenotype by the Th2 cytokine IL-4 also rewire glucose

metabolism in an Akt-mTORC1-dependent manner to supply acetyl-CoA (Ac-CoA) for histone

acetylation and activation of a subset of IL-4-inducbile genes136. Thus, consistent with

observations in other cell types, shifts in macrophage metabolism can regulate transcriptional

responses through modulation of substrates used for chromatin modifications137,138. Elucidating

new regulatory modules and further characterizing known pathways in macrophage metabolism

that are affected downstream of pathogen recognition will illuminate novel targets for therapeutic

intervention to manipulate host responses to microbial infection.

Page 35: Regulation of Inflammatory Macrophage Activation and

21

Thesis summary

This thesis reports on several novel insights into the metabolic underpinnings of macrophage

inflammatory responses, specifically at the levels of glucose oxidation and mitochondrial

respiration. In Chapter 2, we describe histone acetylation as a fundamental mechanism by which

glucose metabolism fuels inflammatory gene induction and underpins inflammatory gene

suppression during LPS activation and tolerance, respectively. The Ac-CoA-producing enzyme

ATP-citrate lyase (ACLY) and the histone acetyltransferase p300 are implicated in this integration

of enhanced glucose metabolism with increased histone acetylation. Furthermore, we provide an

important update and revision to the current paradigm of how mitochondrial respiration supports

inflammatory gene induction, demonstrating that 1) ETC activity critically underpins induction

and suppression of inflammatory genes via regulation of carbon flux through the TCA cycle to

supply substrate to the ACLY-p300 axis for histone acetylation, and 2) that the glycerol phosphate

shuttle and its key regulatory enzyme GPD2, a component of the ETC seldom included in textbook

schemes of the respiratory chain, is a novel LPS-inducible module in macrophage metabolism that

drives enhanced glucose oxidation and thus stands at the epicenter of the metabolic regulation of

macrophage inflammatory responses in vitro and in vivo. Thus we propose a new paradigm for

how glucose utilization underpins M1 macrophage activation and function, defining a model in

which GPD2 and the glycerol phosphate shuttle integrate the extent of microbial stimulation with

glucose oxidation to balance the beneficial and detrimental effects of the inflammatory response.

In Chapter 3, additional experiments that extend and further support this model are reported,

including 1) implication of Akt-dependent signaling in regulating LPS-inducible changes in

respiration and inflammatory gene induction such that pharmacological inhibition of Akt

recapitulates the effects of pharmacological and genetic inhibition of glucose oxidation on

Page 36: Regulation of Inflammatory Macrophage Activation and

22

macrophage LPS activation and tolerance, and 2) discovery of a subset of LPS-inducible genes

that display enhanced Akt-dependent expression in tolerant, restimulated macrophages. Induction

of these primed (P) genes is supported by the activity of Ac-CoA synthetase short-chain family

member 2 (ACSS2), which produces Ac-CoA from acetate, serving as an alternative pathway

fueling histone acetylation in the glucose-oxidation-deficient tolerant state. Finally, Chapter 4

includes further discussion of results and future directions.

Page 37: Regulation of Inflammatory Macrophage Activation and

23

REFERENCES

1. Bone,R.C.etal.Definitionsforsepsisandorganfailureandguidelinesfortheuseofinnovativetherapiesinsepsis.TheACCP/SCCMConsensusConferenceCommittee.AmericanCollegeofChestPhysicians/SocietyofCriticalCareMedicine.Chest101,1644–1655(1992).

2. Iskander,K.N.etal.Sepsis:MultipleAbnormalities,HeterogeneousResponses,and

EvolvingUnderstanding.PhysiologicalReviews93,1247–1288(2013).3. Liu,V.etal.Hospitaldeathsinpatientswithsepsisfrom2independentcohorts.JAMA

312,90–92(2014).4. Hotchkiss,R.S.&Karl,I.E.ThePathophysiologyandTreatmentofSepsis.NewEngland

JournalofMedicine348,138–150(2003).5. Luyt,C.-E.etal.Herpessimplexviruslunginfectioninpatientsundergoingprolonged

mechanicalventilation.Am.J.Respir.Crit.CareMed.175,935–942(2007).6. Limaye,A.P.etal.Cytomegalovirusreactivationincriticallyillimmunocompetent

patients.JAMA300,413–422(2008).7. Ertel,W.etal.Downregulationofproinflammatorycytokinereleaseinwholeblood

fromsepticpatients.Blood85,1341–1347(1995).8. Wolk,K.,Döcke,W.D.,vonBaehr,V.,Volk,H.D.&Sabat,R.Impairedantigen

presentationbyhumanmonocytesduringendotoxintolerance.Blood96,218–223(2000).

9. Monneret,G.etal.Theanti-inflammatoryresponsedominatesaftersepticshock:

associationoflowmonocyteHLA-DRexpressionandhighinterleukin-10concentration.Immunol.Lett.95,193–198(2004).

10. Hotchkiss,R.S.etal.Apoptoticcelldeathinpatientswithsepsis,shock,andmultiple

organdysfunction.Crit.CareMed.27,1230–1251(1999).11. Adib-Conquy,M.&Cavaillon,J.-M.Compensatoryanti-inflammatoryresponse

syndrome.Thromb.Haemost.101,36–47(2009).12. Boomer,J.S.etal.Immunosuppressioninpatientswhodieofsepsisandmultipleorgan

failure.JAMA306,2594–2605(2011).

Page 38: Regulation of Inflammatory Macrophage Activation and

24

13. Abraham,E.etal.Efficacyandsafetyofmonoclonalantibodytohumantumornecrosisfactoralphainpatientswithsepsissyndrome.Arandomized,controlled,double-blind,multicenterclinicaltrial.TNF-alphaMAbSepsisStudyGroup.JAMA273,934–941(1995).

14. Cohen,J.&Carlet,J.INTERSEPT:aninternational,multicenter,placebo-controlledtrial

ofmonoclonalantibodytohumantumornecrosisfactor-alphainpatientswithsepsis.InternationalSepsisTrialStudyGroup.Crit.CareMed.24,1431–1440(1996).

15. Fisher,C.J.etal.Treatmentofsepticshockwiththetumornecrosisfactorreceptor:Fc

fusionprotein.TheSolubleTNFReceptorSepsisStudyGroup.N.Engl.J.Med.334,1697–1702(1996).

16. Fisher,C.J.etal.Recombinanthumaninterleukin1receptorantagonistinthe

treatmentofpatientswithsepsissyndrome.Resultsfromarandomized,double-blind,placebo-controlledtrial.PhaseIIIrhIL-1raSepsisSyndromeStudyGroup.JAMA271,1836–1843(1994).

17. Opal,S.M.etal.Confirmatoryinterleukin-1receptorantagonisttrialinseveresepsis:a

phaseIII,randomized,double-blind,placebo-controlled,multicentertrial.TheInterleukin-1ReceptorAntagonistSepsisInvestigatorGroup.Crit.CareMed.25,1115–1124(1997).

18. Angus,D.C.etal.E5murinemonoclonalantiendotoxinantibodyingram-negative

sepsis:arandomizedcontrolledtrial.E5StudyInvestigators.JAMA283,1723–1730(2000).

19. Hotchkiss,R.S.,Monneret,G.&Payen,D.Immunosuppressioninsepsis:anovel

understandingofthedisorderandanewtherapeuticapproach.LancetInfectDis13,260–268(2013).

20. Payen,D.,Monneret,G.&Hotchkiss,R.Immunotherapy-apotentialnewwayforward

inthetreatmentofsepsis.CriticalCare17,118(2013).21. Janeway,C.A.&Medzhitov,R.InnateImmuneRecognition.AnnualReviewof

Immunology20,197–216(2002).22. Brubaker,S.W.,Bonham,K.S.,Zanoni,I.&Kagan,J.C.InnateImmunePattern

Recognition:ACellBiologicalPerspective.AnnualReviewofImmunology33,257–290(2015).

23. Akira,S.&Takeda,K.Toll-likereceptorsignalling.Nat.Rev.Immunol.4,499–511(2004).

Page 39: Regulation of Inflammatory Macrophage Activation and

25

24. Hashimoto,C.,Hudson,K.L.&Anderson,K.V.TheTollgeneofDrosophila,requiredfordorsal-ventralembryonicpolarity,appearstoencodeatransmembraneprotein.Cell52,269–279(1988).

25. Kawai,T.&Akira,S.Theroleofpattern-recognitionreceptorsininnateimmunity:

updateonToll-likereceptors.NatureImmunology11,373–384(2010).26. Poltorak,A.etal.DefectiveLPSsignalinginC3H/HeJandC57BL/10ScCrmice:mutations

inTlr4gene.Science282,2085–2088(1998).27. Mertins,P.etal.AnIntegrativeFrameworkRevealsSignaling-to-TranscriptionEventsin

Toll-likeReceptorSignaling.CellReports19,2853–2866(2017).28. Cao,Z.,Henzel,W.J.&Gao,X.IRAK:AKinaseAssociatedwiththeInterleukin-1

Receptor.Science271,1128–1131(1996).29. Janssens,S.&Beyaert,R.Functionaldiversityandregulationofdifferentinterleukin-1

receptor-associatedkinase(IRAK)familymembers.Mol.Cell11,293–302(2003).30. Li,S.,Strelow,A.,Fontana,E.J.&Wesche,H.IRAK-4:anovelmemberoftheIRAKfamily

withthepropertiesofanIRAK-kinase.Proc.Natl.Acad.Sci.U.S.A.99,5567–5572(2002).31. Kobayashi,K.etal.IRAK-MisanegativeregulatorofToll-likereceptorsignaling.Cell

110,191–202(2002).32. Kawai,T.etal.LipopolysaccharidestimulatestheMyD88-independentpathwayand

resultsinactivationofIFN-regulatoryfactor3andtheexpressionofasubsetoflipopolysaccharide-induciblegenes.J.Immunol.167,5887–5894(2001).

33. Fitzgerald,K.A.etal.IKKepsilonandTBK1areessentialcomponentsoftheIRF3

signalingpathway.Nat.Immunol.4,491–496(2003).34. Horng,T.,Barton,G.M.&Medzhitov,R.TIRAP:anadaptermoleculeintheTollsignaling

pathway.Nat.Immunol.2,835–841(2001).35. Horng,T.,Barton,G.M.,Flavell,R.A.&Medzhitov,R.TheadaptormoleculeTIRAP

providessignallingspecificityforToll-likereceptors.Nature420,329–333(2002).36. Yamamoto,M.etal.EssentialroleforTIRAPinactivationofthesignallingcascade

sharedbyTLR2andTLR4.Nature420,324–329(2002).37. Kagan,J.C.&Medzhitov,R.Phosphoinositide-mediatedadaptorrecruitmentcontrols

Toll-likereceptorsignaling.Cell125,943–955(2006).

Page 40: Regulation of Inflammatory Macrophage Activation and

26

38. Yamamoto,M.etal.RoleofadaptorTRIFintheMyD88-independenttoll-likereceptorsignalingpathway.Science301,640–643(2003).

39. Yamamoto,M.etal.TRAMisspecificallyinvolvedintheToll-likereceptor4-mediated

MyD88-independentsignalingpathway.Nat.Immunol.4,1144–1150(2003).40. Fitzgerald,K.A.etal.LPS-TLR4signalingtoIRF-3/7andNF-kappaBinvolvesthetoll

adaptersTRAMandTRIF.J.Exp.Med.198,1043–1055(2003).41. Oshiumi,H.etal.TIR-containingadaptermolecule(TICAM)-2,abridgingadapter

recruitingtotoll-likereceptor4TICAM-1thatinducesinterferon-beta.J.Biol.Chem.278,49751–49762(2003).

42. Kagan,J.C.etal.TRAMcouplesendocytosisofToll-likereceptor4totheinductionof

interferon-beta.Nat.Immunol.9,361–368(2008).43. Zanoni,I.etal.CD14controlstheLPS-inducedendocytosisofToll-likereceptor4.Cell

147,868–880(2011).44. Schumann,R.R.etal.Structureandfunctionoflipopolysaccharidebindingprotein.

Science249,1429–1431(1990).45. Eckert,J.K.etal.Thecrystalstructureoflipopolysaccharidebindingproteinrevealsthe

locationofafrequentmutationthatimpairsinnateimmunity.Immunity39,647–660(2013).

46. daSilvaCorreia,J.,Soldau,K.,Christen,U.,Tobias,P.S.&Ulevitch,R.J.

Lipopolysaccharideisincloseproximitytoeachoftheproteinsinitsmembranereceptorcomplex.transferfromCD14toTLR4andMD-2.J.Biol.Chem.276,21129–21135(2001).

47. Gioannini,T.L.&Weiss,J.P.RegulationofinteractionsofGram-negativebacterial

endotoxinswithmammaliancells.Immunol.Res.39,249–260(2007).48. Akira,S.&Takeda,K.Functionsoftoll-likereceptors:lessonsfromKOmice.C.R.Biol.

327,581–589(2004).49. Cook,D.N.,Pisetsky,D.S.&Schwartz,D.A.Toll-likereceptorsinthepathogenesisof

humandisease.NatureImmunology5,975–979(2004).50. Liew,F.Y.,Xu,D.,Brint,E.K.&O’Neill,L.A.J.NegativeregulationofToll-likereceptor-

mediatedimmuneresponses.NatureReviewsImmunology5,446–458(2005).

Page 41: Regulation of Inflammatory Macrophage Activation and

27

51. LeBouder,E.etal.SolubleformsofToll-likereceptor(TLR)2capableofmodulatingTLR2signalingarepresentinhumanplasmaandbreastmilk.J.Immunol.171,6680–6689(2003).

52. Iwami,K.I.etal.Cuttingedge:naturallyoccurringsolubleformofmouseToll-like

receptor4inhibitslipopolysaccharidesignaling.J.Immunol.165,6682–6686(2000).53. Thomassen,E.,Renshaw,B.R.&Sims,J.E.IdentificationandcharacterizationofSIGIRR,

amoleculerepresentinganovelsubtypeoftheIL-1Rsuperfamily.Cytokine11,389–399(1999).

54. Wald,D.etal.SIGIRR,anegativeregulatorofToll-likereceptor-interleukin1receptor

signaling.Nat.Immunol.4,920–927(2003).55. Janssens,S.,Burns,K.,Tschopp,J.&Beyaert,R.Regulationofinterleukin-1-and

lipopolysaccharide-inducedNF-kappaBactivationbyalternativesplicingofMyD88.Curr.Biol.12,467–471(2002).

56. Burns,K.etal.Inhibitionofinterleukin1receptor/Toll-likereceptorsignalingthrough

thealternativelyspliced,shortformofMyD88isduetoitsfailuretorecruitIRAK-4.J.Exp.Med.197,263–268(2003).

57. Hardy,M.P.&O’Neill,L.A.J.ThemurineIRAK2geneencodesfouralternativelyspliced

isoforms,twoofwhichareinhibitory.J.Biol.Chem.279,27699–27708(2004).58. Kinjyo,I.etal.SOCS1/JABisanegativeregulatorofLPS-inducedmacrophageactivation.

Immunity17,583–591(2002).59. Nakagawa,R.etal.SOCS-1participatesinnegativeregulationofLPSresponses.

Immunity17,677–687(2002).60. Boone,D.L.etal.Theubiquitin-modifyingenzymeA20isrequiredforterminationof

Toll-likereceptorresponses.Nat.Immunol.5,1052–1060(2004).61. Burns,K.etal.Tollip,anewcomponentoftheIL-1RIpathway,linksIRAKtotheIL-1

receptor.Nat.CellBiol.2,346–351(2000).62. Bulut,Y.,Faure,E.,Thomas,L.,Equils,O.&Arditi,M.CooperationofToll-likereceptor2

and6forcellularactivationbysolubletuberculosisfactorandBorreliaburgdorferioutersurfaceproteinAlipoprotein:roleofToll-interactingproteinandIL-1receptorsignalingmoleculesinToll-likereceptor2signaling.J.Immunol.167,987–994(2001).

63. Zhang,G.&Ghosh,S.Negativeregulationoftoll-likereceptor-mediatedsignalingby

Tollip.J.Biol.Chem.277,7059–7065(2002).

Page 42: Regulation of Inflammatory Macrophage Activation and

28

64. Seeley,J.J.&Ghosh,S.MolecularmechanismsofinnatememoryandtolerancetoLPS.JournalofLeukocyteBiology101,107–119(2017).

65. Beeson,P.B.Developmentoftolerancetotyphoidbacterialpyrogenanditsabolitionby

reticulo-endothelialblockade.Proc.Soc.Exp.Biol.Med.61,248–250(1946).66. Foster,S.L.,Hargreaves,D.C.&Medzhitov,R.Gene-specificcontrolofinflammationby

TLR-inducedchromatinmodifications.Nature447,972–978(2007).67. Brint,E.K.etal.ST2isaninhibitorofinterleukin1receptorandToll-likereceptor4

signalingandmaintainsendotoxintolerance.Nat.Immunol.5,373–379(2004).68. Diehl,G.E.etal.TRAIL-Rasanegativeregulatorofinnateimmunecellresponses.

Immunity21,877–889(2004).69. Chuang,T.-H.&Ulevitch,R.J.Triad3A,anE3ubiquitin-proteinligaseregulatingToll-like

receptors.Nat.Immunol.5,495–502(2004).70. McCartney-Francis,N.,Jin,W.&Wahl,S.M.AberrantTollreceptorexpressionand

endotoxinhypersensitivityinmicelackingafunctionalTGF-beta1signalingpathway.J.Immunol.172,3814–3821(2004).

71. Naiki,Y.etal.Transforminggrowthfactor-betadifferentiallyinhibitsMyD88-dependent,

butnotTRAM-andTRIF-dependent,lipopolysaccharide-inducedTLR4signaling.J.Biol.Chem.280,5491–5495(2005).

72. Muzio,M.etal.Differentialexpressionandregulationoftoll-likereceptors(TLR)in

humanleukocytes:selectiveexpressionofTLR3indendriticcells.J.Immunol.164,5998–6004(2000).

73. Fan,H.&Cook,J.A.Molecularmechanismsofendotoxintolerance.J.EndotoxinRes.10,

71–84(2004).74. Biswas,S.K.&Tergaonkar,V.Myeloiddifferentiationfactor88-independentToll-like

receptorpathway:Sustaininginflammationorpromotingtolerance?Int.J.Biochem.CellBiol.39,1582–1592(2007).

75. Ivashkiv,L.B.Epigeneticregulationofmacrophagepolarizationandfunction.Trendsin

Immunology34,216–223(2013).76. O’Neill,L.A.J.,Kishton,R.J.&Rathmell,J.Aguidetoimmunometabolismfor

immunologists.NatureReviewsImmunology16,553–565(2016).

Page 43: Regulation of Inflammatory Macrophage Activation and

29

77. Pearce,E.L.&Pearce,E.J.MetabolicPathwaysinImmuneCellActivationandQuiescence.Immunity38,633–643(2013).

78. Buck,M.D.,Sowell,R.T.,Kaech,S.M.&Pearce,E.L.MetabolicInstructionofImmunity.

Cell169,570–586(2017).79. Langston,P.K.,Shibata,M.&Horng,T.MetabolismSupportsMacrophageActivation.

FrontiersinImmunology8,(2017).80. Jung,J.,Zeng,H.&Horng,T.Metabolismasaguidingforceforimmunity.Nat.CellBiol.

21,85–93(2019).81. Wang,A.,Luan,H.H.&Medzhitov,R.Anevolutionaryperspectiveon

immunometabolism.Science363,eaar3932(2019).82. Ryan,D.G.etal.CouplingKrebscyclemetabolitestosignallinginimmunityandcancer.

NatureMetabolism1,16(2019).83. Nelson,D.L.,Cox,M.M.&Lehninger,A.L.Lehningerprinciplesofbiochemistry.(W.H.

Freeman,2005).84. Beard,D.A.,Babson,E.,Curtis,E.&Qian,H.Thermodynamicconstraintsfor

biochemicalnetworks.J.Theor.Biol.228,327–333(2004).85. Henry,C.S.,Broadbelt,L.J.&Hatzimanikatis,V.Thermodynamics-basedmetabolicflux

analysis.Biophys.J.92,1792–1805(2007).86. Schuster,S.,Fell,D.A.&Dandekar,T.Ageneraldefinitionofmetabolicpathwaysuseful

forsystematicorganizationandanalysisofcomplexmetabolicnetworks.Nat.Biotechnol.18,326–332(2000).

87. Schilling,C.H.,Letscher,D.&Palsson,B.O.Theoryforthesystemicdefinitionof

metabolicpathwaysandtheiruseininterpretingmetabolicfunctionfromapathway-orientedperspective.J.Theor.Biol.203,229–248(2000).

88. Kauffman,K.J.,Prakash,P.&Edwards,J.S.Advancesinfluxbalanceanalysis.Curr.Opin.

Biotechnol.14,491–496(2003).89. Beard,D.A.,Liang,S.&Qian,H.Energybalanceforanalysisofcomplexmetabolic

networks.Biophys.J.83,79–86(2002).90. Clark,W.M.Oxidation-reductionpotentialsoforganicsystems.(Williams&Wilkins,

1960).

Page 44: Regulation of Inflammatory Macrophage Activation and

30

91. Mitchell,P.Couplingofphosphorylationtoelectronandhydrogentransferbyachemi-osmotictypeofmechanism.Nature191,144–148(1961).

92. Mitchell,P.Chemiosmoticcouplinginoxidativeandphotosyntheticphosphorylation.

BiolRevCambPhilosSoc41,445–502(1966).93. Mitchell,P.Chemiosmoticcouplinginoxidativeandphotosyntheticphosphorylation.

1966.Biochim.Biophys.Acta1807,1507–1538(2011).94. Nicholls,D.G.&Ferguson,S.J.Bioenergetics4.(ElsevierAcademicPress,2013).95. Everts,B.etal.TLR-drivenearlyglycolyticreprogrammingviathekinasesTBK1-IKKɛ

supportstheanabolicdemandsofdendriticcellactivation.NatureImmunology15,323–332(2014).

96. Fayard,E.,Xue,G.,Parcellier,A.,Bozulic,L.&Hemmings,B.A.ProteinkinaseB

(PKB/Akt),akeymediatorofthePI3Ksignalingpathway.Curr.Top.Microbiol.Immunol.346,31–56(2010).

97. Robey,R.B.&Hay,N.IsAktthe“Warburgkinase”?—Akt-energymetabolism

interactionsandoncogenesis.SeminarsinCancerBiology19,25–31(2009).98. Amiel,E.etal.InhibitionofMechanisticTargetofRapamycinPromotesDendriticCell

ActivationandEnhancesTherapeuticAutologousVaccinationinMice.TheJournalofImmunology189,2151–2158(2012).

99. Drapier,J.C.&Hibbs,J.B.Differentiationofmurinemacrophagestoexpressnonspecific

cytotoxicityfortumorcellsresultsinL-arginine-dependentinhibitionofmitochondrialiron-sulfurenzymesinthemacrophageeffectorcells.J.Immunol.140,2829–2838(1988).

100. Cramer,T.etal.HIF-1αIsEssentialforMyeloidCell-MediatedInflammation.Cell112,

645–657(2003).101. Mills,E.L.etal.SuccinateDehydrogenaseSupportsMetabolicRepurposingof

MitochondriatoDriveInflammatoryMacrophages.Cell167,457-470.e13(2016).102. Palsson-McDermott,E.M.etal.PyruvateKinaseM2RegulatesHif-1αActivityandIL-1β

InductionandIsaCriticalDeterminantoftheWarburgEffectinLPS-ActivatedMacrophages.CellMetabolism21,65–80(2015).

103. Sag,D.,Carling,D.,Stout,R.D.&Suttles,J.Adenosine5’-monophosphate-activated

proteinkinasepromotesmacrophagepolarizationtoananti-inflammatoryfunctionalphenotype.J.Immunol.181,8633–8641(2008).

Page 45: Regulation of Inflammatory Macrophage Activation and

31

104. Ip,W.K.E.,Hoshi,N.,Shouval,D.S.,Snapper,S.&Medzhitov,R.Anti-inflammatoryeffectofIL-10mediatedbymetabolicreprogrammingofmacrophages.Science356,513–519(2017).

105. Fukuzumi,M.,Shinomiya,H.,Shimizu,Y.,Ohishi,K.&Utsumi,S.Endotoxin-induced

enhancementofglucoseinfluxintomurineperitonealmacrophagesviaGLUT1.Infect.Immun.64,108–112(1996).

106. Freemerman,A.J.etal.MetabolicReprogrammingofMacrophages:GLUCOSE

TRANSPORTER1(GLUT1)-MEDIATEDGLUCOSEMETABOLISMDRIVESAPROINFLAMMATORYPHENOTYPE.JournalofBiologicalChemistry289,7884–7896(2014).

107. Freemerman,A.J.etal.MyeloidSlc2a1-DeficientMurineModelRevealedMacrophage

ActivationandMetabolicPhenotypeAreFueledbyGLUT1.J.Immunol.202,1265–1286(2019).

108. Krawczyk,C.M.etal.Toll-likereceptor-inducedchangesinglycolyticmetabolism

regulatedendriticcellactivation.Blood115,4742–4749(2010).109. Liu,L.etal.Proinflammatorysignalsuppressesproliferationandshiftsmacrophage

metabolismfromMyc-dependenttoHIF1α-dependent.Proc.Natl.Acad.Sci.U.S.A.113,1564–1569(2016).

110. Rodriguez-Prados,J.-C.etal.SubstrateFateinActivatedMacrophages:AComparison

betweenInnate,Classic,andAlternativeActivation.TheJournalofImmunology185,605–614(2010).

111. Warburg,O.Ontheoriginofcancercells.Science123,309–314(1956).112. Warburg,O.Onrespiratoryimpairmentincancercells.Science124,269–270(1956).113. Buck,M.D.etal.MitochondrialDynamicsControlsTCellFatethroughMetabolic

Programming.Cell166,63–76(2016).114. Jha,A.K.etal.NetworkIntegrationofParallelMetabolicandTranscriptionalData

RevealsMetabolicModulesthatRegulateMacrophagePolarization.Immunity42,419–430(2015).

115. Kornberg,H.L.&Krebs,H.A.SynthesisofcellconstituentsfromC2-unitsbyamodified

tricarboxylicacidcycle.Nature179,988–991(1957).116. Michelucci,A.etal.Immune-responsivegene1proteinlinksmetabolismtoimmunityby

catalyzingitaconicacidproduction.Proc.Natl.Acad.Sci.U.S.A.110,7820–7825(2013).

Page 46: Regulation of Inflammatory Macrophage Activation and

32

117. Cordes,T.etal.ImmunoresponsiveGene1andItaconateInhibitSuccinateDehydrogenasetoModulateIntracellularSuccinateLevels.JournalofBiologicalChemistry291,14274–14284(2016).

118. Mills,E.L.etal.Itaconateisananti-inflammatorymetabolitethatactivatesNrf2via

alkylationofKEAP1.Nature556,113–117(2018).119. Bambouskova,M.etal.Electrophilicpropertiesofitaconateandderivativesregulate

theIκBζ-ATF3inflammatoryaxis.Nature556,501–504(2018).120. Qualls,J.E.etal.Sustainedgenerationofnitricoxideandcontrolofmycobacterial

infectionrequiresargininosuccinatesynthase1.CellHostMicrobe12,313–323(2012).121. Murray,P.J.Aminoacidauxotrophyasasystemofimmunologicalcontrolnodes.Nat.

Immunol.17,132–139(2016).122. Haschemi,A.etal.TheSedoheptuloseKinaseCARKLDirectsMacrophagePolarization

throughControlofGlucoseMetabolism.CellMetabolism15,813–826(2012).123. Tannahill,G.M.etal.SuccinateisaninflammatorysignalthatinducesIL-1βthroughHIF-

1α.Nature496,238–242(2013).124. Kornberg,H.L.AnapleroticSequencesinMicrobialMetabolism.AngewandteChemie

InternationalEditioninEnglish4,558–565(1965).125. Everts,B.etal.CommitmenttoglycolysissustainssurvivalofNO-producing

inflammatorydendriticcells.Blood120,1422–1431(2012).126. VandenBossche,J.etal.MitochondrialDysfunctionPreventsRepolarizationof

InflammatoryMacrophages.CellReports17,684–696(2016).127. West,A.P.etal.TLRsignallingaugmentsmacrophagebactericidalactivitythrough

mitochondrialROS.Nature472,476–480(2011).128. Sander,L.E.&Garaude,J.Themitochondrialrespiratorychain:Ametabolicrheostatof

innateimmunecell-mediatedantibacterialresponses.Mitochondrion41,28–36(2018).129. Lampropoulou,V.etal.ItaconateLinksInhibitionofSuccinateDehydrogenasewith

MacrophageMetabolicRemodelingandRegulationofInflammation.CellMetabolism24,158–166(2016).

130. Garaude,J.etal.Mitochondrialrespiratory-chainadaptationsinmacrophages

contributetoantibacterialhostdefense.Nat.Immunol.17,1037–1045(2016).

Page 47: Regulation of Inflammatory Macrophage Activation and

33

131. Quinlan,C.L.,Perevoshchikova,I.V.,Hey-Mogensen,M.,Orr,A.L.&Brand,M.D.Sitesofreactiveoxygenspeciesgenerationbymitochondriaoxidizingdifferentsubstrates.RedoxBiol1,304–312(2013).

132. Orr,A.L.etal.InhibitorsofROSproductionbytheubiquinone-bindingsiteof

mitochondrialcomplexIidentifiedbychemicalscreening.FreeRadic.Biol.Med.65,1047–1059(2013).

133. Orr,A.L.,Quinlan,C.L.,Perevoshchikova,I.V.&Brand,M.D.Arefinedanalysisof

superoxideproductionbymitochondrialsn-glycerol3-phosphatedehydrogenase.J.Biol.Chem.287,42921–42935(2012).

134. Chandel,N.S.etal.ReactiveoxygenspeciesgeneratedatmitochondrialcomplexIII

stabilizehypoxia-induciblefactor-1alphaduringhypoxia:amechanismofO2sensing.J.Biol.Chem.275,25130–25138(2000).

135. Cheng,S.-C.etal.mTOR-andHIF-1-mediatedaerobicglycolysisasmetabolicbasisfor

trainedimmunity.Science345,1250684–1250684(2014).136. Covarrubias,A.J.etal.Akt-mTORC1signalingregulatesAclytointegratemetabolicinput

tocontrolofmacrophageactivation.Elife5,(2016).137. Gut,P.&Verdin,E.Thenexusofchromatinregulationandintermediarymetabolism.

Nature502,489–498(2013).138. Kaelin,W.G.&McKnight,S.L.InfluenceofMetabolismonEpigeneticsandDisease.Cell

153,56–69(2013).

Page 48: Regulation of Inflammatory Macrophage Activation and

CHAPTER 2

Glycerol phosphate shuttle enzyme GPD2 drives glucose oxidation to regulate

macrophage inflammatory responses

This chapter is a reproduction of a published manuscript with minor modifications, including incorporation of unpublished data. P. Kent Langston1, Aya Nambu1, Jonathan Jung1,2, Munehiko Shibata1, H. Ibrahim Aksoylar1, Jiahui Lei1, Peining Xu3, Mary T. Doan3, Helen Jiang3, Michael R. MacArthur1, Xia Gao4, Yong Kong5, Edward T. Chouchani6, Jason W. Locasale4, Nathaniel W. Synder3, and Tiffany Horng1,7*

1Department of Genetics & Complex Diseases, Harvard T.H. Chan School of Public Health, Boston, MA 2School of Medicine, University of Glasgow, Glasgow, UK 3A.J. Drexel Autism Institute, Drexel University, Philadelphia, PA 4Department of Pharmacology & Cancer Biology, Duke University School of Medicine, Durham, NC 5Depart of Immunobiology, Yale University School of Medicine, New Haven, CT 6Department of Cell Biology, Harvard Medical School, Boston, MA 7School of Life Sciences and Technology, ShanghaiTech University, Shanghai, China

Author contributions P.K.L. designed and performed experiments and analyzed data. P.K.L. prepared figures, and P.K.L. and T.H. wrote the manuscript. T.H. supervised the project, including experimental design and data analysis. A.N. performed TLR signaling immunoblots. M.S. generated Gpd2-/- mice. P.X., M.T.D., H.J., N.S., X.G., and J.L. performed metabolite profiling and provided related technical expertise. M.R.M. assisted in collection of in vivo data. E.T.C. provided technical expertise.

Page 49: Regulation of Inflammatory Macrophage Activation and

35

ABSTRACT

Macrophages are activated during microbial infection to coordinate inflammatory responses and

host defense. Here we show that in macrophages activated by bacterial lipopolysaccharide (LPS),

mitochondrial glycerol 3-phosphate dehydrogenase (GPD2) regulates glucose oxidation to drive

inflammatory responses. GPD2, a component of the glycerol phosphate shuttle (GPS), boosts

glucose oxidation to fuel acetyl-CoA production, histone acetylation and inflammatory gene

induction. While acute LPS exposure drives macrophage activation, prolonged exposure triggers

entry into LPS tolerance, where macrophages orchestrate immunosuppression to limit the

detrimental effects of sustained inflammation. We find that the shift in the inflammatory response

is modulated by GPD2, which coordinates a shutdown of oxidative metabolism that limits acetyl-

CoA availability for histone acetylation at inflammatory genes, thus contributing to suppression

of inflammatory responses. Therefore, GPD2 and the GPS integrate the extent of microbial

stimulation with glucose oxidation to balance the beneficial and detrimental effects of the

inflammatory response.

Page 50: Regulation of Inflammatory Macrophage Activation and

36

INTRODUCTION

Microbial infection elicits an inflammatory response characterized by activation of immune cells

and elevated local and circulating pro-inflammatory cytokines (e.g. IL-6 and IL-1β) and enhanced

antimicrobial activity. Macrophages are key regulators of innate immunity and inflammation,

coupling recognition of infection by detection of Pathogen-Associated Molecular Patterns

(PAMPs) to induction of inflammatory responses. For example, upon detection of the gram-

negative bacterial component lipopolysaccharide (LPS), macrophages are activated (hereafter LPS

activation) to coordinate the induction of innate and adaptive immune responses by producing

chemoattractants and inflammatory cytokines. However, in response to overwhelming infection or

protracted exposure to LPS, activated macrophages shift to a refractory or tolerant state (hereafter

LPS tolerance)1, 2. Tolerant macrophages mount a blunted response to further LPS challenge, and

produce markedly lower levels of inflammatory cytokines, contributing to the induction of whole-

body immunosuppression. Such immunosuppression limits the detrimental effects of sustained

inflammation, which otherwise damages multiple organ systems and leads to mortality3. Therefore,

acute microbial infection elicits an inflammatory response that confers host defense, but persistent

and overwhelming microbial infection triggers inflammation followed by a protracted period of

immunosuppression as a protective mechanism to mitigate multiorgan dysfunction1, 2, 3.

Cellular metabolism is increasingly appreciated for its role in regulating macrophage activation

and functions as well as innate immunity and inflammation. Cellular metabolism interfaces with

signal transduction and other regulatory processes, and genetic and pharmacological modulation

of metabolism can influence macrophage activation and functions4, 5, 6. Many such studies have

Page 51: Regulation of Inflammatory Macrophage Activation and

37

used macrophage activation (or polarization) by polarizing signals like LPS and the cytokine IL-4

as a model to elucidate fundamental mechanisms by which metabolism regulates macrophage

activation. One paradigm that has emerged is that in macrophages activated by IL-4, which

orchestrate Type 2 inflammatory responses like tissue repair and fibrosis, oxidative metabolism is

increased to boost so-called M2 activation4, 6, 7, 8. Increased oxidative metabolism upregulates the

production of citrate, a TCA cycle intermediate, which is used by the enzyme ATP-citrate lyase

(ACLY) to generate a nucleocytosolic pool of acetyl-CoA (Ac-CoA). Because Ac-CoA is the

metabolic substrate for histone acetylation, ACLY-dependent Ac-CoA production links increased

oxidative metabolism to enhanced expression of IL-4-inducible genes during M2 activation9. In

addition, increased oxidative metabolism supports the production of another TCA intermediate

alpha-ketoglutarate (AKG), a cofactor for the histone demethylase Jmjd3, to regulate histone

demethylation and fuel IL-4-inducible gene expression10. It is noteworthy that in addition to M2

activation, increased oxidative metabolism is a cornerstone of cellular activation, differentiation,

and proliferation in many contexts11. In contrast, a metabolic hallmark of LPS-stimulated

macrophages is a robust shutdown of oxidative metabolism4, 5, 12. Such shutdown appears to be

mediated by multiple mechanisms, including induction of iNOS and production of nitric oxide

(NO), which modifies multiple complexes in the electron transport chain (ETC)13, 14; induction of

IRG1 and production of itaconate, a metabolite that inhibits Complex II of the ETC15; and

induction of reverse electron transport (RET), which counters oxidative phosphorylation16. In

addition to respiratory shutdown, another metabolic hallmark of LPS-stimulated macrophages is

upregulation of aerobic glycolysis, the process by which glucose-derived pyruvate is shunted away

from mitochondrial oxidative metabolism towards the production of lactate4, 5, 8, 12, 17. It has been

proposed that induction of aerobic glycolysis, at the expense of mitochondrial glucose oxidation,

Page 52: Regulation of Inflammatory Macrophage Activation and

38

allows for rapid ATP production during infection by fast-replicating microbes. Beyond this

bioenergetic consideration, however, it is unclear how respiration shutdown and aerobic glycolysis

induction contribute to the inflammatory responses of LPS-stimulated macrophages4, 5.

Here we show that contrary to the paradigm that LPS-stimulated macrophages drive aerobic

glycolysis and shut down respiration, macrophages acutely exposed to LPS upregulate glucose

oxidation in the mitochondria. Glucose oxidation supports production of Ac-CoA, thus increasing

substrate availability for histone acetylation at the inflammatory genes Il6 and Il1b. A key regulator

of such glucose oxidation is GPD2, mitochondrial glycerol phosphate dehydrogenase, a rate-

limiting enzyme in the glycerol 3-phosphate shuttle (GPS). LPS signaling increases GPD2 activity

and flux through the GPS, and in macrophages deficient for GPD2, glucose oxidation, histone

acetylation, and Il6 and Il1b gene induction in response to acute LPS exposure is attenuated.

Importantly, we show that there is a metabolic shift from increased to decreased respiration over

the course of LPS exposure as macrophages transition from activation to tolerance, and that in

tolerant macrophages, decreased respiration limits the ability of glucose oxidation to support Ac-

CoA production thus contributing to the inability to induce Il6 and Il1b. Moreover, the shift from

increased to decreased oxidative metabolism is regulated by GPD2-mediated glucose oxidation,

since block of glucose utilization or GPD2 deficiency prevents this shift, leading to rescue of

histone acetylation and Il6 and Il1b induction. We propose that GPD2 boosts oxidative metabolism

to drive inflammatory gene induction during LPS activation but suppresses oxidative metabolism

to contribute to inhibition of inflammatory gene induction during LPS tolerance. Therefore, GPD2

and the GPS appear to couple the duration and extent of microbial stimulation to the balance of

inflammatory response induction and suppression.

Page 53: Regulation of Inflammatory Macrophage Activation and

39

RESULTS

ACLY supports histone acetylation and inflammatory gene induction during LPS activation

In a recent study, we proposed that one mechanism by which increased oxidative metabolism

supports M2 activation is via enhanced production of Ac-CoA, thus bolstering histone acetylation

at IL-4-inducible genes. An important regulator of this process is ACLY, an enzyme that produces

a nucleocytosolic pool of Ac-CoA that can be used for histone acetylation9. We started the current

study by asking about the role of ACLY in macrophages acutely exposed to LPS (hereafter LPS-

activated macrophages).

We found that LPS stimulation of bone marrow derived macrophages (BMDMs) rapidly increased

ACLY S455 phosphorylation, an activating phosphorylation18, indicating that LPS signaling

enhanced ACLY activation (Fig. 2.1a). Next, we asked how ACLY activity influenced Il6 and

Il1b, prototypes of the LPS-inducible inflammatory response, during LPS activation. In this and

all other experiments examining LPS activation in our study, all parameters of LPS activation are

assessed between 0-3 hours after LPS stimulation. We found that treatment with the ACLY

inhibitor SB-204990 attenuated histone acetylation at Il6 and Il1b gene promoters (Fig. 2.1b) and

impaired Il6 and Il1b gene induction (Fig. 2.1c).

Page 54: Regulation of Inflammatory Macrophage Activation and

40

Figure 2.1 ACLY activity supports inflammatory gene induction in LPS-activated macrophages. BMDMs were stimulated with LPS for 0-3 hours with or without ACLY inhibitor SB 204990 (ACLYi) pretreatment, followed by immunoblot analysis of ACLY phosphorylation at Ser455 (a), ChIP-qPCR analysis of histone acetylation at the Il6 and Il1b promoters (b), and qPCR analysis of Il6 and Il1b gene expression (c). *p ≤0.05, **p ≤0.01, ***p ≤0.001, ****p ≤0.0001 (Student’s t-test). Data are from one experiment representative of three experiments (mean and s.e.m. of duplicates (c) or triplicates (b)).

Two structurally distinct ACLY inhibitors similarly reduced LPS-inducible histone acetylation and

expression of Il6 and Il1b (Supplementary Fig. 2.1a,b), suggesting specificity in the effects of

the ACLY inhibitors. Consistent with the idea that histone acetylation and histone

acetyltransferases including CBP/p300 in particular support LPS-inducible gene expression19, 20,

21, 22, 23, inhibition of p300 with the selective inhibitor C646 reduced histone acetylation and

induction of Il6 and Il1b (Supplementary Fig. 2.1c,d). Together these findings support the idea

that ACLY activity fuels histone acetylation and inflammatory gene induction in LPS-activated

macrophages.

0

200

400

600

800

Relative Expression

Il1b**

0

5000

10000

15000

20000Il6

*

0.0

0.5

1.0

1.5**

AcH3 Il6

0

2

4

6AcH3K27 Il6

****

0

1

2

3AcH4 Il6

**

%Input

0

5

10

15

20AcH3K27 Il1b

********

0.0

0.2

0.4

0.6

0.8

1.0AcH3 Il1b

*******

0

5

10

15AcH4 Il1b

*****

%Input

Relative Expression

p-ACLY(S455)

ACLY

LPS (h): 0.5 10

a cbLPSLPS+ACLYi

Naïve

2

α-Tubulin

Page 55: Regulation of Inflammatory Macrophage Activation and

41

Glucose oxidation is enhanced during LPS activation to fuel histone acetylation and

inflammatory gene induction in macrophages

ACLY-dependent Ac-CoA production depends on availability of the TCA cycle intermediate

citrate and oxidative metabolism, therefore the data from our ACLY analysis was in apparent

conflict with the paradigm that LPS-stimulated macrophages shut down respiration. By measuring

oxygen consumption, we found that acute LPS exposure triggered a small but significant and

highly-reproducible increase in mitochondrial oxidative metabolism (Fig. 2.2). While longer LPS

exposure led to a progressive and sustained decrease in respiration, the early increase was transient

and likely explains the many reports that conclude that LPS stimulated macrophages shut down

respiration (Supplementary Fig. 2.2a). Because LPS stimulation is well-known to enhance

glucose utilization (Supplementary Fig. 2.2b and 4), we asked if the early increase in oxidative

metabolism could be fueled by glucose oxidation. Indeed, treatment with the glucose utilization

inhibitor 2-deoxyglucose (2DG) abrogated the early increase in oxidative metabolism (Fig. 2.2).

Figure 2.2 LPS activation increases glucose oxidation in macrophages. Mitochondrial respiration, determined by oxygen consumption rate (OCR), in BMDMs stimulated with LPS for 1h +/- 2DG pretreatment. Basal and maximal respiration were calculated by subtraction of non-mitochondrial respiration (OCR post-Rot/AA) from mean OCR pre-oligomycin and post-FCCP injection, respectively. *p ≤0.05, **p ≤0.01, ***p ≤0.001 (Student’s t-test). Data are from one experiment representative of ten experiments (mean and s.e.m. of octuplets).

LPSNaïve

0 25 50 75 1000

70

140

210

280

350

Time (min)

OCR (pmol/min)

FCCP Rot/AAOligo

LPS+2DG

0

50

100

150

Basal OCR (pmol/min)

******

0

50

100

150

200

250

Max. OCR (pmol/min) ***

LPSLPS+2DG

Naïve

Maximal

Basal

Page 56: Regulation of Inflammatory Macrophage Activation and

42

Figure 2.3 Glucose oxidation supports inflammatory gene induction in LPS-activated macrophages by providing carbon substrate for histone acetylation. BMDMs were stimulated with LPS for 0-3 h, +/- 2DG pretreatment, followed by analysis of the following parameters. 13C6-glucose tracing into citrate-isocitrate (a) and Ac-CoA (b) during LPS activation presented as abundance of m+2 isotopologue relative to all other isotopologues. (c) Incorporation of carbons from 14C6-glucose into histones in naïve and LPS-activated BMDMs. (d) Promoter region histone acetylation and (e) expression of Il6 and Il1b and (f) IL-6 cytokine production during LPS activation +/- 2DG pretreatment. *p ≤0.05, **p ≤0.01, ***p ≤0.001, ****p ≤0.0001 (Student’s t-test). Data are from one experiment representative of three (a-c) or four (d-f) experiments (mean and s.e.m. of duplicates (c,e,f), triplicates (d), or quadruplicates (a,b)).

To further explore the possibility that glucose oxidation supports Ac-CoA production and

inflammatory gene induction in LPS-activated macrophages, we turned to glucose isotope tracing

experiments using uniformly 13C-labeled glucose. Consistent with this possibility, we found that

acute LPS stimulation enhanced labeling of glucose-derived carbons into citrate/isocitrate (Fig.

2.3a). Furthermore, we observed a robust increase in labeling of (m+2) Ac-CoA, indicating that

glucose oxidation fueled Ac-CoA production during LPS activation (Fig. 2.3b). Tracing glucose

a c

d

Naïve

LPS 1.5h

LPS 3h

0

50

100

150

Percent Enrichment Other Isotopologues

m+2********

Naïve

LPS 1h

LPS 2h

LPS 3h

0

50

100

150

Percent Enrichment

*********

Citrate-Isocitrateb

Acetyl-CoA

0

1000

2000

3000

4000

cpm/mg protein LPS

**

0.0

0.5

1.0

1.5

2.0

2.5

%Input

AcH3 Il6******

0

3

6

9AcH3K27 Il6

*****

0

2

4

6

8AcH4 Il6

***

0.0

0.5

1.0

1.5

2.0

%Input

AcH3 Il1b

*****

0

10

20

30AcH3K27 Il1b

******

0

5

10

15

20AcH4 Il1b

*****

e

0

10000

20000

30000

40000Il6

**

0

1000

2000

3000

4000

Relative Expression

Il1b

**

Relative Expression

0

5000

10000

15000

20000

25000

IL-6 Production (pg/mL)

****

f

LPSLPS+2DG

Naïve

Naïve

Page 57: Regulation of Inflammatory Macrophage Activation and

43

carbons all the way into histones, we were also able to demonstrate increased labeling of histones

in LPS-activated macrophages (Fig. 2.3c). Such glucose metabolism supported histone acetylation

at the promoters of Il6 and Il1b, since 2DG treatment attenuated LPS-inducible histone acetylation

at these promoters (Fig. 2.3d). Correlating with effects on promoter histone acetylation, 2DG

treatment attenuated LPS-inducible expression of Il6 and Il1b (Fig. 2.3e). Finally, 2DG treatment

reduced production of IL-6 cytokine in LPS-activated macrophages (Fig. 2.3f), while production

of mature IL-1β protein in BMDMs requires inflammasome activation24 and was not further

examined. Incubation of BMDMs in glucose free media also led to diminished histone acetylation

and induction of Il6 and Il1b, although the effects were more modest compared to 2DG treatment

likely because glycogen stores are being tapped to fuel glycolysis (Supplementary Fig. 2.3).

Together the mitochondrial oxygen consumption, glucose tracing and functional analyses suggest

that Ac-CoA availability can be limiting for histone acetylation and inflammatory gene induction

during LPS activation, and that LPS stimulation boosts glucose oxidation leading to increased Ac-

CoA availability for supporting histone acetylation and inflammatory gene induction.

GPD2 regulates glucose utilization and oxidation in LPS-activated macrophages

In the cytosolic phase of glucose oxidation, glucose metabolism is coupled to NADH production

(at the step catalyzed by GAPDH), which could slow down glucose oxidation if cytosolic NAD+

is not regenerated. The malate-aspartate shuttle (MAS) is thought to be ubiquitously expressed and

deployed during glucose oxidation to regenerate cytosolic NAD+ while transferring glucose-

derived electrons, in the form of NADH, to Complex I of the ETC25. However, the kinetics by

which glucose carbons label histones in our experiment (1.5h; see Fig. 2.3c), relative to steady-

state histone labeling by the relevant carbon substrate in other contexts (~24h)26, 27, suggested an

Page 58: Regulation of Inflammatory Macrophage Activation and

44

alternative mechanism for rapidly mobilizing glucose oxidation in LPS-activated macrophages.

Using metabolomics, we identified the glycerol phosphate shuttle (GPS) as a top metabolic

pathway regulated during LPS activation (Fig. 2.4a). The GPS is a glycolytic shunt that delivers

glucose-derived electrons to the ETC. In this shuttle, the NADH produced by GAPDH is used to

reduce dihydroxyacetone phosphate (DHAP) to glycerol 3-phosphate (G3P) in a reaction catalyzed

by cytoplasmic glycerol 3-phosphate dehydrogenase (cGPDH, or GPD1); G3P then serves as a

source of electrons for the ETC through the activity of mitochondrial glycerol 3-phosphate

dehydrogenase (mGPDH, or GPD2). GPD2 is a non-canonical component of the ETC that couples

oxidation of G3P back to DHAP with reduction of its FAD cofactor to FADH2, from which

electrons flow downward through the ETC to drive mitochondrial respiration (Fig. 2.4b and

Supplementary Fig. 2.4).

Figure 2.4 LPS activation enhances glycerol phosphate shuttle activity in macrophages. (a) Global steady-state metabolite profiling of unstimulated and LPS-stimulated BMDMs, with 10 most-significantly enhanced pathways presented, ranked by p value, and pathways of highest interest in red text. (b) Schematic depicting the spatial and biochemical position of the GPS and GPD2 at the nexus between glycolysis and electron transport. (c) Expression of GPS enzymes Gpd1 and Gpd2 in BMDMs +/- 1.5 h LPS stimulation. *p ≤0.05, **p ≤0.01 (Student’s t-test). Data are from one experiment representative of one (a) or three (c) experiments (mean and s.e.m. of duplicates (c) or quadruplicates (a)).

gluconeogenesis

TCA cycleintracellular FSH signalingphenylacetate metabolismphospholipid biosynthesis

fructose/mannose degradationpentose phosphate pathway

glycerolipid metabolismglycerol 3-phosphate shuttle

glycolysis

20100

p value

3e-01

1e-06

Fold Enrichment

a b c

matrix

FADH2 FAD

Q QH2

GPD2IMS

cytosol

ELECTRON TRANSPORT

III

DHAP

GLYCOLYSIS

G3PGPD1

NADH NAD+

0

5

10

15

NaïveLPS

Gpd2

**

0.0

0.5

1.0

1.5

Relative Expression Gpd1

*

Relative Expression

Page 59: Regulation of Inflammatory Macrophage Activation and

45

Like the MAS, the GPS regenerates cytosolic NAD+ to maintain cytosolic redox balance during

glucose oxidation. Unlike the MAS, GPS activity is restricted to a limited number of tissues

because GPD2 is expressed in a tissue-specific manner28, 29. The physiological role of the GPS is

poorly understood, but it has been proposed that GPD2/GPS activity allows for high rates of

glucose oxidation, given that GPD2 activity enhances glucose-stimulated insulin secretion in

pancreatic β-cells 29, 30. Nothing is known of GPD2 in macrophages except that its biochemical

activity is increased after microbial infection31.

Consistent with the identification of the GPS in our metabolomics analysis, we found that BMDMs

expressed GPD2 and that LPS stimulation further augmented GPD2 transcript and protein, while

transcript levels of GPD1 were very modestly affected by LPS stimulation (Fig. 2.4c and

Supplementary Fig. 2.5a). Furthermore, LPS stimulation increased GPD2 activity, as assessed

by Seahorse extracellular flux analysis of G3P-driven respiration (Supplementary Fig. 2.5b and

see Discussion). Using Crispr-Cas9 genome editing, we generated mice lacking Gpd2 (Fig. 2.5a).

GPD2 KO mice are viable, fertile, and appear normal. BMDMs from GPD2 KO mice develop

normally as indicated by CD11b and F4/80 expression (Fig. 2.5b). However, GPD2 KO BMDMs

lack GPD2 protein (Fig. 2.5c) and activity, as assessed by G3P-driven respiration (Fig. 2.5d).

Page 60: Regulation of Inflammatory Macrophage Activation and

46

Figure 2.5 Generation of GPD2 KO mice for studying the role of glycerol phosphate shuttle activity in macrophage inflammatory responses. (a) Guide RNA (gRNA) targeted to exon 6 in Gpd2 to generate whole-body GPD2 KO mice using CRISPR-Cas9 genome editing. (b) FACS analysis of F4/80 and CD11b expression on BMDMs from WT and GPD2 KO mice. Numbers indicate percent of cells in gate (F4/80+CD11b+). (c) Immunoblot analysis of GPD2 in naïve WT and GPD2 KO BMDMs. (d) G3P-driven respiration in permeabilized naïve WT and GPD2 KO BMDMs. Bar graph represents fold change in OCR after addition of G3P, ADP, and rotenone (Rot). ****p ≤0.0001 (Student’s t-test). Data are from one experiment representative of one (c) or three (d) experiments (mean and s.e.m. of sextuplets (d)).

Using glucose tracing experiments, we found that acute LPS stimulation of WT BMDMs led to a

dramatic increase in the abundance of total and (m+3) G3P, indicative of enhanced flux through

the GPS. In GPD2 KO BMDMs, abundance of total and (m+3) G3P was further increased,

consistent with accumulation of the GPD2 substrate in the absence of GPD2 activity (Fig. 2.6a).

We also found that the early, 2DG-sensitive increase in oxidative metabolism (Fig. 2.2) was

diminished in GPD2 KO BMDMs (Fig. 2.6b), indicating a role for GPD2 in mediating glucose

oxidation in LPS-activated macrophages. Importantly, the well-established ability of LPS to

GPD2

α-Tubulin

KOWT

a

c

dGPD2 KO

Gpd22 3 4 56 78 9 10,11,12,13

14,15,16

171

gRNA

WT

0.0

0.5

1.0

1.5

2.0

Fold Change in OCR

0 10 20 30 40 50 600

50

100

150

200

Time (min)

OCR (pmol/min)

Oligo AAG3P/ADP/Rot

GPD2 KOWT

****

97.3 97.9

CD11b

F4/80

WT GPD2 KOb

Page 61: Regulation of Inflammatory Macrophage Activation and

47

upregulate glucose uptake and utilization was impaired in GPD2 KO BMDMs (Fig. 2.6c),

indicating that GPS activity is rate-limiting for glucose utilization in LPS-activated macrophages.

In line with reduced glucose utilization and oxidation, GPD2 KO BMDMs displayed less LPS-

inducible glucose labeling into many glycolytic and TCA intermediates (Supplementary Fig. 2.6)

and attenuated aerobic glycolysis (Fig. 2.6d). Taken together, our metabolic analyses of GPD2

KO BMDMs identify the GPS as a critical regulator of glucose utilization and oxidation in LPS-

activated macrophages.

Figure 2.6 The GPS enzyme GPD2 regulates glucose oxidation in LPS-activated macrophages. Wild-type (WT) and GPD2 KO BMDMs were stimulated with LPS for 0-3 h, followed by analysis of the following parameters. (a) 13C6-glucose tracing into glycerol 3-phosphate (presented as peak areas for each isotopologue) in WT and GPD2 KO BMDMs. (b) Basal and maximal mitochondrial respiration, and (c) uptake of 3H-deoxy-D-glucose in WT and GPD2 KO BMDMs. (d) Extracellular acidification rate (ECAR), indicative of aerobic glycolysis, in naïve and LPS-activated WT and GPD2 KO BMDMs. *p ≤0.05, **p ≤0.01, ***p ≤0.001, ****p ≤0.0001 (Student’s t-test). Data are from one experiment representative of three (a,c,d) or four (b) experiments (mean and s.e.m. of triplicates (a,c) or sextuplets (b,d)).

b

c

LPS***

WT GPD2 KO0

10

20

30

40

50

3H cpm/mg protein

LPS 0.5hLPS 1hLPS 2h

***

Naïve

Naïve

a

Naïve

LPS 1.5h

LPS 3hNaïve

LPS 1.5h

LPS 3h

0

2 107

4 107

6 107

8 107

Peak Area

Unlabeledm+1m+2m+3

********

WT GPD2 KO

Glycerol 3-Phosphate

WT GPD2 KO0

10

20

30

40

50

Basal OCR (pmol/min)

WT GPD2 KO0

25

50

75

100

Max. OCR (pmol/min)

*

0

2

4

6

8

10

ECAR (mpH/min) ***

NaïveLPS

d

WT GPD2 KO

Page 62: Regulation of Inflammatory Macrophage Activation and

48

GPD2-mediated glucose oxidation fuels inflammatory gene induction during LPS activation

We next sought to determine how reduced glucose utilization and oxidation in GPD2 KO BMDMs

affects LPS activation. We found that in GPD2 KO BMDMs, LPS-inducible labeling of glucose

carbons into citrate (Fig. 2.7a) and Ac-CoA was reduced (Fig. 2.7b). LPS-inducible histone

acetylation at the Il6 and Il1b promoters was also attenuated (Fig. 2.7c), leading to diminished Il6

and Il1b gene induction (Fig. 2.7d). Finally, GPD2 KO BMDMs produced reduced levels of IL-6

(Fig. 2.7e). Collectively, our data indicates that GPS activity boosts glucose oxidation, leading to

enhanced carbon substrate availability for histone acetylation and increased inflammatory gene

induction (Supplementary Fig. 2.7).

WT GPD2 KO0

1

2

3

%Input

*

AcH3 Il6

WT GPD2 KO0

2

4

6

8

10*

AcH3K27 Il6

WT GPD2 KO0

2

4

6

8

****

AcH4 Il6

WT GPD2 KO0

1

2

3

4

5

%Input

**

AcH3 Il1b

WT GPD2 KO0

10

20

30*

AcH3K27 Il1b

WT GPD2 KO0

2

4

6

**

AcH4 Il1b

WT GPD2 KO0

50000

100000

150000

Relative Expression

**

Il6

WT GPD2 KO0

1000

2000

3000

Relative Expression

**

Il1b

WT GPD2 KOIL-6 Production (pg/mL)

c d

ea

LPS

LPS

Naïve

Naïve

Fold Change %Enrichment

WT GPD2 KO

Acetyl-CoA

0.0

0.5

1.0

1.5

2.0***

0

1

2

3

4*** LPS

Naïve

Fold Change %Enrichment

WT GPD2 KO

Citrate-Isocitrate

b

0

2000

4000

6000

8000

10000 ****

Page 63: Regulation of Inflammatory Macrophage Activation and

49

Figure 2.7 GPD2 activity controls inflammatory gene induction in LPS-activated macrophages by regulating Ac-CoA production and histone acetylation. WT and GPD2 KO BMDMs were stimulated with LPS for 0-3 h, followed by analysis of the following parameters. (a,b) 13C6-glucose tracing into (a) citrate and (b) Ac-CoA in WT and GPD2 KO BMDMs, presented as LPS-induced fold change in percent m+2 enrichment. (c) Promoter region histone acetylation and (d) expression of Il6 and Il1b and (e) IL-6 cytokine production in WT and GPD2 KO BMDMs. *p ≤0.05, **p ≤0.01, ***p ≤0.001, ****p ≤0.0001 (Student’s t-test). Data are from one experiment representative of three (a,b) or four (c-e) experiments (mean and s.e.m. of duplicates (d,e), triplicates (a,c), or quadruplicates (b)).

Glucose utilization drives shutdown of respiration and suppression of inflammatory gene

induction in LPS tolerant macrophages

While the data above indicated that the early increase in glucose oxidation fuels inflammatory

gene induction after acute LPS exposure, what could be the rationale and relevance of the shift to

decreased oxidative metabolism during prolonged exposure (Supplementary Fig. 2.2a)?

Importantly, the inflammatory response is well established to shift over the course of LPS exposure

and microbial infection, in vitro and in vivo. Acute LPS stimulation and microbial infection trigger

induction of inflammation responses, while prolonged LPS stimulation and microbial infection

drive entry into a state of immunosuppression. Therefore, we hypothesized that consistent with the

role of increased oxidative metabolism in supporting inflammatory gene induction during LPS

activation, the decreased oxidative metabolism in LPS tolerant macrophages could limit Ac-CoA

availability for histone acetylation, thus contributing to inflammatory gene suppression.

This hypothesis was tested using a well-established model of LPS tolerance, where an extended

exposure to LPS triggered tolerance as indicated by the inability to upregulate Il6 and Il1b after

LPS rechallenge (Fig. 2.8 and 1, 2). In contrast to LPS activation which was examined at early time

points (0-3 h), LPS tolerance throughout this study was examined at late time points, 12-28 h after

LPS stimulation.

Page 64: Regulation of Inflammatory Macrophage Activation and

50

Figure 2.8 Naïve and LPS tolerant macrophages exhibit profound differences in responsiveness to LPS stimulation. (a) Schematic of workflow for in vitro LPS tolerance experiments. BMDMs were left unstimulated (naïve, N) or stimulated with LPS for 24h to induce LPS tolerance (tolerance, T). LPS responsiveness was assessed by stimulating naïve (N+L) or re-stimulating tolerant (T+L) BMDMs with LPS. In the T+L+2DG condition, 2DG was added during tolerance induction and washed out prior to LPS restimulation in the absence of inhibitor. Naïve and Tolerant cells were collected as baseline controls (N and T). (b) Representative induction of Il6 and Il1b genes in BMDMs treated according to (a).

We found that in contrast to naïve macrophages where LPS exposure drove glucose conversion

into citrate and Ac-CoA, in tolerant macrophages LPS exposure failed to drive glucose conversion

into citrate and Ac-CoA (Fig. 2.9a,b). Therefore, tolerance-associated shutdown of respiration was

incompatible with LPS-inducible glucose oxidation and Ac-CoA production.

Figure 2.9 Glucose utilization impairs respiration in tolerant macrophages. 13C6-glucose tracing into citrate-isocitrate (a) and Ac-CoA (b) in N, N+L, T, and T+L conditions. Incorporation of labeled-glucose-derived carbons presented as abundance of m+2 isotopologue relative to all other isotopologues. (c) Maximal respiration after 12h LPS stimulation +/- 2DG pretreatment. *p ≤0.05, **p ≤0.01, ***p ≤0.001, ****p ≤0.0001 (Student’s t-test). Data are from one experiment representative of three experiments (mean and s.e.m. of quadruplicates (a,b) or octuplets (c)).

a

Tolerant (T)

Tolerant + LPS(T+L) 24h LPS

100 ng/mL

24h Unstim

4h LPS10 ng/mL

24h LPS100 ng/mL

4h LPS10 ng/mL

24h Unstim

Naïve (N)

Naïve + LPS(N+L)

T+L+2DG1h 2DG 24h LPS

100 ng/mL4h LPS10 ng/mL

4h Unstim

4h Unstim

AssayWash

0

5000

10000

15000

Relative Expression N

N+LTT+L

***

0

1000

2000

3000

4000 ***

Il6

Il1b

Relative Expression

b

ca b

N N+L T T+L0

50

100

150

Percent Enrichment

****

N N+L T T+L0

50

100

150

***

Other Isotopologues m+2

Citrate-Isocitrate Acetyl-CoA

0 25 50 75 1000

100

200

300

400

Time (min)

OCR (pmol/min)

FCCP Rot/AAOligoLPS LPS+2DGNaïve

0

50

100

150

Basal OCR (pmol/min)

** ****

0

100

200

300

Max. OCR (pmol/min)

****

****

LPSLPS+2DG

Naïve

Page 65: Regulation of Inflammatory Macrophage Activation and

51

Serendipitously, we found that tolerance-associated respiration shutdown was regulated by

glucose utilization. 2DG treatment during LPS tolerance induction led to a striking although

incomplete rescue of respiration, indicating that glucose utilization was necessary for optimal

inhibition of oxidative metabolism (Fig. 2.9c). Furthermore, 2DG-mediated respiration rescue was

associated with a partial recovery of histone acetylation and inflammatory gene induction in

tolerant macrophages. While LPS stimulation of naïve and tolerant BMDMs led to robust and

weak increases in histone acetylation at the Il6 and Il1b promoters respectively (comparing N and

N+L with T and T+L), 2DG treatment during tolerance induction partially restored histone

acetylation at the Il6 and Il1b promoters (comparing T, T+L and T+L+2DG) (Fig. 2.10a).

Moreover, 2DG treatment during tolerance induction led to recovery of Il6 and Il1b induction (Fig.

2.10b) and IL-6 production (Fig. 2.10c).

Figure 2.10 Glucose utilization limits inflammatory gene induction in tolerant macrophages. (a) Promoter region histone acetylation and (b) expression of Il6 and Il1b and (c) IL-6 cytokine production in N, N+L, T, and T+L+2DG conditions. *p ≤0.05, **p ≤0.01, ***p ≤0.001, ****p ≤0.0001 (Student’s t-test). Data are from one experiment representative of three (a) or four (b,c) experiments (mean and s.e.m. of duplicates (b,c) or triplicates (a)).

0

2

4

6

8

10

%Input (relative over HPRT)

****AcH3 Il6

0

2

4

6

8

10AcH3K27 Il6

**

0

1

2

3

4 AcH4 Il6

***

0

50000

100000

150000

Relative Expression

Il6

**

0

500

1000

1500

2000

2500

IL-6 Production (pg/mL)

*

0

2

4

6

%Input (relative over HPRT)

AcH3 Il1b

**

0

5

10

15AcH3K27 Il1b

****

0

2

4

6AcH4 Il1b

***

a b c

0

500

1000

1500

2000 Il1b

*

N N+L T T+LT+L+2DG N N+L T T+LT+L+2DG N N+L T T+LT+L+2DG N N+L T T+LT+L+2DG

N N+L T T+LT+L+2DGN N+L T T+LT+L+2DG N N+L T T+LT+L+2DG N N+L T T+LT+L+2DG N N+L T T+LT+L+2DG

Relative Expression

Page 66: Regulation of Inflammatory Macrophage Activation and

52

Of note, such effects of 2DG treatment could be disassociated from global effects on LPS signaling.

NF-κB and IRF3 are targets of the two major branches of LPS signaling and master regulators of

the LPS-inducible transcriptional response. They are robustly activated by LPS signaling in naïve

macrophages and poorly activated in tolerant macrophages, as indicated by IκBα degradation and

resynthesis and IRF3 activating phosphorylation respectively (Fig. 2.11a,b). Importantly, 2DG

treatment during tolerance induction had marginal effects on NF-κB and IRF3 activation in

tolerant macrophages (Fig. 2.11a,b), indicating that the effects of 2DG in recovering Il6 and Il1b

induction may not be due to general perturbation of LPS signaling.

Figure 2.11 Glucose utilization impairs respiration and limits inflammatory gene induction in tolerant macrophages without affecting TLR signaling. BMDMs were left unstimulated (N) or tolerized (T) by 24h LPS treatment +/- 2DG pretreatment followed by washout of inhibitor and analysis of LPS-inducible IκBα degradation (a) and IRF3 phosphorylation (b). Data are from one experiment representative of three experiments.

Rather, our data is more consistent with glucose utilization regulating oxidative metabolism to

influence inflammatory gene induction. Oxidative metabolism first increases and then decreases

over the course of LPS exposure, supporting an LPS-stimulated increased in glucose oxidation and

Ac-CoA production during activation but not during tolerance. This shift in oxidative metabolism

appears to be modulated by glucose utilization, as 2DG treatment ablates both the early increase

and the delayed decrease of respiration. Furthermore, respiration shutdown appears to contribute

to the inability of LPS to stimulate histone acetylation and inflammatory gene induction in tolerant

α-Tubulin

LPS (min):

α-Tubulin

IκBα

IκBα

0 5 10 20 30 60 120180

N

T

a

α-Tubulin

IκBαT+2DG

α-Tubulin

p-IRF30 1 2

N T+2DGTLPS (h): 3 0 1 2 3 0 1 2 3

b

Page 67: Regulation of Inflammatory Macrophage Activation and

53

macrophages, since 2DG treatment restores oxidative metabolism and LPS-inducible histone

acetylation and inflammatory gene induction to tolerant macrophages.

GPD2 activity drives suppression of respiration and inflammatory genes during tolerance

To more rigorously test the role of glucose utilization—and in particular glucose oxidation—in

influencing LPS tolerance induction, we turned to GPD2 KO BMDMs. We found that the delayed

inhibition of oxidative metabolism was attenuated in GPD2 KO BMDMs, indicating that GPD2

activity was necessary for optimal respiration shutdown (Fig. 2.12a). Tolerance-associated defects

in LPS-inducible histone acetylation and inflammatory gene induction were also partially rescued

by GPD2 deficiency. While LPS stimulation poorly increased Il6 and Il1b histone acetylation in

tolerant WT BMDMs, tolerant GPD2 KO BMDMs displayed enhanced LPS-inducible histone

acetylation (Fig. 2.12b). Correlating with the differences in histone acetylation, tolerant GPD2

BMDMs displayed enhanced LPS-inducible expression of Il6 and Il1b (Fig. 2.12c) and production

of IL-6 (Fig. 2.12d), indicating that GPD2 activity was needed for optimal tolerance-associated

suppression of inflammatory responses.

We also asked if GPD2 influences LPS signaling by examining activation of NF-κB and IRF3.

Importantly, GPD2 deficiency led to very modest effects on activation of NF-κB and IRF3 in either

the activation or tolerance conditions (Fig. 2.12e,f). Therefore, GPD2 does not appear to exert

broad effects on LPS signaling to influence inflammatory gene induction during LPS activation

and tolerance.

Page 68: Regulation of Inflammatory Macrophage Activation and

54

Figure 2.12 GPD2 activity drives suppression of inflammatory responses in tolerant macrophages. (a) Basal and maximal respiration after 12h LPS stimulation in WT and GPD2 KO BMDMs. Promoter histone acetylation (b), Il6 and Il1b gene expression (c), and IL-6 cytokine production (d) in WT and GPD2 KO BMDMs in the N, N+L, T, and T+L conditions. Immunoblot analysis of LPS-inducible IκBα degradation (e) and IRF3 phosphorylation (f) in naïve and tolerant WT and GPD2 KO BMDMs. *p ≤0.05, **p ≤0.01, ***p ≤0.001, ****p ≤0.0001 (Student’s t-test). Data are from one experiment representative of three experiments (mean and s.e.m. of duplicates (c,d), triplicates (b), octuplets (a)).

WT GPD2 KO%Input (relative over HPRT)

WT GPD2 KOIL-6 Production (pg/mL)

c db

0

2

4

6

8AcH4 Il6

***

***

NN+LTT+L

0

1000

2000

3000

4000Relative Expression

Il6

**

0

100

200

300

400

500Il1b

**

*

WT GPD2 KO

WT GPD2 KO

0

1000

2000

3000

4000

5000

*

0

5

10

15

%Input (relative over HPRT)

AcH4 Il1b*

**

WT GPD2 KO

Relative Expression

WT GPD2 KO

Max. OCR (pmol/min)

LPSNaïve

0 25 50 75 1000

50

100

150

200

Time (min)

OCR (pmol/min)

GPD2 KO LPSWT LPSGPD2 KO Naïve

FCCP Rot/AAOligo WT Naïve

0

50

100

150

****

0

20

40

60

Basal OCR (pmol/min)

**

WT GPD2 KO

a

α-Tubulin

LPS (min):

α-Tubulin

IκBα

IκBα

α-Tubulin

p-IRF3

0 5 10 20 30 60 120 1800 5 10 20 30 60 120 180Naïve Tolerant

GPD2 KO

WT

0 60 120Naïve

0 60 120 0 60 120 0 60 120NaïveTolerant Tolerant

WT GPD2 KO

e

fLPS (min):

Page 69: Regulation of Inflammatory Macrophage Activation and

55

We considered alternative mechanisms by which GPD2-mediated glucose oxidation could

contribute to suppression of the inflammatory response in tolerant macrophages. Because 2DG

treatment during tolerance induction protected against respiration shutdown and inflammatory

gene suppression, we hypothesized that GPD2-mediated glucose oxidation could regulate a shift

from increased to decreased oxidative metabolism that contributes to a shift from inflammatory

gene induction to suppression. In LPS-stimulated macrophages, multiple factors are thought to

modulate respiration shutdown including induction of iNOS, IRG1, and reverse electron transport

(RET)13, 14, 15, 16. However, GPD2 deficiency had little effect on Nos2 and Irg1 expression during

LPS stimulation (data not shown), therefore we considered a role in induction of RET.

Intriguingly, the etiology and regulation of RET intimated a role in GPD2-mediated regulation of

inflammatory responses. In contrast to forward electron transport (FET), which is the normal

operation of the ETC, RET is a consequence of overwhelming electron influx into the ETC leading

to saturation of electron transport capacity, creating thermodynamic conditions favoring reverse

electron flow32. We considered that GPD2-driven FET would initially boost oxidative metabolism

to support increased Ac-CoA production and histone acetylation during activation; however

sustained GPD2 activity during prolonged LPS exposure would overwhelm forward electron

transport capacity leading to RET, thus limiting TCA cycling, Ac-CoA production, and

inflammatory gene induction (Supplementary Fig. 2.7). To test this hypothesis, we examined

RET using two complementary approaches. RET through CI leads to increased production of

superoxide that is sensitive to the CI inhibitor rotenone, in contrast to FET where rotenone

enhances superoxide production (33, 34 and Fig. 2.13). Another consequence of RET is attenuation

of NADH oxidation to NAD+ at CI (compared to robust oxidation during FET) (32 and Fig. 2.13).

Page 70: Regulation of Inflammatory Macrophage Activation and

56

Figure 2.13 Schematic depicting FET and RET during LPS activation and tolerance. During acute LPS exposure (LPS activation), electrons from oxidation of metabolic substrates (i.e. glucose), flow forward through the ETC (green dotted line), creating a proton motive force for ATP production and also returning electron acceptor molecules (i.e. NAD+) for continued oxidation of metabolic substrates. LPS-induced GPD2 activity initially boosts forward electron transport (FET; left) to support an increase in glucose oxidation for acetyl-CoA (Ac-CoA) synthesis and induction of inflammatory genes by enhanced histone acetylation. However, sustained GPD2 activity may overwhelm the ubiquinone pool (Q), the common sink for electrons from CI, CII, and GPD2, leading to a thermodynamic environment that permits electron backflow (red dotted line). Such reverse electron transport (RET) decreases return of NAD+ molecules to the TCA cycle, impairing glucose oxidation for Ac-CoA synthesis and histone acetylation at inflammatory genes. Therefore, we propose that GPD2 activity turns the ETC into a rheostat for inflammatory gene induction in BMDMs, linking the duration of LPS exposure to the directionality of electron transport to control glucose oxidation and balance induction and suppression of inflammatory responses.

First, we found that while rotenone treatment increased superoxide levels in naïve WT BMDMs

as expected, it decreased superoxide levels in tolerant WT BMDMs indicative of RET after

prolonged LPS exposure (Fig. 2.14a,b). Importantly, GPD2 deficiency diminished the ability of

rotenone to reduce superoxide levels during tolerance, indicating that GPD2 activity contributed

to tolerance-associated RET (Fig. 2.14a,b). The effects of 2DG treatment were even more striking,

allowing rotenone to stimulate superoxide levels indicating complete rescue of tolerance-

associated RET (Fig. 2.14a,b). Second, we used an assay based on detection of NAD(P)H

autofluorescence in real-time, where addition of rotenone acutely blocks CI activity to reveal

NADH oxidation at CI35. A robust increase in NADH levels after rotenone treatment is indicative

!"#$

%%

&'"(')*+,!"#-

"&.

$- $- $- $-

$/0

% 1 %%% %2 2')3(*

4.#/

566,

7(8

,897

:6,

0/;<

!"#$

%%

&'"(')*+,!"#-

.5-"#. "&.

$- $- $- $-

!"0/ $/0

% 1 %%% %2 2')3(*

4.#/

566,

7(8

,897

:6,

4=. #$".

>?"#@ >?"#@

A.B("*35C:35D6E?D7F:7G(H+,*37D6(&7:6IJD73(>?H&@

A.B(&D+,7:6*,EK,C,7I,( H+,*37D6(&7:6IJD73(>KH&@

4=. #$".

!"0/

.5-"#.

Page 71: Regulation of Inflammatory Macrophage Activation and

57

of FET at CI, while a modest increase is indicative of diminished FET and/or enhanced RET, since

the net NADH level is determined by the relative magnitudes of opposing FET and RET.

Figure 2.14 GPD2 activity drives induction of reverse electron transport (RET) in tolerant macrophages. (a,b) WT or GPD2 KO BMDMs were left unstimulated or stimulated 12h with LPS +/- 2DG; in some conditions, rotenone was included in the final 3h of culture (+Rot) to assess sensitivity of mitochondrial superoxide production to CI inhibition. (a), representative experiment showing mitochondrial superoxide production as measured by MitoSox mean fluorescence intensity (MFI); (b), data compiled from three representative experiments, with bar graph representing percent change in MitoSox MFI after Rot treatment relative to no Rot. (c) Real-time NAD(P)H autofluorescence in WT BMDMs unstimulated (Naive) or stimulated with LPS for 12h (LPS) before and after treatment with rotenone (Rot). (d) WT or GPD2 KO BMDMs were left unstimulated or stimulated 12 h with LPS +/- 2DG, followed by analysis of change in NAD(P)H autofluorescence after addition of Rot. *p ≤0.05, **p ≤0.01, ***p ≤0.001, ****p ≤0.0001 (Student’s t-test). Data are from three experiments (b), two experiments (c,d), or one experiment representative of three (a) experiments (mean and s.e.m. of duplicates (a,d) or measurements from ≥100 cells per condition at each time during assay (c)).

We found that in WT macrophages, rotenone differentially modulated NADH levels in naïve and

tolerant conditions, robustly enhancing NADH levels in the naïve state indicative of robust FET

while modestly enhancing NADH levels in the tolerant state indicative of reduced FET and/or

increased RET (Fig. 2.14c,d). Importantly, rotenone did not differentially modulate NADH levels

a

101 102 103 104

MitoSox

Count

WT GPD2 KO

NaïveNaïve+Rot

-100

-50

0

50

100

150

%Change in MitoSox

by Rotenone

LPS+RotNaïve+Rot

WT GPD2KO WT+2DG% Change in NAD(P)H

autofluorescence

after Rotenone

LPS+RotNaïve+Rot

c

101 102 103 104 101 102 103 104

WT+2DG

176

276

271

216

291

410

301

377

447

662

412

625

LPSLPS+Rot

WT GPD2 KO WT+2DG

0

5

10

15

*

*** ns

ns

ns ns

b

Naïve

-200 -150 -100 -50 0 50 1000.9

1.0

1.1

1.2

Mean NAD(P)H

autofluorescence

(relative to no rotenone)

Rot

Time (relative to rotenone, s)

****

LPS

d

Page 72: Regulation of Inflammatory Macrophage Activation and

58

between the naïve and tolerant states in GPD2 KO BMDMs or 2DG-treated WT BMDMs (Fig.

2.14d), indicating that GPD2 deficiency or 2DG treatment mitigated reduction of FET and/or

induction of RET during tolerance. An orthogonal approach to measure NAD+ and NADH levels

also indicated a reduction of the NAD+/NADH ratio during LPS tolerance consistent with RET

induction, and attenuation by GPD2 deficiency indicative of a role for GPD2 in tolerance-

associated RET induction (Supplementary Fig. 2.8). Reversal of RET-induced redox imbalance

by treatment with the NAD+ precursor nicotinamide (NAM) rescued LPS-inducible histone

acetylation at the Il6 and Il1b promoters in tolerant WT BMDMs (Supplementary Fig. 2.9a),

leading to complete rescue of Il6 and Il1b gene induction (Supplementary Fig. 2.9b). Such rescue

by NAM suggests that RET may contribute to suppression of inflammatory gene induction in

tolerant macrophages by driving accumulation of NADH at the expense of NAD+ return to the

TCA cycle, restricting substrate oxidation and thus limiting Ac-CoA production for histone

acetylation and inflammatory gene induction.

Together these findings allow us to propose a model whereby GPS activity uses the ETC as a

rheostat to dynamically influence inflammatory gene induction and suppression. During acute LPS

exposure, increased GPS activity initially increases FET to boost oxidative metabolism and Ac-

CoA production, leading to enhanced histone acetylation and inflammatory gene induction.

However, sustained GPS activity during prolonged LPS exposure eventually overwhelms the ETC

to trigger RET and redox imbalance (i.e., imbalance of the NAD+/NADH ratio), leading to reduced

Ac-CoA production that limits histone acetylation for inflammatory gene induction in tolerant

macrophages.

Page 73: Regulation of Inflammatory Macrophage Activation and

59

GPD2 activity suppresses inflammatory responses during septic shock

Because GPD2 suppresses inflammatory responses in an in vitro model of LPS tolerance (Fig.

2.12c,d), we asked if GPD2 could similarly regulate inflammatory response suppression in vivo.

We used an established model of in vivo LPS tolerance, in which a sub-lethal dose of LPS protects

against septic shock triggered by a subsequent, lethal dose of LPS (Fig. 2.15a). In contrast to mice

that did not receive the sub-lethal LPS pretreatment, mice that received the pretreatment

profoundly suppressed IL-6 production during septic shock, as expected. Importantly, such

suppression was attenuated in GPD2 KO mice leading to enhanced IL-6 levels (Fig. 2.15b). Since

GPD2 is expressed at high levels in only a few tissue and cell types (including brown adipose

tissue (BAT), pancreas, and brain in addition to macrophages)28, 29, increased IL-6 production was

likely due to cell-autonomous GPD2 deficiency in macrophages and related myeloid cells, and

GPD2 deficiency did not affect some clinical parameters such as drop in blood glucose levels and

lymphocyte numbers (Supplementary Fig. 2.10). Importantly, the drop in body temperature

characteristic of septic shock was exaggerated in GPD2 KO mice (Fig. 2.15b). Finally, WT mice

receiving sublethal LPS pretreatment survived subsequent lethal LPS challenge while GPD2 KO

mice succumbed, demonstrating a profound defect in LPS tolerance in the absence of GPD2 (Fig.

2.15c). Collectively the data indicates that GPD2 activity suppresses inflammatory responses to

protect against systemic physiological derangement multiorgan dysfunction and mortality during

septic shock.

Page 74: Regulation of Inflammatory Macrophage Activation and

60

Figure 2.15 GPD2 activity supports LPS tolerance in vivo. (a) In vivo LPS tolerance protocol. WT and GPD2 KO mice were injected with a sublethal dose of LPS (3 mg/kg) or vehicle (saline) followed 24 h later by a lethal dose of LPS (30 mg/kg IP). (b) Serum levels of IL-6 assessed after lethal LPS challenge. p=0.0155 (n=9) and p=0.0237 (n=10) respectively for mice without or with sublethal LPS pretreatment; 2way ANOVA with Sidak’s multiple comparisons test. (c) Body temperature after lethal LPS challenge. (d) Survival of mice after lethal LPS challenge in the in vivo LPS tolerance protocol. p =0.0012; WT n=20, KO n=19. *p ≤0.05, **p ≤0.01 (Student’s t-test). Data are from two experiments (mean and s.d. shown).

a

d

GPD2 KOWT

Time (hours)

Percent Survival

WT GPD2 KO

Serum IL-6 (pg/mL)

Body Temp. (ºC)

Serum IL-6Body TemperatureSurvival

24h Vehicle

24h LPS(Sublethal)

LPS (Lethal)

b c

SublethalLethal

0

100000

200000

300000

400000 **

– ++

– +++ +

WT GPD2 KO

SublethalLethal

– ++

– +++ +

––

0 24 48 720

50

100

20

25

30

35

40

Page 75: Regulation of Inflammatory Macrophage Activation and

61

DISCUSSION

In this study, we make several findings regarding the role of glucose utilization in controlling the

fate of macrophages exposed to LPS. First, although glucose oxidation has been reported in LPS-

stimulated dendritic cells in a limited number of studies14, 36, the paradigm is that LPS-stimulated

macrophages divert glucose to aerobic glycolysis and shut down oxidative metabolism.

Importantly, how such glucose utilization influences macrophage responses to LPS remains not

well understood. Using detailed glucose tracing experiments, we demonstrate that in LPS-activated

macrophages, oxidative metabolism is increased to fuel Ac-CoA production and provide carbon

substrate for histone acetylation at the Il6 and Il1b promoters. Conversely, 2DG treatment

attenuates oxidative metabolism, histone acetylation, and Il6 and Il1b induction. These findings

define a new role for glucose utilization in LPS-stimulated macrophages—namely glucose

oxidation—as well as an underpinning basis by which such glucose utilization drives LPS

activation (Supplementary Fig. 2.7). We also identify the Ac-CoA producing enzyme ACLY as

a regulator of this process, showing that its LPS-inducible activity supports histone acetylation and

inflammatory gene induction during LPS activation.

We further identify GPD2 and the GPS as key and hitherto unknown regulators of glucose

oxidation in LPS-stimulated macrophages. GPD2 is expressed in a tissue-restricted fashion, but is

expressed at high levels in macrophages and further upregulated by LPS stimulation. LPS

stimulation dramatically increases flux through the GPS (Fig. 2.6a). This is likely due to multiple

factors, including increased glucose utilization stimulating substrate delivery to the GPS, as well

as increased GPD2 expression. We also considered the possibility that LPS stimulation could

Page 76: Regulation of Inflammatory Macrophage Activation and

62

increase the intrinsic activity of GPD2 to enhance GPS flux, since GPD2 contains two EF hands

that are thought to be activated by Ca2+ signaling and LPS stimulation mobilizes Ca2+ signaling29,

37. In our experiments, LPS stimulation increased intrinsic GPD2 activity, but only at late time

points (Supplementary Fig. 2.5b). One possibility is that this assay is not robust and/or sensitive

enough to detect a smaller increase in GPD2 activity acutely after LPS stimulation. Alternatively,

LPS-inducible intrinsic GPD2 activity drives GPS activity during tolerance induction rather than

LPS activation, and the latter is driven mainly by increased glucose utilization and substrate flux.

In any case, our data indicates that LPS stimulation mobilizes the GPS to bolster inflammatory

responses during LPS activation, such that GPD2 deficiency attenuates LPS-inducible glucose

oxidation and reduces Ac-CoA production, histone acetylation, and Il6 and Il1b induction. Of note,

the effects of GPD2 deficiency are not quite as pronounced as that of 2DG treatment, which may

reflect the ability of glucose utilization to drive inflammatory gene induction independent of

GPD2-mediated glucose oxidation, and/or compensatory adaptations in the genetic model due to

constitutive loss of GPD2 activity. Complementing our analysis of GPD2 KO BMDMs, acute

inhibition of GPD2 activity with a pharmacological inhibitor recapitulates our key findings in the

genetic model, including effects on respiration, histone acetylation and inflammatory gene

induction (Supplementary Fig. 2.11). Together these findings identify the GPS as a critical

regulator of glucose oxidation and inflammatory gene induction in LPS-activated macrophages.

Given the primacy of glucose utilization in LPS-stimulated macrophages, future studies to examine

the GPS in this context are warranted.

Third, although respiration shutdown is a metabolic hallmark of LPS-stimulated macrophages, its

role if any in modulating macrophage responses to LPS had been unclear4, 5, 38. Here we show that

Page 77: Regulation of Inflammatory Macrophage Activation and

63

respiration shutdown occurs only at delayed time points after LPS stimulation, after the initial

increase in oxidative metabolism. Such shutdown limits the oxidation of glucose (and of other

carbon substrates), thus reducing Ac-CoA production and histone acetylation to reinforce

inflammatory gene suppression in tolerant macrophages. Little is known of the metabolic

underpinnings of LPS tolerance, and we propose that shutdown of oxidative metabolism is central

to metabolic reprogramming during LPS tolerance. The other metabolic hallmark of LPS-

stimulated macrophages, induction of aerobic glycolysis, is presumably a compensatory

mechanism to support cellular bioenergetics (i.e., ATP production) in the face of respiration

shutdown.

Fourth, GPD2 and the GPS are important regulators of tolerance-associated respiration shutdown,

since inhibiting glucose utilization or GPD2 deficiency is sufficient for partial restoration of

oxidative metabolism. This leads to a partial recovery of histone acetylation and inflammatory

gene induction in tolerant macrophages, as well as increased IL-6 production and mortality in a

mouse model of septic shock. We propose a model in which GPS activity influences oxidative

metabolism to modulate a shift from inflammatory response induction to suppression. During acute

LPS exposure, GPS activity drives FET and Ac-CoA production to enhance histone acetylation

and inflammatory gene induction, but during prolonged LPS exposure, sustained GPS activity

eventually saturates electron transport capacity to trigger RET and limit Ac-CoA production for

histone acetylation and inflammatory gene induction (Supplementary Fig. 2.12). In support of

this model, we found that rotenone-sensitive superoxide and NADH production increases in the

tolerant state indicative of RET induction, and that these events are impaired in GPD2 deficiency

and upon 2DG treatment. Our findings are the first to describe a physiological role for GPD2-

Page 78: Regulation of Inflammatory Macrophage Activation and

64

driven RET, since GPD2-mediated oxidation of G3P has been shown to induce RET only in

isolated mitochondria.

Note that in our model for GPS-mediated regulation of inflammatory responses, GPS activity

triggers RET as a consequence of its initial boost of FET. Such GPS activity allows the GPS to

integrate the duration of LPS exposure with ETC activity, supporting inflammatory response

induction after acute microbial exposure but contributing to inflammatory response suppression

after prolonged microbial exposure (Supplementary Fig. 2.12). Inflammation confers host

defense at the expense of tissue damage and must be carefully regulated to balance its beneficial

and detrimental effects, and GPS activity, by integrating the extent of microbial exposure to

inflammatory response induction versus suppression, may provide one such balancing mechanism.

As discussed above, our detailed metabolic analyses implicate histone acetylation at inflammatory

genes as a key target by which GPD2 regulates inflammatory gene expression. Complementing

the metabolic analyses, our analysis of LPS signaling indicates no broad effects of 2DG treatment

or GPD2 deficiency on NF-kB and IRF3 activation; furthermore, our RNA-seq analysis of LPS-

activated WT and GPD2 KO BMDMs indicates that GPD2 regulates a small subset of LPS-

inducible genes (data not shown). In this regard two points are worthy of discussion. First, while

GPD2 does not globally modulate LPS signaling or LPS activation, GPD2 activity could impinge

on specific aspects of LPS signaling that contribute to its regulation of inflammatory gene

expression. For example, GPD2-mediated glucose oxidation could influence the expression or

activity of specific signaling components or transcription factors, via effects on promoter histone

acetylation and protein acetylation respectively. Such a role for GPD2 would be non-exclusive

Page 79: Regulation of Inflammatory Macrophage Activation and

65

with the one that we propose here—modulation of histone acetylation at inflammatory genes—

and could be investigated in future studies. Second, how GPD2-mediated glucose oxidation and

Ac-CoA production can specifically modulate histone acetylation and the expression of a subset

of LPS-inducible genes remains unclear. One possibility is that ACLY is recruited to chromatin,

producing a local pool of Ac-CoA that regulates histone acetylation in a gene-specific manner. In

support of this possibility, ACLY can be found in the nucleus39, and precedence for such gene-

specific association and regulation has been established by the metabolic machinery that produces

S-adenosylmethionine, the substrate for histone and DNA methylation40.

One paradigm that has emerged from the macrophage metabolism field is that LPS-stimulated

macrophages shut down oxidative metabolism while M2 macrophages increase oxidative

metabolism. Our findings indicate that increased oxidative metabolism is a common feature of

LPS-activated and M2 macrophages, and that one consequence of this process is provision of A-

CoA for histone acetylation at inducible genes. While a role for histone acetylation in bolstering

gene induction has long been appreciated, a seminal study showed that in tumor cells, Ac-CoA

availability can be limiting for histone acetylation thus influencing gene expression39. A similar

principle appears to be true in macrophages, thus allowing levels of carbon substrates like glucose

to modulate the transcriptional basis of macrophage activation. Furthermore, because Ac-CoA

levels can be limiting, polarizing signals like LPS and IL-4 mobilize oxidative metabolism, by

impinging on key regulators like GPD2, to drive optimal gene induction and macrophage

activation. In the case of LPS-stimulated macrophages, this process is further exploited by LPS

signaling to eventually influence respiration shutdown and inflammatory gene suppression.

Page 80: Regulation of Inflammatory Macrophage Activation and

66

It is noteworthy that LPS-stimulated macrophages upregulate flux through the GPS by increasing

GPD2 expression and activity, given that macrophages can deploy the malate/aspartate shuttle to

maintain redox balance during glucose oxidation41. We propose that GPD2/GPS-mediated glucose

oxidation has a unique, non-redundant role in optimal control of inflammatory gene induction and

suppression. In particular, because GPD2 is situated in the ETC directly bridging cytosolic glucose

metabolism with ETC flux, GPS activity may allow for very robust glucose oxidation. Indeed,

available evidence in the literature30 and the rapid kinetics of our glucose labeling experiments

support this notion. We propose that in LPS-stimulated macrophages, rapid and robust

mobilization of glucose oxidation drives ETC flux and inflammatory gene expression, and in the

event of sustained LPS exposure, contributes to RET induction. In contrast, the malate-aspartate

shuttle may support more modest rates of glucose oxidation that does not allow for a comparable

increase in ETC activity, and would be fundamentally incompatible with RET induction (given

that NADH derived from this shuttle is delivered to Complex I). Therefore, deployment of the

GPS in the high metabolic state of LPS-stimulated macrophages seems to enable appropriate

modulation of inflammatory response induction and suppression. We propose that the GPS is a

hitherto unappreciated and unique regulator of LPS responses, using the ETC as a rheostat to

couple duration of LPS exposure to ETC directionality to modulate a balance between the

induction and suppression of inflammatory responses.

Page 81: Regulation of Inflammatory Macrophage Activation and

67

MATERIALS AND METHODS

BMDM culture. BMDMs from male and female mice were differentiated and cultured as

described previously42. For LPS activation (0-3h post-LPS), BMDMs were stimulated with 100

ng/mL LPS (LPS Ultra-Pure, Sigma-Aldrich). All of the metabolic parameters associated with

LPS activation, including glucose uptake and oxidation, glucose tracing into TCA intermediates

and Ac-CoA, ACLY phosphorylation, glucose labeling into histones, and promoter histone

acetylation, were assessed at early time points after LPS stimulation (0-3 h), while the downstream

consequences of such metabolic events, i.e., inflammatory gene induction and cytokine production,

were assessed at 3 h after LPS stimulation. For LPS tolerance (12-28h post-LPS), BMDMs were

challenged with 100 ng/mL LPS for 24h to induce tolerance (T), washed, and then rechallenged

with 10 ng/mL LPS (T+L) for the desired amount of time. In such experiments, naïve BMDMs

were incubated in media for 24h (N), washed, and challenged with 10 ng/mL LPS (N+L) to provide

responsive-cell controls. Pharmacological inhibition of ACLY (ACLYi, 80-160 µM; SB-204990,

Tocris, United Kingdom), glucose utilization (2DG, 5-10 mM; 2-deoxy-glucose, Sigma-Aldrich),

and GPD2 (GPD2i, 300 µM; iGP-1, Calbiochem) was achieved by pretreating BMDMs for 1h

prior to LPS activation. For rescue of tolerance, BMDMs were pretreated for 1h prior to activation,

but drugs were not added back after wash before restimulation. Additional drugs include the ACLY

inhibitors BMS-303141 and Medica16, the p300 inhibitor C646, and the glutamine and fatty acid

oxidation inhibitors BPTES and Etomoxir.

Page 82: Regulation of Inflammatory Macrophage Activation and

68

Seahorse assays. Basal and LPS-induced changes in oxygen consumption (OCR) and extracellular

acidification (ECAR) rates were measured with a Seahorse XF96 Extracellular Flux Analyzer

(Agilent). For experiments in intact cells, assays were performed in Seahorse XF Assay Medium

supplemented with glucose (11 mM) and adjusted to pH 7.4 with NaOH. Mix-wait-measure

durations were set to 3, 2, and 3 minutes, respectively. For the MitoStress Test, oligomycin, FCCP,

and rotenone/antimycin A were sequentially injected to achieve final concentrations of 1, 1.5, and

2 µM. Glycolysis Tests were performed in unsupplemented XF Assay Medium, and glucose,

oligomycin, and 2DG were sequentially injected to achieve final concentrations of 11 mM, 1 µM,

and 500 mM, respectively. Non-mitochondrial OCR was subtracted from the mean pre-oligomycin

OCR and the mean post-FCCP OCR to calculate Basal and Maximal mitochondrial respiration,

respectively. For experiments in permeabilized cells43, assays were performed in mannose sucrose

buffer supplemented with 0.1% BSA and adjusted to pH 7.2 with KOH (MAS-BSA Buffer). Cells

were permeabilized by injection of perfringolysin O (PMP reagent, Agilent) in the XF96 Analyzer,

and specific mitochondrial enzyme activities were determined by measuring State 3 respiration in

the presence of substrate/inhibitors/ADP, followed by sequential injection of oligomycin and

antimycin A. GPD2 activity was measured as the change in OCR induced by injection of 10 mM

sn-glycerol 3-phosphate (G3P, Sigma-Aldrich), 1 mM ADP, and 2 uM Rotenone. After each

experiment, OCR and ECAR values were normalized to DNA content in each well by Hoechst

33342 staining (Life Technologies, Carlsbad, CA).

Page 83: Regulation of Inflammatory Macrophage Activation and

69

Steady-state metabolomics and relative metabolic flux. For steady-state analysis of metabolic

networks, BMDMs were stimulated for the desired period of time and then lysed and polar

metabolites extracted using 80:20 methanol:water (HPLC-grade, Sigma-Aldrich). Extracts were

dried under nitrogen gas followed by LC-MS (Beth Israel Deaconess Medical Center Mass

Spectrometry Facility) to monitor targeted metabolites from a library of 300 species, including

those involved in glycolysis, the glycerol 3-phosphate shuttle, and the TCA cycle. MS peak areas

were normalized to median and sample protein concentrations, and LPS-induced fold-change data

was run through the Metaboanalyst 2.0 web-based pipeline to visualize metabolic pathway

enrichment. For relative metabolic flux analysis, naïve and LPS-stimulated BMDMs were washed

with glucose-free medium and then labeled with 2 g/L 13C6-gluose (Cambridge Isotope

Laboratories) in complete medium (10% dialyzed FBS) for 30 min. Metabolites were extracted,

dried, and run as in44. Normalized peak data was analyzed for percent labeling of glucose into each

possible isotopologue of species in glycolysis, the glycerol 3-phosphate shuttle, and the TCA cycle.

Acyl-CoA mass spectrometry. Naïve and LPS-stimulated BMDMs were washed with glucose-

free medium and then labeled with 2 g/L 13C6-gluose (Cambridge Isotope Laboratories) in

complete medium (10% dialyzed FBS) for 30 min. Cells were then lysed in 800 uL ice-cold 10%

trichloroacetic acid and extracted as previously described45. Analysis by LC-high resolution MS

was conducted on an Ultimate 3000 UHPLC coupled to a Q Extractive Plus operating in the

positive ion mode as previously described46. Isotopologue analysis was conducted using FluxFix47.

Page 84: Regulation of Inflammatory Macrophage Activation and

70

Glucose tracing into histones. Glucose tracing into histones was done following an established

protocol48. In the last 1.5h of stimulation, BMDMs were labeled with 14C6-glucose followed by

acid extraction of acetylated histones using a method previously described27. Radiolabel

incorporation was assessed by scintillation counting and normalized to total protein concentration.

Glucose uptake. Basal and LPS-inducible glucose uptake in BMDMs was measured using 3H-

deoxy-D-glucose (2-3H[G]) (PerkinElmer) in KRBH buffer as described previously9. Counts per

minute (cpm) values were normalized to protein concentration in each sample.

NAD+/NADH analyses. For NADH autofluorescence, BMDMs were excited at 340/26 nm on a

Nikon Eclipse Ti-S microscope. The emitted signal was collected using a 460/80 nm bandpass

filter. Autofluorescence was analyzed with Fiji image processing software.�Measurements of the

NAD+/NADH ratio were done using a colorimetric kit (Biovision) according to the manufacturer's

protocol.

Chromatin immunoprecipitation. Acetylation at inflammatory gene promoters was measured by

chromatin immunoprecipitation (ChIP) assay as described previously9. Promoters are defined as

the region between -500 bp and the transcription start site. Antibodies used were as follows:

acetylated H3 (Millipore 06–599), acetylated H3K27 (Abcam ab4729), acetylated H4 (Millipore

06–866), and IgG (Santa Cruz, Dallas, TX, SC-2027). Fold enrichment was calculated as ChIP

signals normalized to input.

Page 85: Regulation of Inflammatory Macrophage Activation and

71

Gene expression. RNA was isolated using RNA-Bee (Tel-Test) according to the manufacturer’s

protocol. cDNA synthesis was performed using a High Capacity cDNA Reverse Transcription Kit

(Applied Biosystems). qPCRs were run on a Bio-Rad C1000 Thermalcycler, and expression of

target genes was calculated by the ddCT method using CFX Manger Software (Bio-Rad). Relative

expression values for target genes were calculated by normalization to expression of hypoxanthine

phosphoribosyltransferase (HPRT).

Generation of GPD2 KO mice. GPD2 KO mice were generated on the C57BL/6J background

using a gRNA targeting exon 6 of mGPD2 (oligo 1: CACCGCGTACCGTCATAGTAGACAA,

oligo 2: CGCATGGCAGTATCATCTGTTCAAA). gRNA and Cas9 RNA were injected into eggs

followed by implantation into recipient female mice at the BWH Transgenic Core. Mice were

backcrossed two times to C57BL/6 mice. BMDMs from two independently generated lines were

used with similar results.

Mice. Wild-type C57BL/6J mice (The Jackson Laboratory) and whole-body Gpd2-/- (KO) mice

were maintained under specific pathogen-free conditions at the Harvard T.H. Chan School of

Public Health in accordance with Institutional Animal Care and Use Committee (IACUC)

guidelines. Male and female mice were used for in vivo LPS tolerance experiments and as a source

of bone marrow for BMDM culture between 6 and 12 weeks of age. All protocols were approved

by the IACUC of Harvard Medical School. In the in vivo LPS tolerance experiments, WT and

GPD2 KO mice were injected with vehicle (saline) or 3 mg/kg LPS via intraperitoneal injection

(IP) followed by 30 mg/kg LPS IP 24h later. Serum IL-6 and internal body temperature (TH-5

Thermalert Clinical Monitoring Thermometer, Physitemp) were measured for 6 h after lethal LPS

Page 86: Regulation of Inflammatory Macrophage Activation and

72

challenge. Survival of WT and GPD2 KO mice that received both challenges was recorded as an

indication of differential endotoxin tolerance in vivo. Blood glucose and lactate levels in septic

mice were measured using standard glucose and lactate meters (Clarity BG1000, Clarity

Diagnostics; Lactate Plus, Nova Biomedical). Numbers of circulating lymphocytes and monocytes

were obtained using a Hemavet 950 (Drew Scientific).

Immunoblot and enzyme-linked immunosorbent assays. For immunoblotting, BMDMs were

lysed in 1% NP-40 buffer or RIPA buffer, and protein concentration was determined using the

Bradford method or micro BCA assay (Pierce). Primary antibodies used were anti-phospho-

ACLY(S455) and anti-ACLY (1:1000, Cell Signaling), anti-GPD2 (1:800, ProteinTech), and anti-

a-Tubulin (1:5000, Sigma-Aldrich). For ELISA, cell culture supernatants and mouse whole blood

were spun down (5 min at 400 g and 20 min at 13,000 g, respectively), diluted appropriately in 5%

BSA, and assayed for IL-6 using BioLegend Standard ELISA kits according to the manufacturer’s

protocols. For in vitro LPS tolerance experiments, cell culture supernatants in the T and T+L

conditions were collected after LPS washout and rest or restimulation for 4h, allowing for

measurement of IL-6 produced during tolerance rather than activation.

Flow cytometry. Efficiency of macrophage differentiation from bone-marrow-derived

progenitors was determined by staining D7 BMDMs with anti-F4/80 (BM8) and anti-CD11b

(M1/70) antibodies (BD Bioscience) for 30 min at 4°C in FACS buffer (2% FCS), followed by

washes and fixation in 2% ultrapure paraformaldehyde. Mitochondrial superoxide was measured

by MitoSox (Life Technologies) staining according to the manufacturer's protocol. Briefly,

BMDMs were left unstimulated or stimulated with LPS for the indicated times and then labeled

Page 87: Regulation of Inflammatory Macrophage Activation and

73

with 2.5 uM MitoSox Red in the final 15 minutes of treatment, followed by washout and no

fixation. Mean fluorescence intensity of MitoSox in the PE channel was measured to determine

levels of mitochondrial superoxide at the time of labeling. All samples were acquired on a BD

FACS Fortessa flow cytometer, and FlowJo software (v10.4) was used to analyze data.

Statistical analysis. Statistical significance was determined by a two-tailed Student’s t-test or

2way ANOVA with Sidak’s multiple comparisons test using Prism software (GraphPad). *, **,

***, and **** in figures denote p value ≤0.05, 0.01, 0.001, and 0.0001, respectively.

Acknowledgments

J.J. was supported by the British Society for Immunology and T.H. by NIH grant R01AI102964.

N.W.S. was supported by NIH R03 HD092630, X.G. by the Canadian Institutes of Health

Research (CIHR) reference number 146818, and E.T.C. by the Claudia Adams Barr Program. We

thank Dr. J.F. Mohan for helpful suggestions for Crispr-Cas9 genome editing, Dr. L.B. Sullivan

for technical advice on the GPD2 Seahorse assay, and Dr. R.L.S. Goncalves for technical advice

on the NAD(P)H autofluorescence assay.

Competing Interests

The authors declare no competing interests.

Page 88: Regulation of Inflammatory Macrophage Activation and

74

Supplementary Figure 2.1 ATP-citrate lyase (ACLY) and p300 activity are necessary for supporting pro-inflammatory gene induction during LPS activation in BMDMs. Il6 and Il1b promoter region histone acetylation (a,d) and gene expression (b,c) in BMDMs unstimulated (Naive) or stimulated with LPS (LPS) in the presence or absence of the ACLY inhibitors BMS 303141 (LPS+BMS) or Medica16 (LPS+Medica16) or the p300 inhibitor C646. *p ≤0.05, **p ≤0.01, ***p ≤0.001, ****p ≤0.0001 (Student’s t-test). Data are from one experiment representative of three experiments (mean and s.e.m. of duplicates (b,c) or triplicates (a,d)).

a

0.0

0.7

1.4

2.1

2.8

3.5

%Input (relative over HPRT) AcH4 Il6

****

0

5

10

15

20 AcH4 Il1b NaïveLPSLPS+BMSLPS+Medica16

********

0.0

0.2

0.4

0.6

0.8

1.0

%Input

AcH3 Il6

******

0

2

4

6

8

10 AcH3K27 Il6

********

0

2

4

6 AcH4 Il6

******

0.0

0.1

0.2

0.3

0.4

%Input

AcH3 Il1b

***

0

2

4

6

8 AcH3K27 Il1b

****

0.0

0.5

1.0

1.5

2.0

2.5 AcH4 Il1b

****

0

200

400

600

800

1000

Relative Expression

Il6*

0

50

100

150

200

Relative Expression

Il1b *

LPS+p300iLPSNaïve

0

100

200

300

400

Relative Expression

Il6***

0

200

400

600

800

1000Il1b

**

b

c d

Page 89: Regulation of Inflammatory Macrophage Activation and

75

Supplementary Figure 2.2 LPS exposure modulates mitochondrial respiration and glucose uptake as a function of time. (a) Oxygen consumption rate (OCR) was measured using Seahorse Extracellular Flux Analysis on BMDMs unstimulated (Naive) or stimulated with LPS for the indicated times. Maximal OCR was calculated as the mean uncoupled OCR after FCCP injection, minus the non-mitochondrial OCR remaining after rotenone/antimycin A (Rot/AA) injection. Acute LPS exposure induces a transient but highly-reproducible increase in maximal respiration, followed by a gradual and sustained shutdown during prolonged LPS exposure. This dynamic change in respiration in BMDMs stimulated with LPS traces the metabolic shift underpinning the transition from activation to tolerance. (b) Glucose uptake in naïve and LPS-activated BMDMs measured by 3H-deoxy-D-glucose uptake (cpm = counts/min). *p ≤0.05, **p ≤0.01, ***p ≤0.001, ****p ≤0.0001 (Student’s t-test). Data are from one experiment representative of three (b) or ten (a) experiments (mean and s.e.m. of duplicates (b) or octuplets (a)).

Supplementary Figure 2.3 Glucose starvation attenuates pro-inflammatory gene induction and promoter region histone acetylation in LPS-activated BMDMs. Expression (a) and promoter region histone acetylation (b) of Il6 and Il1b in naïve and LPS-activated BMDMs pre-incubated in glucose-free media for 24h. *p ≤0.05, **p ≤0.01, ***p ≤0.001 (Student’s t-test). Data are from one experiment representative of three experiments (mean and s.e.m. of duplicates (a) or triplicates (b)).

Max. OCR (pmol/min) Naïve

LPS 1hLPS 12h

0

50

100

150

200

250 ********

0

10

20

30

40

50

3H cpm/mg protein

****

*******

LPS 0.5hLPS 1hLPS 2h

Naïve

a b

LPSNaïve

0

2000

4000

6000

Relative Expression

Il6

0

200

400

600

Relative Expression

Il1b

+Glucose -Glucose0

2

4

6

8

%Input (relative over HPRT) AcH3 Il6

**

+Glucose -Glucose0

1

2

3AcH3K27 Il6

**

+Glucose -Glucose0

1

2

3

4 AcH4 Il6

*

+Glucose -Glucose0

1

2

3

4

5

%Input (relative over HPRT) AcH3 Il1b

**

+Glucose -Glucose0

2

4

6 AcH3K27 Il1b**

+Glucose -Glucose0

2

4

6

8 AcH4 Il1b***

+Glucose -Glucose

+Glucose -Glucose

a b

Page 90: Regulation of Inflammatory Macrophage Activation and

76

Supplementary Figure 2.4 Depiction of the spatial and biochemical position of GPD2 at the nexus between glycolysis and electron transport. Glyceraldehyde 3-phosphate dehydrogenase (GAPDH) generates NADH from oxidation of glucose in the cytosol, supplying electrons for reduction of dihydroxyacetone phosphate (DHAP) to glycerol 3-phosphate (G3P) by cytosolic glycerol 3-phosphate dehydrogenase (GPD1). Electrons from G3P are then passed directly to the electron transport chain (ETC) through the FAD cofactor of mitochondrial glycerol 3-phosphate dehydrogenase (GPD2), reducing ubiquinone (Q) to ubiquinol (QH2) and driving the proton motive force through the remaining parts of the ETC to fuel oxidative phosphorylation. As the glycerol 3-phosphate shuttle (GAPDH-GPD1-GPD2) directly links glycolysis to the ETC and resupplies NAD+ to the former, increased GPD2 activity serves as a metabolic node by which the ETC may regulate the rate of glucose utilization. LPS activation increases GPD2 expression, permitting the acquisition of a high metabolic state characterized by increased glucose oxidation and G3P-drive mitochondrial respiration.

matrix

FADH2 FAD

Q QH2

GPD2IMS

cytosol

ELECTRON TRANSPORT

III

DHAP

GLYCOLYSIS

G3PGPD1

NADH NAD+

1,3-bisphosphoglycerateGAPDH

Glyceraldehyde3-phosphate

GLYCEROLPHOSPHATESHUTTLE

Page 91: Regulation of Inflammatory Macrophage Activation and

77

Supplementary Figure 2.5 LPS activation increases expression and activity of GPD2 in BMDMs. (a) Expression and (b) activity of GPD2 in naïve and LPS-activated WT and GPD2 KO BMDMs, measured by Western blot and Seahorse. Activity was calculated as the oxygen consumption rate (OCR) in permeabilized BMDMs after injection of G3P, ADP, and rotenone relative to naïve.

0

50

100

150

200

250

Relative OCR (%)

LPS 6hLPS 12h

**

*

GPD2Naïve

GPD2

α-Tubulin

LPS (h): 0 1.5 3 6 9

a b

Page 92: Regulation of Inflammatory Macrophage Activation and

78

Supplementary Figure 2.6 GPD2 regulates glucose flux through glycolysis and the TCA cycle. 13C6-glucose tracing into intermediates of glycolysis and TCA cycle in WT and GPD2 KO BMDMs during LPS activation, presented as LPS-induced fold change in percent m+6 (1,2), m+3 (3-7), or m+2 (8-11) enrichment. Data are from one experiment representative of two experiments (mean and s.e.m. of triplicates).

LPS 1.5hLPS 3h

Naïve

Fold Change %Enrichment

WT GPD2 KO0

1

2

3

4 *******

0

10

20

30

40

50 ********

0

5

10

15 ***

0

2

4

6

8 ********

WT GPD2 KOWT GPD2 KO

WT GPD2 KO

(1)

WT GPD2 KO0.0

0.5

1.0

1.5

2.0

2.5

Fold Change %Enrichment

WT GPD2 KO0

2

4

6

8

WT GPD2 KO0

1

2

3

4

5

WT GPD2 KO0

1

2

3

4

WT GPD2 KO0

2

4

6

WT GPD2 KO0

1

2

3

4

5

Fructose6-phosphate (1)

Fructose1,6-bisphosphate (2)

Glucose6-phosphate (1)

Glucose

Glyeraldehyde3-phosphate (3)

3-Phosphoglycerate (5)

Pyruvate (6) Lactate (7)

Citrate (8)

Fumarate (10)

Malate (11)

Succinate (9)

Oxaloacetate

Isocitrate (8)

Fold Change %Enrichment (2)

Fold Change %Enrichment (5)Fold Change %Enrichment (3)

Dihydroxyacetonephosphate (4)

Fold Change %Enrichment (6)

Fold Change %Enrichment (7)

(8)

Fold Change %Enrichment (9)

Fold Change %Enrichment (10)

Fold Change %Enrichment (11) Acetyl-CoA

WT GPD2 KO

Fold Change %Enrichment (4)

0

1

2

3

****

**p=0.06

*****

Page 93: Regulation of Inflammatory Macrophage Activation and

79

Supplementary Figure 2.7 Model integrating GPD2-regulated mitochondrial oxidative metabolism with LPS activation. GPD2-dependent transport of electrons from the glycerol phosphate shuttle (GPS) through the ETC is upregulated by LPS activation to support an increase in glucose-derived carbon flux through glycolysis and the TCA cycle, providing increased availability of citrate and Ac-CoA. Coupled with an LPS-induced increase in ACLY activity, enhanced production of Ac-CoA drives histone acetylation and induction of inflammatory genes like Il6 and Il1b. As oxidation of carbon substrates in the TCA cycle depends on the availability of electron acceptors such as NAD+, the cyclic reduction and oxidation of these molecules serves as an important link between oxidative metabolism in the TCA cycle and electron transport chain (ETC) activity.

Supplementary Figure 2.8 Prolonged LPS stimulation disrupts forward electron transport, leading to a shift in mitochondrial redox balance. NAD+/NADH ratio in WT and GPD2 KO BMDMs unstimulated (Naïve) or stimulated with LPS for 12h (LPS). Data are from one experiment representative of three experiments (mean and s.e.m. of duplicates).

!"#$%&'

!"($%"(&)&

*+,

!"#$%

!!! !& &

-$.,%-

+,-

/)&0%1'2-$'0("30)%1

456

'($)

!"($'7%"8.9:%&9:30'6:#00"'

-,4;!!

'($*

WT GPD2 KO0.0

0.2

0.4

0.6

0.8

1.0

NAD+ /NADH ratio

**

*****

LPSNaïve

Page 94: Regulation of Inflammatory Macrophage Activation and

80

Supplementary Figure 2.9 Nicotinamide (NAM) reverses RET-redox imbalance to rescue pro-inflammatory gene induction and promoter region histone acetylation in tolerant BMDMs. Promoter region histone acetylation (a) and expression (b) of Il6 and Il1b in naïve (N), LPS activated (N+L), tolerant (T), and tolerant rechallenged BMDMs +/- NAM during tolerization (T+L+NAM). *p ≤0.05, **p ≤0.01, ***p ≤0.001 (Student’s t-test). Data are from one experiment representative of three experiments (mean and s.e.m. of duplicates (b) or triplicates (a)).

Supplementary Figure 2.10 GPD2 deficiency does not affect some clinical parameters altered during sepsis. Blood glucose (a) and lactate (b), which decreased and modestly increased during sepsis, were not dependent on the presence of GPD2. WT and GPD2 KO mice experienced equal changes in levels of circulating lymphocytes (c) and monocytes (d) during sepsis.

0

20000

40000

60000

80000

100000 ****

Relative Expression

Il6NN+LTT+LT+L+NAM

0

500

1000

1500 Il1b

*

0

1

2

3

4

5

0

2

4

6

8 H4Ac Il1bH4Ac Il6

%Input (relative over HPRT)

********

baBlood Glucose (mg/dL)

Blood Lactate (mM)

# Lymphocytes (x10

3 /μL)

# Monocytes (x10

3 /μL)

0

100

200

300

0

2

4

6

8

0

1

2

3

4

0.0

0.1

0.2

0.3

WT GPD2 KO

SublethalLethal

– ++

– +++ +

––

WT GPD2 KO

– ++

– +++ +

––

WT GPD2 KO

– ++

– +++ +

––

WT GPD2 KO

– ++

– +++ +

––

SublethalLethal

a b

c d

Page 95: Regulation of Inflammatory Macrophage Activation and

81

Supplementary Figure 2.11 Pharmacological inhibition of GPD2 modulates LPS activation and tolerance. Inhibiting the increase in GPD2 activity induced by LPS activation (0-3h LPS) using iGP-1 (GPD2i) attenuated enhanced respiration (a), histone acetylation at Il6 and Il1b promoters (b), and expression of Il6 and Il1b (c). GPD2 inhibition partially protected against respiration shutdown during prolonged LPS exposure (12-24h LPS) (d). Cotreatment with GPD2i during induction of tolerance, followed by restimulation in the absence of inhibitor (T+L+GPD2i) demonstrated rescue of promoter histone acetylation (e) and expression of Il6 and Il1b (f).

0

2

4

6

8 ***AcH3 Il6

0

1

2

3

4AcH3K27 Il6

**

0.0

0.5

1.0

1.5

2.0

2.5AcH4 Il6

***

0

1

2

3

4

%Input (relative over HPRT) AcH3 Il1b

**

0

1

2

3

4

5AcH3K27 Il1b

***

0

1

2

3

4

5AcH4 Il1b

***

0

1000

2000

3000

4000

5000Il1b

**

0

5000

10000

15000

20000

Relative Expression

Il6

***

0

50

100

150 Il6

*

0

50

100

150

Relative Expression

Il1b

*

0

50

100

150

200

Max. OCR (pmol/min) ** ***

0

100

200

300

400

Max. OCR (pmol/min)

***

***

0.0

0.5

1.0

1.5

2.0

2.5

%Input

*******AcH3 Il6

0

3

6

9AcH3K27 Il6

***

0

2

4

6

8AcH4 Il6

***

0

5

10

15

%Input

AcH3 Il1b******

0

6

12

18AcH3K27 Il1b

****

0.0

0.5

1.0

AcH4 Il1b

*

Relative Expression

LPSLPS+GPD2i

Naïve

N N+L T T+LT+L+GPD2i%Input (relative over HPRT)

LPS LPS+GPD2iNaïve

N N+L T T+LT+L+GPD2i N N+L T T+LT+L+GPD2i

N N+L T T+LT+L+GPD2i N N+L T T+LT+L+GPD2i N N+L T T+LT+L+GPD2i

N N+L T T+LT+L+GPD2iN N+L T T+LT+L+GPD2i

LPS+GPD2iLPS

a b c

d e

f

Page 96: Regulation of Inflammatory Macrophage Activation and

82

Supplementary Figure 2.12 Model of the metabolic adaptations underpinning the biphasic functional response to LPS exposure. Duration of LPS exposure regulates the transition from inflammatory gene induction to suppression (blue histogram) in macrophages. Acute LPS exposure, or activation, induces an increase in GPD2-dependent oxidative metabolism, which enhances the production of Ac-CoA from glucose to fuel histone acetylation and thus promote expression of the inflammatory genes Il6 and Il1b. Prolonged LPS exposure produces a state of tolerance in which induction Il6 and Il1b expression is attenuated due to decreased respiration limiting flux of glucose through the TCA cycle into Ac-CoA for histone acetylation.

!"#$%!

&'()*+%,-%.-/01-234%5'(6

7+5*-!"8$!

9:'"%56

;,.:)

<<)*%(=->654%

,56

),?->654+()*+%

,

!"#$%!

7+5*-!"8$!

9:'"%56

!$8;@!8;AB 8A/2>!B$2

90&CD28$90&CD28$

Page 97: Regulation of Inflammatory Macrophage Activation and

83

REFERENCES

1. Foster, S.L. & Medzhitov, R. Gene-specific control of the TLR-induced inflammatoryresponse.ClinImmunol130,7-15(2009).

2. Seeley,J.J.&Ghosh,S.MolecularmechanismsofinnatememoryandtolerancetoLPS.J

LeukocBiol101,107-119(2017).3. Hotchkiss, R.S., Monneret, G. & Payen, D. Sepsis-induced immunosuppression: from

cellulardysfunctionstoimmunotherapy.NatRevImmunol13,862-874(2013).4. O'Neill, L.A.&Pearce, E.J. Immunometabolismgovernsdendritic cell andmacrophage

function.JExpMed213,15-23(2016).5. Langston, P.K., Shibata,M.&Horng, T.Metabolism SupportsMacrophage Activation.

FrontImmunol8,61(2017).6. Jung,J.,Zeng,H.&Horng,T.Metabolismasaguidingforceforimmunity.NatCellBiol21,

85-93(2019).7. Vats, D. et al. Oxidativemetabolism and PGC-1beta attenuatemacrophage-mediated

inflammation.CellMetab4,13-24(2006).8. Lacy-Hulbert, A. & Moore, K.J. Designer macrophages: oxidative metabolism fuels

inflammationrepair.CellMetab4,7-8(2006).9. Covarrubias,A.J.etal.Akt-mTORC1signalingregulatesAclytointegratemetabolicinput

tocontrolofmacrophageactivation.Elife5(2016).10. Liu,P.S.etal.α-ketoglutarateorchestratesmacrophageactivationthroughmetabolicand

epigeneticreprogramming.NatImmunol18,985-994(2017).11. Weinberg,F.&Chandel,N.S.Mitochondrialmetabolismandcancer.AnnNYAcadSci

1177,66-73(2009).12. Hard,G.C.Somebiochemicalaspectsoftheimmunemacrophage.BrJExpPathol51,97-

105(1970).13. Drapier,J.C.&Hibbs,J.B.Differentiationofmurinemacrophagestoexpressnonspecific

cytotoxicity for tumor cells results in L-arginine-dependent inhibitionofmitochondrialiron-sulfurenzymesinthemacrophageeffectorcells.JImmunol140,2829-2838(1988).

14. Everts,B.etal.CommitmenttoglycolysissustainssurvivalofNO-producinginflammatory

dendriticcells.Blood120,1422-1431(2012).

Page 98: Regulation of Inflammatory Macrophage Activation and

84

15. Lampropoulou, V. et al. Itaconate Links Inhibition of Succinate Dehydrogenase withMacrophageMetabolicRemodelingandRegulationofInflammation.CellMetab24,158-166(2016).

16. Mills, E.L. et al. Succinate Dehydrogenase Supports Metabolic Repurposing of

MitochondriatoDriveInflammatoryMacrophages.Cell167,457-470.e413(2016).17. Cheng, S.C. et al. Broad defects in the energy metabolism of leukocytes underlie

immunoparalysisinsepsis.NatImmunol17,406-413(2016).18. Berwick,D.C.,Hers,I.,Heesom,K.J.,Moule,S.K.&Tavare,J.M.TheidentificationofATP-

citratelyaseasaproteinkinaseB(Akt)substrateinprimaryadipocytes.JBiolChem277,33895-33900(2002).

19. Sacta, M.A. et al. Gene-specific mechanisms direct glucocorticoid-receptor-driven

repressionofinflammatoryresponsegenesinmacrophages.Elife7(2018).20. Hargreaves, D.C., Horng, T. &Medzhitov, R. Control of inducible gene expression by

signal-dependenttranscriptionalelongation.Cell138,129-145(2009).21. Smale,S.T.&Natoli,G.Transcriptionalcontrolof inflammatoryresponses.ColdSpring

HarbPerspectBiol6,a016261(2014).22. Ghisletti, S. et al. Identification and characterization of enhancers controlling the

inflammatorygeneexpressionprograminmacrophages.Immunity32,317-328(2010).23. Smale, S.T., Tarakhovsky,A.&Natoli,G. Chromatin contributions to the regulationof

innateimmunity.AnnuRevImmunol32,489-511(2014).24. Broz,P.&Dixit,V.M.Inflammasomes:mechanismofassembly,regulationandsignalling.

NatRevImmunol16,407-420(2016).25. Chandel,N.S.NavigatingMetabolism.ColdSpringHarborLaboratoryPress,2015.26. Wong,B.W.etal.Theroleoffattyacidβ-oxidationinlymphangiogenesis.Nature542,49-

54(2017).27. Lee, J.V. et al. Akt-dependentmetabolic reprogramming regulates tumor cell histone

acetylation.CellMetab20,306-319(2014).28. Koza,R.A. etal. Sequenceand tissue-dependentRNAexpressionofmouseFAD-linked

glycerol-3-phosphatedehydrogenase.ArchBiochemBiophys336,97-104(1996).

Page 99: Regulation of Inflammatory Macrophage Activation and

85

29. Mráček, T., Drahota, Z. & Houštěk, J. The function and the role of themitochondrialglycerol-3-phosphatedehydrogenaseinmammaliantissues.BiochimBiophysActa1827,401-410(2013).

30. Eto,K.etal.RoleofNADHshuttlesysteminglucose-inducedactivationofmitochondrial

metabolismandinsulinsecretion.Science283,981-985(1999).31. Garaude,J.etal.Mitochondrialrespiratory-chainadaptationsinmacrophagescontribute

toantibacterialhostdefense.NatImmunol17,1037-1045(2016).32. Chouchani, E.T. et al. AUnifyingMechanism forMitochondrial SuperoxideProduction

duringIschemia-ReperfusionInjury.CellMetab23,254-263(2016).33. Votyakova,T.V.&Reynolds,I.J.DeltaPsi(m)-Dependentand-independentproductionof

reactiveoxygenspeciesbyratbrainmitochondria.JNeurochem79,266-277(2001).34. Barrientos,A.&Moraes,C.T.TitratingtheeffectsofmitochondrialcomplexIimpairment

inthecellphysiology.JBiolChem274,16188-16197(1999).35. Chouchani, E.T. et al. Ischaemic accumulation of succinate controls reperfusion injury

throughmitochondrialROS.Nature515,431-435(2014).36. Everts, B. et al. TLR-driven early glycolytic reprogramming via the kinases TBK1-IKKɛ

supports the anabolic demands of dendritic cell activation.Nat Immunol15,323-332(2014).

37. Beleznai,Z., Szalay, L.& Jancsik,V.Ca2+andMg2+asmodulatorsofmitochondrial L-

glycerol-3-phosphatedehydrogenase.EurJBiochem170,631-636(1988).38. Weinberg,S.E.,Sena,L.A.&Chandel,N.S.Mitochondriaintheregulationofinnateand

adaptiveimmunity.Immunity42,406-417(2015).39. Wellen, K.E. et al. ATP-citrate lyase links cellular metabolism to histone acetylation.

Science324,1076-1080(2009).40. Sivanand,S.,Viney,I.&Wellen,K.E.SpatiotemporalControlofAcetyl-CoAMetabolismin

ChromatinRegulation.TrendsBiochemSci43,61-74(2018).41. Sanin,D.E.etal.MitochondrialMembranePotentialRegulatesNuclearGeneExpression

inMacrophagesExposedtoProstaglandinE2.Immunity49,1021-1033.e1026(2018).42. Byles,V.etal.TheTSC-mTORpathwayregulatesmacrophagepolarization.NatCommun

4,2834(2013).

Page 100: Regulation of Inflammatory Macrophage Activation and

86

43. Salabei,J.K.,Gibb,A.A.&Hill,B.G.Comprehensivemeasurementofrespiratoryactivityinpermeabilizedcellsusingextracellularfluxanalysis.NatProtoc9,421-438(2014).

44. Liu, X., Ser, Z. & Locasale, J.W. Development and quantitative evaluation of a high-

resolutionmetabolomicstechnology.AnalChem86,2175-2184(2014).45. Snyder,N.W.etal.Productionofstableisotope-labeledacyl-coenzymeAthioestersby

yeaststableisotopelabelingbyessentialnutrientsincellculture.AnalBiochem474,59-65(2015).

46. Frey,A.J.etal.LC-quadrupole/Orbitraphigh-resolutionmassspectrometryenablesstable

isotope-resolved simultaneous quantification and (13)C-isotopic labeling of acyl-coenzymeAthioesters.AnalBioanalChem408,3651-3658(2016).

47. Trefely,S.,Ashwell,P.&Snyder,N.W.FluxFix:automaticisotopologuenormalizationfor

metabolictraceranalysis.BMCBioinformatics17,485(2016).48. Schoors, S. etal. Fattyacid carbon is essential fordNTP synthesis inendothelial cells.

Nature520,192-197(2015).

Page 101: Regulation of Inflammatory Macrophage Activation and

CHAPTER 3

Additional metabolic modules regulating macrophage responses to microbial encounter:

roles for Akt and nucleocytosolic acetyl-CoA synthetase ACSS2

This chapter is a collection of unpublished data. It is presented here as a completed work to facilitate understanding and serve as a reference for further investigation.

P. Kent Langston1, Jonathan Jung1,2, Yong Kong3, Ibrahim Aksoylar1, and Tiffany Horng1,4*

1Department of Genetics & Complex Diseases, Harvard T.H. Chan School of Public Health, Boston, MA 2School of Medicine, University of Glasgow, Glasgow, UK 3Depart of Immunobiology, Yale University School of Medicine, New Haven, CT 4School of Life Sciences and Technology, ShanghaiTech University, Shanghai, China

Author contributions P.K.L. designed and performed experiments and analyzed data. P.K.L. prepared figures. T.H. supervised the project, including experimental design and data analysis. J.J., Y.K., and I.A. helped with gene expression, RNA-seq analysis, and LPS dose immunoblots, respectively.

Page 102: Regulation of Inflammatory Macrophage Activation and

88

ABSTRACT

Macrophages are activated during microbial infection to coordinate inflammatory responses and

host defense. Here we show that in macrophages activated by bacterial lipopolysaccharide (LPS),

the serine/threonine kinase Akt regulates glucose oxidation driven by mitochondrial glycerol 3-

phosphate dehydrogenase (GPD2) to control histone acetylation and inflammatory gene induction.

While acute LPS exposure drives macrophage activation, prolonged exposure triggers entry into

LPS tolerance, where macrophages orchestrate immunosuppression to limit the detrimental effects

of sustained inflammation. We find that Akt couples the strength and duration of LPS exposure to

a shutdown of oxidative metabolism, limiting histone acetylation at inflammatory genes and thus

contributing to suppression of inflammatory responses. In LPS tolerant macrophages re-exposed

to LPS, Akt also regulates the induction of a subset of highly-responsive, primed (P) genes,

important for anti-microbial activity and mitochondrial respiration. Such induction of P genes is

fueled by acetate-derived Ac-CoA production catalyzed by nucleocytosolic acetyl-CoA synthetase

(ACSS2) rather than citrate-derived Ac-CoA from GPD2-dependent glucose oxidation and ATP-

citrate lyase (ACLY) activity. Therefore, Akt and ACSS2 act as additional regulatory modules

integrating TLR signaling with shifts in macrophage metabolism to induce gene-specific responses

to microbial encounter.

Page 103: Regulation of Inflammatory Macrophage Activation and

89

INTRODUCTION

Macrophages, innate immune cells of the myeloid lineage, are the cornerstone of the host’s

response to microbial encounter. Expression of a repertoire of pathogen recognition receptors

(PRRs), enables macrophages to sense invading microbes through detection of various pathogen-

associated molecular patterns (PAMPs), including lipopolysaccharide (LPS) from gram-negative

bacteria. Recognition of LPS by the PRR Toll-like receptor 4 (TLR4) polarizes macrophages to an

inflammatory phenotype, characterized by upregulation of pro-inflammatory cytokine and

chemokine production, expression of co-stimulatory molecules to enhance adaptive immunity, and

increased phagocytosis and killing of bacteria. Although this robust response is critical for

conquering infection, over-exuberant inflammatory responses result in tissue damage that may

become fatal to the host, as observed in patients with multi-organ dysfunction as a consequence of

systemic inflammation during septic shock1–3. As a mechanism of protection against

inflammation-mediated organ damage, macrophages exposed to LPS for a prolonged period shift

into a state of reduced responsiveness, or tolerance, to further activation4,5. Such tolerance

contributes to the whole-body immunosuppression observed in patients that survive septic shock,

mitigating multi-organ dysfunction6,7.

In recent years, metabolism has come to be appreciated as an important layer of regulation in host

defense8–11. Indeed, specific metabolic shifts within innate and adaptive immune cells have been

shown to influence each of the four tasks of the immune system: recognition, effector functions,

regulation, and memory. Mechanisms underpinning such regulation are context-dependent and

range from integration with signaling to gene-specific chromatin modifications. The current

Page 104: Regulation of Inflammatory Macrophage Activation and

90

paradigm for metabolic instruction of inflammatory macrophage activation is that shutdown of

oxidative phosphorylation and induction of aerobic glycolysis are important adaptations coupling

TLR signaling with enhanced inflammatory cytokine production9,12. In our previous study, we

revised this conceptual framework, by demonstrating that oxidative metabolism actually supports

induction of macrophage inflammatory responses, such that impairment drives inflammatory gene

suppression and LPS tolerance (see Chapter 2). Specifically, we reported that the glycerol

phosphate shuttle (GPS), a glycolytic shunt, is a novel LPS-inducible metabolic module that boosts

glucose oxidation to drive inflammatory gene induction during LPS activation but suppresses

oxidative metabolism to contribute to inhibition of inflammatory gene induction during LPS

tolerance. Consistent with the well-established link between oxidative metabolism and regulation

of transcriptional responses13, we identified availability of nucleocytosolic acetyl-CoA (Ac-CoA)

as a node coupling glucose oxidation to histone acetylation at the inflammatory genes Il6 and Il1b.

Thus, the model proposed in our previous study provides much needed clarity to the relationship

between glucose metabolism and induction of macrophage inflammatory responses; however, the

nexus between TLR signaling and glucose oxidation remains incompletely understood.

Additionally, metabolic regulation of other macrophage effector functions in the context of

bacterial infection has not been well-investigated and thus constitutes a largely untapped and

important solution space.

Here we show that the serine/threonine kinase Akt (protein kinase B, PKB), a known target

downstream of PI3K14 and TBK1/IKKε15, integrates glucose oxidation with the strength and

duration of LPS stimulation to control the amplitude and duration of inflammatory gene induction.

We also identify a subset of genes, constituting 8% of the LPS-inducible transcriptome picked up

Page 105: Regulation of Inflammatory Macrophage Activation and

91

by our analysis, that become hyper-responsive following prolonged LPS exposure, in contrast with

the hypo-responsiveness of inflammatory genes that is characteristic of the LPS tolerant state.

These primed genes, or “P genes,” encode molecules important for biological processes distinct

from tolerant genes, including phagocytosis, mitochondrial ATP production, and ion transport and

also show regulation by Akt. Moreover, consistent with the shift from increased to decreased

oxidative metabolism over the course of LPS exposure as macrophages transition from activation

to tolerance, P genes are not directly supported by glucose oxidation since block of glucose

utilization or attenuation of citrate conversion to Ac-CoA by inhibition of ATP-citrate lyase

(ACLY) does not abolish induction of P genes in tolerant bone marrow derived macrophages

(BMDMs). Instead, we show that induction of P genes is supported by acetate-derived Ac-CoA

production catalyzed by nucleocytosolic acetyl-CoA synthetase (AceCS1, or ACSS2) and

highlight histone deacetylation as a source of acetate fueling this process. Although such regulation

is comparable to observations in other contexts of impaired Ac-CoA production from glucose

oxidation16,17, this is the first evidence of a role for ACSS2 in macrophage biology. Finally, we

propose that in addition to a role of Akt in the induction and suppression of inflammatory genes

through control of oxidative metabolism and thus citrate-derived Ac-CoA levels, Akt may also

regulate P genes through control of ACSS2 by phosphorylation at conserved serine residues. Thus,

Akt and ACSS2 act as additional regulatory modules integrating TLR signaling with shifts in

macrophage metabolism to induce gene-specific responses to microbial encounter.

Page 106: Regulation of Inflammatory Macrophage Activation and

92

RESULTS

Magnitude and rate of metabolic shifts reflect strength of macrophage LPS stimulation

In our previous study, we demonstrated that the ETC acts as a rheostat to couple duration of LPS

exposure to ETC directionality to modulate a balance between the induction and suppression of

inflammatory responses. During this investigation, we also observed that such metabolic shifts are

coupled not only to the duration but also the strength of LPS exposure. Indeed, activation of bone

marrow derived macrophages (BMDMs) with low or high dose LPS resulted in distinct changes

in the rates of extracellular acidification (ECAR) and oxygen consumption (OCR), reflective of

the strength of stimulation (Fig. 3.1a,b). Further titration of LPS dose used for BMDM stimulation

followed by analysis of acute inflammatory gene induction revealed similar dose-dependent effects

on induction of Il6 (Fig. 3.1c) and many other inflammatory genes (data not shown). We also

observed that respiration shutdown was positively correlated with the strength of LPS activation

(Fig. 3.1d) and the strength of tolerance to LPS restimulation (Fig. 3.1e), which was assessed using

the strategy described in Fig. 2.8a. Thus, consistent with our previous study and building on the

link between mitochondrial metabolism and responsiveness of inflammatory genes, ETC activity

couples dose and time of LPS exposure to magnitude and duration of inflammatory gene induction.

Page 107: Regulation of Inflammatory Macrophage Activation and

93

Figure 3.1 Dose-dependent metabolic shifts couple strength of LPS stimulation to amplitude and duration of inflammatory gene induction in macrophages. (a,b) Seahorse extracellular flux analysis of extracellular acidification rate (ECAR) and oxygen consumption rate (OCR) in real-time by BMDMs unstimulated (Naïve) or LPS-activated with a low (LPS 10 ng/mL) or high dose (LPS 100 ng/mL). (c) BMDMs were unstimulated (Naïve) or stimulated with the indicated doses of LPS for 3 hours followed by qPCR analysis of Il6 and Il1b gene expression. (d) Maximal mitochondrial respiration in BMDMs unstimulated (Naïve) or stimulated with indicated doses of LPS for 12h. Maximal respiration was calculated by subtraction of non-mitochondrial respiration (OCR post-Rot/AA) from mean uncoupled respiration (OCR post-FCCP). (e) Representative induction of Il6 and Il1b genes in BMDMs tolerized with different doses of LPS. BMDMs were treated according to the schematic in Fig. 2.8a. Concentrations indicate dose of LPS used during 24h tolerization. All BMDMs were rechallenged with the same dose, as indicated in Fig. 2.X. *p ≤0.05, **p ≤0.01, ***p ≤0.001, ****p ≤0.0001 (Student’s t-test). Data are from one experiment representative of three experiments (mean and s.e.m. of duplicates (c,e) or octuplets (a,b,d)).

Akt signaling integrates strength of LPS stimulation with glucose oxidation to control

macrophage inflammatory responses

Our previous observations demonstrate that deployment of the glycerol phosphate shuttle (GPS)

is an early event during LPS stimulation that critically drives ETC activity to support enhanced

production of acetyl-CoA (Ac-CoA) for histone acetylation and inflammatory gene induction in

macrophages. Is has also been reported that activation of Akt downstream of TBK1-IKKε

signaling is necessary for rapid induction of glucose utilization through glycolysis in LPS-

0

1000

2000

3000

4000

5000

Relative Expression

******

**

*Il6

NaïveLPS 1 ng/mLLPS 10 ng/mLLPS 50 ng/mLLPS 100 ng/mL

0 30 60 900

20

40

60

80

Time (min)

ECAR (mpH/min)

Media orLPS

0 25 50 75 100 125050100150200250300 Oligo FCCP

Media orLPS

Time (min)OCR (pmol/min)

LPS 100 ng/mLLPS 10 ng/mLNaïve

a b c

0

50

100

150 Il1b NN+LTT+L 100 ng/mLT+L 10 ng/mLT+L 1 ng/mL

Relative Expression

*

0

100

200

300 NaïveLPS 1 ng/mLLPS 3 ng/mLLPS 10 ng/mLLPS 100 ng/mL

****

*******

***

Max. OCR (pmol/min)

d e

Page 108: Regulation of Inflammatory Macrophage Activation and

94

stimulated dendritic cells (DCs)15. Given that low and high doses of LPS induce different rates of

glycolysis and magnitudes of respiration in BMDMs during activation, we postulated that

differences in Akt signaling may underpin this response. Analysis of Akt activation over a

timecourse by immunoblot assay for Akt phosphorylation at S473 and T308, modifications both

necessary for full activity14, revealed differences in kinetics supportive of a role for Akt in

integrating strength of LPS stimulation with rate and extent of glucose utilization (Fig. 3.2a).

Consistent with this, pharmacological inhibition of Akt (AKTi) completely blunted the early

increase in GPD2-dependent glucose oxidation (Fig. 3.2b), and thus also attenuated LPS-induced

promoter region histone acetylation and expression of Il6 and Il1b (Fig. 3.2c,d).

Figure 3.2 Akt signaling integrates LPS dose with glucose oxidation to regulate LPS activation in BMDMs. (a) Immunoblot analysis of activating phosphorylation of Akt at S473 and T308 during BMDM activation induced by the indicated doses of LPS. (b) Maximal mitochondrial respiration in BMDMs unstimulated (Naïve) or stimulated with LPS for 1h +/- pretreatment with the Akt inhibitor MK-2206 (AKTi). BMDMs were stimulated with LPS for 0-3h +/- AKTi followed by analysis of Il6 and Il1b promoter region histone acetylation (c) and gene expression (d). *p ≤0.05, **p ≤0.01, ***p ≤0.001 (Student’s t-test). Data are from one experiment representative of three experiments (mean and s.e.m. of duplicates (d), triplicates (c), or octuplets (b)).

0.0

0.5

1.0

1.5

%Input

AcH3 Il6

0

2

4

6AcH3K27 Il6

0

2

4

6AcH4 Il6

0.0

0.5

1.0

1.5

%Input

AcH3 Il1b

0

5

10

15AcH3K27 Il1b

0

7

14

21AcH4 Il1b

LPSLPS+AKTi

Naïve

***** **********

***** **********

0 20 40LPS (min): 6010 240120 0 20 40 6010 240120

10 ng/mL LPS 100 ng/mL LPS

p-AKT(S473)

AKT

p-AKT(T308)

0

50

100

150

200

Max. OCR (pmol/min)

* **

0

5000

10000

15000

0

100

200

300**

**

Relative Expression

Relative Expression Il6

Il1b

b c d

a

Page 109: Regulation of Inflammatory Macrophage Activation and

95

The early increase in glucose oxidation necessary to support inflammatory gene induction during

LPS activation ultimately leads to respiration shutdown, driving inflammatory gene suppression

and blunted responsiveness, or tolerance, to subsequent LPS encounters. Inhibition of Akt

protected against shutdown of respiration and suppression of Il6 and Il1b expression during

prolonged LPS exposure (Fig. 3.3a,b) and was comparable to that observed by impairing glucose

oxidation by 2DG cotreatment or GPD2 deletion (Fig. 2.9, 2.12, and Supplementary Fig. 3.1).

Figure 3.3 Akt signaling leads to impaired glucose oxidation and inflammatory gene suppression during prolonged LPS exposure. BMDMs were unstimulated (Naïve) or stimulated with LPS for 12h +/- the Akt inhibitor MK-2206 (AKTi) followed by analysis of the following parameters. (a) Mitochondrial respiration, determined by oxygen consumption rate (OCR). Basal and maximal respiration were calculated by subtraction of non-mitochondrial respiration (OCR post-Rot/AA) from mean OCR pre-oligomycin and post-FCCP, respectively. (b) Il6 and Il1b gene expression. *p ≤0.05, **p ≤0.01, ***p ≤0.001, ****p ≤0.0001 (Student’s t-test). Data are from one experiment representative of three experiments (mean and s.e.m. of duplicates (b) or octuplets (a)). Such protection was attributed to effects of AKTi on glucose oxidation in part because inhibition

of Akt resulted in significant attenuation of LPS-induced Gpd2 expression (Supplementary Fig.

3.2 and see Discussion). As expected, protection of respiration by AKTi cotreatment also rescued

histone acetylation and inflammatory gene induction in tolerant macrophages (Fig 3.4a,b). Thus,

0

1500

3000

4500

6000

7500

0

200

400

600

800Il6

Relative Expression

Il1b

**

a

b0 25 50 75 100050100150200250300350400

Time (min)

OCR (pmol/min)

LPS LPS+AKTiNaïveFCCP Rot/AAOligo

0

50

100

150

Basal OCR (pmol/min)

** ****

0

100

200

300

Max. OCR (pmol/min)

****

****

LPSLPS+AKTi

Naïve

LPSLPS+AKTi

Naïve

Page 110: Regulation of Inflammatory Macrophage Activation and

96

glucose oxidation via GPD2 and the GPS may be a novel target of Akt in LPS-activated

macrophages, serving as a critical node at the nexus of TLR signaling and metabolism to regulate

the magnitude and duration of inflammatory responses.

Figure 3.4 Inhibition of Akt protects against induction of LPS tolerance in macrophages. (a) Promoter region histone acetylation and (b) expression of Il6 and Il1b in N, N+L, T, and T+L+AKTi conditions. T+L+AKTi indicates cotreatment with Akt inhibitor (MK-2206) during tolerization followed by washout and LPS rechallenge without inhibitor (compare to Fig. 2.X). *p ≤0.05, **p ≤0.01, ***p ≤0.001, ****p ≤0.0001 (Student’s t-test). Data are from one experiment representative of three experiments (mean and s.e.m. of duplicates (b) or triplicates (a)).

Tolerant macrophages have a large repertoire of responsive genes that are regulated by Akt

Although pro-inflammatory macrophage effector molecules, such as IL-6 and IL-1b, that are

acutely induced and later suppressed following TLR stimulation by various PAMPs are the most

commonly studied components of the innate response to microbial encounter, the LPS-inducible

transcriptome of macrophages encodes a vast array of molecules important for myriad biological

functions, including phagocytosis, ion transport, response to unfolded proteins, and more4,18,19. We

performed RNA-seq analysis on naïve and tolerant BMDMs unstimulated or stimulated with LPS

0

3

6

9 AcH3 Il6

0

1

2

3

4

5 AcH3K27 Il6

0

1

2

3 AcH4 Il6

0

1

2

3

4 AcH3 Il1b

0

1

2

3

4

5 AcH3K27 Il1b

0

1

2

3

4

5 AcH4 Il1b

NN+LTT+LT+L+AKTi

*** *** ***

*** ******

%Input (relative over HPRT)

%Input (relative over HPRT)

0

1000

2000

3000

4000

0

5000

10000

15000

20000

Relative Expression

Relative Expression

*

*

Il6

Il1b

a b

Page 111: Regulation of Inflammatory Macrophage Activation and

97

as in Fig. 2.8a and grouped all LPS-inducible genes into three classes based on expression patterns

across conditions. Proportions of each class within the total LPS-inducible gene set, as well as

representative expression of genes from each class are presented in Fig. 3.5. Tolerized “T” genes

(i.e. Il6 and Il1b) were defined as those that showed impaired induction in the tolerant state

compared to the responsive state. Non-Tolerized “NT” genes (i.e. Lcn2 and Rantes) did not appear

tolerant but were also not further induced in tolerant BMDMs after LPS restimulation. Finally,

Primed “P” genes (i.e. Fpr1 and Irg1) were defined as those that, like NT genes, did not exhibit

tolerance but also displayed strong induction in tolerant, LPS-restimulated BMDMs.

Figure 3.5 The LPS-inducible transcriptome of macrophages falls into three classes based on expression after repeated LPS stimulation. (a) Proportions of Tolerized (T), Non-Tolerized (NT), and Primed (P) genes within the LPS-inducible gene set of BMDMs treated as in Fig. 2.8a, measured by RNA-seq analysis. Specific analysis is described in the Materials and Methods. (b) Gene expression of Il6, Lcn2, and Fpr1, representing T, NT, and P genes. *p ≤0.05, **p ≤0.01, ***p ≤0.001, ****p ≤0.0001 (Student’s t-test). Data are from one experiment representative of one (a) or three experiments (b) (mean and s.e.m. of duplicates (b) or number of genes based on fold change calculations from triplicate fpkm values (a)).

This third class, although the smallest of the three subsets, constituting only 8% of all LPS-

inducible genes detected by our analysis, was of greatest interest, as induction of such P genes

must be supported by an alternative source of Ac-CoA to that derived from citrate from glucose

oxidation. Furthermore, because the expression pattern of P genes is antithetical to the concept of

tolerance, P genes must belong to distinct and important functional clusters. Gene ontology

0

10000

20000

30000

40000

Relative Expression

Il6

0

1000

2000

3000Lcn2

0

500

1000

1500 Fpr1

a b

5581

Total LPS-Inducible=7379 Tolerized(T) Genes

Non-Tolerized (NT) Genes

Primed(P) Genes

NN+LTT+L

T genes P genesNT genes

1204

594

****ns

****

Page 112: Regulation of Inflammatory Macrophage Activation and

98

analysis was performed to further classify P genes based on the biological processes and cellular

components to which they belong (Supplementary Figs. 3.3).

Although glucose oxidation is impaired in tolerant macrophages, we first turned to Akt in our

investigation into the metabolic underpinnings of induction of P genes, as Akt activity is still

induced by LPS stimulation of tolerant BMDMs (Fig. 3.6a), and relevant metabolic targets of Akt

in control of macrophage activation remain incompletely characterized. Strikingly, treatment of

tolerant BMDMs with AKTi before LPS restimulation (Fig. 3.6b) completely attenuated

reinduction of some P genes (Fig. 3.6c). However, consistent with suppression of T genes due to

impaired glucose oxidation and carbon flux into Ac-CoA for histone acetylation and gene

induction (Fig. 2.9 and 2.10), inhibition of glucose utilization and ACLY activity by 2DG and

ACLYi treatment between challenges did not lead to strong attenuation of P genes

(Supplementary Fig. 3.4). It is worth noting that the additional rest between LPS washout and

LPS restimulation was selected and incorporated to minimize the contribution of transcript

stability to P gene expression levels measured in T and T+L conditions so that reinduction could

be assessed with greater precision. This duration of rest between challenges was shown to be

insufficient for natural reversal of tolerance, based on no recovery of T genes in the T+L condition

(data not shown). Thus, Akt regulates induction of P genes upon restimulation of tolerant

macrophages through a mechanism independent of enhanced ACLY-mediated conversion of

glucose-derived citrate to Ac-CoA.

Page 113: Regulation of Inflammatory Macrophage Activation and

99

Figure 3.6 LPS-inducible Akt signaling regulates induction of P genes in tolerant BMDMs. (a) Immunoblot analysis of activating phosphorylation of Akt at S473 and T308 during LPS restimulation of tolerant BMDMs. (b) Schematic of workflow for analysis of P gene regulation. BMDMs were left unstimulated (naïve, N) or stimulated with LPS for 24h to induce LPS tolerance (tolerance, T). BMDMs were washed and left without LPS for 4h. LPS responsiveness was assessed by stimulating naïve (N+L) or re-stimulating tolerant (T+L) BMDMs with LPS. In the T+L+inhibitor condition, drug was added in the final hour of rest between tolerance induction and restimulation. Naïve and Tolerant cells were collected as baseline controls (N and T). (c) Representative induction of Fpr1 and Ptges genes in BMDMs treated according to (b) with the Akt inhibitor MK-2206 used in the T+L+inhibitor condition (T+L+AKTi). ***p ≤0.001 (Student’s t-test). Data are from one experiment representative of two (a) or three experiments (c) (mean and s.e.m. of duplicates (c)).

Induction of responsive genes in tolerant macrophages is supported by ACSS2 activity

Under conditions of impaired conversion of citrate to Ac-CoA such as in yeast lacking ACLY20–

22 and in cancer cells with truncated glucose oxidation due to preferential shunting of pyruvate

towards lactate16, acetyl-CoA synthetase enzymes (AceCS1/2) can provide Ac-CoA through an

ATP-dependent reaction between acetate and CoA23. Therefore, we measured protein levels of the

T+L+inhibitor1h inhibitor24h LPS

100 ng/mL4h LPS10 ng/mL

3h Unstim

0 30 60LPS (min): 24015

p-AKT(S473)

AKT

p-AKT(T308)

Tolerant

Assay

0

2000

4000

6000

Relative Expression

Fpr1***

0

200

400

600

800

1000Ptges

***NN+LTT+LT+L+AKTi

a c

b

T+L24h LPS100 ng/mL

4h LPS10 ng/mL

Assay

4h Unstim

Wash

Tolerant (T)24h LPS100 ng/mL

4h Unstim

T+L24h LPS100 ng/mL

4h LPS10 ng/mL

4h Unstim

Naive (N)24h Unstim 4h Unstim

Assay

Assay

Assay

Page 114: Regulation of Inflammatory Macrophage Activation and

100

nucleocytosolic AceCS1 enzyme (ACSS2) in tolerant and tolerant restimulated BMDMs and

found expression in both conditions (Fig. 3.7a). Pharmacologic inhibition of ACSS2 (ACSS2i)

between challenges completely attenuated reinduction of P genes (Fig. 3.7b). To our knowledge,

we are the first to report a role for ACSS2 in the regulation of LPS-inducible genes.

Figure 3.7 ACSS2 is expressed in tolerant macrophages and regulates induction of P genes. (a) Immunoblot analysis of nucleocytosolic acetyl-CoA synthetase (ACSS2) in tolerant BMDMs restimulated with LPS for the indicated times. (b) Gene expression of Fpr1 and Bpifc in BMDMs treated as in Fig. 3.6b, using an inhibitor of ACSS2 in the interval between tolerization and restimulation (T+L+ACSS2i). of workflow for analysis of P gene regulation. *p ≤0.05, **p ≤0.01, ***p ≤0.001, ****p ≤0.0001 (Student’s t-test). Data are from one experiment representative of three experiments (mean and s.e.m. of duplicates (b)).

The known major sources of acetate in mammalian cells are uptake from the extracellular

environment24,25, conversion from pyruvate via activity of keto acid dehydrogenases12, and

deacetylation of acetylated proteins such as histones12,13. Given the well-established role for

enhanced histone deacetylase (HDAC) activity in LPS tolerant macrophages4 and the failure of

2DG treatment during the window between challenges to attenuate P genes, we focused on histone

deacetylation as a source of acetate fueling ACSS2 activity in tolerant restimulated macrophages.

Specifically, we postulated that suppression and induction of T and P genes in tolerant

macrophages may be linked through a futile cycle of histone deacetylation and hyperacetylation,

respectively. Such a mechanism would serve as a temporal regulatory circuit, allowing for hyper

induction of P genes only after induction and suppression of T genes. This phasic response would

also store carbons from glucose oxidation in acetylated chromatin at T genes during activation to

Relative Expression

Fpr1 Bpifc

0

500

1000

1500 *

0

50

100

150

200**

α-Tubulin

ACSS20LPS (h): 4Tolerant

NN+LTT+LT+L+ACSS2i

a b

Page 115: Regulation of Inflammatory Macrophage Activation and

101

be used for acetylation of P genes in a setting of impaired glucose oxidation, thus turning the

chromatin into a metabolic capacitor capable of storing acetyl groups and therefore acetylation

potential. Consistent with this concept, treating BMDMs with the HDAC inhibitors sodium

butyrate and trichostatin A (TSA) during the tolerizing LPS challenge rescued T genes and

attenuated P genes (Fig. 3.8a).

Figure 3.8 Histone deacetylation supports induction of P genes in tolerant BMDMs. Gene expression of Il1b, Fpr1, and Ptges in BMDMs treated as in Fig. 3.6b, with the HDAC inhibitors sodium butyrate (Butyrate) and trichostatin A (TSA) added during prior to tolerizing LPS stimulation (a) or prior to LPS restimulation (b), as indicated in by the corresponding schematics. *p ≤0.05, **p ≤0.01, ***p ≤0.001, ****p ≤0.0001 (Student’s t-test). Data are from one experiment (mean and s.e.m. of duplicates (a,b)).

Unexpectedly, treatment with butyrate and TSA in the window between challenges also attenuated

P genes while having only modest effects on T genes (Fig. 3.8b). Taken together, these data

illuminate a role of deacetylation in supporting the induction of P genes in tolerant macrophages.

0

1000

2000

3000

4000

5000

Relative Expression

Il1b

**

0

1000

2000

3000

4000 Fpr1

**

0

100

200

300

400 Ptges **

NN+LTT+LT+L+ButyrateT+L+TSA

T+L+HDACi1h inhibitor24h LPS

100 ng/mL4h LPS10 ng/mL

3h UnstimAssay

a

0

2000

4000

6000

0

500

1000

1500Il1b Fpr1 Ptges

Relative Expression

T+L+HDACi24h LPS100 ng/mL

4h LPS10 ng/mL

AssayWash

1h inhibitor 4h Unstim

Wash

b

*ns

****

0

100

200

300 **

Page 116: Regulation of Inflammatory Macrophage Activation and

102

Furthermore, although inhibiting HDAC activity during tolerization suggests that P gene induction

is linked in series to T gene suppression, attenuation of P gene induction by HDAC inhibition after

acquisition of a tolerant state raises to question whether induction may be supported in parallel by

deacetylation of histones at other gene sets.

Finally, we asked how ACSS2 activity is being regulated in tolerant macrophages. Canonically,

ACSS2 is activated by deacetylation catalyzed by Sirt128, possibly as a mechanism of sustaining

histone acetylation under nutrient-limited conditions. Inhibition of Sirt1 in the window between

challenges did not attenuate P gene induction in tolerant BMDMs, suggesting a non-canonical

mode of ACSS2 regulation (Fig. 3.9). Given the enhanced activity of Akt in this setting and the

effects of AKTi on P gene induction (Fig. 3.6), we investigated whether ACSS2 could be a target

of Akt. A database search revealed several residues in ACSS2 that could be targeted by Akt-

dependent phosphorylation, specifically S30, S263, and S267 (UniProtKB: Q9QXG4). Future

studies will be focused on determining whether ACSS2 is activated by Akt in tolerant macrophages.

This is covered in more detail in the Discussion section that follows.

Figure 3.9 Sirt1 does not regulate the induction of P genes in tolerant macrophages. (a) Gene expression of Fpr1 and Ptges in BMDMs treated as in Fig. 3.6b, using an inhibitor of Sirt1 in the interval between tolerization and restimulation (T+L+Sirt1i). *p ≤0.05 (Student’s t-test). Data are from two experiments (mean and s.e.m. of duplicates).

0

2000

4000

6000

8000

Relative Expression

Fpr1*

0

200

400

600

800

1000Ptges

ns NN+LTT+LT+L+Sirt1i

Page 117: Regulation of Inflammatory Macrophage Activation and

103

DISCUSSION

In this study, we make several findings regarding the metabolic regulation of macrophage effector

functions during LPS activation and tolerance. First, we found that the strength of macrophage

activation and tolerance, measured by induction of Il6 and Il1b, reflect the dose of LPS exposure

and correlate with early dose-dependent shifts in glycolysis, mitochondrial respiration, and Akt

activation. These observations are consistent with previous reports that TLR stimulation results in

dose-dependent metabolic adaptations in human monocytes29 and that Akt is activated downstream

of TBK1/IKKε in LPS-activated dendritic cells to promote rapid induction of glycolysis15.

However, based on our previous study showing the importance of oxidative metabolism in

supporting inflammatory macrophage activation and on our present findings illuminating parallels

between LPS dose-dependent Akt signaling and changes in mitochondrial respiration, we chose to

investigate the potential for Akt to regulate activation through control of glucose oxidation. We

found that pharmacological inhibition of Akt during LPS stimulation blunted the early increase in

mitochondrial respiration driven by enhanced GPD2-dependent glucose oxidation (see Chapter

2), thus also attenuating LPS-induced promoter region histone acetylation and expression of Il6

and Il1b. Inhibition of Akt during prolonged LPS exposure also recapitulated the protection against

respiration shutdown, inflammatory gene suppression, and LPS tolerance observed through

impairing glucose oxidation by 2DG treatment or GPD2 deletion. Because these adaptations are a

consequence of enhanced GPD2-driven glucose oxidation contributing to a switch from forward

to reverse electron transport (FET to RET) during the transition from LPS activation to tolerance,

we asked whether Akt could regulate GPD2 and found that AKTi treatment strongly attenuated

LPS-inducible Gpd2 expression. Taken together, these data support an updated version of our

previous model (Supplementary Fig. 2.7), placing Akt as an additional regulatory node

Page 118: Regulation of Inflammatory Macrophage Activation and

104

integrating TLR signaling with GPD2-dependent glucose oxidation to control inflammatory gene

induction and suppression in macrophages (Supplementary Fig. 3.5). In addition to regulation of

glucose oxidation via modulation of glycolytic flux and Gpd2 expression, Akt likely supports

activation through direct positive regulation of ACLY by phosphorylation at S473 as has been

reported in cancer30 and alternatively activated macrophages31. This interaction is included in our

updated working model (Supplementary Fig. 3.5a), but it will be important to validate the Akt-

ACLY axis during LPS activation by immunoblot assay for ACLY S473 phosphorylation under

conditions of inhibited Akt signaling. It is also noteworthy that the attenuation of LPS-inducible

Gpd2 expression by AKTi treatment was observed at times later than the early increase in GPD2-

dependent mitochondrial respiration. This is not inconsistent with a critical role for Akt in driving

inflammatory gene suppression and tolerance through enhanced GPD2 activity, as the FET-to-

RET transition occurs late during LPS stimulation and depends on continued glucose oxidation

under conditions of impaired respiration (induced by many mechanisms including inhibition by

nitric oxide32 and itaconate33–35). Akt-dependence of the GPD2-driven respiration after 1h of LPS

stimulation could be a result of increased flux through upper glycolysis and the glycerol phosphate

shuttle due to enhanced hexokinase II (HK-II) activity permitted by its Akt-dependent

translocation to the mitochondrial outer membrane15,36,37. Given the delayed induction of

glycolytic genes downstream of Akt-HIF-1α, enhanced glucose oxidation fueled by increased

substrate flux through HK-II and the GPS is an attractive mechanism.

Second, we define a subset of LPS-inducible genes exhibiting enhanced responsiveness, or

priming, following tolerization and identify a novel role for ACSS2 as a regulator of these genes

in tolerant macrophages. These primed (P) genes were an interesting case study on the metabolic

Page 119: Regulation of Inflammatory Macrophage Activation and

105

regulation of macrophage effector functions because their pattern of expression was seemingly

inconsistent with the concept of tolerance and our understanding of metabolic regulation of gene

induction in LPS-stimulated macrophages. However, gene ontology analysis revealed that P genes

belong to functional clusters distinct from pro-inflammatory cytokine and chemokine production,

providing insight into which biological processes are prioritized in tolerant macrophages. As

discussed in Chapter 1, the biochemical reactions in the metabolic networks of living organisms

take on measurable changes in flux in response to the application of stress. “Steady-state” is the

metabolic status as a function of the sum stress imposed by processes required for a cell’s existence

(i.e. energy production, maintenance of ion gradients, and biosynthesis to balance molecular

turnover), and any shifts in cellular metabolism in response to additional stress are calibrated to

support specific biological functions to counter that stress and return to “steady-state.” Thus, given

the existence of P genes in tolerant macrophages, we focused on uncovering the specific changes

in metabolism calibrated to support induction and execution of P gene-related functions.

As oxidative metabolism is gradually shut down during tolerization, we postulated that P genes

must be supported by an alternative pathway. Inhibitor experiments designed to block glucose

utilization and citrate-derived Ac-CoA production between tolerization and LPS restimulation did

not show strong attenuation of P genes, supporting our hypothesis. Strikingly, blocking acetate-

derived Ac-CoA production by inhibiting ACSS2 completely attenuated P genes, suggesting a

compensatory switch in response to impaired glucose oxidation. This switch from reliance on

ACLY- to ACSS2-mediated production of Ac-CoA for gene induction is consistent with

observations in other conditions of impaired conversion of citrate to Ac-CoA16,20–22. Of course,

future experiments measuring the size and substrate-dependency of the Ac-CoA pool in ACSS2-

Page 120: Regulation of Inflammatory Macrophage Activation and

106

deficient macrophages are needed, as well as analysis of promoter region histone acetylation at P

genes, similar to what was performed in Chapter 2.

Although isotope tracing experiments have not yet been done in this setting, inhibition of HDACs

during tolerization resulted in complete rescue of T genes and attenuation of P genes, supporting

histone deacetylation as a source of acetate as has been previously reported16,17,27. Moreover, our

data showing failure of 2DG treatment between LPS challenges to attenuate P gene induction

excludes dependency on acetate derived from glucose through the activity of alpha-keto acid

dehydrogenases that has recently been proposed as an important source of acetate in other cell

types26. Given the known role of HDACs in the induction of LPS tolerance4, we interpret this data

as pointing towards a temporal regulatory circuit in which a futile cycle of histone deacetylation

and hyperacetylation allows for hyper-induction of P genes after induction and suppression of T

genes. Such a phasic response would store carbons from glucose oxidation in acetylated histones

at T genes during activation to be used for acetylation of P genes in a setting of impaired glucose

oxidation, thus turning chromatin into a metabolic capacitor capable of storing acetyl groups and

therefore acetylation potential. Inhibition of HDACs after tolerization did not rescue Il1b but

completely attenuated the P genes Fpr1 and Ptges, suggesting that there may be a larger pool of

acetate from deacetylation of other genes that can also fuel P gene hyperacetylation and induction.

Future experiments addressing this question will include ChIP-seq analysis of a larger set of T and

P genes to determine whether they are regulated in a closed circuit or if other regions of chromatin

can store acetylation potential. Isotope tracing experiments will also be performed to directly test

whether histone acetylation at P genes is supported by acetate derived from deacetylation of

histones that were previously acetylated downstream of glucose oxidation (see Chapter 4).

Page 121: Regulation of Inflammatory Macrophage Activation and

107

It is also noteworthy that the importance of P genes in assembly of CI and CV of the ETC

(Supplementary Fig. 3.3) provides an additional layer of logic to the existence of a metabolic

circuit linking the sequential induction of T and P genes. Indeed, as glucose oxidation critically

supports T gene induction, we postulate that a rationale for the existence of P genes is to serve as

a mechanism of reversing impaired mitochondrial oxidative capacity to restore responsiveness of

inflammatory T genes after elimination of microbial infection. This conceptual framework is

further supported by the natural recovery from tolerance that occurs after several days of LPS

exposure4, coincident with a recovery of oxidative metabolism (data not shown). We predict that

autocrine IL-10 may play a role in this time-dependent reversal38.

Third, we demonstrate that Akt regulates P genes and may do so through phosphorylation of

ACSS2, constituting a novel mode of regulation of ACSS2. Our inhibitor experiments show that

inhibition of Akt but not Sirt1 between LPS challenges results in attenuation of P gene induction

comparable to blocking ACSS2 and HDACs. This observation was not recapitulated in BMDMs

with constitutive Akt1/2-deficiency (data not shown), highlighting the importance of using

methods conferring punctual inhibition in future studies investigating regulation of ACSS2 activity

and induction of P genes. As additional evidence supporting Akt’s potential to directly control

ACSS2, we found three serine residues in ACSS2 that are conserved phospho-targets of Akt

(positions 30, 263, and 267). We are also aware of evidence for ACSS2 phosphorylation in situ,

and future experiments will be focused on investigating such regulation in tolerant macrophages.

Thus, in addition to a role of Akt in the induction and suppression of inflammatory genes through

control of glucose oxidation, Akt may also regulate P genes through control of ACSS2 by direct

phosphorylation (Supplementary Fig. 3.5b).

Page 122: Regulation of Inflammatory Macrophage Activation and

108

Taken together, the observations presented here extend our model for how GPD2-dependent

glucose oxidation couples the duration of LPS exposure to induction and suppression of

inflammatory genes, providing the first evidence of a role for ACSS2 in macrophage biology and

highlighting Akt and ACSS2 as additional regulatory modules integrating TLR signaling with

shifts in macrophage metabolism to induce gene-specific responses to microbial encounter.

Page 123: Regulation of Inflammatory Macrophage Activation and

109

MATERIALS AND METHODS

BMDM culture. BMDMs from male and female mice were differentiated and cultured as

described previously39. For LPS activation (0-3h post-LPS), BMDMs were stimulated with 1-100

ng/mL LPS (LPS Ultra-Pure, Sigma-Aldrich). All of the metabolic parameters associated with

LPS activation, including glucose oxidation, Akt phosphorylation, and promoter histone

acetylation, were assessed at early time points after LPS stimulation (0-3 h), while the downstream

consequences of such metabolic events, i.e., inflammatory gene induction and cytokine production,

were assessed at 3 h after LPS stimulation. For LPS tolerance (12-28h post-LPS), BMDMs were

challenged with 100 ng/mL LPS for 24h to induce tolerance (T), washed, and then rechallenged

with 10 ng/mL LPS (T+L) for the desired amount of time. In such experiments, naïve BMDMs

were incubated in media for 24h (N), washed, and challenged with 10 ng/mL LPS (N+L) to provide

responsive-cell controls. In Fig. 3.1, tolerance was induced using the concentrations indicated in

the legend. When measuring induction of P genes, BMDMs were washed and media was replaced

after 24h incubation in media or LPS, followed by a 4h rest before assay (N and T) or LPS

rechallenge (N+L and T+L). Pharmacological inhibition of Akt (AKTi, 8 µM; MK-2206, Tocris,

United Kingdom), ACLY (ACLYi, 80-160 µM; SB-204990, Tocris, United Kingdom), ACSS2

(ACSS2i, 25-50 µM; N-(2,3-di-2-thienyl-6-quinox- alinyl)-N0-(2-methoxyethyl)16, ChemBridge),

histone deacetylases (HDACi, sodium butyrate 3 mM and trichostatin A (TSA) 50-100 µM;

Sigma-Aldrich), Sirt1 (Sirt1i, 0.5-1 mM; EX-527, Tocris, United Kingdom), and glucose

utilization (2DG, 5-10 mM; 2-deoxy-glucose, Sigma-Aldrich), was achieved by pretreating

BMDMs for 1h prior to LPS activation or in the final hour of 4h rest prior to LPS restimulation.

Page 124: Regulation of Inflammatory Macrophage Activation and

110

Seahorse assays. Basal and LPS-induced changes in oxygen consumption (OCR) and extracellular

acidification (ECAR) rates were measured with a Seahorse XF96 Extracellular Flux Analyzer

(Agilent). Assays were performed in Seahorse XF Assay Medium supplemented with glucose (11

mM) and adjusted to pH 7.4 with NaOH. Mix-wait-measure durations were set to 3, 2, and 3

minutes, respectively. For the MitoStress Test, oligomycin, FCCP, and rotenone/antimycin A were

sequentially injected to achieve final concentrations of 1, 1.5, and 2 µM. Non-mitochondrial OCR

was subtracted from the mean pre-oligomycin OCR and the mean post-FCCP OCR to calculate

Basal and Maximal mitochondrial respiration, respectively. For measuring LPS-induced changes

in OCR in real-time, media or LPS was injected after establishing basal OCR, followed by

oligomycin (1 µM) and FCCP (1.5 µM). Real-time measurement of LPS-induced changes in

ECAR was performed by injection of media or LPS followed by sequential injections of

oligomycin (1 µM) and 2DG (500 mM) rather than oligomycin and FCCP. After each experiment,

OCR and ECAR values were normalized to DNA content in each well by Hoechst 33342 staining

(Life Technologies, Carlsbad, CA).

Chromatin immunoprecipitation. Acetylation at inflammatory gene promoters was measured by

chromatin immunoprecipitation (ChIP) assay as described previously. Promoters are defined as

the region between -500 bp and the transcription start site. Antibodies used were as follows:

acetylated H3 (Millipore 06–599), acetylated H3K27 (Abcam ab4729), acetylated H4 (Millipore

06–866), and IgG (Santa Cruz, Dallas, TX, SC-2027). Fold enrichment was calculated as ChIP

signals normalized to input.

Page 125: Regulation of Inflammatory Macrophage Activation and

111

Gene expression. RNA was isolated using RNA-Bee (Tel-Test) according to the manufacturer’s

protocol. cDNA synthesis was performed using a High Capacity cDNA Reverse Transcription Kit

(Applied Biosystems). qPCRs were run on a Bio-Rad C1000 Thermalcycler, and expression of

target genes was calculated by the ddCT method using CFX Manger Software (Bio-Rad). Relative

expression values for target genes were calculated by normalization to expression of hypoxanthine

phosphoribosyltransferase (HPRT).

RNA-seq. Strand-specific libraries were generated using 500ng RNA input using TruSeq library

preparation kit (Illumina, San Diego, CA). cDNA libraries were multiplexed using specific unique

adaptors and sequenced using Illumina NextSeq 500 under single end 75bp read length parameters.

Raw reads were trimmed off sequencing adaptors and low quality regions by btrim23. Trimmed

reads were mapped to mouse genome (GRCm38) by tophat241. After the counts were collected,

differential expression analysis was done by DEseq242, which calculated the fold changes and

adjusted p-values. The fpkm values were calculated by DEseq2. The cutoff for defining the total

LPS-inducible gene set was fold change >1.2 for N+L over N. Subsets of LPS-inducible genes

were defined as follows: fold change >1.2 for N+L over T+L, labeled T genes; fold change

between 0.8 and 1.2 for N+L over T+L, labeled NT genes; fold change >1.2 for T+L over N+L

and fold change >1.2 for T+L over T, labeled P genes. Enrichment analysis of Gene Ontology (GO)

terms in the P genes subset of LPS-inducible genes was performed using Enrichr43,44. Heatmaps

were created using Prism software.

Mice. Wild-type C57BL/6J mice (The Jackson Laboratory) and whole-body Gpd2-/- (KO) mice

(generated as reported in Chapter 2) were maintained under specific pathogen-free conditions at

Page 126: Regulation of Inflammatory Macrophage Activation and

112

the Harvard T.H. Chan School of Public Health in accordance with Institutional Animal Care and

Use Committee (IACUC) guidelines. Male and female mice were used for endotoxic shock

experiments and as a source of bone marrow for BMDM culture between 6 and 12 weeks of age.

All protocols were approved by the IACUC of Harvard Medical School. In the endotoxic shock

experiments, WT and GPD2 KO mice were injected with vehicle (saline) or 3 mg/kg LPS via

intraperitoneal injection (IP), and serum IL-6 and IL-1β were measured 6 h later.

Immunoblot and enzyme-linked immunosorbent assays. For immunoblotting, BMDMs were

lysed in 1% NP-40 buffer or RIPA buffer, and protein concentration was determined using the

Bradford method or micro BCA assay (Pierce). Primary antibodies used were anti-phospho-

AKT(S473), anti-phospho-AKT(T308), and anti-AKT (1:1000, Cell Signaling), anti-ACSS2

(1:1000, Abcam), and anti-α-Tubulin (1:5000, Sigma-Aldrich). For ELISA, cell culture

supernatants and mouse whole blood were spun down, diluted in 5% BSA, and assayed for IL-6

and IL-1β using BioLegend Standard ELISA kits according to the manufacturer’s protocols.

Statistical analysis. Statistical significance was determined by a two-tailed Student’s t-test using

Prism software (GraphPad). *, **, ***, and **** in figures denote p value ≤0.05, 0.01, 0.001, and

0.0001, respectively.

Acknowledgements

J.J. was supported by the British Society for Immunology and T.H. by NIH grant R01AI102964.

Competing Interests

The authors declare no competing interests.

Page 127: Regulation of Inflammatory Macrophage Activation and

113

Supplementary Figure 3.1 Attenuating glucose oxidation impairs suppression of inflammatory gene expression and cytokine production in macrophages and mice. Expression of Il6 and Il1b and IL-6 cytokine production during prolonged LPS stimulation (12-24h) in WT BMDMs +/- 2DG (a,b) and in GPD2 KO BMDMs (c,d). Serum levels of IL-6 and IL-1β in 8-wk-old male WT and GPD2 KO mice injected IP with a high dose of LPS. *p ≤0.05, **p ≤0.01, ***p ≤0.001, ****p ≤0.0001 (Student’s t-test). Data are from one experiment representative of one (e) or three (a-d) experiments (mean and s.e.m. of duplicates (a-d) or n=4 mice per group (e)).

Supplementary Figure 3.2 LPS-induced expression of Gpd2 is regulated by Akt. (a) Gene expression of Gpd2 in BMDMs unstimulated (Naïve) or stimulated with LPS for 4h +/- Akt inhibitor (AKTi). *p ≤0.05 (Student’s t-test). Data are from one experiment representative of two experiments (mean and s.e.m. of duplicates).

WT GPD2 KO0

20

40

60

80

Relative Expression

**

Il6

WT GPD2 KO0

20

40

60

80***

Il1b

WT GPD2 KO0

5000

10000

15000

20000

25000

IL-6 production (pg/mL)

**** LPSNaïve

0

10000

20000

30000

40000

IL-6 Production (pg/mL) ****

0

1000

2000

3000

4000 Il6

**

0

20

40

60

Relative Expression

Il1b

* LPSLPS+2DG

Naïve

a b

c

WT GPD2 KO0

20000

40000

60000

80000

Serum IL-6 (pg/mL) **

WT GPD2 KO0

2000

4000

6000

8000

10000

Serum IL-1β (pg/mL) *

d

e

0

2

4

6

8

10 Gpd2

Relative Expression * LPS

LPS+AKTi

Naïve

Page 128: Regulation of Inflammatory Macrophage Activation and

114

Supplementary Figure 3.3 P genes belong to an array of gene ontologies. Gene set enrichment analysis performed on P genes using Enrichr GO Biological Process (a) and GO Cellular Component (b) reference lists. Pathways of interest are shown and ranked by p-value (Fisher exact test).

Supplementary Figure 3.4 Inhibiting the glucose-ACLY-Ac-CoA axis in tolerant macrophages does not attenuate induction of P genes. Expression of the P genes Ptges and Fpr1 in BMDMs treated as in Fig. 3.6b, using inhibitors of glucose utilization (2DG, a) and ATP citrate lyase (ACLYi, b) in the interval between tolerization and restimulation. *p ≤0.05 (Student’s t-test). Data are from one experiment representative of two experiments (mean and s.e.m. of duplicates).

p-value

p-value1e-02

3e-01

1e-02

5e-02

a

b

protein localization to nucleus

regulation of protein depolymerization

nucleic acid transport

immune response-activating cell surface receptor signaling pathway

positive regulation of IRE1-mediated unfolded protein response

positive regulation of phagocytosis

positive regulation of Ras protein signal transduction

regulation of Rac protein signal transduction

mitochondrial ATP synthesis coupled electron transport

positive regulation of endocytosis

cation transmembrane transport

positive regulation of Na+ transmembrane transporter activity

mitochondrial proton-transporting ATP synthase complex

CuI3-RING ubiquitin ligase complex

T cell receptor complex

transcription factor TFIIIC complex

nucleolar ribonuclease P complex

mitochondrial respiratory chain complex I

Golgi stack

HFE-transferrin receptor complex

autolysosome

core mediator complex

0

100

200

300

Relative Expression

PtgesNN+LTT+LT+L+2DG

0

2000

4000

6000

Relative Expression

Fpr1

* NN+LTT+LT+L+ACLYi

ba

Page 129: Regulation of Inflammatory Macrophage Activation and

115

Supplementary Figure 3.5 Integration of Akt and ACSS2 into previous model of metabolic regulation of gene induction in LPS stimulated macrophages. (a) During LPS activation, GPD2-dependent transport of electrons from the glycerol phosphate shuttle (GPS) through the ETC is upregulated to support an increase in glucose-derived carbon flux through glycolysis and the TCA cycle, providing increased availability of citrate and Ac-CoA. Increased glucose oxidation is also supported by increased activity and expression of glycolytic genes by Akt signaling. Coupled with an LPS-induced increase in ACLY activity downstream of Akt, enhanced production of Ac-CoA drives histone acetylation and induction of inflammatory genes like Il6 and Il1b. As oxidation of carbon substrates in the TCA cycle depends on the availability of electron acceptors such as NAD+, the cyclic reduction and oxidation of these molecules serves as an important link between oxidative metabolism in the TCA cycle and electron transport chain (ETC) activity. (b) During LPS re-stimulation of tolerant macrophages, glucose oxidation is impaired due to mitochondrial NAD+/NADH imbalance driven by GPD2-dependent reverse electron transport at complex I. Induction of primed genes (P) in this setting is fueled by Ac-CoA produced from acetate via the activity of ACSS2. Deacetylation of histones at other gene promoters, such as those of tolerized genes like Il6 and Il1b (T), contributes to the pool of acetate fueling histone acetylation and induction of P genes such as Fpr1 and Ptges.

!"#$%$#

&'(")*#

&'+")'+*,*

-./

!"#$%

!!! !& &

!"0/)!

./!

1,*$)2#3!"#$+'%$,)2

456

'($)

&'+"#7)'809:)*9:%$#6:($$'#

!!

'($*

!;.

. 5

!/66<

!"#$%&'

!"($%"(&)&

*+,

!"#$%

!!! !& &

-$.,%-

+,-

/)&0%1'2-$'0("30)%1

456

'($)

!"($'7%"8.9:%&9:30'6:#00"'

-,4;!!

'($*

-<+

+ 5

a

b

Page 130: Regulation of Inflammatory Macrophage Activation and

116

REFERENCES

1. Hotchkiss,R.S.&Karl,I.E.ThePathophysiologyandTreatmentofSepsis.NewEnglandJournalofMedicine348,138–150(2003).

2. Boomer,J.S.etal.Immunosuppressioninpatientswhodieofsepsisandmultipleorgan

failure.JAMA306,2594–2605(2011).3. Hotchkiss,R.S.etal.Sepsisandsepticshock.NatRevDisPrimers2,16045(2016).4. Foster,S.L.,Hargreaves,D.C.&Medzhitov,R.Gene-specificcontrolofinflammationby

TLR-inducedchromatinmodifications.Nature447,972–978(2007).5. Foster,S.L.&Medzhitov,R.Gene-specificcontroloftheTLR-inducedinflammatory response.ClinicalImmunology130,7–15(2009).6. Hotchkiss,R.S.,Monneret,G.&Payen,D.Sepsis-inducedimmunosuppression:from

cellulardysfunctionstoimmunotherapy.NatureReviewsImmunology13,862–874(2013).

7. Hotchkiss,R.S.,Coopersmith,C.M.,McDunn,J.E.&Ferguson,T.A.Thesepsisseesaw: tiltingtowardimmunosuppression.Nat.Med.15,496–497(2009).8. O’Neill,L.A.J.,Kishton,R.J.&Rathmell,J.Aguidetoimmunometabolismfor immunologists.NatureReviewsImmunology16,553–565(2016).9. Langston,P.K.,Shibata,M.&Horng,T.MetabolismSupportsMacrophageActivation. FrontiersinImmunology8,(2017).10. Jung,J.,Zeng,H.&Horng,T.Metabolismasaguidingforceforimmunity.Nat.CellBiol.

21,85–93(2019).11. Ryan,D.G.etal.CouplingKrebscyclemetabolitestosignallinginimmunityandcancer. NatureMetabolism1,16(2019).12. O’Neill,L.A.J.&Pearce,E.J.Immunometabolismgovernsdendriticcelland

macrophagefunction.TheJournalofExperimentalMedicine213,15–23(2016).13. Kaelin,W.G.&McKnight,S.L.InfluenceofMetabolismonEpigeneticsandDisease.Cell 153,56–69(2013).14. Fayard,E.,Xue,G.,Parcellier,A.,Bozulic,L.&Hemmings,B.A.ProteinkinaseB

(PKB/Akt),akeymediatorofthePI3Ksignalingpathway.Curr.Top.Microbiol.Immunol.346,31–56(2010).

Page 131: Regulation of Inflammatory Macrophage Activation and

117

15. Everts,B.etal.TLR-drivenearlyglycolyticreprogrammingviathekinasesTBK1-IKKɛsupportstheanabolicdemandsofdendriticcellactivation.NatureImmunology15,323–332(2014).

16. Comerford,S.A.etal.AcetateDependenceofTumors.Cell159,1591–1602(2014).17. Bulusu,V.etal.AcetateRecapturingbyNuclearAcetyl-CoASynthetase2PreventsLoss

ofHistoneAcetylationduringOxygenandSerumLimitation.CellReports18,647–658(2017).

18. Mertins,P.etal.AnIntegrativeFrameworkRevealsSignaling-to-TranscriptionEventsin

Toll-likeReceptorSignaling.CellReports19,2853–2866(2017).19. Seeley,J.J.etal.InductionofinnateimmunememoryviamicroRNAtargetingof

chromatinremodellingfactors.Nature559,114–119(2018).20. DeVirgilio,C.etal.Cloninganddisruptionofagenerequiredforgrowthonacetatebut

notonethanol:theacetyl-coenzymeAsynthetasegeneofSaccharomycescerevisiae.Yeast8,1043–1051(1992).

21. Takahashi,H.,McCaffery,J.M.,Irizarry,R.A.&Boeke,J.D.Nucleocytosolicacetyl-

coenzymeasynthetaseisrequiredforhistoneacetylationandglobaltranscription.Mol.Cell23,207–217(2006).

22. vandenBerg,M.A.etal.Thetwoacetyl-coenzymeAsynthetasesofSaccharomyces cerevisiaedifferwithrespecttokineticpropertiesandtranscriptionalregulation.J.Biol. Chem.271,28953–28959(1996).23. Lipmann,F.,Jones,M.E.,Black,S.&Flynn,R.M.ThemechanismoftheATP-CoA-

acetatereaction.JCellPhysiolSuppl41,109–112(1953).24. Kamphorst,J.J.,Chung,M.K.,Fan,J.&Rabinowitz,J.D.Quantitativeanalysisofacetyl-

CoAproductioninhypoxiccancercellsrevealssubstantialcontributionfromacetate.Cancer&Metabolism2,23(2014).

25. McDonnell,E.etal.LipidsReprogramMetabolismtoBecomeaMajorCarbonSourcefor HistoneAcetylation.CellReports17,1463–1472(2016).26. Liu,X.etal.AcetateProductionfromGlucoseandCouplingtoMitochondrial

MetabolisminMammals.Cell175,502-513.e13(2018).27. McBrian,M.A.etal.HistoneAcetylationRegulatesIntracellularpH.MolecularCell49,

310–321(2013).

Page 132: Regulation of Inflammatory Macrophage Activation and

118

28. Hallows,W.C.,Lee,S.&Denu,J.M.Sirtuinsdeacetylateandactivatemammalian acetyl-CoAsynthetases.PNAS103,10230–10235(2006).29. Lachmandas,E.etal.MicrobialstimulationofdifferentToll-likereceptorsignalling pathwaysinducesdiversemetabolicprogrammesinhumanmonocytes.Nature Microbiology2,(2017).30. Lee,J.V.etal.Akt-dependentmetabolicreprogrammingregulatestumorcellhistone acetylation.CellMetab.20,306–319(2014).31. Covarrubias,A.J.etal.Akt-mTORC1signalingregulatesAclytointegratemetabolic

inputtocontrolofmacrophageactivation.Elife5,(2016).32. Drapier,J.C.&Hibbs,J.B.Differentiationofmurinemacrophagestoexpressnonspecific

cytotoxicityfortumorcellsresultsinL-arginine-dependentinhibitionofmitochondrialiron-sulfurenzymesinthemacrophageeffectorcells.J.Immunol.140,2829–2838(1988).

33. Cordes,T.etal.ImmunoresponsiveGene1andItaconateInhibitSuccinate

DehydrogenasetoModulateIntracellularSuccinateLevels.JournalofBiologicalChemistry291,14274–14284(2016).

34. Lampropoulou,V.etal.ItaconateLinksInhibitionofSuccinateDehydrogenasewith

MacrophageMetabolicRemodelingandRegulationofInflammation.CellMetabolism24,158–166(2016).

35. Mills,E.L.etal.SuccinateDehydrogenaseSupportsMetabolicRepurposingof

MitochondriatoDriveInflammatoryMacrophages.Cell167,457-470.e13(2016).36. John,S.,Weiss,J.N.&Ribalet,B.SubcellularlocalizationofhexokinasesIandIIdirects

themetabolicfateofglucose.PLoSONE6,e17674(2011).37. Miyamoto,S.,Murphy,A.N.&Brown,J.H.Aktmediatesmitochondrialprotectionin

cardiomyocytesthroughphosphorylationofmitochondrialhexokinase-II.CellDeathDiffer.15,521–529(2008).

38. Ip,W.K.E.,Hoshi,N.,Shouval,D.S.,Snapper,S.&Medzhitov,R.Anti-inflammatory

effectofIL-10mediatedbymetabolicreprogrammingofmacrophages.Science356,513–519(2017).

39. Byles,V.etal.TheTSC-mTORpathwayregulatesmacrophagepolarization.NatCommun

4,2834(2013).

Page 133: Regulation of Inflammatory Macrophage Activation and

119

40. Kong,Y.Btrim:afast,lightweightadapterandqualitytrimmingprogramfornext-generationsequencingtechnologies.Genomics98,152–153(2011).

41. Kim,D.etal.TopHat2:accuratealignmentoftranscriptomesinthepresenceof

insertions,deletionsandgenefusions.GenomeBiol.14,R36(2013).42. Love,M.I.,Huber,W.&Anders,S.Moderatedestimationoffoldchangeanddispersion

forRNA-seqdatawithDESeq2.GenomeBiol.15,550(2014).43. Chen,E.Y.etal.Enrichr:interactiveandcollaborativeHTML5genelistenrichment

analysistool.BMCBioinformatics14,128(2013).44. Kuleshov,M.V.etal.Enrichr:acomprehensivegenesetenrichmentanalysiswebserver 2016update.NucleicAcidsRes.44,W90-97(2016).

Page 134: Regulation of Inflammatory Macrophage Activation and

CHAPTER 4

Discussion and Future Directions

Page 135: Regulation of Inflammatory Macrophage Activation and

121

Summary and Conclusions

Systemic microbial infection in vertebrate animals elicits an acute systemic inflammatory response

syndrome, characterized by activation of the innate and adaptive arms of the immune cells and

elevated local and circulating pro-inflammatory cytokines (e.g. IL-6, IL-1β, and TNFα) and

enhanced antimicrobial activity1. This robust pro-inflammatory response is followed by prolonged

immunoparalysis, characterized by general immunosuppression due to decreased pro-

inflammatory cytokine production, impaired antigen presentation and costimulation, and marked

T cell anergy2–5. Immunosuppression in this context is believed to be a compensatory mechanism

to limit damage to the host. However, protracted immunosuppression contributes a high percentage

of sepsis-related hospital deaths6. Central to coordinating the transition from activation to

suppression are macrophages, which sense pathogen associated molecular patterns (PAMPs) such

as lipopolysaccharide (LPS). Macrophages mount an initial pro-inflammatory response followed

by a persistent state of tolerance to subsequent bacterial challenges. Phase-two clinical trials aimed

at suppressing inflammation during sepsis have been unsuccessful, turning attention toward

preventing immunoparalysis7. Thus, elucidating the mechanisms underpinning the critical

inflection point in the biphasic response curve during sepsis could have significant therapeutic

potential. Given the recent understanding that specific metabolic shifts critically support

immunological recognition of infection and execution of effector functions, we pursued a

metabolic underpinning to the biphasic response to microbial encounter.

Here, we uncovered and characterized several metabolic modules regulating the induction and

suppression of gene-specific macrophage inflammatory responses to microbial encounter in vitro

and in vivo. In Chapter 2, we carefully measured the time-dependent changes in metabolite

Page 136: Regulation of Inflammatory Macrophage Activation and

122

concentrations and mitochondrial respiration over the course of macrophage exposure to LPS and

revised the current paradigm for how mitochondrial metabolism regulates inflammatory cytokine

production, demonstrating that glucose oxidation initially supports but later suppresses induction

of inflammatory genes through a metabolic pathway previously undescribed in macrophages. First,

we showed that the link between glucose oxidation and transcription was by modulation of

glucose-derived carbon flux through the TCA cycle and into the nucleocytosolic pool of acetyl-

CoA (Ac-CoA), which serves as the substrate for transcription-activating histone acetylation at

inflammatory gene promoters. We also show that the increase in glucose-derived Ac-CoA during

activation depends on the enhanced activity of the enzyme ATP-citrate lyase (ACLY), which

catalyzes the conversion of mitochondria-derived citrate to Ac-CoA and oxaloacetate. Next, we

identified and characterized a novel role for the glycerol phosphate shuttle (GPS) and its limiting

enzyme GPD2 (mitochondrial glycerol 3-phosphate dehydrogenase, mGPDH) in supporting the

biphasic response to LPS by driving the acute LPS-inducible increase in glucose oxidation while

also contributing to the shutdown of respiration characteristic of macrophages subjected to

prolonged LPS exposure. Based on data obtained from experiments in bone-marrow-derived

macrophages (BMDMs) from newly-generated Gpd2-/- (KO) mice, we attributed the switch from

increased to decreased glucose oxidation during the transition from LPS activation to tolerance to

GPD2-dependent induction of reverse electron transport (RET). Indeed, based on specific changes

in levels of complex I (CI)-associated superoxide and NADH, we proposed a model in which GPS

activity, by delivering glucose-derived electrons to the electron transport chain (ETC), initially

boosts oxidative metabolism to support increased acetyl-CoA (Ac-CoA) production and histone

acetylation during LPS activation, while sustained GPS activity during prolonged LPS exposure

overwhelms forward electron transport capacity leading to RET, thus reducing TCA cycling to

Page 137: Regulation of Inflammatory Macrophage Activation and

123

limit Ac-CoA production and inflammatory gene induction (Supplementary Fig. 2.12). Finally,

the importance of GPD2-dependent regulation in the host response to microbial infection was

demonstrated by using an in vivo model of LPS tolerance, in which a subletal dose of LPS protects

against septic shock triggered by a subsequent, lethal dose of LPS. In contrast to mice that did not

receive the sub-lethal LPS pretreatment, mice that receive the pretreatment profoundly suppress

IL-6 production during septic shock. Such suppression was attenuated in GPD2 KO mice, leading

to enhanced IL-6 levels and exaggerated endotoxin-induced hypothermia. GPD2 KO mice that

received sublethal LPS preconditioning also succumbed to subsequent lethal LPS challenge, while

no lethality was observed in WT mice. Therefore, we illuminate the GPS is a hitherto

unappreciated and unique regulator of LPS responses, using the ETC as a rheostat to couple

duration of LPS exposure to ETC directionality to modulate a balance between the induction and

suppression of inflammatory responses.

In Chapter 3, we extended our model for how GPD2-dependent glucose oxidation couples the

duration of LPS exposure to induction and suppression of inflammatory genes by introducing an

additional regulatory node at the nexus between TLR signaling and metabolism. We also identified

a subset of LPS-inducible genes that displayed an expression pattern opposite to tolerance and

were supported by a parallel metabolic module to glucose oxidation. Specifically, we showed that

the serine/threonine kinase Akt (protein kinase B, PKB), a known target downstream of TLR

signaling8, integrated glucose oxidation with the strength and duration of LPS stimulation to

control the amplitude and duration of inflammatory gene induction. We also identified a subset of

genes, constituting 8% of the LPS-inducible transcriptome picked up by our analysis, that became

hyper-responsive following prolonged LPS exposure, in contrast with the hypo-responsiveness of

Page 138: Regulation of Inflammatory Macrophage Activation and

124

inflammatory genes that is characteristic of the LPS tolerant state. These primed genes, or “P

genes,” encode molecules important for biological processes distinct from tolerant genes.

Moreover, consistent with the shift from increased to decreased oxidative metabolism over the

course of LPS exposure as macrophages transition from activation to tolerance, P genes were not

directly supported by glucose oxidation; rather, we showed that induction of P genes was supported

by acetate-derived Ac-CoA production catalyzed by nucleocytosolic acetyl-CoA synthetase

(AceCS1, or ACSS2). Moreover, we highlighted histone deacetylation as a source of acetate

fueling this process. Finally, we proposed that in addition to a role of Akt in the induction and

suppression of inflammatory genes through control of oxidative metabolism and thus citrate-

derived Ac-CoA levels, Akt may also regulate P genes through control of ACSS2 by

phosphorylation at conserved serine residues (Supplementary Fig. 3.5). Thus, we provide the first

evidence of a role for ACSS2 in macrophage biology and highlight Akt and ACSS2 as additional

regulatory modules integrating TLR signaling with shifts in macrophage metabolism to induce

gene-specific responses to microbial encounter.

Lessons from a worked example in immunometabolism

Exploration into how shifts in intermediary metabolism support the four tasks of the immune

system – recognition, effector functions, regulation, and memory – has widened the lens through

which we view host defense and has led to the genesis of the relatively-new subfield of

immunology called immunometabolism9. However, immunometabolism still largely sits at the

fringes of its two parent disciplines, not biochemical enough for the metabolic biologists and not

immunobiologically-important enough for classical immunologists. Of course there are papers that

stand as exceptions. We like to think that our work is included with these exceptions.

Page 139: Regulation of Inflammatory Macrophage Activation and

125

Recall that in Chapter 1, the foundational relationship between metabolism and cell functions was

described from a thermodynamic perspective. It was iterated that animate organisms are

constrained by the same laws that govern the behavior of inanimate chemicals and that the

reactions of such heterogeneous substances in an organism’s global metabolic network take on

measurable changes in flux in response to the application of stress. Modern reluctance to accept

the critical importance of metabolic shifts in the optimal execution of specific stress responses is

partially a consequence of the focus of early work in biochemistry and partially a result of poor

education in metabolism. The 19th and 20th centuries brought tremendous progress towards our

understanding of biochemistry, as basic principles and tools of chemistry were applied to the study

of heterogeneous substances within biological systems and how specific behaviors of such

substances could support life. Such understanding has been crammed into a conceptual closet

labeled “cell maintenance” by most until recent years (excluding perhaps the physiologists and

some cancer biologists). Our worked example of how glucose oxidation regulates macrophage

inflammatory responses is one with the potential to change this way of thinking, as it takes well-

described macrophage effector functions during LPS activation and tolerance and systematically

demonstrates regulation by modules in intermediary metabolism (Ac-CoA synthesis, TCA cycle,

and GPS). A summary of this example and its broader application to other fields of study follows.

In our example, consider the macrophage as an open system. LPS is the imposed stress on the

system, and the gene-specific inflammatory response is the system’s functional requirement of

highest priority. TLR signaling, although of absolute necessity for transducing the stress into the

functional response in this system (see Chapter 1), is excluded from this example because of our

data demonstrating glucose oxidation as a layer of regulation downstream of TLR signaling acting

Page 140: Regulation of Inflammatory Macrophage Activation and

126

as a rheostat for the functional response rather than as a regulator of TLR signaling itself. At the

foundation of this functional response to stress are the chromatin modifications that have been

described as necessary events during gene-specific activation and tolerance10–12. We demonstrated

that such modifications depend on the metabolic status of the system, specifically on the flux of

external glucose-derived carbons into the system, through glycolysis and the TCA cycle, into the

nucleocytosolic pool of Ac-CoA, and into histones. Furthermore, we demonstrated that selectively

blocking this flux path by pharmacologic or genetic manipulation results in attenuation of the gene-

specific functional response. Similarly, tolerance was revealed to depend on this flux path such

that rescue of responsiveness of the system to stress was achieved by protection of glucose

oxidation and gene-specific histone acetylation. Such regulation of specific functional responses

through the influence of metabolism on chromatin via changes in acetylation is a mode of

metabolic control seen in other systems13–15 and in response to other stresses16–19. Moreover,

modification of chromatin by metabolites other than Ac-CoA has also been shown to integrate

metabolism with functional shifts in the face of imposed stress, including histone demethylation

by FAD-dependent LSDs and AKG-dependent JmjC demethylases and DNA demethylation by

TET proteins20–22. Although not investigated in the work in Chapters 2 and 3, post-translational

modifications of non-histone proteins by spontaneous and enzyme-catalyzed reactions with

metabolites is a general principle from biochemistry and is an important mode of regulation to

consider when thinking about how metabolism might influence a system’s specific stress

response23. Indeed, examples in inflammatory macrophages have recently been reported24,25.

The enzyme- and substrate-dependent synthesis of metabolites fueling modifications to histones

and other proteins is another layer of metabolic regulation of specific functional responses

Page 141: Regulation of Inflammatory Macrophage Activation and

127

described by our work and others. In Chapter 2, we show that activation of the enzyme ACLY by

phosphorylation at S455 is necessary for supporting the induction of inflammatory genes in

response to LPS. Such regulation has previously been described in alternatively activated

macrophages16 and is likely dependent on Akt, given our data in Chapter 3 and work in cancer

cells26. Moreover, we demonstrate that different substrate-enzyme-product axes may support

distinct functional responses, supported by our evidence that acetate-ACSS2-Ac-CoA rather than

citrate-ACLY-Ac-CoA regulates the induction of P genes. Such differences in substrate

dependency may seem esoteric; however, they provide valuable insight into the system at the time

of observation, highlighting prioritized functions (i.e. those of P genes versus T genes) and

metabolic status (i.e. glucose oxidation off versus on). Illuminating these system properties

informs the design of interventions to support or inhibit functions of the system in the

corresponding context. For example, one might target acetate metabolism in cancer cells that have

switched to acetate-dependent histone acetylation under nutrient-limited conditions resembling a

tumor microenvironment27–30. In our macrophage example, limiting glucose oxidation strongly

attenuated inflammatory gene induction during activation and enhanced responsiveness to

restimulation in tolerant macrophages, underpinning a profound lack of protection against a lethal

dose of LPS administered to GPD2 KO mice even after induction of tolerance using a sublethal

dose of LPS. The observations from our mouse experiments are comparable to the protective

effects of carbohydrate restriction and calorie restriction in mice given a systemic bacterial

infection31.

Further upstream of Ac-CoA production, our work revealed that GPD2 and the GPS drive glucose

oxidation to induce different functional states at different times following application of stress in

Page 142: Regulation of Inflammatory Macrophage Activation and

128

the form of LPS, initially supporting activation and later contributing to tolerance. This mechanism

provides an example of a single metabolic module regulating a dynamic response over time. Such

regulation is consistent the concept of metabolic shifts being calibrated to support a stress-specific

functional demand placed on a system. GPD2-dependent glucose oxidation follows this thinking,

as induction of pro-inflammatory genes is initially supported by increased GPS activity to enhance

Ac-CoA production and histone acetylation, while suppression of inflammatory gene induction is

achieved by respiration shutdown to limit the availability of citrate for ACLY-dependent Ac-CoA

production and thus attenuate histone acetylation. Temporal regulation of gene-specific functions

by metabolism is also found in the tolerant, restimulated system, as HDAC activity appears to fuel

induction of P genes after suppression of T genes. Again, this metabolic shift reflects a change in

priority of cellular functions. It is also possible that the increase in HDAC activity in the tolerant

state serves the added role of countering the intracellular acidification that results from enhanced

lactate production32. Finally, we observed that many P genes encode subunits of CI and CV of the

ETC (Supplementary Fig. 3.3). We postulate that this is a mechanism of reversing impaired

mitochondrial oxidative capacity to restore responsiveness of inflammatory genes after complete

elimination of microbial infection. Thus, our work provides an example of how a single metabolic

module and clusters of metabolic modules support time-dependent changes in the functional

requirements of a system in the face of a prolonged stress. If the goal of one’s work is to reorganize

the hierarchy of functions in a system over time (i.e. restore responsiveness of inflammatory genes),

then identification and targeting of such metabolic modules is an option. Another recent example

of exploiting this modality was in the reversal of LPS tolerance by treating human monocytes with

β-glucan ex vivo after tolerization in vivo. Rescue was achieved presumably by reversal of

itaconate-mediated impairment of respiration, although additional evidence is needed33.

Page 143: Regulation of Inflammatory Macrophage Activation and

129

Finally, the dynamic regulation of inflammatory responses by glucose metabolism in macrophages

assigns novel biological importance to fundamental principles of biochemistry and bioenergetics

that have not been implicated in many settings but constitute foundational mechanisms that may

play a role in a wide array of biological systems. Specifically, we showed the critical importance

of sufficient electron acceptor molecules in the oxidation of carbon substrates to support induction

of histone acetylation and expression of inflammatory genes. Indeed, deployment of the GPS was

necessary for maximal induction of glucose uptake and oxidation in LPS-activated BMDMs,

consistent with its role as a redox shuttle in other tissues requiring high rates of glucose

utilization34,35. Moreover, histone acetylation and inflammatory gene induction were rescued in

tolerant macrophages treated with the NAD+ precursor nicotinamide (NAM) (Supplementary Fig.

2.9), supporting the need for electron acceptors for TCA cycling proposed by Williamson and

Krebs36 and demonstrated recently in cancer cells37. Additionally, we demonstrated that GPD2-

driven respiration can lead to RET in macrophages, underpinning its anti-inflammatory activity in

the tolerant state. Indeed, GPD2-driven RET limits inflammatory gene induction by inducing

accumulation of NADH at the expense of NAD+ return to the TCA cycle, thus impinging on

substrate oxidation in the mitochondrial matrix. The switch from FET to RET is a consequence of

thermodynamics and could actually be calculated if substrate oxidation rates and ETC component

concentrations in the system were known. As mentioned in Chapter 1, the mitochondrial

respiratory chain can be ordered based on the Gibbs energies of its components; however, more

conventional is organization based on redox potentials (Eh). RET is favored under conditions of

high proton motive force (Δp), due to an increase in one or both of its components. Shown in the

mathematical representation of Δp at 37°C in Eq. 4.1, these components are chemical (ΔpH) and

Page 144: Regulation of Inflammatory Macrophage Activation and

130

electrical (ΔΨ) and result from differences in protons and differences in electrical potential across

the mitochondrial inner membrane, respectively38.

∆" = ∆Ψ − 61∆pH (Equation 4.1)

Prolonged exposure to LPS increases Δp in macrophages, as measured by increased ΔΨ (data not

shown and 39); however, as GPD2 is poorly coupled to the proton circuit across the inner membrane

because its redox activity is not coupled to translocation of protons, it acts as a continuous source

of electron flow from oxidation of glycerol 3-phosphate (G3P) despite high Δp. This leads to over-

reduction of the mobile electron carrier ubiquinone (Q), resulting in an Eh for Q/QH2 that favors

reverse electron flow through CI to reduce NAD+ to NADH. Although the enhanced expression

and activity of GPD2 in LPS-stimulated macrophages favors RET through this mechanism, RET

can be driven by other ETC components feeding into Q, as well (except for CI)40. In addition to

modulation of the NAD+/ NADH ratio, RET is also a source of superoxide through electron leak

at the Q-facing site of CI, or site IQ41,42. Thus, induction of RET by substrate feeding into the Q

pool under conditions of high Δp represents a mechanism by which oxidative metabolism and

bioenergetics may act on specific biological functions. Although isolated mitochondria from

several cells and tissues have been shown to produce RET ROS43–47, there are currently only a few

settings in mammalian cells in which RET has been implicated, including LPS-stimulated

macrophages (Fig. 2.13, 2.14, and 39), tissues subjected to ischemia-reperfusion injury40,48, and in

primary astrocytes49. Therefore, RET remains an under-investigated mechanism for metabolic

regulation of cellular and tissular stress responses.

Page 145: Regulation of Inflammatory Macrophage Activation and

131

Concluding Thoughts and Future Directions

Every scientist can appreciate the joy of being the first to observe a phenomenon of the natural

world. To conceive, test, and realize an idea is a euphoric experience that justifies the negative

emotions elicited by “failed” experiments, lost sleep, and monetary and technical hurdles. In the

present case, the work shown and discussed in Chapters 2 and 3 started with the simple thesis,

“There must be a metabolic basis of endotoxin tolerance.” This idea was born out of an obsession

with immunological memory that likely stemmed from the first primary papers50–53 I read during

my undergraduate training with Jason M. Grayson. To have realized this thesis and to have, in the

process, uncovered the elegant mechanisms of metabolic regulation of inflammatory macrophage

activation and tolerance discussed in previous Chapters, has been a remarkable experience.

However, exploration into the unknown illuminates many more areas in need of investigation. The

sections that follow will focus on describing these areas in addition to providing suggestions for

future experiments to strengthen the conceptual framework built by the present work.

When discussing the observations from Chapter 2 in seminars or with colleagues, it is often asked

“Is this true for stimuli besides LPS?” This is a fine question and one that, at present, can only be

answered with a soft “yes.” In a collaboration with Linden Hu’s Laboratory at Tufts University

School of Medicine, we have shown that glucose oxidation is initially enhanced and later repressed

following exposure of BMDMs to the TLR2 ligand Pam3Cys-Ser-(Lys)4 (PamCSK) or to the live

bacteria known for causing Lyme Disease, Borrelia burgdorferi (Bb). Changes in promoter region

histone acetylation are important for inflammatory gene induction and suppression during

activation and tolerance following PamCSK or Bb stimulation of BMDMs, and impairing glucose

oxidation by inhibiting glucose utilization (2DG) or GPS activity (GPD2 KO) results in rescue of

Page 146: Regulation of Inflammatory Macrophage Activation and

132

histone acetylation and inflammatory gene induction in tolerant BMDMs comparable to that

observed in our LPS tolerance experiments. Much of this work was performed by Urmila Powale

at Tufts and is now a primary focus in the Hu Laboratory. Although these observations suggest

that glucose oxidation is a conserved metabolic requirement for inflammatory gene induction

downstream of TLR signaling, it will be necessary to test other TLR ligands. Additionally, whether

glucose oxidation is a requirement for optimal induction of macrophage transcriptional responses

downstream of sensing through other pathogen recognition receptors (PRRs) or cytokine receptors

(i.e. IL-4R, IL-13R, IL-10R) is also of great interest. Indeed, the phenotypes of macrophages

exposed to different signals fall along a spectrum as a function of the specific signal54,55 and the

shifts in metabolism56,57, but just as transcriptional profiling has revealed cores of genes conserved

between subsets and phenotypes of immune cells, there must also be core metabolic pathways

shared between macrophages exposed to different signals. Supportive of this concept, BMDMs

alternatively activated by IL-4 stimulation have an acute increase in Akt-dependent glucose

utilization similar to that observed in macrophages classically activated by LPS16. In contrast,

modulation of whole-body metabolism by dietary manipulation results in opposing effects of

fasting and feeding on antibacterial and antiviral responses in mice31. As GPS activity is primarily

controlled by selective expression of Gpd234, discovery of additional settings in which enhanced

glucose oxidation is enhanced could be initiated by screening a library of PRR ligands and

cytokines for the ability to induce Gpd2 expression in BMDMs at different times after stimulation.

Based on the examples provided above, a reductionist approach would need to be applied in each

case of enhanced Gpd2 expression to verify that GPS activity is also increased and supports

specific effector functions (i.e. glucose tracing into glycerol 3-phosphate (G3P), Seahorse analysis

of glucose oxidation, and measurement of effector functions in GPD2-deficient BMDMs and mice).

Page 147: Regulation of Inflammatory Macrophage Activation and

133

In addition to exploring the role of GPD2-dependent glucose oxidation in macrophages polarized

to other states, the importance of GPD2 in other cells is also an area of investigation illuminated

by our work. As stated above, GPD2 and GPS activity are regulated by tissue-specific expression

of Gpd234. There are currently only a few biological contexts in mammals and mammalian cells

in which GPD2 has been implicated, including insulin secretion by pancreatic β-cells58, liver

gluconeogenesis59,60, regulation of blood glucose and serum glycerol and triglyceride levels in

female mice61, and the growth of certain cancers62. There is also evidence of time-dependent

changes in Gpd2 expression in skeletal muscle following contraction63–66 and in T cells following

activation67. In the majority of these settings, enhanced GPD2 activity resulting from increased

expression is functionally linked to specific biological functions through modulation of

mitochondrial reactive oxygen species (ROS) levels.

The oxidation of G3P was first shown in tissue preparations by Meyerhof in 1919, while the

glycerol phosphate shuttle was not proposed until 195868,69 after characterization of the cytosolic

and mitochondrial glycerol 3-phosphate dehydrogenases. Tissue-restricted expression of GPD2 is

thought to be an adaptation to minimize ROS production in tissues that do not require high rates

of glucose oxidation35. Although it is unclear how ROS produced at different sites within the ETC

lead to distinct biological functions, it is worth noting that G3P oxidation by GPD2 can drive ROS

production at multiple locations, including its own F and Q domains, as well as site IQ from RET43.

One may postulate that production of superoxide towards the outer versus inner side of the

mitochondrial inner membrane (i.e. by GF versus GQ, respectively) could have different biological

effects based on the differential availability and chemical character of substrates and detoxifying

enzymes. However, recall that the importance of GPD2 in the regulation of macrophage

Page 148: Regulation of Inflammatory Macrophage Activation and

134

inflammatory responses was not attributed to RET ROS but rather to induction of redox imbalance

due to reduction of NAD+ at CI, a lesser-appreciated yet fundamental consequence of RET70. This

mechanism of regulation was further supported by the observation that treating BMDMs with site

IQ electron leak inhibitors (S1QELs)45,49 during tolerization to LPS did not rescue inflammatory

gene induction upon restimulation (data not shown). Preventing RET in LPS-stimulated BMDMs

by ectopic expression of the alternative oxidase (AOX) as was done previously39, but effects on

tolerance were not measured. As AOX activity would prevent both RET ROS and NAD+ reduction

at CI, it is likely that AOX-expressing macrophages would be resistant to LPS-induced tolerance.

It would be exciting to induce endotoxin tolerance in WT and AOX-expressing mice and determine

their level of protection against a subsequent lethal dose of LPS, similar to what was done in GPD2

KO mice in Chapter 2.

Of course, changes in carbon flux and in ROS levels as a consequence of enhanced GPD2 activity

could play concomitant roles in coupling changes in oxidative metabolism with the execution of

specific effector functions. Taking again the example of time-dependent changes in GPD2 activity

in the skeletal muscle following contraction, the acute increase63,64 and subsequent decrease65 in

activity could be to support enhanced mitochondrial ATP production during the period in which

contractions are being elicited, and such oxidative metabolism could enhance ROS from multiple

sites in the ETC, inducing a hormetic response through redox-sensitive signaling networks.

Thermogenesis by BAT is another context in which GPD2 activity is presumed to be important –

cold exposure increases GPD2 expression34, and mice lacking GPD2 have 15% lower resting

energy expenditure at thermoneutrality61 – however, a mechanism for how GPD2 is supporting the

Page 149: Regulation of Inflammatory Macrophage Activation and

135

thermogenic activity of brown adipose tissue (BAT) is lacking71. Thus, elucidating the importance

and regulatory interaction of GPD2 within other biological contexts is an open area of investigation.

Current knowledge on the transcriptional regulation of GPD2 is limited. GPD2 is encoded by a

single gene on chromosome 2 of humans and 3 of rodents, containing approximately 2432 base

pairs and 17 exons34,35. There are three different first exons of Gpd2: 1a, found in brain; 1b, found

in all tissues; and 1c, found predominantly in the testis35. The 1b exon contains a thyroid hormone

receptor responsive element, which may explain the modulation of GPD2 activity in BAT and

skeletal muscle in rodents given thyroid hormones and during hibernation72. GPD2 protein

contains 727 amino acid residues, which constitute FAD-, G3P-, and calcium-binding domains34.

In Chapter 3, we proposed that LPS-inducible GPD2 and GPS activity may be Akt-dependent, as

pretreatment with AKTi attenuated attenuated Gpd2 expression during LPS activation. This result

should be validated in genetic models of macrophage Akt-deficiency, as have been used in the

study of polarization previously73,74. As a more comprehensive investigation into the path

connecting TLR signaling with induction of Gpd2 expression, it would be useful to systematically

test the importance of known modules coupling TLR signaling with transcription events75 through

mutagenesis or knock-out studies. Such studies could even be performed by mining existing

publically-available sequencing data. It is worth noting that uncovering common signaling

modules integrating TLR engagement with enhanced Gpd2 expression and GPS activity would

allow generalization across ligands and between immune cell subsets.

As final commentary on the biology surrounding the GPS, our discovery of the importance of

GPD2 in control of macrophage inflammatory responses presents a novel regulatory module that

Page 150: Regulation of Inflammatory Macrophage Activation and

136

can be targeted by an array of interventions to modulate the host’s response to microbial encounter.

Most appealing is the off-label use of the diabetes drug metformin as an inhibitor of GPD2. Indeed,

although metformin’s glucose-lowering effects have been attributed to several mechanisms,

including inhibition of CI76, some reports suggest that GPD2 may also be inhibited by

metformin59,60. Therefore, the well-known anti-inflammatory effects of metformin77–80 may be

underpinned by antagonism against GPD2. It would be useful to test this in a cell autonomous

system by performing Seahorse extracellular flux analysis of G3P-driven respiration in

permeabilized BMDMs untreated or treated with metformin and stimulated with LPS for different

times. As oral administration of metformin leads to GPD2 inhibition in the liver, where it is first

metabolized, alternative modes of delivery would better suit its anti-inflammatory application. For

example, a slow-release injectable formulation was recently developed and tested for antitumor

effects81. Injection of metformin into the joints of an animal or human with Lyme arthritis due to

late-stage Bb infection may have significant anti-inflammatory effects on resident immune cells

by inhibiting GPD2 activity and glucose oxidation.

As an alternative approach to inhibiting GPD2 activity directly, proximal processes affected by

increased GPD2 activity may also be therapeutically targeted to modulate inflammatory responses.

In the case of GPD2-dependent inflammatory gene suppression, we identified GPD2-driven RET

as a mechanism underpinning shutdown of respiration and restriction of TCA cycling due to

NAD+/NADH imbalance at CI. Recall that RET involves electron flow from a source of oxidized

substrate (i.e. G3P) to the Q/QH2 pool, which at a sufficient Eh under high Δp is able to donate

electrons thermodynamically “uphill” to CI, driving reduction of NAD+ and electron leak onto O2

at IQ to produce superoxide. Although not measured, it is possible that the concentration of

Page 151: Regulation of Inflammatory Macrophage Activation and

137

available Q molecules in the inner membrane space is limiting during prolonged LPS exposure

and that modulation of the Q pool may protect against induction of RET by sustained GPD2

activity. Indeed, it has been proposed that Q deficiency plays a role in a number of human diseases

and that supplementation with coenzyme Q10 (CoQ) can correct such deficiency and ameliorate

disease symptoms82. Delivery of CoQ is difficult due to its exceptionally high hydrophobicity.

Preliminary experiments in our cell autonomous system for studying macrophage LPS tolerance

did not show protection of inflammatory gene induction after prolonged LPS exposure when high-

dose CoQ was supplied (data not shown), even when handled as described previously83. It will be

important to quantify the pool size and redox status of the Q/QH2 pair in macrophages over an

LPS timecourse to determine whether there is a deficiency that could be corrected by CoQ

supplementation, as such knowledge would inform the design of future animal experiments and

would also strengthen the model of LPS-induced RET in macrophages. Moreover, if CoQ

supplementation is shown to prevent RET and thus protect against LPS tolerance in vitro and/or

in vivo, it would suggest that RET may play a role in pathologies resulting from CoQ deficiency

in other settings such as diet-induced obesity84.

Finally, regarding the availability of Ac-CoA for the induction of transcriptional responses

following LPS stimulation, there are many additional experiments required for improving clarity,

especially in the case of ACSS2-dependent P genes. A common question in response to the work

described in Chapter 2 is, “What determines the gene specificity?” Although we demonstrated

that the histone acetyltransferase (HAT) p300 was involved in coupling the increase in glucose-

derived Ac-CoA with enhanced histone acetylation in the promoter region of pro-inflammatory

genes, we were surprised to find that add-back of Ac-CoA by supplying a high concentration of

Page 152: Regulation of Inflammatory Macrophage Activation and

138

exogenous acetate did not rescue induction of pro-inflammatory genes in tolerant macrophages

(data not shown). It is noteworthy that induction of P genes was not measured in these experiments

but presents an important avenue for further exploration, as this subset of genes was shown to be

profoundly dependent on ACSS2 in Chapter 3, while pro-inflammatory T genes may be

specifically regulated by citrate-derived Ac-CoA. On the ACLY-dependency of T gene induction

and lack of rescue by acetate, it is possible that spatial regulation of ACLY activity supports

compartment-specific control of Ac-CoA85. Production of a localized pool of Ac-CoA at T genes

by spatially-associated ACLY activity is an interesting area for further investigation86. Similarly,

substrate source fueling ACSS2-dependent activity also requires many additional experiments.

Most critical is the measurement of Ac-CoA levels and P gene expression levels in macrophages

with silenced ACSS2 activity. Substrate dependency should then be elucidated by tracing of

isotopically-labeled acetate into Ac-CoA and histones as was done for glucose in Chapter 2.

Pulse-capture experiments in which isotopically-labeled glucose is transiently supplied during

activation and then removed followed by measure of HDAC-inhibitor-sensitive labeled acetate

will also provide insight into the proportion of acetate that is available for induction of P genes in

tolerant macrophages. Of course, as mentioned previously, identification of the upstream signaling

proteins responsible for enhancing ACSS2 activity in the tolerant state is also an area in need of

further exploration. Although we have preliminary data showing that AKTi recapitulates the

effects of ACSS2i in P gene induction, direct evidence for Akt-dependent phosphorylation of

ACSS2 in this setting is needed.

Investigation into the areas discussed above will undoubtedly extend our appreciation for the many

important roles of metabolism in supporting the context-specific functional requirements of

Page 153: Regulation of Inflammatory Macrophage Activation and

139

eukaryotic organisms. The phenotypes one can observe through experiments in immunobiology

and physiology are often striking and very enjoyable to study. Metabolism provides a foundation

of regulation and support of these phenotypes that tethers the complex biological phenomena to

the organized set of first principles explaining the behavior of chemical species. Such an

understanding of biological responses from a chemical perspective is not only clarifying but also

illuminates many powerful points of intervention to exert control over the behavior of organisms.

It is exciting to think on what wonderful phenomena lie ahead as continued exploration pushes

back the barrier of the unknown to uncover new knowledge.

Page 154: Regulation of Inflammatory Macrophage Activation and

140

REFERENCES

1. Bone,R.C.etal.Definitionsforsepsisandorganfailureandguidelinesfortheuseofinnovativetherapiesinsepsis.TheACCP/SCCMConsensusConferenceCommittee.AmericanCollegeofChestPhysicians/SocietyofCriticalCareMedicine.Chest101,1644–1655(1992).

2. Ertel,W.etal.Downregulationofproinflammatorycytokinereleaseinwholebloodfrom

septicpatients.Blood85,1341–1347(1995).3. Wolk,K.,Döcke,W.D.,vonBaehr,V.,Volk,H.D.&Sabat,R.Impairedantigenpresentation

byhumanmonocytesduringendotoxintolerance.Blood96,218–223(2000).4. Monneret,G.etal.Theanti-inflammatoryresponsedominatesaftersepticshock:

associationoflowmonocyteHLA-DRexpressionandhighinterleukin-10concentration.Immunol.Lett.95,193–198(2004).

5. Hotchkiss,R.S.etal.Apoptoticcelldeathinpatientswithsepsis,shock,andmultipleorgan

dysfunction.Crit.CareMed.27,1230–1251(1999).6. Boomer,J.S.etal.Immunosuppressioninpatientswhodieofsepsisandmultipleorgan

failure.JAMA306,2594–2605(2011).7. Payen,D.,Monneret,G.&Hotchkiss,R.Immunotherapy-apotentialnewwayforwardin

thetreatmentofsepsis.CriticalCare17,118(2013).8. Everts,B.etal.TLR-drivenearlyglycolyticreprogrammingviathekinasesTBK1-IKKɛ

supportstheanabolicdemandsofdendriticcellactivation.NatureImmunology15,323–332(2014).

9. Mathis,D.&Shoelson,S.E.Immunometabolism:anemergingfrontier.Nat.Rev.Immunol.

11,81(2011).10.Foster,S.L.,Hargreaves,D.C.&Medzhitov,R.Gene-specificcontrolofinflammationby

TLR-inducedchromatinmodifications.Nature447,972–978(2007).11. Ivashkiv,L.B.Epigeneticregulationofmacrophagepolarizationandfunction.Trendsin

Immunology34,216–223(2013).12.Seeley,J.J.&Ghosh,S.MolecularmechanismsofinnatememoryandtolerancetoLPS.

JournalofLeukocyteBiology101,107–119(2017).13.Verdin,E.&Ott,M.50yearsofproteinacetylation:fromgeneregulationtoepigenetics,

metabolismandbeyond.Nat.Rev.Mol.CellBiol.16,258–264(2015).

Page 155: Regulation of Inflammatory Macrophage Activation and

141

14.Corbet,C.etal.AcidosisDrivestheReprogrammingofFattyAcidMetabolisminCancerCellsthroughChangesinMitochondrialandHistoneAcetylation.CellMetab.24,311–323(2016).

15.Donohoe,D.R.etal.TheWarburgEffectDictatestheMechanismofButyrate-Mediated

HistoneAcetylationandCellProliferation.MolecularCell48,612–626(2012).16.Covarrubias,A.J.etal.Akt-mTORC1signalingregulatesAclytointegratemetabolicinputto

controlofmacrophageactivation.Elife5,(2016).17.Cheng,S.-C.etal.mTOR-andHIF-1-mediatedaerobicglycolysisasmetabolicbasisfor

trainedimmunity.Science345,1250684–1250684(2014).18.Saeed,S.etal.Epigeneticprogrammingofmonocyte-to-macrophagedifferentiationand

trainedinnateimmunity.Science345,1251086–1251086(2014).19.Cameron,A.M.,Lawless,S.J.&Pearce,E.J.Metabolismandacetylationininnateimmune

cellfunctionandfate.SeminarsinImmunology28,408–416(2016).20.Donohoe,D.R.&Bultman,S.J.Metaboloepigenetics:Interrelationshipsbetweenenergy

metabolismandepigeneticcontrolofgeneexpression.JournalofCellularPhysiology227,3169–3177(2012).

21.Gut,P.&Verdin,E.Thenexusofchromatinregulationandintermediarymetabolism.

Nature502,489–498(2013).22.Kaelin,W.G.&McKnight,S.L.InfluenceofMetabolismonEpigeneticsandDisease.Cell

153,56–69(2013).23.Choudhary,C.,Weinert,B.T.,Nishida,Y.,Verdin,E.&Mann,M.Thegrowinglandscapeof

lysineacetylationlinksmetabolismandcellsignalling.Nat.Rev.Mol.CellBiol.15,536–550(2014).

24.Mills,E.L.etal.Itaconateisananti-inflammatorymetabolitethatactivatesNrf2via

alkylationofKEAP1.Nature556,113–117(2018).25.Galván-Peña,S.etal.MalonylationofGAPDHisaninflammatorysignalinmacrophages.

NatCommun10,338(2019).26.Lee,J.V.etal.Akt-dependentmetabolicreprogrammingregulatestumorcellhistone

acetylation.CellMetab.20,306–319(2014).27.Comerford,S.A.etal.AcetateDependenceofTumors.Cell159,1591–1602(2014).

Page 156: Regulation of Inflammatory Macrophage Activation and

142

28.Kamphorst,J.J.,Chung,M.K.,Fan,J.&Rabinowitz,J.D.Quantitativeanalysisofacetyl-CoAproductioninhypoxiccancercellsrevealssubstantialcontributionfromacetate.Cancer&Metabolism2,23(2014).

29.Schug,Z.T.etal.Acetyl-CoASynthetase2PromotesAcetateUtilizationandMaintains

CancerCellGrowthunderMetabolicStress.CancerCell27,57–71(2015).30.Bulusu,V.etal.AcetateRecapturingbyNuclearAcetyl-CoASynthetase2PreventsLossof

HistoneAcetylationduringOxygenandSerumLimitation.CellReports18,647–658(2017).31.Wang,A.etal.OpposingEffectsofFastingMetabolismonTissueToleranceinBacterialand

ViralInflammation.Cell166,1512-1525.e12(2016).32.McBrian,M.A.etal.HistoneAcetylationRegulatesIntracellularpH.MolecularCell49,310–

321(2013).33.Domínguez-Andrés,J.etal.TheItaconatePathwayIsaCentralRegulatoryNodeLinking

InnateImmuneToleranceandTrainedImmunity.CellMetabolism(2018).doi:10.1016/j.cmet.2018.09.003

34.Koza,R.A.etal.SequenceandTissue-DependentRNAExpressionofMouseFAD-Linked

Glycerol-3-PhosphateDehydrogenase.ArchivesofBiochemistryandBiophysics336,97–104(1996).

35.Mráček,T.,Drahota,Z.&Houštěk,J.Thefunctionandtheroleofthemitochondrial

glycerol-3-phosphatedehydrogenaseinmammaliantissues.BiochimicaetBiophysicaActa

(BBA)-Bioenergetics1827,401–410(2013).36.Williamson,D.H.,Lund,P.&Krebs,H.A.Theredoxstateoffreenicotinamide-adenine

dinucleotideinthecytoplasmandmitochondriaofratliver.Biochem.J.103,514–527(1967).

37.Sullivan,L.B.etal.SupportingAspartateBiosynthesisIsanEssentialFunctionofRespiration

inProliferatingCells.Cell162,552–563(2015).38.Nicholls,D.G.&Ferguson,S.J.Bioenergetics4.(ElsevierAcademicPress,2013).39.Mills,E.L.etal.SuccinateDehydrogenaseSupportsMetabolicRepurposingofMitochondria

toDriveInflammatoryMacrophages.Cell167,457-470.e13(2016).40.Chouchani,E.T.etal.AUnifyingMechanismforMitochondrialSuperoxideProduction

duringIschemia-ReperfusionInjury.CellMetabolism23,254–263(2016).

Page 157: Regulation of Inflammatory Macrophage Activation and

143

41.Chance,B.&Hollunger,G.Theinteractionofenergyandelectrontransferreactionsinmitochondria.I.Generalpropertiesandnatureoftheproductsofsuccinate-linkedreductionofpyridinenucleotide.J.Biol.Chem.236,1534–1543(1961).

42.Chance,B.Theinteractionofenergyandelectrontransferreactionsinmitochondria.II.

Generalpropertiesofadenosinetriphosphate-linkedoxidationofcytochromeandreductionofpyridinenucleotide.J.Biol.Chem.236,1544–1554(1961).

43.Orr,A.L.,Quinlan,C.L.,Perevoshchikova,I.V.&Brand,M.D.Arefinedanalysisof

superoxideproductionbymitochondrialsn-glycerol3-phosphatedehydrogenase.J.Biol.Chem.287,42921–42935(2012).

44.Quinlan,C.L.etal.MitochondrialcomplexIIcangeneratereactiveoxygenspeciesathigh

ratesinboththeforwardandreversereactions.J.Biol.Chem.287,27255–27264(2012).45.Orr,A.L.etal.InhibitorsofROSproductionbytheubiquinone-bindingsiteofmitochondrial

complexIidentifiedbychemicalscreening.FreeRadic.Biol.Med.65,1047–1059(2013).46.Quinlan,C.L.,Perevoshchikova,I.V.,Hey-Mogensen,M.,Orr,A.L.&Brand,M.D.Sitesof

reactiveoxygenspeciesgenerationbymitochondriaoxidizingdifferentsubstrates.RedoxBiol1,304–312(2013).

47.Goncalves,R.L.S.,Quinlan,C.L.,Perevoshchikova,I.V.,Hey-Mogensen,M.&Brand,M.D.

SitesofSuperoxideandHydrogenPeroxideProductionbyMuscleMitochondriaAssessedexVivounderConditionsMimickingRestandExercise.JournalofBiologicalChemistry290,209–227(2015).

48.Chouchani,E.T.etal.Ischaemicaccumulationofsuccinatecontrolsreperfusioninjury

throughmitochondrialROS.Nature515,431–435(2014).49.Brand,M.D.etal.SuppressorsofSuperoxide-H2O2ProductionatSiteIQof

MitochondrialComplexIProtectagainstStemCellHyperplasiaandIschemia-ReperfusionInjury.CellMetabolism24,582–592(2016).

50.Araki,K.etal.mTORregulatesmemoryCD8T-celldifferentiation.Nature460,108–112

(2009).51.Kaech,S.M.,Hemby,S.,Kersh,E.&Ahmed,R.MolecularandFunctionalProfilingof

MemoryCD8TCellDifferentiation.Cell111,837–851(2002).52. Intlekofer,A.M.etal.EffectorandmemoryCD8+TcellfatecoupledbyT-betand

eomesodermin.NatureImmunology6,1236–1244(2005).

Page 158: Regulation of Inflammatory Macrophage Activation and

144

53.Masopust,D.,Kaech,S.M.,Wherry,E.J.&Ahmed,R.TheroleofprogramminginmemoryT-celldevelopment.CurrentOpinioninImmunology16,217–225(2004).

54.Mosser,D.M.&Edwards,J.P.Exploringthefullspectrumofmacrophageactivation.

NatureReviewsImmunology8,958–969(2008).55.Xue,J.etal.Transcriptome-BasedNetworkAnalysisRevealsaSpectrumModelofHuman

MacrophageActivation.Immunity40,274–288(2014).56.Langston,P.K.,Shibata,M.&Horng,T.MetabolismSupportsMacrophageActivation.

FrontiersinImmunology8,(2017).57.Buck,M.D.,Sowell,R.T.,Kaech,S.M.&Pearce,E.L.MetabolicInstructionofImmunity.

Cell169,570–586(2017).58.Eto,K.etal.RoleofNADHShuttleSysteminGlucose-InducedActivationofMitochondrial

MetabolismandInsulinSecretion.Science283,981–985(1999).59.Madiraju,A.K.etal.Metforminsuppressesgluconeogenesisbyinhibitingmitochondrial

glycerophosphatedehydrogenase.Nature510,542–546(2014).60.Madiraju,A.K.etal.Metformininhibitsgluconeogenesisviaaredox-dependentmechanism

invivo.NatureMedicine24,1384–1394(2018).61.Alfadda,A.,DosSantos,R.A.,Stepanyan,Z.,Marrif,H.&Silva,J.E.Micewithdeletionofthe

mitochondrialglycerol-3-phosphatedehydrogenasegeneexhibitathriftyphenotype:effectofgender.AmericanJournalofPhysiology-Regulatory,IntegrativeandComparative

Physiology287,R147–R156(2004).62.RoyChowdhury,S.K.,Raha,S.,Tarnopolsky,M.A.&Singh,G.Increasedexpressionof

mitochondrialglycerophosphatedehydrogenaseandantioxidantenzymesinprostatecancercelllines/cancer.FreeRadicalResearch41,1116–1124(2007).

63.Rasmussen,U.F.,Krustrup,P.,Bangsbo,J.&Rasmussen,H.N.Theeffectofhigh-intensity

exhaustiveexercisestudiedinisolatedmitochondriafromhumanskeletalmuscle.PflugersArch.443,180–187(2001).

64.Sacktor,B.,Wormser-Shavit,E.&White,J.I.DIPHOSPHOPYRIDINENUCLEOTIDE-LINKED

CYTOPLASMICMETABOLITESINRATLEGMUSCLEINSITUDURINGCONTRACTIONANDRECOVERY.J.Biol.Chem.240,2678–2681(1965).

65.Zambon,A.C.etal.Time-andexercise-dependentgeneregulationinhumanskeletal

muscle.GenomeBiol.4,R61(2003).

Page 159: Regulation of Inflammatory Macrophage Activation and

145

66.Casimiro-Lopes,G.,Ramos,D.,Sorenson,M.M.&Salerno,V.P.Redoxbalanceandmitochondrialglycerolphosphatedehydrogenaseactivityintrainedrats.EuropeanJournalofAppliedPhysiology112,3839–3846(2012).

67.Kamiński,M.M.etal.TcellActivationIsDrivenbyanADP-DependentGlucokinaseLinking

EnhancedGlycolysiswithMitochondrialReactiveOxygenSpeciesGeneration.CellReports2,1300–1315(2012).

68.Bücher,T.&Klingenberg,M.WegedesWasserstoffsinderlebendigenOrganisation.

AngewandteChemie70,552–570(1958).69.Estabrook,R.W.&Sacktor,B.α-GlycerophosphateOxidaseofFlightMuscleMitochondria.

J.Biol.Chem.233,1014–1019(1958).70.Rottenberg,H.&Gutman,M.Controloftherateofreverseelectrontransportin

submitochondrialparticlesbythefreeenergy.Biochemistry16,3220–3227(1977).71.Mills,E.L.etal.Accumulationofsuccinatecontrolsactivationofadiposetissue

thermogenesis.Nature560,102(2018).72.Berrada,W.,Naya,A.,Ouafik,L.&Bourhim,N.Effectofhibernation,thyroidhormonesand

dexamethasoneoncytosolicandmitochondrialglycerol-3-phosphatedehydrogenasefromjerboa(Jaculusorientalis).Comp.Biochem.Physiol.B,Biochem.Mol.Biol.125,439–449(2000).

73.Arranz,A.etal.Akt1andAkt2proteinkinasesdifferentiallycontributetomacrophage

polarization.PNAS109,9517–9522(2012).74.Byles,V.etal.TheTSC-mTORpathwayregulatesmacrophagepolarization.NatCommun4,

2834(2013).75.Mertins,P.etal.AnIntegrativeFrameworkRevealsSignaling-to-TranscriptionEventsin

Toll-likeReceptorSignaling.CellReports19,2853–2866(2017).76.Wheaton,W.W.etal.MetformininhibitsmitochondrialcomplexIofcancercellstoreduce

tumorigenesis.eLife3,(2014).77.Bergheim,I.etal.Metforminpreventsendotoxin-inducedliverinjuryafterpartial

hepatectomy.J.Pharmacol.Exp.Ther.316,1053–1061(2006).78.Tsoyi,K.etal.MetformininhibitsHMGB1releaseinLPS-treatedRAW264.7cellsand

increasessurvivalrateofendotoxaemicmice.Br.J.Pharmacol.162,1498–1508(2011).

Page 160: Regulation of Inflammatory Macrophage Activation and

146

79.Zhou,Z.-Y.etal.Metforminexertsglucose-loweringactioninhigh-fatfedmiceviaattenuatingendotoxemiaandenhancinginsulinsignaling.ActaPharmacol.Sin.37,1063–1075(2016).

80.Kim,J.etal.MetforminSuppressesLipopolysaccharide(LPS)-inducedInflammatory

ResponseinMurineMacrophagesviaActivatingTranscriptionFactor-3(ATF-3)Induction.JournalofBiologicalChemistry289,23246–23255(2014).

81.Baldassari,S.etal.DevelopmentofanInjectableSlow-ReleaseMetforminFormulationand

EvaluationofItsPotentialAntitumorEffects.ScientificReports8,3929(2018).82.Stefely,J.A.&Pagliarini,D.J.BiochemistryofMitochondrialCoenzymeQBiosynthesis.

TrendsinBiochemicalSciences42,824–843(2017).83.Anderson,C.M.etal.DependenceofbrownadiposetissuefunctiononCD36-mediated

coenzymeQuptake.CellRep10,505–515(2015).84.Fink,B.D.etal.Metaboliceffectsofamitochondrial-targetedcoenzymeQanaloginhigh

fatfedobesemice.PharmacologyResearch&Perspectives5,e00301(2017).85.Shi,L.&Tu,B.P.Acetyl-CoAandtheregulationofmetabolism:mechanismsand

consequences.CurrentOpinioninCellBiology33,125–131(2015).86.Sivanand,S.,Viney,I.&Wellen,K.E.SpatiotemporalControlofAcetyl-CoAMetabolismin

ChromatinRegulation.TrendsBiochem.Sci.43,61–74(2018).