190
Time to eat: Links between neuronal function and cellular phagocytosis Elizabeth Stone Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy under the Executive Committee of the Graduate School of Arts and Sciences COLUMBIA UNIVERSITY 2015

Time to eat

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Time to eat

Time to eat:

Links  between  neuronal  function  and  cellular  phagocytosis  

   

Elizabeth Stone

Submitted in partial fulfillment of the requirements for the degree of

Doctor of Philosophy under  the  Executive  Committee  

of the Graduate School of Arts and Sciences

 COLUMBIA UNIVERSITY

2015

Page 2: Time to eat

©2015

Elizabeth Stone

All rights reserved

Page 3: Time to eat

ABSTRACT

Time to eat: Links between neuronal function and cellular phagocytosis

Elizabeth Stone

   How do the brain and the immune system interact, and what are the consequences of this

interaction on the physiology of an organism during infection? The main focus of my

thesis is neuroimmune interaction, as studied in the following: (1) circadian regulation of

immune system function, specifically phagocytosis by immune cells during bacterial

infection; (2) the impact of circadian-regulated metabolism and feeding behavior on

immunity and host tolerance of bacterial infection; and (3) immune system function in

the context of Fragile X syndrome, a neurological disease known to cause circadian

dysregulation. To investigate the interactions between these complex physiologies, I use

the well-characterized and genetically tractable Drosophila melanogaster animal model.

Each topic is briefly introduced in Chapter 1. Chapter 2 focuses on the body of work

identifying the circadian regulation of the immune system, particularly phagocytosis, by

immune cells during bacterial infection. Chapter 3 highlights findings regarding how diet

and host metabolic state impact survival after infection. Chapter 4 illustrates phagocytic

immune cell defects both systemically and in the brain in the Drosophila model of

Fragile X syndrome. Lastly the conclusions discuss how these three works have built on

our fund of knowledge of neuroimmune interactions and the future implications for these

results.

Page 4: Time to eat

i

TABLE OF CONTENTS LIST OF GRAPHS, IMAGES, AND ILLUSTRATIONS ........................................... iv

ACKNOWLEDGEMENTS .......................................................................................... viii

DEDICATION ................................................................................................................. xi

CHAPTER 1: INTRODUCTION .................................................................................... 1

An introduction to the fly .............................................................................................................. 2

An introduction to circadian regulation ........................................................................................ 3

The molecular clock is the timekeeper for circadian rhythms ...................................................... 5

Clock neurons and the master clock ............................................................................................. 9

An introduction to the innate immune system of the fly ............................................................. 12

Circadian regulation of immunity ............................................................................................... 16

Circadian regulation of metabolism and immunity ..................................................................... 18

An introduction to Fragile X syndrome ...................................................................................... 19

FMRP is an RNA-binding protein that functions as a translational repressor ............................ 21

Synaptic plasticity is altered in Fragile X syndrome .................................................................. 23

Novel therapeutics to treat Fragile X syndrome ......................................................................... 24

CHAPTER 2: The circadian clock protein Timeless regulates phagocytosis of

bacteria in Drosophila ..................................................................................................... 26

Summary ..................................................................................................................................... 27

Abstract ....................................................................................................................................... 30

Introduction ................................................................................................................................. 32

Materials and Methods ............................................................................................................ 34

Results/Discussion .................................................................................................................... 39

Page 5: Time to eat

ii

Acknowledgements ..................................................................................................................... 58

CHAPTER 3: Host tolerance of infection is promoted by dietary glucose and amino

acids through insulin and TORC2 signaling ................................................................ 59

Summary ..................................................................................................................................... 60

Abstract ....................................................................................................................................... 63

Introduction ................................................................................................................................. 65

Materials and Methods ............................................................................................................ 68

Results ......................................................................................................................................... 77

Discussion ................................................................................................................................... 94

Supplemental Materials ............................................................................................................... 97

Author Contributions ................................................................................................................ 102

Acknowledgments ..................................................................................................................... 103

CHAPTER 4: Phagocytic immune cell defects in the Drosophila model of Fragile X

syndrome ........................................................................................................................ 104

Summary ................................................................................................................................... 105

Abstract ..................................................................................................................................... 107

Introduction ............................................................................................................................... 108

Materials and Methods .......................................................................................................... 112

Results ..................................................................................................................................... 118

Discussion ................................................................................................................................. 130

Chapter 5: CONCLUSIONS ........................................................................................ 135

Phagocytosis is circadian-regulated in Drosophila ................................................................... 136

The circadian regulation of feeding and metabolism affects survival of infection in

Drosophila ................................................................................................................................ 104

Page 6: Time to eat

iii

The Drosophila model of Fragile X syndrome exhibits defects in phagocytosis ..................... 105

REFERENCES .............................................................................................................. 145

APPENDIX I ................................................................................................................. 163

APPENDIX II: Phagocytosis is upregulated in dFMR1 hypomorphs ..................... 164

Abstract ..................................................................................................................................... 165

Materials and Methods .......................................................................................................... 166

Results ..................................................................................................................................... 169

Page 7: Time to eat

iv

LIST OF GRAPHS, IMAGES, AND ILLUSTRATIONS

CHAPTER 1: INTRODUCTION .................................................................................... 1

Fig. 1. The molecular clock is comprised of negative transcriptional feedback loops ................. 7

Fig. 2. Drosophila use three main immune mechanisms to fight pathogens .............................. 13

CHAPTER 2: The circadian clock protein Timeless regulates phagocytosis of

bacteria in Drosophila ..................................................................................................... 26

Fig. 1. Tim activity has significant consequences for immunity. .............................................. 40

Fig. 2. Tim regulates resistance to infection by S. pneumoniae and S. marcescens .................. 42

Fig. 3. Resistance against S. pneumoniae correlates with Tim activity. .................................... 43

Fig. 4. Tim protein does not regulate melanization or AMP gene expression. .......................... 47

Fig. 5. Tim protein regulates an early stage of phagocytosis of bacteria by hemocytes. ........... 50

Fig. 6. Bead inhibition of phagocytosis in wild type flies eliminates the difference in survival

kinetics with Tim mutants after S. pneumoniae infection. .......................................................... 53

Fig. S1. Tim protein does not regulate AMP gene expression. ................................................. 56

CHAPTER 3: Host tolerance of infection is promoted by dietary glucose and amino

acids through insulin and TORC2 signaling ................................................................ 59

Fig. 1. Period mutants exhibit greater tolerance than isogenic controls during infection with B.

cepacia ........................................................................................................................................ 78

Fig. 2. Per mutants have lower metabolic resources and eat more than wild type flies. ........... 80

Fig. 3. Dietary restriction does not increase infection tolerance of either Per mutants or wild

type. ............................................................................................................................................. 83

Fig. 4. Dietary glucose and amino acids increase tolerance of B. cepacia

infection. ..................................................................................................................................... 85

Page 8: Time to eat

v

Fig. 5. Insulin-like signaling mediates the glucose contribution to infection tolerance and the

TORC2 complex mediates the effects of dietary amino acids. ................................................... 88

Fig. 6. The effect of amino acids on survival of infection is facilitated by gut

microbiota. ..................................................................................................................... 92

Fig. S1. Antimicrobial peptides (AMPs) are induced by B. cepacia infection to similar

expression levels in Per mutants and wild type. ........................................................... 97

Fig. S2. Rapamycin reduces survival of B. cepacia infection. ..................................... 98

Fig. S3. Expression of Tsc1 and Tsc2 transcripts are induced by heat- shock. ............. 99

Fig. S4. Amino acid-stimulated tolerance is facilitated by the gut microbiota. .......... 100

Table S1. Primer sequences ........................................................................................ 101

CHAPTER 4: Phagocytic immune cell defects in the Drosophila model of Fragile X

syndrome ........................................................................................................................ 104

Fig. 1. dFMR1 mutants are less resistant to infections with specific pathogens. ...................... 119

Fig. 2. dFMR1 mutants exhibit altered immune system parameters. ....................................... 122

Fig. 3. Clearance of severed axons is delayed in dFMR1 mutants relative to wild-type

controls ...................................................................................................................................... 126

Fig. 4. Fewer activated glia are seen in dFMR1 mutants than in wild-type controls after axonal

severing ..................................................................................................................................... 127

Fig. 5. Time course of Draper expression after axonal severing shows that glial phagocytosis is

delayed but not abolished in dFMR1 mutants ........................................................................... 128

APPENDIX I ................................................................................................................. 163

Fig. 1. Per mutants are more tolerant of infection with P. aeruginosa, and starvation eliminates

this survival advantage .............................................................................................................. 163

Page 9: Time to eat

vi

APPENDIX II: Phagocytosis is upregulated in dFMR1 hypomorphs ..................... 164

Fig. 1. dFMR1 hypomorphs survive longer than wild-type controls after infection with certain

pathogens. ................................................................................................................................. 169

Fig. 2. dFMR1 hypomorphs are more resistant to infection with S. pneumoniae and S.

marcenscens compared to wild-type controls ........................................................................... 170

Fig. 3. Antimicrobial peptide synthesis is not altered by reduced levels of dFMR1 ............... 172

Fig. 4. dFMR1 hypomorphs exhibit increased phagocytosis of bacteria by hemocytes ........... 173

Fig. 5. dFMR1 hypomorphs do not exhibit circadian variation in phagocytosis ...................... 174

Fig. 6. Inhibiting phagocytosis abolishes the survival benefit observed in dFMR1 hypomorphs

with S. pneumonia infection. ................................................................................................... 175

Page 10: Time to eat

vii

ACKNOWLEDGEMENTS

Before anyone else, I need to thank my PhD mentor, Dr. Mimi Shirasu-Hiza. When I

walked into her office on that propitious day—almost five years ago now—to discuss

possible rotation projects, I had no forewarning that her scientific vision and ideas would

be so thoroughly compatible with my own. Since then, barely a week has gone by when I

have not popped into Mimi’s office simply to offer a new idea for an experiment or to

seek some feedback, only to find myself bouncing thoughts around with her for

sometimes hours on end. Without Mimi’s innovative perspectives, effective collaboration

and unwavering support, there is no way that I could have completed this work. I am

lucky to call Mimi my mentor and friend. Many, many thanks to you, Mimi.

To the Shirasu-Hiza lab members past and present, thank you for your continued

support, both scientifically and socially. Thanks to my fellow grad students, Victoria and

Vanessa, for your wildly entertaining stories and your help over the years with writing

and suggesting experiments. And Vanessa, my many thanks for your help with the

Listeria infections and melanization while I was pregnant. My daughter Rebecca thanks

you too. And Emily, your help with the hemocyte phagocytosis for the dFMR1 project

was crucial. I could not have gotten so far without you. And lastly, to all of the techs

and undergrads past and present, I can’t thank you enough for your help with ordering,

stocks, and all of the other things that come up in lab. Adam, Albert, Paul, Charlotte, and

especially Reed, you guys are awesome. And Reed, I can’t thank you enough for taking

over the dFMR1 project and putting so much blood, sweat, and tears into it. I know you

have had a love-hate relationship with this project since you were an undergrad, but see?

Isn’t it exciting?? Take care of my (non-biological) baby. It’s now in your very capable

Page 11: Time to eat

viii

hands, and I know you will make the most of the project in my absence. And Jenn

Ziegenfuss—I know you’re not a member of the Shirasu-Hiza lab, but thank you so much

for all of your help troubleshooting glial phagocytosis with us. The project wouldn’t have

gotten to this point without your help.

I am so very grateful to my thesis committee members, Dr. Wes Grueber, Dr. Ai

Yamamoto, and Dr. Ron Liem. The three of you have had invaluable suggestions and

have kept me from spinning my wheels for too long on one aspect of my project or

another. Each of you has offered a unique contribution. There are many reasons why I

am especially thankful for Wes. Not only did you share your Drosophila and imaging

expertise and let me use your confocal, but you also for agreed to serve as a co-mentor

for my NRSA fellowship. Wes, you have been unbelievably supportive throughout my

PhD; thank you for your continuing mentorship. I also owe a big thanks to Dr. Tom

Jongens for agreeing to serve as my outside reader, serving as a co-mentor for my NRSA

fellowship, and offering vital advice on the immunity aspect of my dFMR1 project.

My many thanks to the MD-PhD program here at Columbia University. The

support from Dr. Patrice Spitalnik, Dr. Ron Liem, and Dr. Michael Shelanski has been

crucial to my success in the program. Thank you for all that you do. Jeffery Brandt is also

one of the best and most capable administrators that I have every encountered, and I can’t

thank you enough for your assistance with everything in the program, especially grant

submissions. Also, a huge thank you to my MD/PhD cohort and graduate friends outside

of the Shirasu-Hiza lab for your continuing love and support.

I have spent the last 26+ years either learning in school or working as a research

technician for a University, and I have to thank many educators throughout the years who

Page 12: Time to eat

ix

have supported and encouraged my love of learning. Starting early, Mrs. Judy Lieb, my

Seminar teacher at Lynnewood Elementary School, was instrumental in fostering a

curiosity for puzzles and mysteries. I will never forget how Mrs. Sharon Levin and Mr.

Mark Levy from Akiba Hebrew Academy had confidence in me, even if I didn’t. It was

at Barnard College where my professors and classmates created an environment for me to

thrive and learn the confidence that I carry with me today. I am also very grateful to Dr.

Jen Darnell, for whom I was a research technician for three years after college. Jen

taught me the importance of controls and how to conduct the most rigorous experiment

for a particular question.

Without the continuing love and support from my family and friends, I would not

be here today. To my parents—I love you both very much, and I cannot thank you

enough for all that you have done for me at every stage of my life. For as long as I can

remember, my dad has nurtured my interest in science, and although he tried his hardest

to deter me, I am following in his footsteps. My mom has been enormously encouraging

and has taught me the importance of inclusiveness and compassion for others and to be

generous with my time. While both of my parents have pushed me to try my hardest,

they also allowed me to choose my own path with my hobbies and my career. Thank

you, thank you, thank you, Mom and Dad. To my brothers, Eric and Martin—Eric, thank

you for giving me unrelenting tough love from the age of five until you went off to

college; you’ve taught me to stand up for myself. And Martin, thank you for protecting

me from Eric’s onslaught; you taught me that I could rely on the help of others. I am

extremely lucky to have you both in my life. Thank you, bros, and I love you both. Many

Page 13: Time to eat

x

thanks to my Grandmom, Selma Kohn, who has been my number one fan and who has

encouraged me every step of the way.

Last but not least, I thank my husband, Zachary Jacobs, for his incredible support

through the years and my daughter, Rebecca Jacobs, for sharing Mommy with her

passion for science and medicine. It hasn’t always been easy, but Zachary has always

been there for me. From late night studying in medical school, editing my qualifying

exam and my grants, trying to make sense of my science, listening to endless practice

talks, and putting up with late nights and weekends in lab, I could not ask for a more

supportive partner. Zachary and Rebecca, I love both of you to the moon and back, and I

can’t thank the two of you enough.

Page 14: Time to eat

xi

DEDICATION I dedicate this work to my daughter, Rebecca Jacobs. I hope that this work inspires you to

choose your own adventure, to try your hardest to be the best person you can be, and to

make an impact. I love you, munchkin.

Page 15: Time to eat

1

CHAPTER 1:

INTRODUCTION

Page 16: Time to eat

2

How do the brain and the immune system interact, and what are the consequences of this

interaction on the physiology of an organism during infection? The main focus of my

work is neuroimmune interaction. The three major projects that I will discuss here are:

(1) circadian regulation of immune system function, specifically phagocytosis by immune

cells during bacterial infection; (2) the impact of circadian-regulated metabolism and

feeding behavior on immunity and host tolerance of bacterial infection; and (3) immune

system function in the context of Fragile X syndrome, a neurological disease known to

cause circadian dysregulation. I used the well-characterized Drosophila melanogaster

animal model to investigate these projects. These projects are highly representative of the

major interests of the Shirasu-Hiza lab, which utilizes Drosophila as a model organism to

understand novel interaction between complex systemic physiologies. I briefly introduce

each topic in Chapter 1. Chapter 2 focuses on the body of work identifying the circadian

regulation of the immune system, particularly phagocytosis, by immune cells during

bacterial infection. Chapter 3 highlights findings on how diet and host metabolic state

impact survival after infection. In Chapter 4, I illustrate two phagocytic immune cell

defects in the Drosophila model of Fragile X syndrome. Lastly, I will discuss how these

three works have built on our fund of knowledge of neuroimmune interactions in the

conclusions.

An introduction to the fly

We use Drosophila melanogaster as a tool to study neuroimmune interactions. There is a

rich tradition of using Drosophila to explore heredity and genetic interactions, beginning

with Charles W Woodworth and Thomas Hunt Morgan in the early 20th century. Flies are

Page 17: Time to eat

3

genetically tractable and there are very powerful genetic tools available at our disposal.

In addition to fast generation times, a large number of animals are available per

experiment, increasing the power for statistical significance. Many of the applicable

circadian, metabolic, immunity, and neurological disease pathways are highly conserved,

both molecularly and functionally, between Drosophila and mammals [1-4]. Moreover,

these pathways and physiologies are often simplified in Drosophila. For example,

Drosophila has only an innate immune system while mammals have both an adaptive and

innate immune system, and typically there are also fewer gene family members [3]. Thus,

the fly is an ideal organism to address the complex questions that we ask in lab.

An introduction to circadian regulation

Circadian rhythms are ancient, evolutionarily conserved 24-hour fluctuations in

molecular, physiological, and behavioral processes. These daily fluctuations, while

influenced by external cues such as sunlight, continue even in the absence of external

cues, suggesting an internal “clock” that keeps time for the organism. The first historical

account of circadian regulation was reported in the fourth century B.C.E. when

Adrosthenes, a ship captain for Alexander the Great, recorded that the leaves of the

tamarind tree exhibited daily, regular movements [5]. It was not until approximately 2100

years later when the first recorded experiment testing endogenous rhythms took place; in

1729, Jean-Jacques d'Ortous de Mairan, an astronomer, observed that when he placed a

Mimosa pudica plant in the dark, the leaves still opened and closed with the same rhythm

as a plant subjected to natural light [5]. Plants exhibit diurnal rhythms by opening leaves

and flowers to maximize sunlight exposure during the day while closing their leaves and

Page 18: Time to eat

4

flowers at night to minimize the surface area and water loss. Each species of animal has

its own rhythm that controls when the animal wakes, eats, and sleeps. Humans have very

specific circadian regulation of meal times, waking activity, and sleeping. Some of the

physiological and behavioral processes regulated by circadian rhythms include sleep,

learning and memory, feeding, metabolism, and immune system function [6-8]. The

molecular mechanisms governing the circadian regulation of these processes are still not

well understood.

Circadian regulation is set by numerous zeitgebers, or exogenous synchronizers

[9]. Zeitgebers are thought to reset the internal body clock to synchronize the internal

body clock to the environment. One powerful external cue is the cycling wavelengths and

brightness of light [9]. For diurnal animals, including humans, squirrels, butterflies, and

most lizards, the presence of light stimulates wakefulness. However, raccoons, owls, and

other nocturnal animals sleep when it is light and are active at night [10]. These animals

are presumed to have adapted to nocturnal schedules either to avoid predators or to hunt

prey that have adapted to wakefulness at night. Still other animals are most active at

dawn and at dusk but can display wakefulness during the day or night; these animals are

said to exhibit crepuscular rhythms. Some examples of animals that display crepuscular

rhythms are various insects, including Drosophila and mosquito species, bats in the

suborder Megachiroptera, pygmy rabbits, and species of deer [11-15]. Temperature

fluctuations and food intake are two other strong zeitgebers [16,17], and in humans and

Drosophila, it is has been hypothesized that social interactions can be as strong as the

other external cues described [18], but in humans it may be the effect of altered light

exposure instead of social interactions on the circadian cycle [18,19].

Page 19: Time to eat

5

The molecular clock is the timekeeper for circadian rhythms

Drosophila have been crucial for our understanding of the molecular clock. The first

circadian regulatory gene was isolated through a forward genetic screen in Drosophila. In

their seminal report published in 1971, Seymour Benzer and Ronald Konopka described

three Drosophila mutants with abnormal eclosion rhythms and altered circadian

locomotor activity[20]. In populations of wild-type flies, eclosion rhythms and circadian

locomotor activity have periods of approximately 24 hours; Benzer and Konopka found

that one of their mutants was completely arrhythmic, one mutant had a short period

averaging 19 hours, and one mutant had a long period averaging 28 hours. Using elegant

classical genetics techniques, they showed that these three mutations mapped to the same

locus on the X chromosome, and they concluded that these are three different alleles of

the same gene, which they named period. These experiments were the first demonstration

of a circadian regulatory gene in any organism.

Since these experiments, PERIOD (PER) has been shown to be a critical part of

the basic, evolutionarily conserved four component molecular clock, which acts in a

negative transcriptional feedback loop to regulate circadian rhythm. In Drosophila, the

four essential components for circadian regulation are CLOCK (CLK), CYCLE (CYC),

PERIOD (PER) and TIMELESS (TIM) [21]. Rhythms are generated by negative

transcriptional feedback loops by the molecular clock. Briefly, the transcriptional

activators CLK and CYC, which have basic helix-loop-helix (bHLH) and PAS domains,

important for DNA binding and protein-protein interactions, respectively, activate the

expression of PER and TIM, along with many target genes, slowly throughout the day by

Page 20: Time to eat

6

binding to enhancer elements (E-box) of their promoter regions [22]. Although TIM does

not have any appreciable DNA-binding domain, PER has a PAS domain, which is known

to interact with bHLH domains [23]. At night, when levels of PER and TIM are at their

peak, PER/TIM translocate to the nucleus together where they repress the binding of

CLK/CYC to the PER and TIM promotors [21], essentially inhibiting their own

transcription. TIM is then degraded in the presence of light [24], resulting in the release

of the inhibition of the transcription of per and tim. The levels of these proteins oscillate

with a period of approximately 24 hours, and light entrains, or regulates the period of, the

cycling of these proteins (Fig 1A).

There are many other enzymes and proteins that interact with the CLK and CYC

(Fig 1A). CLK/CYC also promote the transcription of many other genes. The proteins of

two of these genes, vrille (vri) and PAR domain protein 1ε (pdp1ε), act as a repressor and

activator, respectively, of CLK transcription [25]. VRI is a basic leucine zipper

transcription factor, while PDP1ε has a transcriptional activating domain [25]. These two

proteins work in opposition in a second negative feedback loop to regulate the mRNA

transcript levels of clk throughout the course of the day [26], but with little effect on CLK

protein expression. Clockwork orange (CWO) is another transcription factor that is

circadian regulated and is a bHLH repressor, preventing the transcription of the target

genes of CLK/CYC [26] (Fig 1B).

Numerous proteins, including kinases and phosphatases, either promote or reduce

the stability of PER and TIM (Fig 1B). For instance, doubletime (DBT) is a kinase that

phosphorylates PER and leads to its degradation by the proteasome [27]. Interestingly,

when PER is associated with TIM, DBT is unable to phosphorylate PER, preventing the

Page 21: Time to eat

7

Figure 1. The molecular clock is comprised of negative transcriptional feedback loops. (A) Here is the basic negative feedback loop that makes up the molecular clock in Drosophila, which has the transcriptional regulators CLOCK (CLK) and CYCLE (CYC) binding to the promoters of the proteins PERIOD (PER) and TIMELESS (TIM). TIM and PER then block the association of CLK and CYC with the per and tim promoters. CLK and CYC also bind to the promoters of PAR domain protein 1ε (pdp1ε) and vrille (vri); the protein products of these genes, PDP1ε and VRI, promote or inhibit the transcription of clk, respectively, thus further modifying the feedback loop. Adapted from Zheng and Sehgal, 2008 [28].

Page 22: Time to eat

8

(B) The post-translational modification of PER and TIM further regulate the circadian clock. TIM binds PER and prevents PER degradation. In the presence of light during the day, cryptochrome (CRY) binds TIM, targeting TIM for degradation through the E3 ligase JETLAG. The absence of TIM during the day causes PER to be phosphorylated by doubletime (DBT), and the E3 ligase supernumerary limbs (SLIMB) then targets PER for degradation by the proteasome. Protein phosphatase 2A (PP2A) and protein phosphatase I (PP1) dephosphorylate PER and TIM, respectively, stabilizing them from degradation. Additionally, the transcription factor and bHLH repressor clockwork orange (CWO) binds blocks the association of CLK and CYC to their DNA targets. Adapted from Zheng and Sehgal, 2008 [28]. (C) The mammalian homologs of the molecular clock are regulated in a similar fashion. Mammalian CLK and BMAL (the CYC homology) drive the expression of mPer and mCry (the TIM homolog). Casein Kinase I, the functional homolog of DBT, phosphorylates BMAL, mPer, and mCry, targeting them for degradation. The F Box E3 ligase FBXL3 further targets mCry for degradation by the proteasome. Protein phosphatase 1 (PP1), on the other hand, dephosphorylates mPer and mCry, stabilizing them from degradation. Adapted from Ko and Takahashi, 2006 [29].

degradation of PER [30]; thus TIM stabilizes PER. The E3 ligase supernumerary limbs

(SLIMB) targets phosphorylated PER for degradation by the proteasome [27]. Shaggy

(SGG), the fly homolog of glycogen synthase kinase 3β, phosphorylates TIM [31]. Two

phosphatases, protein phosphatase 2 (PP2) [32] and protein phosphatase 1 (PP1) [33]

dephosphorylate and stabilize PER and TIM, respectively, which prevents their

Page 23: Time to eat

9

degradation. In the presence of blue light, the photoreceptor crytochrome (CRY) signals

for the degradation of TIM through JETLAG, an E3 ligase, and the proteasome [34].

These proteins all contribute to the length and timing of the rhythms of the molecular

clock.

Many of the same players are also involved in the molecular clock in mammals

(Fig 1C). For instance, CLOCK and BMAL1 (the mammalian homolog of CYC)

promote the transcription of the mammalian PER homologues (mPer1 and mPer2) and

crytochromes (mCry1 and mCry2), in the place of TIM. Casein kinase 1 (CK1), most

homologous to DBT, phosphorylates BMAL1 [35], mPer [36], and mCry [35] and targets

them for degradation. PP1 dephosphorylates and promotes the stability of mPer [37].

FBXL3, an E3 ligase F box protein, targets mCry for degradation by the proteasome [38].

Just as circadian regulation in Drosophila is made up of two negative

transcriptional feedback loops, the mammalian clock also has a second transcriptional

loop, in which CLK and BMAL1 drive the transcription of two retinoic acid-related

orphan nuclear receptors, Rorα and Rev-erbα [29]. RORα and REV-ERBα then

translocate into the nucleus and bind to segments of the Bmal1 promoter containing

retinoic acid related orphan response elements (ROREs). RORα promotes the

transcription of Bmal1 while REV-ERBα inhibits Bmal1 transcription.

Clock neurons and the master clock

Although the molecular clock is expressed in the brain and in many other tissues, specific

neurons in the brain are the master timekeeper [39]. In the brain, clock expression is

restricted to specific regulatory areas. These neurons are collectively called the central

Page 24: Time to eat

10

clock, and they are thought to synchronize and regulate all other clocks throughout the

body, ultimately entraining an organism to the exogenous environment. The central clock

in Drosophila is governed by 150 so-called clock neurons [39]. Clock neurons are

centrally located in the brain and express the molecular clock described above. There are

7 groups of clock neurons in the brain: dorsal neurons (DN1, DN2, and DN3), small and

large ventral lateral neurons (sLNv and lLNv), dorsal lateral neurons (LNd), and the

lateral posterior neurons (LPN).

A small subset of clock neurons in the Drosophila brain is important for

regulating increased activity in anticipation of the change in light during dawn and dusk.

The morning peak is controlled by the sLNv [24], which are active in the morning, while

the DN1 and LNd control the evening peak; these populations of neurons are referred to

as the M and E cells, respectively [40]. In addition to the molecular clock, the sLNv

secrete the neuropeptide pigment dispersing factor PDF and signal to the E cells in the

morning, while the E cells signal back to the M cells in the evening [24]. The presence of

PDF in the M cells as well as CRY drives the morning activity, while the PDF-negative

and CRY-positive E cells direct the evening activity [40]. PDF expression by the sNLv

has two other important roles beside dictating the timing of the morning anticipation:

PDF is used both to synchronize clock neurons within the brain to ensure that all neurons

are cycling with the same period and to regulate the output of clock neurons to the rest of

the body. For example, one output of PDF-secreting neurons sends projections to the

mushroom body, the fly structure analogous to the hippocampus, which is important for

the sleep-wake cycle of the fly and for learning and memory [21,41,42].

Similar to Drosophila, mammalian circadian regulation is directed by a central

Page 25: Time to eat

11

clock and synchronizing signals regulate peripheral clocks [43]. The mammalian central

clock is located in the suprachiasmatic nucleus (SCN) of the hypothalamus. In the

presence of light, the retina signals to the SCN to entrain the molecular oscillators in the

SCN to the exogenous environment. The SCN then communicates to peripheral tissues

through endocrine signaling [44,45], neuronal signaling via the autonomic nervous

system [46], and temperature changes [47,48]. Interestingly, unlike peripheral tissues,

which require input from endocrine signaling and temperature regulation from the SCN,

intercellular synchronization in the SCN makes its molecular oscillator remarkably

resistant to disruptions from endocrine signaling or temperature changes [49]. Thus, the

job of the SCN is to be in sync with external light cues, while the clocks of peripheral

tissues are more malleable and can respond to the local environment [50].

With the exception of the testis [51], almost every other tissue of the body that has

been tested expresses a cycling clock [47]; an incomplete list of tissues with peripheral

clocks includes the heart, adrenal gland, ovary, muscle, kidney, spleen, lung, and liver.

As an example, the liver is entrained by four different mechanisms [52]. First, autonomic

nervous system innervation of the liver contributes to the cycling of Per1 and Per2 in

addition to two glucose enzymes, PEPCK and GLuT2 [46]. Second, hormonal secretion

modifies downstream targets in the liver. For instance, rhythmic secretion of

glucocorticoids alters the expression of approximately 60% of the transcriptome of the

liver through the transcription factor HNF4α [53]. Additionally, the expression of many

nuclear hormone receptors is cyclic in the liver [54], affecting metabolic function at

different times of the day. Third, the temperature of the animal can change the cycling of

circadian genes in the liver [55] and all other peripheral clocks that have been tested [56].

Page 26: Time to eat

12

Finally, feeding during the organism’s sleep phase, but not during the organism’s wake

cycle, strongly shifts the cycling of the liver, but not the SCN [57].

An introduction to the innate immune system of the fly

Drosophila innate immunity has at least three well-characterized branches that work

together to fight pathogens [58]: (1) antimicrobial peptide production, (2) reactive

oxygen species synthesis resulting in melanization, and (3) phagocytosis, or the

engulfment of pathogens by hemocytes. Induction of each of these arms of the

Drosophila innate immune system occurs following infections with different pathogens.

After systemic infection with bacteria or fungi, two different signaling cascades,

the Toll and imd pathways, culminate in the induction of the Drosophila antimicrobial

peptides in the fat body, analogous to the mammalian liver [58]. The Toll pathway, first

discovered in Drosophila as a developmental patterning gene [59], was published in the

pivotal work by Bruno Lemaitre and Jules Hoffman, where they reported that the Toll

pathway is important for immune defense against both fungal and bacterial infections

[60]. Fungi and bacteria expressing lys-type peptidoglycan (generally Gram positive

bacteria) preferentially activate the Toll pathway via specific peptidoglycan receptor

proteins (PGRPs) and activation of the receptor Toll, while bacteria expressing DAP-type

peptidoglycan (generally Gram negative bacteria) activate the immunity deficiency (imd)

pathway [60] via different PGRP receptors and other downstream effectors [61] (Fig 2A).

The two downstream NF-κb family members, Dif/Cactus and Relish, respectively, and

result in the transcription of antimicrobial peptides and other genes. Typically,

Page 27: Time to eat

13

   

   

       Figure 2. Drosophila use three main immune mechanisms to fight pathogens. (A) Antimicrobial peptides, produced in response to Toll and imd signaling, block pathogen growth by disrupting pathogen membranes [61]. In the presence of Gram positive bacteria and fungi, the Toll signaling cascade is

vertebrates

DD DD

DD dMyD88

Tube Pelle

DIF

Cactus

Drosomycin Other AMPs

DD DD

MyD88

IRAK

NF- B

I B

Proinflammatory cytokines

Drosophila

Toll TLR

Nucleus

PGRP

DD

IMD

DREDD dFADD

Relish

Diptericin Other AMPs

Proinflammatory cytokines

vertebrates Drosophila

Kenny (dIKK- ) Ird5

(dIKK- )

TNFR

DD

RIP

Caspase8 FADD

IKK-

IKK-

TNF-

NF- B

I B

A Toll/TLR signaling IMD/TNF- signaling

Adapted from Hoffmann and Reichhart, 2002

lysosome

Microbe recognition

ROS production Melanin deposition

Phenol Quinones

Phenoloxidase

B Melanization

fly

Microbe recognition

Internalization

Phagosome maturation

Immune cell

lysosome

C Phagocytosis

Page 28: Time to eat

14

activated through the Toll-induced signaling complex (TISC), comprised of the Toll receptor (mammalian homologs are the Toll-like receptors, or TLRs) and 3 death domain (DD) containing proteins: dMyD88 (homologous to the mammalian MyD88), Tube, and Pelle, which is an IL-IR-associated kinase (IRAK) protein (homologous to mammalian IRAK). Cactus (mammalian homolog IκB), which is normally associated with Dif (mammalian homolog NF-κB), is phosphorylated and degraded, freeing Dif to translocate to the nucleus and transcribe certain antimicrobial peptides, including Drosomycin. Gram negative bacteria, on the other hand, bind to various PGRP receptors (homologous to mammalian TNF-α receptor, or TNFR), which binds to the DD-containing IMD (mammalian homolog RIP), along with DREDD (mammalian homolog Caspase 8) and dFADD (mammalian homolog FADD) to activate the imd pathway. Together with Kenny (mammalian homolog IKK-γ) and Ird5 (mammalian homolog IKK-β), DREDD and dFADD target Relish (another Drosophila homolog of the mammalian NF-κB) for cleavage. Cleaved Relish translocates to the nucleus to transcribe numerous antimicrobial peptides, including Diptericin. Adapted from Hoffmann and Reinhhart, 2002 [3]. (B) The melanization cascade produces reactive oxygen species, which are toxic to pathogens, and culminates in the deposition of melanin. The enzyme phenoloxidase over many steps converts phenol into quinones that are toxic to bacteria and results in the deposition of melanin, which is also used for encapsulation of parasites and clot formation in wounded flies [62]. (C) The process of phagocytosis has three basic steps: microbe recognition, internalization or engulfment of the pathogen into a phagosome, and the maturation of the phagosome by fusion with lysosomes. The resultant phagolysosome compartment is acidified, and the pathogen is destroyed.

researchers in the field infect Drosophila with M. luteus, which is Gram positive, and E.

coli, which is Gram negative, to induce the Toll pathway and the imd pathway,

respectively [63].

The mammalian homologues, the Toll-Like Receptor family (TLRs), were then

discovered in mice and their innate immune functions were investigated in the influential

work by Medzhitov et, al (1997) [64], Poltorak et al (1998) [65], and Hashino et al (1999)

[66]. The mammalian homologues of these pathways, the TLR pathway and interleukin-1

receptor (IL-1R) pathway, which are most similar to the Drosophila Toll pathways, and

the TNFα pathway, which is analogous to the Drosophila imd pathway, are very well

conserved (Fig 2A). In mammals, these pathways represent a crucial link between the

innate and adaptive arms of the immune system.

Melanization is an immune mechanism used by the fly to sequester and kill

pathogens by a combination of encapsulation and the generation of reactive oxygen

species (ROS) [67]. After wounding or infection, serine proteases, including MP1 and

Page 29: Time to eat

15

MP2, activate a proteolytic cascade in which the enzyme Hayan cleaves pro-

phenoloxidase is into phenoloxidase, which then oxidizes phenols into quinones in the

hemolymph of the fly [68,69] . The quinones themselves and other ROS intermediates of

the proteolytic cascade are toxic to pathogens. Additionally, the cascade ends with the

deposition of melanin, which is not only toxic to pathogens, but it also isolates pathogens

from spreading throughout the hemolymph (Fig. 2B). Because the process of

melanization is also toxic to the fly, the induction of melanization is under tight control

[68]. There are three different protease inhibitors called serpins (Spn27A, Spn28D, and

Spn77Ba) that inhibit phenoloxidase activation to keep melanization in check. To induce

melanization in the lab, we use L. monocytogenes and S. typhimurium, and we can

observe melanin deposits through the cuticle [70].

And lastly, phagocytosis by hemocytes, or specialized immune cells in the

circulatory system, results in the engulfment and subsequent degradation/digestion of

pathogens and injured or dead cells in Drosophila [71]. Briefly, specialized receptors on

the hemocyte’s surface recognize the pathogen and signal for actin cytoskeletal

rearrangement to begin engulfment of the pathogen into an intracellular vesicle called the

phagosome. When the pathogen is completely engulfed, the phagosome fuses with

cellular lysosomes, acidifying the resulting phagolysosome compartment. The contents of

the resulting phagolysosome are then digested and broken down (Fig 2C). In lab, we use

fluorescent beads injected into flies to measure phagocytic activity by hemocytes [63].

The molecular mechanisms of phagocytosis are highly conserved between

vertebrates and lower organisms. Indeed, much of what we know about the molecular

mechanisms of phagocytosis comes from studies of invertebrates, unicellular organisms,

Page 30: Time to eat

16

or tissue culture cells in vitro [71]. For example, RNAi screens of S2 cells, which are

immortalized phagocytic immune cells from Drosophila, have identified many of the

transmembrane receptors, cytoskeletal components, and signaling pathways required cell-

autonomously for phagocytosis by immune blood cells [72].

Circadian regulation of immunity

Sleep and infection have long been thought to be associated with one another. For

instance, sleep is crucial for healing after an illness or stressor [73]. During acute illness,

patients often sleep longer with an irregular cycle [73]. It has been demonstrated in

mammals that many cytokines affect sleep; pro-inflammatory cytokines generally

promote sleep while anti-inflammatory cytokines inhibit sleep [74], suggesting a link

between sleep and immune parameters. For example, the inflammatory cytokines IL-1α,

IL-1β, IL-6, and TNF-α all promote sleep in mammals [75,76]. Additionally, sleep

deprivation and sleep fragmentation lead to altered immune parameters in mammals, with

altered cytokine and chemokine expression and changes in the circadian regulation of

circulating leukocytes [73,77]. There have been a limited number studies examining how

levels of circulating blood cells in humans are affected after sleep deprivation. Two of

these studies demonstrated that the maximum number of circulating B and T cells and

monocytes was delayed significantly and that natural killer cells cycled out of phase with

a much smaller amplitude of oscillation when compared with controls with normal

amounts of sleep [78,79]. These studies suggest a link between sleep and immune

parameters.

The powerful genetic tools available in Drosophila make the organism an ideal

Page 31: Time to eat

17

model to study the molecular interaction between sleep and immunity. Numerous genes

involved in Toll and imd signaling were found to be upregulated in a sleep deprivation

model, including Relish, the NF-κB homolog [80]. In addition, the circadian and sleep

mutant cyc01 exhibited increased levels of many immunity genes, including Relish, even

when their immune systems were not challenged [80]. Lastly, Relish mutants were found

to have disrupted sleep [80]. These results all suggest that sleep and immunity are linked

in Drosophila.

Sleep is a major behavioral output of circadian regulation but is certainly not the

only physiology regulated by the molecular clock. Although there have been many

reports demonstrating an association between sleep and cytokine expression/circulating

leukocyte numbers in mammals and sleep and immunity gene expression in Drosophila,

the molecular and functional connections between the immune system and circadian

components themselves have not been well understood. In a paper published in 2007, Dr.

Shirasu-Hiza, while working in Dr. David Schneider’s lab, found that when wild-type

flies were challenged with pathogens, these flies lost their circadian regulation [81].

Moreover, Tim01 and Per01 mutant flies lacking essential components of the molecular

clock were found to die more quickly than wild-type controls when infected with two

different pathogens, S. pneumoniae and L. monocytogenes. These results illustrate that

the lack of circadian regulation leads to altered immune system function. I expand on

these results in Chapter two, examining the specific immune mechanism underlying

defects in survival of infection of Tim01 mutants. In the Chapter summary, I discuss my

specific contributions to this work.

Page 32: Time to eat

18

Circadian regulation of metabolism and immunity

There are two types of mechanisms that organisms employ to fight pathogens:

resistance and tolerance [82]. An organism uses resistance mechanisms to kill and clear

bacteria, resulting in a reduction of the pathogenic burden on the organism. Resistance

mechanisms are generally thought of as the traditional immune system defenses; for

example, the Drosophila immune system defenses that I described above, antimicrobial

peptides, melanization, and phagocytosis, are all resistance mechanisms. Tolerance

mechanisms give an organism the ability to withstand the pathogenic affects of an

infection, either from damage accumulated by the microbe or the defenses mounted by

the organism itself [83].

Tolerance is not very well characterized, but two examples of tolerance

mechanisms are feeding behaviors and metabolism. Feeding can affect an organism’s

survival to particular pathogens. For instance, in Drosophila, reducing nutrient intake

improves survival when flies are infected with S. typhimurium, E. coli, and E. caratova

but decreases survival rates when flies are infected with L. monocytogenes [84,85].

Additionally, when infected with M. marinum or L. monocytogenes, reduced survival

rates are associated with a decrease in metabolic stores [86,87]. In general, the

contributions of specific nutrients and metabolic pathways on tolerance mechanisms have

yet to be elucidated.

Feeding behavior and the expression of metabolic genes are circadian-regulated;

both Drosophila and mouse circadian mutants display changes in metabolism and altered

feeding behaviors [88,89]. Although we and others have demonstrated that resistance

phenotypes against particular pathogens is circadian-regulated [63,81,90], it is unknown

Page 33: Time to eat

19

whether circadian regulation of these two tolerance mechanisms, feeding and

metabolism, will alter immunity against infection.

Chapter 3 contains a manuscript in revisions describing work in which we

examine how altering feeding behavior and diet in Drosophila can affect tolerance to the

fast and lethal infection with the human pathogen, B. cepecia. In the Chapter summary, I

describe my specific contributions to this manuscript.

Introduction to Fragile X Syndrome

Many neurological diseases are associated with circadian dysregulation, including

Alzheimer disease, Parkinson Disease, epilepsy, stroke, multiple sclerosis, and autism. It

is not yet clear how circadian dysregulation or immune system function affects disease

progression or symptoms. Here I studied immune system function in a Drosophila model

for Fragile X syndrome, a disorder known be associated with circadian dysregulation.

This section will introduce Fragile X syndrome, while Chapter 4 will describe my results.

Fragile X syndrome is the most common monogenic cause of intellectual

disability and autism, occurring in approximately 1:4000 males and 1:8000 females [91].

About 70% of patients with Fragile X syndrome also have autistic features [92], and

patients with Fragile X syndrome make up about 5-8% of all causes of autism [93,94].

Fragile X syndrome often presents in patients with difficulties in learning and memory;

behavioral problems, including anxiety, aggression, and attention; circadian

dysregulation; seizures; and other non-neurological symptoms, such as frequent ear

infections and gastrointestinal symptoms[92,95,96]. Patients with the disease often have

long facies, long limbs, and prominent ears; male patients with the disease also have

macroorchidism [96]. Loss of FMR1 in humans and animal models causes neurons to

Page 34: Time to eat

20

display excessive growth of dendritic spines and defects in synaptic plasticity, but the

mechanism underlying the neuronal dysfunction is unknown [97,98].

In the vast majority of cases, Fragile X syndrome is caused by a triplicate repeat

CGG expansion in the promoter region of the gene FMR1. When there are more than 200

CGG repeats, the promoter region is hypermethylated and the FMR1 gene is

subsequently silenced [99]. Two other disorders occur from the premutation at this gene

locus from an intermediate number, between 55 and 200, of CGG repeats: Fragile X-

associated Tremor/Ataxia Syndrome (FXTAS), which is characterized by intellectual

disabilities and motor symptoms; and Fragile X-associated Premature Ovarian

Insufficiency (FXPOI), which results in the termination of menstruation in a woman

younger than 40 years old. FXTAS and FXPOI have been reviewed by Wheeler, et al

(2013) [100]. The FMR1 gene encodes an RNA-binding protein called Fragile X Mental

Retardation Protein (FMRP), which regulates the translation of many targets. FMRP has

two mammalian homologues, FXR1 and FXR2 (for Fragile X Related 1 and 2), and these

two proteins have some overlapping RNA-binding properties with FMRP [101]. FMRP is

ubiquitously expressed, but is enriched in neurons and in the testis. The role of FMRP has

been widely studied and fairly well characterized in neurons.

Fragile X syndrome was first described as an X-linked disorder when Martin and

Bell examined the phylogeny of a large family affected with the disease (1943) [102].

FMR1 was first cloned in 1991 by Verkerk, et al [103], and was found to be an RNA

binding protein, which was suggested by the presence of the RNA-binding KH domains

and RGG box [104-106]. In 1996, two independent groups showed that FMRP is

associated with ribosomes [107,108], and FMRP was later found to be a translational

Page 35: Time to eat

21

repressor [109]. The vast majority of subsequent studies have focused on the absence of

FMRP on protein synthesis in neurons; the role that FMRP plays in other tissues and cell

types has not been well characterized.

The Drosophila and mouse models for Fragile X syndrome are often utilized to

study molecular, physiological, and behavior phenotypes caused by the absence of

FMRP, as well as the consequences of pharmacological interventions. Both of these

animal models recapitulate many of the symptoms seen in human patients with Fragile X

syndrome, including defects in learning and memory and circadian dysregulation

[110,111]. Mice, like humans, express FMRP and the two homologs, FXR1 and FXR2

[112]. Drosophila, on the other hand, has only one FMRP family member: dFMR1[4].

Drosophila dFMR1 has a high degree of similarity and identity [4]. FMRP shares

functional overlap with dFMR1, as driving the expression of human FMRP in a

Drosophila dFMR1 null animal rescued protein levels in the brain and synaptic structure

[113]. Most of the discoveries discussed below were conducted using these animal

models.

FMRP is an RNA-binding protein that functions as a translational repressor

FMRP has three RNA-binding domains that are crucial for RNA binding: 2 KH domains

(KH1 and KH2) and an RGG box. Although in vitro studies have illustrated that the KH

domains have been shown to bind to RNA kissing complexes [114] and the RGG box

binds to G-quartets [115], an in vivo assay examining where FMRP binds its mRNA

targets has shown that FMRP binds without any particular bias to the RNA coding

sequences [109]. In neurons, FMRP is closely associated with polyribosomes and

Page 36: Time to eat

22

functions to stall polyribosomal translation of the mRNAs to which it binds [109]. The

RNA-binding ability of the KH2 domain is critical for this association, as a missense

mutation in the KH2 domain, I304N [116], abrogates the association of FMRP with

polyribosomes in sole human patient with this mutation[117] and a mouse model of this

mutation [116,118]. This patient has a particularly severe case of the disease [117]. In the

absence of FMRP, there is an increase in protein synthesis in the brain in general [119],

and, administration of an inhibitor of protein translation rescued long-term memory

dysfunction in a Drosophila model of Fragile X syndrome [120]. These findings

highlight the importance of FMRP’s role in stalling polyribosomal translation on

neuronal function.

To try to understand FMRP function in neurons, many studies have been

conducted to elucidate the mRNA targets to which FMRP binds. The first study

examining FMRP targets was conducted by Brown, et al (2001) and used

immunoprecipitation and microarrays to find that roughly 4% of mRNA transcripts in the

brain are potential FMRP targets [121]. Another early study found that dFMR1 binds

futsch mRNA [122], which is the Drosophila homolog of MAP1B, one of the targets that

Brown, et al highlighted. Zhang, et al (2001) determined that the absence of dFMR1

causes increased levels of FUTSCH protein, and that double dfmr1 null/futsch

hypomorph rescues defects in synaptic architecture in the fly [122]. While working as a

technician in the Darnell lab prior to my PhD, we found 842 mRNA targets enriched in

the pre- and postsynaptic compartments to which FMRP binds in vivo by utilizing a

stringent crosslinking-immunoprecipitation (CLIP) protocol and high-throughput RNA

sequencing [109,123]. FMRP binds to a selective group of targets. For instance, FMRP

Page 37: Time to eat

23

interacts with NMDA receptor and mGluR subunits and downstream partners, but FMRP

does not bind to the AMPA receptor or its downstream affectors [109]. These results

suggest that ameliorating protein synthesis may rescue the morphological and behavioral

phenotypes observed in the absence of FMRP.

Synaptic plasticity is altered in Fragile X syndrome

On the cellular level, patients with Fragile X syndrome and animal models of the disease

have immature, thin dendritic spines on their post-synaptic neurons in contrast to the

more mature, mushroom-shaped spines found in controls [97,124]. The dendritic spine

density seen in Fragile X syndrome models and in patients suggests that synapses are not

being appropriately strengthened or matured.

Neurons rely on local protein translation in dendrites and axon terminals for

proper synaptic plasticity. Two forms of synaptic plasticity, long-term potentiation

(LTP), which strengthens synapses through activity, and long-term depression (LTD),

which weakens synaptic strength, both require rapid protein synthesis in postsynaptic

dendrites [125,126]. As FMRP acts as a translational repressor [127] that interacts with

many mRNAs in the pre- and postsynaptic compartment [109], it follows that the absence

of FMRP leads to disordered synaptic plasticity. While LTP is not affected in the absence

of FMRP [128], LTD was enhanced in the absence of FMRP [129]. In wild-type

conditions, excitatory metabotropic glutamate receptor (mGluR) stimulation, one

mechanism important for LTD, causes an increase in the translation of FMRP [130].

Indeed, Darnell et al (2011) found that the mGluR receptor itself is an FMRP target,

along with many downstream affecters of mGluR signaling [109]. Interestingly, even

Page 38: Time to eat

24

though LTD generally requires protein synthesis to occur, Nosyreva and Huber found

that in the presence of a translational inhibitor, FMRP mutants still exhibited enhanced

LTD [131]. As FMRP knockouts exhibit increased LTD, these results suggest that the

absence of FMRP causes excessive protein translation downstream of mGluR signaling,

causing the increased LTD and altered synaptic plasticity. These results led to the mGluR

theory of Fragile X syndrome [132].

Novel therapeutics to treat Fragile X syndrome

In light of the recent studies examining how the loss of FMRP affects neurons, there are

some new therapeutics that are being tested on animal models and on human patients.

One drug in particular, CTEP, is an mGluR inhibitor that has shown to rescue a number

of phenotypes observed in mice, including, learning and memory and the dendritic spine

defect [133]. It is notable that treatment with CTEP in this study began in 4-5 week old

mice, which is after the mice already began to exhibit the Fragile X phenotypes [133].

These results suggest that CTEP may be a viable drug for patients with Fragile X

syndrome, even after they develop symptoms. In fact, there are a number of mGluR

inhibitors that are currently in drug trials [133].

Another avenue for therapeutics is general protein synthesis inhibition. When

dFMR1 mutant flies were treated with the protein synthesis inhibitors cyclohexamide and

puromycin, the defect observed in learning was ameliorated [120]. Although these drugs

are too toxic for use in humans, these results demonstrate that inhibiting protein synthesis

in general is a possible route for drug treatment. One such drug in clinical trials is

minocycline, the tetracycline analog that is able to cross the blood brain barrier

Page 39: Time to eat

25

[134,135]. Minocycline is known to work to inhibit MMP activity and is speculated to

reduce protein translation by stalling ribosomes [111]. Minocycline treatment was found

to reduce MMP-9 in mammals [136] and MMP-1 in flies [137], which are increased in

the murine and Drosophila models of Fragile X syndrome, respectively. Treatment with

minocycline rescued neuronal architecture in both mice and flies and behavioral

phenotypes in mice in the absence of FMRP. These results suggest that the increase in

MMP activity contributes to symptoms in Fragile X syndrome.

While much is known about neuronal defects in Fragile X syndrome, defects in

other cell types, including cells important for immune system function, have not been

well-characterized. Using the Drosophila model of Fragile X syndrome, I examined both

systemic immune system function and the function of glia, the resident immune cells in

the brain. Chapter 4 contains a manuscript in preparation where I describe my results that

the absence of dFMR1 causes altered systemic immune system function and immune

system dysfunction in the brain. In the Chapter summary, I describe my specific

contributions to this manuscript.

Page 40: Time to eat

26

CHAPTER 2:

The circadian clock protein Timeless regulates phagocytosis of bacteria

in Drosophila

Published:

Stone, E. F., Fulton, B. O., Ayres, J. S., Pham, L. N., Ziauddin, J., & Shirasu-Hiza, M. M.

(2012). The Circadian Clock Protein Timeless Regulates Phagocytosis of Bacteria in

Drosophila. PLoS Pathogens, 8(1), e1002445. doi:10.1371/journal.ppat.1002445.s001

Page 41: Time to eat

27

SUMMARY

Shift work, jet lag, and sleep disturbances from noise and light pollution are an

unfortunate consequence of living in a modern society. Shift work and jet lag lead to

altered circadian rhythms, which are the natural, endogenous oscillations of many

molecular, physiological, and behavioral processes with a period of approximately 24

hours. While it seems intuitive that disruptions of sleep and circadian rhythm affect our

immune system function and overall health, the underlying immune mechanisms are not

understood. One focus of our lab is to investigate circadian regulation of immune system

function using Drosophila as a model. We asked what happens to the fly’s immune

system function when the molecular clock, which is responsible for proper circadian

regulation, is genetically silenced.

The molecular components responsible for circadian regulation were first

discovered in Drosophila. Briefly, the transcription factors Clock and Cycle (Clk and

Cyc), translocate into the nucleus and bind to the promoter regions of the Period and

Timeless (Per and Tim) genes, leading to their expression. As the Per and Tim proteins

are expressed, they also heterodimerize and translocate to the nucleus, blocking the

association of Clk and Cyc with the promoter regions for Per and Tim genes. These four

transcription factors cycle in every cell and cause the expression of many other genes to

oscillate throughout the course of the day. When any component of the molecular clock is

silenced, this negative transcriptional feedback loop no longer functions, leading to

circadian dysregulation.

In this study, we examined the well-characterized Tim01 mutant, a null mutant

exhibiting complete loss of circadian regulation, to characterize the impact of loss of the

Page 42: Time to eat

28

molecular clock on the three main branches of the fly’s innate immune system function in

Tim mutants.

In addition to data analysis and intellectual contributions to writing the paper, I

made three major scientific contributions to this paper. First I analyzed the immune

mechanism of melanization in Tim mutants and wild-type controls (Fig. 4A)

Melanization is one of the three main mechanisms of the fly’s innate immune system.

Briefly, when flies are wounded or infected with bacteria or parasites, there is induction

of a proteolytic cascade resulting in the production of reactive oxygen species, which are

toxic to bacteria and parasites, and the deposition of melanin. I examined melanization in

Tim mutants and their wild-type controls at different times during the day, and I found

that Tim mutants do not exhibit differences in melanization compared to wild-type

controls. Additionally, I found that wild-type flies do not exhibit differences in

melanization after infection during their subjective day compared to their wild-type flies

infected during their subjective night. These results suggest that the immune mechanism

of melanization is not circadian-regulated in Drosophila.

Finally, I conducted the critical experiment demonstrating that the immune

mechanism of phagocytosis underlies differences in survival of infection between Tim

mutants and wild-type controls. We had previously found that Tim mutants are more

sensitive to S. pneumoniae infection. Other studies have illustrated that Drosophila fight

S. pneumoniae infection with phagocytosis, or the engulfment of bacteria by hemocytes,

which are immune cells in the fly [138]. In this paper, we showed that Tim mutants

exhibit reduced phagocytosis compared to wild-type controls. We then hypothesized that

the absence of Tim reduces phagocytosis of S. pneumoniae, resulting in reduced survival

Page 43: Time to eat

29

after infection. To test this, I inhibited phagocytosis in wild-type and Tim mutant flies by

by injection of 0.2 µm polystyrene beads, which are phagocytosed and inhibit subsequent

phagocytosis. As a control, I injected identical cohorts with buffer alone (PBS). A few

days later, I infected both bead-injected and PBS-injected controls with S. pneumoniae.

The major result of these studies was that bead inhibition of phagocytosis completely

ablated the difference in survival typically observed between Tim mutants and wild-type

controls (Fig 6). These results suggest that the survival defect of Tim mutants after S.

pneumoniae infection is due to a defect in phagocytosis.

My contributions to this work were published in the following article: Stone, E.

F., Fulton, B. O., Ayres, J. S., Pham, L. N., Ziauddin, J., & Shirasu-Hiza, M. M. (2012).

The Circadian Clock Protein Timeless Regulates Phagocytosis of Bacteria in Drosophila.

PLoS Pathogens, 8(1), e1002445. doi:10.1371/journal.ppat.1002445.s001

Page 44: Time to eat

30

ABSTRACT

Survival of bacterial infection is the result of complex host-pathogen interactions. An

often-overlooked aspect of these interactions is the circadian state of the host.

Previously, we demonstrated that Drosophila mutants lacking the circadian regulatory

proteins Timeless (Tim) and Period (Per) are sensitive to infection by S. pneumoniae.

Sensitivity to infection can be mediated either by changes in resistance (control of

microbial load) or tolerance (endurance of the pathogenic effects of infection). Here we

show that Tim regulates resistance against both S. pneumoniae and S. marcescens. We

set out to characterize and identify the underlying mechanism of resistance that is

circadian-regulated. Using S. pneumoniae, we found that resistance oscillates daily in

adult wild-type flies and that these oscillations are absent in Tim mutants. Drosophila

have at least three main resistance mechanisms to kill high levels of bacteria in their

hemolymph: melanization, antimicrobial peptides, and phagocytosis. We found that

melanization is not circadian-regulated. We further found that basal levels of AMP gene

expression exhibit time-of-day oscillations but that these are Tim-independent; moreover,

infection-induced AMP gene expression is not circadian-regulated. We then show that

phagocytosis is circadian-regulated. Wild-type flies exhibit up-regulated phagocytic

activity at night; Tim mutants have normal phagocytic activity during the day but lack

this night-time peak. Tim appears to regulate an upstream event in phagocytosis, such as

bacterial recognition or activation of phagocytic hemocytes. Interestingly, inhibition of

phagocytosis in wild type flies results in survival kinetics similar to Tim mutants after

infection with S. pneumoniae. Taken together, these results suggest that loss of circadian

Page 45: Time to eat

31

oscillation of a specific immune function (phagocytosis) can have significant effects on

long-term survival of infection.

Page 46: Time to eat

32

INTRODUCTION

Survival of bacterial infection is the result of complex host-pathogen interactions. An

often-overlooked aspect of these interactions is the circadian state of the host. Most

metazoans undergo daily, dynamic changes in their physiology that are regulated by well-

characterized circadian machinery. At its core, the circadian clock is composed of four

transcriptional regulators paired as two distinct heterodimers. In Drosophila, Period and

Timeless form one heterodimer and Clock and Cycle form the other. These two

heterodimers engage in an autoregulatory negative feedback loop that causes circadian

oscillations in the expression levels of target genes [139]. Microarray analyses of

different vertebrate tissues, such as heart, liver, or spleen-derived macrophages, have

shown that approximately 5-10% of total gene expression in each tissue is circadian-

regulated [140,141]. These oscillations in transcription are thought to underlie circadian

changes in the organism’s physiology and behavior.

Reflecting the pervasiveness of circadian regulation, the Drosophila circadian

mutants Timeless (Tim) and Period (Per) have pleiotropic phenotypes. They exhibit loss

of circadian rhythms in locomotor activity, eclosion, male courtship, and oxidative stress

response [142-144]. These circadian mutants are also sensitive to infection by at least

two bacterial pathogens and resistant to infection by another [81,145]. The circadian-

regulated mechanisms underlying these changes in immunity were not known.

Two types of mechanisms can affect survival after infection: resistance and

tolerance [146]. Resistance mechanisms control microbial growth, while tolerance

mechanisms allow the host to endure the pathogenic effects of infection, including the

host’s response to infection. In Drosophila, known resistance mechanisms to control

Page 47: Time to eat

33

microbial growth in the hemolymph include antimicrobial peptide synthesis, reactive

oxygen species generation, and phagocytosis by immune cells. Previously, microarray

analysis comparing wild type and circadian mutants suggested that expression of immune

signaling molecules that induce resistance mechanisms such as Imd (Immune deficient)

may be circadian-regulated [147,148]. Here we test the hypothesis that circadian mutants

are sensitive to infection due to changes in mechanisms of resistance and investigate

specific resistance mechanisms for circadian regulation.

It has been recently shown that in mice, many immune cell functions, such as

cytokine secretion by macrophages and natural killer (NK) cells, undergo oscillatory

time-of-day variation [141,149]. Phagocytosis by immune cells such as macrophages and

neutrophils has been reported to change with time of day, though it is not clear if this

oscillation is diurnal (due to photoperiod) or circadian (regulated by circadian proteins)

[150-153]. Phagocytosis of photoreceptors by pigment cells in the vertebrate retina has

also been shown to be circadian-regulated [154]. Thus we further hypothesized that

phagocytosis of bacteria by immune cells in Drosophila is also circadian-regulated and

that decreased phagocytic activity may significantly contribute to the sensitivity of

circadian mutants to certain types of bacterial infection.

Page 48: Time to eat

34

MATERIALS AND METHODS Fly and bacterial strains

The Canton S (CS) wild-type and CS tim01 lines have been previously described [81].

Infections were performed with the following bacteria as described [155]: Streptococcus

pneumoniae strain SP1, a streptomycin-resistant variant of D39 and gift from Elizabeth

Joyce in Stan Falkow’s laboratory at Stanford University [156]; Listeria monocytogenes

strain 10403S, gift from Julie Theriot at Stanford University [157]; Serratia marcescens

strain DB1140, gift from Man Wah Tan at Stanford University [158]; Salmonella

typhimurium strain SL1344, gift from Falkow lab at Stanford University [159];

Burkholderia cepacia ATCC strain 2541; Escheria coli strain DH5-alpha; and

Micrococcus luteus ATCC strain 4698. S. pneumoniae was frozen at OD600 0.15 in 1 ml

aliquots with 10% glycerol, pelleted upon thawing, rediluted three-fold in fresh BHI

(Brain Heart Infusion media, Difco), and incubated (standing) for 1.5-2 hour at 37˚C with

5% CO2 to OD600 0.15 for injection into flies, which were incubated at 29˚C.

L. monocytogenes was grown in standing BHI, overnight at 37˚C, and diluted to OD600 of

0.1 for injection into flies, which were incubated at 25˚C. S. marcescens was grown in

shaking BHI, overnight at 37˚C, and diluted to OD600 ranging from 0.1 to 0.6 for

injection into flies, which were incubated at 29˚C. S. typhimurium was grown in standing

LB, overnight at 37˚C, and diluted to OD600 of 0.1 for injection into flies, which were

incubated at 29˚C. B. cepacia was grown in standing BHI, overnight at 29˚C, and diluted

to OD600 of 0.01 to 0.001 for injection into flies, which were incubated at 18˚ or 25˚C. E.

coli was grown in shaking LB, overnight at 37˚C, and diluted to OD600 0.1 for injection

Page 49: Time to eat

35

into flies, which were incubated at 25˚C. M. luteus was grown in standing LB, overnight

at 30˚C, and diluted to OD600 0.1 for injection into flies, which were incubated at 25˚C.

Injection

All experiments, including injections, were performed with male flies, 5-7 days post-

eclosion. Flies were raised at 25°C, 55-65% humidity on yeasted molasses food in a 12h

light/dark cycle. For injections, flies were anesthetized with CO2. Injections were

carried out with a pulled glass capillary needle. A Picospritzer III (Parker-Hannifan) or

custom-made Tritech microinjector (Tritech) was used to inject 50 nL of liquid into each

fly. Volume was calibrated by measuring the diameter of the expelled drop under oil.

Injections comparing ZT07 (day) and ZT19 (night) were conducted in a dark room using

red safety lights as light sources.

Survival assays

60-85 flies per genotype per condition were assayed for each survival curve and placed in

3 vials of dextrose food with approximately 20 flies each. In each experiment, 20-40

flies of each line were also injected with media as a wounding control. Death was

recorded daily. Data was converted to Kaplan-Meier format using custom Excel-based

software called Count the Dead. Survival curves are plotted as Kaplan-Meier plots and

statistical significance is tested using log-rank analysis using GraphPad Prism software.

All experiments were performed at least three times and yielded similar results.

CFU determination

Page 50: Time to eat

36

Following challenge with microbes, six individual flies were collected at each time point.

These flies were homogenized, diluted serially and plated on appropriate media (tryptic

soy blood agar for S. pneumoniae, LB for all others). Statistical significance was

determined using non-parametric two-tailed t-tests. All experiments were performed at

least three times and yielded similar results.

Melanization assays

Flies were injected with S. pneumoniae (OD600 0.1) or L. monocytogenes (OD600 0.1) and

incubated at 29˚C. On day 2 of the S. pneumoniae infection or day 3 of the L.

monocytogenes infection, flies were visually inspected for deposits of melanin resulting

from the proteolytic cascade that generates reactive oxygen species. Injection with media

typically causes a wound-induced melanization response, a small deposit of melanin at

the site of injection within 24 hours. Infection with certain pathogenic bacteria such as L.

monocytogenes or S. typhimurium also causes a disseminated melanization response, with

large spots of melanin deposition observed elsewhere than the site of injection: under the

cuticle or in deeper tissue on both dorsal and ventral sides of the abdomen, thorax, and

head. Flies were quantified for both number and size of spots.

Antimicrobial peptide gene expression analysis

Flies were injected with 50 nL of bacterial culture at OD600 0.10 (S. pneumoniae, M.

luteus, or E. coli), or media alone (BHI, Difco). Following injection, flies were placed in

vials containing molasses food and incubated at 25˚C for five hours at the same light:

dark cycle to which they had been entrained. Three groups of six flies were homogenized

Page 51: Time to eat

37

in Trizol and stored at -80˚C until processed. RNA was isolated using a standard Trizol

preparation and the samples treated with DNAse (Invitrogen). Fermentas RevertAid First

Strand cDNA synthesis kit was used to produce the cDNA following the manufacturer’s

protocol. Quantitative RT-PCR was performed using a Stratagene Mx3000 qPCR

machine, with Roche FastStart Universal SYBR Green Master (Rox) mix and the

following primer sets: Drosomycin, Diptericin, Rpl1 [160]. Total mRNA concentration

was normalized using Rpl1 expression. Differences in the infection-induced gene

expression were calculated by normalizing to basal gene expression in uninfected CS

(day sample) taken at the same time point; p-values were obtained by Mann-Whitney

test. The data shown in Figures 4 and S1 pool the results of three independent trials, each

having three biological replicates of six flies in each sample for each condition.

Phagocytosis assays

5-7 day old male flies were injected with 50 nL of 20 mg/ml pHrodo-labeled S. aureus or

E. coli in water (Molecular Probes, cat# A10010 and P35361). The flies were allowed to

phagocytose the particles for 30-60 min. The wings of the flies were removed and flies

were pinned with a minutien pin onto a silicon pad. Fluorescence images were taken of

the dorsal surface using epifluorescent illumination with a Leica MZ3 microscope fitted

with an ORCA camera (Hamamatsu). Images were captured with Openlab (Improvision)

software. The exposure was set such that the brightest images had a very small number

of saturated pixels. The experiment was repeated three times with 6-10 flies for each

treatment.

Page 52: Time to eat

38

Bead inhibition of phagocytosis

Fluorescent 0.2 um polystyrene beads (Molecular Probes, cat# F13080) were injected

into hemolymph to inhibit phagocytosis as previously described [161]. Briefly, 200 ul of

beads in solution were washed three times in sterile water and resuspended in 20 ul final

volume. 5-7 day old male flies were injected with 50 nL of bead solution. To confirm

that phagocytosis was inhibited with this protocol, the in vivo phagocytosis assay was

performed as described above. Phagocytosis was completely inhibited within 2 days of

bead injection. Flies were then injected with S. pneumoniae at OD600 ranging from 0.05-

0.15 (see Survival Assays, above).

Accession numbers

Accession numbers are listed below for the following genes, which were examined in this

study. Accession numbers were obtained from www.uniprot.org: Timeless (P49021);

Drosomycin (P41964); Diptericin (P24492)

Page 53: Time to eat

39

RESULTS / DISCUSSION

Circadian immune phenotypes are pathogen-specific

The fly immune response to infection is highly complex and pathogen-specific.

Previously, we showed that Timeless (Tim) null mutants die more quickly than wild-type

flies when infected with S. pneumoniae [81]. Another group showed that Tim mutants

die less quickly than wild-type flies when infected with P. aeruginosa, suggesting a

complex immune phenotype that is pathogen-specific [162]. To further understand the

immune phenotype of Tim mutants, we infected Tim mutants with three other pathogenic

bacteria (S. marcescens, B. cepacia, and S. typhimurium). We compared Tim mutant

survival time with wild-type flies and confirmed a complex immune phenotype (Figure 1,

A-D). Tim mutants died more quickly than wild-type flies when infected with S.

marcescens and died with wild-type kinetics when infected with S. typhimurium and B.

cepacia. Tim mutants showed no difference in sensitivity to wounding alone. Thus, these

data suggest that Tim activity has different consequences for immunity against different

pathogens.

Tim regulates resistance to specific bacterial infection

We next asked if Tim regulates mechanisms of resistance or tolerance against specific

bacterial pathogens [70]. We functionally distinguished between these two types of

immune defense by comparing the bacterial loads of Tim mutant and wild-type flies after

infection with four different bacterial pathogens (Figure 2, A-D). We found that loss of

Tim protein causes a loss of resistance against two pathogens. Tim mutants die faster

than wild-type flies when infected with S. pneumoniae and S. marcescens and have high

Page 54: Time to eat

40

Figure 1. Tim activity has significant consequences for immunity. Shown here are Kaplan-Meier survival curves comparing Tim null mutants (orange circles) and wild-type control flies (green squares). Tim mutants die more quickly after infection with (A) S. pneumoniae (p<0.0001) or (B) S. marcescens (p<0.0001). Tim mutants die at the same rate as wild-type flies after infection with (C) B. cepacia (p=0.1001), (D) S. typhimurium (p=0.1707), or (E) medium alone (p=0.9836). p-values were obtained by log-rank analysis.

B

Media alone:Tim mutant is like wild type

0 10 20 30 40Time (days)

S. marcescens:Tim mutant is more sensitive

A

WildtypeTim mutant

S. pneumoniae:Tim mutant is more sensitive

E

Time (days)

Perc

ent S

urvi

val

100

75

50

25

00 2 4 6 8 10 12

p < 0.0001***

p = 0.9836n.s.

Time (days)

p < 0.0001***

0 2 4 6 8 10 12 14

Perc

ent S

urvi

val

100

75

50

25

0

D B. cepacia:Tim mutant is like wild type

p = 0.1707n.s.

C S. typhimurium:Tim mutant is like wild type

Time (days)

p = 0.1001n.s.

Perc

ent S

urvi

val

5 10 15 20 250

100

75

50

25

0

Time (days)

Perc

ent S

urvi

val

100

75

50

25

00 2 4 6 8 10 12

Perc

ent S

urvi

val

100

75

50

25

0

Page 55: Time to eat

41

bacterial loads. This suggests that Tim mutants are less able to control microbial growth

and lack resistance to these two pathogens. Tim mutants had no effect on survival or

bacterial growth during infection with S. typhimurium and B. cepacia.  

These results demonstrate that Tim regulates resistance mechanisms for specific

bacterial pathogens. Different physiologies will be important for defense against each

type of pathogen; thus specific microbes can be used to probe different aspects of the

immune system. Circadian proteins like Tim are known to regulate many different types

of behavior and up to 5-10% of the transcriptome of different tissues. If specific

physiologies that contribute to the defense against certain pathogens are circadian-

regulated in Drosophila, then Tim mutants will present immune phenotypes when

infected with those pathogens. Thus we can use our knowledge about the defenses

required for different pathogens to identify those that are circadian-regulated.

Light-induced degradation of Tim protein in wild-type flies causes sensitivity to S.

pneumoniae.

We chose to focus on Tim mutants and their dramatic defect in resistance against S.

pneumoniae. We hypothesized that Tim mutants might lack resistance because of a

developmental defect (such as malformation of immune tissues) or because of a direct

effect of circadian disruption on immunity in the adult fly. To distinguish between these

two possibilities, we disrupted Tim protein in genetically wild-type adult flies using

constant light (Figure 3A). Blue light (~380 to 475 nm) causes complete and rapid

ubiquitin-mediated degradation of Tim protein [163]; normally, this allows the internal

Page 56: Time to eat

42

molecular clock to be synchronized with external light cues [142]. Flies reared in

constant

  Figure 2. Tim regulates resistance to infection by S. pneumoniae and S. marcescens. Mutants lacking resistance, or bacterial killing, have shorter survival times and greater bacterial loads than wild-type flies. Tim null mutants lack resistance to (A) S. pneumoniae and (B) S. marcescens. Tim mutants have bacterial loads similar to wild-type flies when infected with (C) B. cepacia and (D) S. typhimurium. p-values were obtained by two-tailed, unpaired t-test; all experiments were performed with day-time injections.    light exhibit complete loss of circadian rhythm [164]. Here we use constant exposure to

light to induce a Tim-null phenocopy in adult wild-type flies, circumventing any

developmental requirements for Tim protein. Wild-type flies were infected with S.

Col

ony-

Form

ing

Uni

tsPe

r Fly

(CFU

s)

B

D

S. marcescens:Tim mutant lacks resistance

107

105

103

106

105

104

103

Col

ony-

Form

ing

Uni

tsPe

r Fly

(CFU

s)

n.s. not significant p < 0.001 p < 0.0001*****

Days Post Injection

WildtypeTim mutant

1 20

*** **n. s.

B. cepacia:Tim mutant is like wild type

S. pneumoniae:Tim mutant lacks resistance

Col

ony-

Form

ing

Uni

tsPe

r Fly

(CFU

s)

A

C

1 20

*** **n. s.

Days Post Injection0 1 2 3

106

104

102

108

Days Post Injection

Col

ony-

Form

ing

Uni

tsPe

r Fly

(CFU

s) 106

104

102

108 n. s. n. s. n. s. n. s.

S. typhimurium:Tim mutant is like wild type

n. s. n. s. n. s. n. s.

Days Post Injection0 1 2 3

Page 57: Time to eat

43

pneumoniae and incubated either in darkness or constant light. We found that flies

incubated after infection in constant light died faster than those incubated in darkness

   

A

C

Time (days)

Perc

ent S

urvi

val

100

75

50

25

00 1 2 3 4 5 6

Light-induced degradation of Tim protein makesWT flies sensitive to infection

B

Time (days)

sleep/wakecycle

time

Light-induced sleep/wake rhythms do not rescueTim mutant sensitivity to infection

Time (days)

expectedTim protein

levels

time

WT flies injected during the day, when Tim proteinis low, are sensitive to infection

expectedTim protein

levels

injection

time

WT flies in DARKWT flies in LIGHTTim mutants in DARKTim mutants in LIGHT

LightDark

Photoperiod

0 1 2 3 5 6 74

Perc

ent S

urvi

val

100

75

50

25

0

WT inj in DARK phaseWT inj in LIGHTphaseTim inj in DARK phaseTim inj in LIGHT phase

Perc

ent S

urvi

val

100

75

50

25

0masked*

WT in DARKWT in LIGHT:DARKTim in DARK

*Tim in LIGHT:DARK0 1 2 3 5 6 74

injection

injection

Page 58: Time to eat

44

Figure 3. Resistance against S. pneumonia correlates with Tim activity. Shown here are Kaplan-Meier survival curves comparing wild-type flies and Tim mutants infected with S. pneumoniae and subjected to different light conditions as detailed in the schematic diagrams. (A) Light-induced degradation of Tim protein in wild-type flies results in increased sensitivity to S. pneumoniae. Infected wild-type flies incubated in constant light were more sensitive than wild-type flies incubated in darkness (p<0.0001). Tim mutants incubated in constant light died as rapidly as did Tim mutants incubated in darkness (p=0.3404). (B) Wild-type flies were more sensitive to S. pneumoniae when infected at a time of day when Tim protein is low. Wild-type flies were less sensitive when infected 6 hours after lights off (DARK) than 6 hours after lights on, when Tim protein is minimally and maximally expressed, respectively (p<0.0001). Tim mutant sensitivity was not rescued by infection 6 hours after lights off (DARK) (p=0.4134). (C) Light-induced sleep-wake rhythms (a startle response called “masking”) do not rescue Tim mutant sensitivity to infection. Tim mutant sensitivity was not rescued by incubation in cycling light:dark conditions compared to incubation in the dark. In fact, both Tim and wild-type flies were more sensitive to infection in cycling light:dark conditions than in the dark (p=0.0125 and p<0.0001 respectively).    

(p<0.0001). As a negative control, we treated Tim null mutants and found that they were

not further sensitized by constant light, suggesting that Tim is epistatic to and therefore

downstream of light treatment. These results are consistent with the hypothesis that Tim

protein regulates resistance to S. pneumoniae in the adult fly and also demonstrate that

environmental and genetic disruption of circadian regulation alters immunity.  

Wild-type flies are more sensitive to S. pneumoniae when infected at a time of day

when Tim protein expression is low.

If Tim protein positively regulates resistance in the adult fly and if Tim protein levels

oscillate over the circadian day, we predicted that the resistance of wild-type flies should

also oscillate over the circadian day. To test this, we infected two sets of flies, light-

entrained in anti-phase with each other, at times that correspond to minimal and maximal

Tim protein expression. Tim protein levels are lowest approximately 7 hours after lights

turn on (Zeitgeber 07 or ZT07/day) and highest at approximately 7 hours after lights off

Page 59: Time to eat

45

(ZT19/night) [165]. Consistent with minimal and maximal Tim expression, wild-type

flies infected during the day (ZT07) died significantly faster than those infected at night

(ZT19) (Figure 3B). Tim null mutants showed no difference in survival time based on

time of infection. These results suggest that increased Tim expression at night positively

regulates a mechanism(s) of resistance required to fight infection by S. pneumoniae.

Loss of resistance seen in Tim mutants cannot be rescued by a sleep/wake cycle

induced by environmental light cues

“Masking” is a phenomenon in which flies lacking endogenous circadian rhythm exhibit

circadian-like, light-stimulated locomotor activity and sleep-wake patterns due to external

light cues (startle response) [166]. Thus Tim mutants in masking light/dark conditions

exhibit a locomotor response to light stimulus even in the absence of Tim protein. We set

out to ask if masking would restore the immune function of Tim mutants back to that of

wild-type flies. To test this, we infected wild-type flies and Tim mutants with S.

pneumoniae and incubated them either in the dark (relying on endogenous circadian

rhythm) or in circadian light-dark conditions (masking) (Figure 3C). Masking conditions

did not rescue the immune phenotype of Tim mutants infected with S. pneumoniae,

suggesting that this phenotype is not rescued by light-stimulated locomotor activity but

requires the normal expression of Tim protein.

The melanization response is not circadian-regulated

We next examined the three main Drosophila immune responses for circadian regulation:

melanization (generation of reactive oxygen species, or ROS), anti-microbial peptide

Page 60: Time to eat

46

(AMP) synthesis, and phagocytosis by hemocytes. We tested each immune response for

time-of-day differences in wild type and Tim mutants and found evidence for circadian

regulation of phagocytosis but not melanization or AMP gene induction. This is

consistent with our survival data for S. pneumonia, as phagocytic hemocytes are crucial

to control the growth of S. pneumonia in Drosophila [138].

We first tested if the melanization response is circadian-regulated. Insects react to

injury and some bacterial infections with an enzymatic cascade that generates toxic

reactive oxygen species (ROS) and results in the deposition of melanin, visible through

the cuticle as dark black spots [70]. If melanization is circadian-regulated, we would

expect to see a difference in melanized spot formation between wild-type flies and Tim

mutants as well as between wild-type flies at different times of day, but not between Tim

mutants at different times of day. We assayed melanization in wild type and Tim mutants

infected with S. pneumoniae, but did not observe a systemic melanization response,

consistent with published reports [138,167]. We then assayed the melanization response

of flies infected with L. monocytogenes. Flies exhibit a strong melanization response

after L. monocytogenes infection and require melanization enzymes for resistance against

this pathogen [70]. We found no significant difference in melanized spot formation

between wild-type flies and Tim mutants or between wild-type flies at different times of

day (Figure 4A). These data suggest that melanization is not regulated by Tim.

AMP gene expression induced by septic injury is not circadian-regulated

We next tested if antimicrobial peptide (AMP) gene expression is circadian-regulated.

Expression of AMP genes in response to infection is the best characterized immune

Page 61: Time to eat

47

response of Drosophila; these small peptides secreted by the fat body are thought to

control microbial growth by mechanisms such as disrupting bacterial membranes.

Previously, others had found that expression of several immunity signaling genes

   Figure 4. Tim protein does not regulate melanization or AMP gene expression. (A) Wild type and Tim mutants injected with L. monocytogenes at ZT07 (DAY) or ZT19 (NIGHT) have similar levels of melanization. p-values for pair-wise comparisons were not significant: WT (DAY) vs. Tim (DAY), p=0.3551; WT (NIGHT) vs. Tim (NIGHT), p=1.000; WT (DAY) vs. WT (NIGHT), p=0.4031; Tim (DAY) vs. Tim (NIGHT), p=0.8717. p-values were obtained by unpaired, two-tailed t-test; n = 18-20 flies per genotype per condition. (B) Basal (uninduced) levels of Drosomycin expression were increased at ZT11 relative to ZT05, ZT17, ZT23 in both wild type and Tim mutants, suggesting that this change in expression is not mediated by Tim. (C) Infection-induced levels of Drosomycin expression were not significantly different after injection of wild type and Tim mutants with M. luteus at ZT05 (DAY) or ZT 17 (NIGHT). Every pair-wise comparison resulted in p-values greater than 0.05 (not significant) by Mann-Whitney test. (D) Infection-induced levels of Diptericin expression were not significantly different after injection of wild type and Tim mutants with E. coli at ZT05 (DAY) or ZT17 (NIGHT). Every pair-wise comparison resulted in p-values greater than 0.05 (not significant) by Mann-Whitney test.      

A

C

Mel

anin

spo

ts/fl

yR

elat

ive

Expr

essi

on(D

roso

myc

in)

WTDAY

TimDAY

WTNIGHT

TimNIGHT

40

0

10

20

30

WTDAY

TimDAY

WTNIGHT

TimNIGHT

6

4

2

0

B

Rel

ativ

e ex

pres

sion

(Dro

som

ycin

)

WildtypeTim mutant

0 5 10 15 20

6

4

2

0

DR

elat

ive

expr

essi

on(D

ipte

ricin

)

WTDAY

TimDAY

WTNIGHT

TimNIGHT

2000

6000

4000

0

10000

8000

ZT

Page 62: Time to eat

48

controlling AMP gene expression such as Rel and Imd is circadian-regulated

[145,147,148,162]. Here we measured the basal (uninduced) and pathogen-induced

expression of two representative AMP genes, Drosomycin (Drs) and Diptericin (Dpt),

which are widely used as reporters for induction of the Toll and imd signaling pathways,

respectively [168].  

We first compared the basal (uninduced) expression of Drs and Dpt in wild type

and Tim mutants at four time points around the circadian cycle using qRT-PCR: ZT05,

ZT11, ZT17, and ZT23. We found that both wild type and Tim mutants exhibited higher

basal expression of the Toll-regulated AMP Drs at ZT11 than other time points (Figure

4B). This result suggests that basal expression of Drs varies with time of day but that this

difference in expression is not mediated by Tim protein. This time-of-day difference in

basal expression is statistically significant but very small relative to infection-induced

expression levels (Figure 4 C, D). Basal expression of Dpt was not significantly different

at any time of day in either wild type or Tim mutants (Figure S1A).

We next measured Drs and Dpt expression after S. pneumoniae infection in wild

type and Tim mutants at ZT05 and ZT17, approximately when wild-type flies exhibit

differences in survival and Tim mutants do not (see Figure 3). If AMP induction were

circadian-regulated, we would expect to see a difference in AMP expression between

wild-type flies at ZT05 and ZT17 and between wild type and Tim mutants at ZT17 but

not between Tim mutants at ZT05 and ZT17. We found that S. pneumoniae-induced

expression of Drs and Dpt was not significantly different at either time of day in either

wild type or Tim mutants (Figure S1B, C).

Page 63: Time to eat

49

S. pneumoniae is not typically used to induce AMP expression and did not cause

the dramatic induction of AMP gene expression seen with other bacteria relative to media

alone [169]. Thus we next examined circadian-regulation of AMP expression in response

to two bacteria, M. luteus and E. coli, that have been most extensively used to probe

AMP induction and more generally, to quantify signaling through the Toll pathway

(activated by M. luteus) or the Imd pathway (activated by E. coli) [168]. We examined

Drs and Dpt induction in wild-type flies and Tim mutants infected at either ZT07 or ZT19

with either M. luteus or E. coli and collected six hours later for qRT-PCR analysis.

Again, we found that infection-induced expression levels of Drs and Dpt were not

significantly different at these time points in wild type and Tim mutants (Figure 4C,D).

Taken together, these data suggest that, for both Drs and Dpt, basal levels of expression

and short-term expression induced by these bacteria (S. pneumonia, M. luteus, E. coli) are

not circadian-regulated.

Tim regulates phagocytosis of pathogens by immune cells

Finally, we tested for circadian regulation of the Drosophila immune response of

phagocytosis—the physical engulfment and destruction of bacteria by specialized

immune cells. Phagocytosis is typically assayed by measuring the internalization of

fluorescently-labeled bacteria such as S. aureus or E. coli. Here, we measured phagocytic

activity at different times of day in both wild-type flies and Tim mutants. If phagocytosis

is regulated by Tim protein, we predict that phagocytic activity will be high at night and

low during the day in wild-type flies but will not vary with time of day in Tim mutants.

We assayed phagocytic activity in adult flies by injecting dead S. aureus labeled with

Page 64: Time to eat

50

pHrodo, a pH-sensitive rhodamine dye that fluoresces in acidic environments. When

pHrodo-labeled bacteria are phagocytosed and processed into acidic lysosomes, the

phagocytic cell emits red fluorescence and can be imaged through the dorsal surface of

live, intact flies. We found that phagocytic activity was significantly

   Figure 5. Tim protein regulates an early stage of phagocytosis of bacteria by hemocytes. (A) Wild-type flies in the night phase of the circadian cycle have more phagocytic activity than

WT day WT night

WT assayed in DAYWT assayed at NIGHT Tim assayed in NIGHT

Tim assayed in DAY

A

B

pHrodo-labeledS. aureus

0

400

800

1200

1600

WT Tim WT TimS. aureus E. coli

p = 0.51 p = 0.74

Phag

ocyt

ic a

ctiv

ity a

sar

ea o

f flu

ores

cenc

e (p

ixel

s)

p = 0.63

p < 0.001**

n. s. n. s. n. s.

Page 65: Time to eat

51

wild-type flies in the day phase. Shown here are images of the dorsal surface of representative flies after injection of dead S. aureus labeled with the fluorophore, pHrodo, which emits fluorescence in acidic environments such as the lysosomes. Hemocytes that have phagocytosed these bacteria become fluorescent. Dashed lines indicate the region enlarged in the inset. (B) Phagocytic activity for wild type and Tim mutant flies was quantified by measuring areas of fluorescence. Wild-type flies exhibited significant circadian differences in phagocytosis of S. aureus (p<0.001) but not E. coli (p=0.74). This circadian difference is not present in Tim mutant flies with either S. aureus (p=0.51) or E. coli (p=0.63). p-values were obtained by unpaired, two-tailed t-test.  higher at ZT19 (night) than ZT07 (day) in wild-type flies. Phagocytic activity did not

vary with the circadian cycle in Tim mutants (Figure 5A, B). Thus, phagocytic activity

oscillates with circadian rhythm in vivo in wild-type flies but not in Tim mutants,

consistent with the hypothesis that Tim protein up-regulates phagocytosis of S. aureus at

night. If phagocytosis is required to clear S. aureus, these data are consistent with

previous studies demonstrating that wild-type flies exhibit higher rates of survival when

infected with S. aureus at night than when infected during the day [162].  

To test if circadian-regulated phagocytosis is bacteria-specific, we then assayed

phagocytic activity with a different type of bacteria. We injected wild type and Tim

mutants with pHrodo-labeled E. coli at ZT07 (day) and ZT19 (night) (Figure 5B). In

contrast to injections of S. aureus, wild-type flies did not exhibit a circadian difference in

phagocytosis of E. coli. Tim mutants also exhibited no circadian difference in

phagocytosis of E. coli. These results suggest that Tim regulates bacteria-specific

phagocytosis by immune cells.

Phagocytosis can be generally described as three steps: receptor-mediated

substrate recognition and binding, particle engulfment, and phagosome maturation.

Fluorescently-labeled S. aureus and E. coli have often been used in phagocytosis assays

to determine whether specific phagocytic components discriminate between different

types of bacteria. Many cellular steps of phagocytosis subsequent to substrate recognition

Page 66: Time to eat

52

do not appear to be bacteria-specific. For example, inhibition of b-COP, thought to be

involved in phagosome maturation, leads to decreased phagocytosis of both S. aureus and

E. coli [170]. In contrast, some molecular components responsible for phagocyte

recognition of bacteria are bacteria-specific. For example, inhibition of the phagocytic

receptor PGRP-LC leads to decreased binding and phagocytosis of E. coli but not S.

aureus [170]. Another phagocytic receptor, PGRP-Sc1a, mediates phagocytosis of S.

aureus but not E. coli [171]. Thus our finding that Tim up-regulates phagocytosis of S.

aureus but not E. coli suggests that Tim regulates a bacteria-specific step of this process

such as substrate recognition or binding.

Bead inhibition of phagocytosis eliminates the survival difference between wild type

and Tim mutants after infection with S. pneumoniae

Because phagocytosis is crucial in defense against S. pneumoniae [138], these results

suggest that differences in phagocytic activity might contribute to the difference in

survival after S. pneumoniae infection between wild type and Tim mutants. Thus we

tested if inhibition of phagocytic activity would decrease these survival differences.

Phagocytosis can be inhibited in vivo by injection of polystyrene beads; phagocytes

engulf these beads and are unable to phagocytose subsequent injections of fluorescently-

labeled bacteria [161]. Thus we compared the survival kinetics of wild type and Tim

mutants infected with S. pneumoniae with or without bead pre-injection (Figure 6). As

described above, Tim mutants are highly sensitive to S. pneumoniae infection relative to

wild type flies. Consistent with published results, we found that bead pre-injection

increased sensitivity of wild type flies. Bead pre-injection of wild type flies decreased

Page 67: Time to eat

53

survival rate to those similar to Tim mutants, suggesting that inhibition of phagocytosis is

sufficient to recapitulate Tim mutant sensitivity. Consistent with this, we also found that

bead pre-injection did not increase the sensitivity of Tim mutants to S. pneumoniae,

suggesting that phagocytosis is already impaired in these flies. Interestingly, in some

   Figure 6. Bead inhibition of phagocytosis in wild type flies eliminates the difference in survival kinetics with Tim mutants after S. pneumoniae infection. Shown here are Kaplan-Meier survival curves comparing wild-type flies and Tim mutants infected with S. pneumoniae with and without pre-injection of beads to inhibit phagocytosis. Infected wild-type flies pre-injected with beads were more sensitive than wild-type flies without pre-injection (p<0.0001) and had similar survival kinetics as Tim mutants (p=0.0502). Infected Tim mutants had identical survival kinetics with or without bead pre-injection (p=0.1327). In this experiment infected, bead-injected wild-type flies were more sensitive than infected, bead-injected Tim mutants (p<0.0001); however, this result varied from experiment to experiment. experiments, bead-inhibited wild type flies appear to be more sensitive than bead-

inhibited Tim mutants, suggesting that the presence of Tim protein in the absence of

phagocytosis may have a negative effect on survival. Taken together, these results

suggest that Tim-mediated phagocytosis plays an important role in survival of S.

pneumoniae and that loss of Tim is equivalent to total inhibition of phagocytosis, though

Tim mutants are still able to phagocytose. Inhibition of phagocytosis is thought to inhibit

Wild type + beadsTim + beadsWild type, no beadsTim, no beads

Perc

ent S

urvi

val

0 1 2 3 540

25

100

75

50

Time (days)

Page 68: Time to eat

54

phagocytes’ ability to signal the presence of infection; perhaps Tim also plays a role in

these downstream signaling events. Importantly, our results do not rule out the possibility

of multiple roles for Tim protein in immunity against bacterial infection, including S.

pneumoniae infection.  

Circadian regulation of immunity has long been reported anecdotally in humans and other

animals. Here we show that Tim protein has complex, pathogen-specific effects on

immunity, regulating resistance to infection in Drosophila. Resistance against S.

pneumoniae is upregulated by Tim in the adult fly and is not simply a diurnal response

(i.e., activated by light/dark cycles even in the absence of circadian regulatory protein

function). We also identified phagocytosis as a specific circadian-regulated immune

mechanism. This is the first demonstration of a link between Tim protein and the cellular

immune response in vivo. Our data suggests that circadian regulation affects phagocytic

immune cells at an early stage of specific pathogen recognition and makes a significant

contribution to survival of infection.

This work has implications for people whose occupations disrupt their circadian

regulation such as night-shift workers, flight attendants, and hospital staff. Recently, it

was shown that chronic jet lag (shifts in circadian rhythm) in mice did not cause loss of

sleep or increased stress but did cause circadian dysregulation of the innate immune

system and dramatic vulnerability to LPS-induced endotoxemic shock [172].

Furthermore, mice infected with S. pneumoniae during their rest phase died significantly

less quickly than mice infected during their active phase [173]. In modern society,

circadian dysregulation and exposure to bacterial infection at night are likely not

Page 69: Time to eat

55

uncommon. This work demonstrates that disruption of circadian oscillations in

phagocytosis can have significant adverse effects on resistance against infection.

Page 70: Time to eat

56

SUPPLEMENTAL INFORMATION

Figure S1. Tim protein does not regulate AMP gene expression. (A) Basal (uninduced) levels of Diptericin expression were very low and not significantly different at different times of day in wild type and Tim mutants. Every pair-wise combination resulted in p-value greater than

AR

elat

ive

Expr

essi

on(D

ipte

ricin

)4

0

1

2

3

WildtypeTim mutant

0 5 10 15 20ZT

B

Rel

ativ

e Ex

pres

sion

(Dro

som

ysin

)

WTDAYS. p.

TimDAYS. p.

WTNIGHTS. p.

TimNIGHTS. p.

WTDAYBHI

TimDAYBHI

WTNIGHT

BHI

TimNIGHT

BHI

30

0

10

20

C

Rel

ativ

e Ex

pres

sion

(Dip

teric

in)

8000

0

2000

4000

6000

WTDAYS. p.

TimDAYS. p.

WTNIGHTS. p.

TimNIGHTS. p.

WTDAYBHI

TimDAYBHI

WTNIGHT

BHI

TimNIGHT

BHI

Page 71: Time to eat

57

0.05 (not significant) by Mann-Whitney test. (B) Drosomycin expression levels induced by injection of S. pneumoniae or media (BHI) are not circadian-regulated. Wild type and Tim mutants were injected at ZT05 (DAY) or ZT 17 (NIGHT). Pair-wise comparisons of flies injected with S. pneumoniae resulted in p-values greater than 0.05 (not significant) by Mann-Whitney test. Tim mutants injected with media during the day exhibited significantly less Drosomycin expression than wild-type flies injected with media during the day or night or Tim mutants injected at night; all other pair-wise comparisons of flies injected with media did not show significant differences by Mann-Whitney test. S. pneumoniae-infected flies exhibited higher levels of Drosomycin expression than BHI-injected controls; p<0.01 by Mann-Whitney test. (C) Infection-induced levels of Diptericin expression were not significantly different after injection of wild type and Tim mutants with E. coli at ZT05 (DAY) or ZT17 (NIGHT). Pair-wise comparisons within the set of flies injected with S. pneumoniae resulted in p-values greater than 0.05 (not significant) by Mann-Whitney test. S. pneumoniae-infected and BHI-injected flies did not exhibit significant differences in Diptericin expression; p>0.05 by Mann-Whitney test.

Page 72: Time to eat

58

ACKNOWLEDGMENTS

We thank Elizabeth Joyce, Stan Falkow, Man Wah Tan, and Julie Theriot for providing

bacteria; we thank Gerard Karsenty and the Karsenty lab for use of their qRT-PCR

machine, Gary Struhl for his fluorescence microscope, and Laura Johnston and Andrew

Tomlinson for microscopy equipment; and we thank Joel Shirasu-Hiza for help with

computer programming. David Schneider, Joe Takahashi, Jen Zallen, Tarun Kapoor, and

Julie Canman provided insightful comments on this manuscript. Marc Dionne, Paul

Taghert, and Amita Sehgal provided invaluable discussion, as did members of the

Schneider lab and other members of the Shirasu-Hiza lab (Victoria Allen, James Cheong,

Albert Kim, Paul Kim, Yang “Sunny” Li, Reed O’Connor, Adam Ryan). We thank the

reviewers for useful comments and experimental suggestions.

Page 73: Time to eat

59

CHAPTER 3:

Host tolerance of infection is promoted by dietary glucose and amino

acids through insulin and TORC2 signaling

This manuscript is being revised for resubmission to Current Biology:

Victoria W. Allen, Reed M. O’Connor, Clarice G. Zhou, Vanessa M. Hill, Elizabeth F.

Stone, Royce B. Park, Keith R. Murphy, Julie C. Canman, William W. Ja, and Mimi M.

Shirasu-Hiza

Page 74: Time to eat

60

SUMMARY

Sepsis is extraordinarily life threatening, with a mortality rate spanning 35-56% [174].

Sepsis and its related syndromes, severe sepsis and septic shock, are prevalent

worldwide, and accounted for 4.2% of hospital stays in 2009 in the US alone [175]. The

proper treatment for patients with sepsis is still debated and rife with controversy,

possibly contributing to the high morbidity and mortality rates [176]. While it is known

that sepsis syndrome results from the body’s response to numerous infections, the role of

the circadian regulation and the host’s metabolism on sepsis syndrome remains unknown.

To test the circadian regulation of a sepsis-like infection and the metabolic

consequences, we created two models for sepsis by infecting wild-type and Period (Per)

mutant Drosophila with high bacterial loads of P. aeruginosa and B. cepecia, creating an

overwhelming disease burden in the fly, to investigate the contribution of the host

metabolic state on survival after infection. These results are currently being revised for

resubmission to Current Biology.

I made four contributions to this paper. First, I pioneered the sepsis model in lab

by infecting Drosophila with a high bacterial load of P. aeruginosa. When I infected Per

mutants and wild-type controls with P. aeruginosa, I found that Per mutants survive

significantly longer after infection compared to wild-type controls (Appendix I, Fig. 1A-

B). Because this is a fast infection and the fly’s feeding behavior is circadian-regulated,

we speculated that timing of feeding contributed to the improved survival of Per mutants.

We hypothesized that Per mutants irregular feeding behaviors contributed to their

improved survival after infection. When I placed wild-type and Per flies on a restricted

diet and then infected the flies with P. aeruginosa, diet-restricted Per mutants survived

Page 75: Time to eat

61

like diet-restricted wild-type flies (Appendix I, Fig 1C-D). My second contribution

demonstrated that feeding is the key physiology impacting Per tolerance to infection.

Although we had decided to use B. cepecia for infections in the final paper, the P.

aeruginosa results provided a framework and a phenotype for the paper on which the rest

of the paper was modeled.

The following two experiments that I had conducted were included in the paper.

For my third contribution, I had investigated whether phagocytosis was the mechanism

behind the tolerance phenotype seen in Per mutants after infection with B. cepecia (Fig

1H). I injected wild-type and Per mutant flies with either small polystyrene beads to

inhibit phagocytosis or PBS alone as a control. The fly’s hemocytes phagocytose these

beads and are subsequently unable to phagocytose bacteria for the duration of the fly’s

life. These flies were then injected with B. cepecia two or three days later after bead

injection, and survival was counted. While bead-inhibition reduced survival times in

both wild-type and Per mutants, we found that bead inhibition did not abolish the survival

advantage that we had observed in Per mutants flies. These results suggest that there is

not a phagocytosis advantage in Per mutants with B. cepecia infection.

We had found that Per mutants have reduced energy stores, Per mutants eat more

than wild-type controls, and that Per mutants have reduced survival with a restricted diet.

We examined which nutrient signaling pathways contribute to survival after infection.

One of the major nutrient signaling pathways in the fly is insulin-like signaling, and the

phosphorylation of Akt (p-Akt), which is downstream of insulin-like signaling, is readout

for glucose-mediated insulin-like signaling. For my final contribution, I tested the

presence of p-Akt in wild-type and Per mutants by Western blot analysis, and I found that

Page 76: Time to eat

62

Per mutants have increased p-Akt compared to wild-type controls, suggesting increased

insulin-like signaling (Fig 5C). These results demonstrated that increased insulin-like

signaling is correlated with improved survival after B. cepecia infection.

Those two experiments showing that phagocytosis is not important for the

increased tolerance in Per mutants after infection with B. cepecia and that Per mutants

have increased p-Akt signaling were included in the following manuscript, which is

currently being revised for resubmission to Current Biology: Allen V. W., O’Connor R.

M., Zhou C. G., Hill V. M., Stone E. F., Park R. B., Murphy K. R., Canman J. C., Ja W.

W., and Shirasu-Hiza M. M. Host tolerance of infection is promoted by dietary glucose

and amino acids through insulin and TORC2 signaling.

Page 77: Time to eat

63

ABSTRACT

Background: Nearly all metazoans undergo dynamic circadian-regulated changes in their

behavior and physiology. Currently it is not known how circadian-regulated behaviors

impact immunity against infection. There are two broad categories of defense against

bacterial infection: resistance mechanisms to control microbial growth and tolerance

mechanisms to withstand the pathogenic effects of infection.

Results: Our study of the Drosophila circadian mutant Period, which is behaviorally

arrhythmic, led us to identify a novel link between increased feeding behavior and

increased host tolerance of infection by B. cepacia, a bacterial pathogen of rising

importance in hospital-acquired infection. We further found that infection tolerance in

wild-type animals is stimulated by transient exposure to dietary glucose and amino

acids. Glucose-stimulated tolerance is mediated by the insulin-like signaling pathway

and can be induced by feeding or direct glucose injection. Using glucose injections, we

identified a surprisingly narrow window for glucose stimulation of tolerance, between 2

hours before and after infection. Amino acids-stimulated tolerance is mediated by TOR

Complex 2 (TORC2), not TORC1--an unexpected result considering the known role of

TORC1 as an amino acid sensor. In contrast to glucose, amino acids must be

administered by ingestion to stimulate tolerance and is facilitated by the host microbiota.

Conclusion: Taken together, this work demonstrates that circadian mutants can be used

as a tool to dissect physiologies important for survival of infection and indicates that the

nature, quantity, and timing of dietary intake on the day of infection by B. cepacia can

Page 78: Time to eat

64

make a significant difference in long-term survival.

Page 79: Time to eat

65

INTRODUCTION

Nearly all metazoans undergo daily, dynamic changes in their behavior and physiology

regulated by well-characterized, evolutionarily conserved circadian mechanisms [139].

Currently it is not known how circadian-regulated behaviors may impact immunity

against infection. The circadian clock is composed of four transcriptional regulators

paired as two heterodimers engaged in an autoregulatory transcriptional negative

feedback loop [142]. In Drosophila, Clock and Cycle form one heterodimer and Timeless

(Tim) and Period (Per) form the other. Clock and Cycle proteins are transcriptional

activators that activate expression of Tim and Per as well as hundreds of tissue-specific

target genes [21,139,177]. Circadian oscillations in gene expression are thought to cause

circadian oscillations in physiological function and, ultimately, organismal behavior.

One evolutionarily conserved, circadian-regulated physiology is immunity against

infection [63,81]. For both flies and vertebrates, innate immunity is the first line of

defense against infection. Drosophila lack adaptive immune components such as T cells

and B cells and rely solely on innate immune responses to survive infection [178].

Evolutionary conservation extends to the two primary Drosophila immune signaling

pathways, the Toll and imd (immune deficiency) pathways [179]. Flies and vertebrates

employ several similar innate immune effectors to kill bacteria, including: phagocytosis

by immune cells; reactive oxygen species generation, which in the fly is caused by

melanization; and secreted antimicrobial peptides (AMPs) with bacteriocidal properties.

Resistance is only one type of immune mechanism by which host organisms

combat bacterial infection. Resistance mechanisms control bacterial proliferation during

the course of infection, thus reducing the effects of infection by reducing the host’s

Page 80: Time to eat

66

pathogen burden. A second distinct and complementary mechanism of immunity is

termed tolerance [146,180]. Tolerance mechanisms allow the organism to survive the

pathological effects of infection—whether caused by the microbe itself or by the host

immune response—without necessarily decreasing bacterial load [181,182].

Infection tolerance mechanisms are not well understood but include feeding and

metabolism. In Drosophila, decreased survival of infection is associated with decreased

metabolic stores during infection with two different bacterial pathogens, M. marinum or

L. monocytogenes [86,183]. The effect of feeding behavior on infection is also pathogen-

specific: decreased feeding increases survival of S. typhimurium, E. coli, and E. caratova

infections but decreases survival of L. monocytogenes infection [184,185]. In many

cases, the precise nutrients important for survival and the underlying molecular signaling

pathways have not been identified.

Both feeding behavior and expression of metabolic genes are known to be

circadian-regulated and both fly and mouse circadian mutants exhibit metabolic disorders

and changes in feeding behavior [186,187]. While we and others have shown previously

that host resistance against specific pathogens is circadian-regulated, it has not yet been

shown whether loss of circadian regulation of metabolism and feeding behavior will

affect immunity against infection [63,81,90].

Here we exploit a rapid, lethal infection of Drosophila with the human pathogen

B. cepacia to examine how small, acute differences in feeding behavior and diet can

impact infection tolerance. We show that infection tolerance is stimulated by an acute

flux of both dietary glucose and amino acids at the time of infection and that these effects

are respectively mediated by the insulin-like signaling pathway and the TORC2 signaling

Page 81: Time to eat

67

pathway. Injected glucose can stimulate tolerance when administered within two hours

before or after infection; in contrast, amino acids must be ingested and their effects are

mediated by the fly's microbiota. This work suggests that what and how much the host

eats on the day of infection can make a significant difference in their long-term survival.

Page 82: Time to eat

68

MATERIALS AND METHODS

All infection experiments were performed at least three times and yielded similar results.

p-values were obtained by unpaired t-test unless otherwise noted.

Fly lines

Per01mutants were compared to an isogenic white-eyed (w1118) Canton S strain. Oregon

R flies were used in experiments examining specific dietary components. dFOXO

mutants were trans heterozygotes generated by crossing dFOXO21 and dFOXO25

heterozygous parentals (gift of Marc Dionne [86]). UAS-Tsc1/Tsc2 (gift of Marc Tatar

[188]) homozygous males were crossed to w1118; Gal80-ts; tub-Gal4/TM6c (gift of Marc

Dionne) virgins and maintained at 18º through development. To induce activation of the

transgenes, flies were placed at 29º between 24 and 48 hours prior to infection.

Transgene induction after 24 hours at 29º was confirmed by qRT-PCR (see below, also

Fig. S3). RicTOR null mutants (gift of Stephen Cohen [189]) were derived by imprecise

excision of a p-element inserted into the RicTOR coding sequence. Controls for these

mutants were derived by precise excision of the same p-element. Axenic (germ free) OR

flies were generated as described below.

Fly cultures and media

Flies were bred and raised on standard food containing cornmeal, molasses, and yeast.

2.10 kg of cornmeal, 0.96 kg of agar, and 1.1 L of molasses were combined with 36 L

(summer) or 40 L (winter) distilled water. The mixture was boiled for 25 minutes, then

let cool for 45 minutes, or until the temperature reached 65°C or less. Methylparaben

Page 83: Time to eat

69

(Tegosept, 50 g in 300 mL ethanol) and 180 mL of propionic acid were then added as

antifungal agents. Sugar-only foods contained 1% (w/v) agar (Fisher #A360-500 or Alfa

Aesar #A10752) and different concentrations (w/v) of glucose (Fisher #BP350-1 or

#BP350-500) dissolved in ddH20. Amino acid-containing food contained 1% agar, 5%

glucose, and 2% (w/v) amino acid mixture per the manufacturer’s instruction (Sunrise

Science 1360-030). For experiments involving diets other than standard food, flies were

switched to the experimental food ~24 hours before infection.

Bacterial cultures

Infections were performed as previously described [155] with Burkholderia

cepacia (ATCC strain #25416). B. cepacia was grown overnight at 29°C in 5 mL

standing Brain Heart Infusion (Teknova), resuspended in sterile PBS (Invitrogen

#003000), and diluted to an OD600 of 0.025 or 0.0275 for injection into flies (short

infection at 29°C) or OD600 of 0.1 (long infection at 18°C).

Injections

All experiments, including injections, were performed with age-matched male flies, 5–10

days post-eclosion. All flies except Gal80-ts; tub-Gal4>UAS-Tsc1/Tsc2 flies were raised

at 25°C, 55–65% humidity on standard food and entrained to a 12h light/dark cycle in a

Darwin Chambers incubator for at least 3 days prior to infection. Gal80-ts; tub-

Gal4>UAS-Tsc1/Tsc2 flies were raised at 18ºC through development. For transgene

induction, flies were shifted to 29ºC for 24 hours prior to infection. For injections, flies

were lightly anesthetized with CO2. Injections were carried out with a 10µL Drummond

Page 84: Time to eat

70

Scientific glass capillary needle (#3-000-210- G), machine-pulled by a Sutter Instrument

Co. machine (Model P-30). A custom-modified Tritech microinjector was used to inject

50 nL of liquid into each fly. Volume was calibrated by measuring the diameter of an

expelled drop in halocarbon oil 700 (Sigma, H8898) under a layer of mineral oil (Sigma,

M8410). All injections were performed between Zeitgeber time (ZT) 7.5 and ZT 10.5 to

minimize variability from circadian effects on immunity. After injection with B. cepacia

flies were incubated at 29°C, except for long infections and heatshock control

experiments in which flies were incubated at 18°C. For co-injection of amino acids, B.

cepacia was resuspended in either PBS or a solution of amino acids dissolved in PBS

(low dose 11 µg/mL or high dose 220 µg/ml). For co-injection of rapamycin, flies were

injected with B. cepacia suspended in either buffer alone or in buffer containing 0.2

mg/mL rapamycin (CALBIOCHEM #553210).

Survival assays

Except where otherwise noted, approximately 60 male flies per genotype per condition

were assayed for each survival curve and divided into vials with approximately 20 flies

per vial. In each experiment, approximately 20 flies of each line were also injected with

media to control for the effects of wounding alone. Flies were injected between ZT 7.5

and ZT 10.5 and incubated at 29˚C in 12:12 light:dark conditions. Death was assayed by

visual inspection and recorded the next day every hour or more frequently. Data were

converted to Kaplan-Meier format using custom Excel-based software called Count the

Dead (J. Shirasu-Hiza). Survival curves were plotted as Kaplan-Meier survival plots and

statistical significance was tested by log-rank analysis using GraphPad Prism software.

Page 85: Time to eat

71

Median survival times were compared in order to determine percentage-differences in

survival time.

Starvation assay

Using the DAM5 system (TriKinetics), age-matched 5-7 day old male flies were

incubated at 25º, 55-65% humidity in a 12:12 LD:DL cycle on media containing 1% agar

alone. For each genotype, n=16 flies per experiment. Flies were counted as dead

following an absence of beam crossings for the remainder of the experiment. p-values

for survival curves were obtained by log-rank analysis. Similar results were obtained

when starvation experiments were performed using visual inspection instead of DAM5 to

confirm death.

Bacterial load quantitation

Following challenge with microbes, approximately six individual flies were collected at

each time point. Time points chosen typically included just after infection (0h); mid-

infection (6h); and late in infection (12-16h) just before flies began to die. Flies were

homogenized in PBS, diluted serially and plated on standard LB agar plates. Statistical

significance was determined using unpaired t tests for 0 hour time points, while

subsequent time points were tested with non-parametric Mann-Whitney U tests, which

does not assume that samples have normal distribution. To ensure that flies were not

already infected with bacteria and that the buffer was not contaminated, 3 PBS-injected

flies were collected at the beginning of the experiment, homogenized and plated on LB

agar plates. Typically no colonies were observed for uninfected flies.

Page 86: Time to eat

72

Quantitative real-time RT-PCR

RNA was extracted from three or four samples of 5-6 adult flies using TRIZOL

(Invitrogen) according to the manufacturer’s directions. RNA was treated with DNAse I

(Invitrogen, #18068-015), followed by reverse transcription using the First Strand cDNA

Synthesis Kit (Thermo Scientific, #K1622), priming with random hexamers.

Quantitative PCR was performed in 25 µL reactions with Roche Fast Start Universal

SYBR Green Master Mix (#04913850001) on a Stratagene Mx3000P or a Bio-Rad CFX

Connect Real-Time System. The cycling conditions used were as follows: Hold 95°C for

10 minutes, then 40 cycles of 95°C for 15s, 58°C for 30s, 72°C for 30s, then a final

sequence of 95°C for 1 minute, 57° for 30s, and 95° for 30s. All calculated gene

expression values were initially normalized to the transcript level of ribosomal protein 4,

Rpl1, prior to further analysis. Primer sequences are listed in Supplemental Table 1.

Melanization assays

Two groups of flies were injected with B. cepacia (OD600 0.005 – 0.05) and incubated at

18°C. One day after infection, flies from one group were each visually inspected for

wound site melanization, a small deposit of melanin at the site of injection. Three to

twelve days after infection, flies from the second group were visually inspected for the

systemic melanization response, characterized by one or more melanin deposits located

anywhere on the body other than at the injection site. p-values were obtained by paired t-

test for three independent trials.

Page 87: Time to eat

73

Bead inhibition of phagocytosis

Phagocytosis was inhibited by injecting fluorescent 1.0 µm polystyrene beads (Molecular

Probes, cat# F13080) into hemolymph as previously described [63]. 200 µl of beads in

solution were washed three times in sterile PBS and resuspended in PBS for a final

volume of 30 µl. 7-10 day old male flies were injected with either 100 nL of bead

solution or buffer (PBS) as a control. To confirm that phagocytosis was inhibited with

this protocol, an in vivo phagocytosis assay was performed 2-3 days post bead inhibition

on a small sample of bead- and PBS-injected flies. Briefly, flies were injected with 50 nL

of 20 mg/mL pHrodo-labeled S. aureus in PBS (Molecular probes, cat# A10010). Flies

phagocytosed the bacteria for 20-30 minutes. The dorsal aspects of the flies were

superglued to coverslips, and the dorsal surface was imaged using epifluorescence

illumination with a Nikon upright fluorescent microscope with HQ CCD using Nikon

Elements software. Phagocytosis was completely inhibited within 2 days of bead

injection. Flies were then injected with B. cepecia at an OD600 of 0.025 (see Survival

Assays, above). p-values for survival curves were obtained by log-rank analysis.

Metabolic storage assays

For metabolic assays, samples were prepared by homogenizing 8 adult male flies (5-10

days old) in 200 ul Tris-EDTA (TE; 10 mM Tris, 1 mM EDTA, pH 8) + 0.1% Triton X-

100, then freezing at -20ºC. Triglyceride levels were measured as previously described

[190]. Triglyceride levels were blanked against reactions lacking lipase, which measured

the baseline levels of glycerol in the samples. Glucose and glycogen were measured as

described [86]. Briefly, the glucose and trehalose levels were measured using a glucose

Page 88: Time to eat

74

oxidase reagent (Pointe Scientific) dissolved in ddH2O. Glycogen was measured using

the same reagent supplemented with 1 U/mL amyloglucosidase (Sigma, 10115-1G-F).

This reaction was blanked against the glucose reaction for each given sample. All

reactions were performed on 96-well tissue culture plates with 0.2 ml reaction mixture +

0.02 ml sample. Plates were incubated at 37º for one hour, and then the absorbance for

each well was measured at 490 nm using a BioRad Model 680 Microplate Reader. The

values obtained for each metabolic reserve were normalized to the average weight of flies

of that given genotype. Average weight was determined by measuring the mass of 8

samples of 50 adult male flies from each genotype. The resulting ratio of wild type to

Per mass was 1.054; in no case did this adjustment convert a result from being

insignificant to significant, nor vice versa. These values were then normalized to the

mean obtained from the wild type values and plotted with the normalized SEM.

Feeding assays

CAFE assays and 32P feeding assays were performed as previously described [191-193].

Briefly, for the radiolabeled feeding assay, flies were maintained on medium

supplemented with α-32P-dCTP for 24 h and then collected for liquid scintillation to

measure accumulated 32P. Aliquots of radiolabeled food were used to convert

scintillation counts to volume. For the CAFE assay, flies were fed liquid food solutions

of yeast extract and sucrose in glass microcapillaries that facilitated the direct

measurement of volume consumed.

Protein extraction and Western blotting

Page 89: Time to eat

75

Four whole flies were homogenized in 80 µL of extraction buffer (10 mM HEPES, 100

mM KCl, 5% glycerol, 0.1% Triton, and 0.5 mM phenylmethylsulfonyl fluoride).

Samples were stored with 20 µL of loading buffer (20% glycerol, 10% SDS, 250 mM

Tris-HCl pH 6.8, 5% 2-mercaptoethanol, bromphenol blue), and processed as previously

described [194]. Briefly, samples were resolved on a 4-15% SDS polyacrylamide tris-

glycine gel, transferred overnight at 4°C at 22 volts, and immunoblotted using standard

procedures. Blots were blocked using 5% Carnation Instant Nonfat Dry Milk in PBS.

Antibodies were diluted in 5% milk in PBS. p-Akt, α-mouse, and α-rabbit antibodies

(Cell Signaling #4054S, #7076, #7074S) were used at a dilution of 1:1000. The actin

antibody was purchased from Abcam (ab8224) and used at a concentration of 1:10,000.

Immobilon Western (Millipore #WBKLS0500) was used as a chemiluminiscent

substrate. Fiji (a distribution of Image J) quantification of phospho-Akt signal was

performed by measuring the mean intensities of the phospho-Akt bands, which were

normalized to the mean intensities of the actin controls for each sample.

Preparation of germ-free flies

Standard food was autoclaved, added sterilely to autoclaved vials, and allowed to solidify

under UV light for several hours. Vials were then covered with parafilm and left under

UV light overnight, then plugged with autoclaved flugs the next day. These vials were

stored in a plastic tray that was treated with 70% ethanol, 10% bleach, and UV light

which was placed in a similarly treated Sterilite plastic storage bin. Axenic (germ-free)

eggs were prepared using a modified version of a previously published method [195].

Briefly, 5-8 males and 25-40 females were allowed to mate on petri dishes containing

Page 90: Time to eat

76

food prepared by autoclaving 20g agar (Fisher #A360-500), 50g sucrose (source), 1L

ddH2O, and 8-10 drops blue food coloring (source) for 30 minutes. This food was

allowed to cool overnight under UV lights and was then partially covered with yeast

pellets before flies were added. These flies were allowed to breed overnight in a 25º

incubator with a 12:12 light:dark cycle. Before 10 am the following morning, flies were

removed from the dishes using bleached tools to prevent contamination. In a biohazard

hood, agar plates were washed with either 10% bleach (axenic) or ddH2O (conventional)

for approximately 2.5 minutes. Next, the contents of each dish were added to a nylon

mesh filter basket that was treated with bleach and ethanol. Embryos were then washed

with either 10% bleach (axenic) or ddH2O (conventional). Both sets were then washed

with ddH2O to remove bleach and yeast pellets. Embryos were then transferred to sterile

food using a sterile paintbrush. Both conventional and axenic flies were grown and

maintained under sterile conditions.

Page 91: Time to eat

77

RESULTS

Period mutants exhibit increased tolerance when infected with B. cepacia.

We found that arrhythmic Per mutants survived longer than wild type flies when infected

with the human pathogen Burkholderia cepacia, which has previously been characterized

using Drosophila [155,196-198]. This increased survival was seen following both fast

and slow infection conditions: either a low bacterial dose at low temperature or a high

dose at high temperature (Fig.1A and 1B, p<0.0001 for both). To determine if this

increased survival was due to changes in resistance or tolerance, we measured bacterial

loads in individual flies at different time points for both fast and slow infections. For

both conditions at each time point, we found that wild type and Per mutants carried

equivalent bacterial loads (Fig.1C and 1D, p>0.05 for each time point). This result

suggests that the enhanced survival of Per mutants is not due to greater resistance, but

due to greater tolerance.

Known resistance mechanisms do not explain increased survival of infection.

To confirm that Per mutants do not have increased resistance to B. cepacia, we

quantitatively analyzed specific resistance mechanisms following infection. We

compared wild type and Per mutants infected with B. cepacia for the three well-

characterized mechanisms of resistance in Drosophila: antimicrobial peptide (AMP)

induction, melanization, and phagocytosis. We found no significant differences in B.

cepacia-induced AMP expression between wild type and Per mutants (Fig. 1E, 1F,

Supplemental Fig. 1A-E). We also found no difference in the B. cepacia-induced

melanization response between wild type and Per mutants (Fig. 1G). Wound-site

Page 92: Time to eat

78

melanization was unchanged and, as previously published, B. cepacia does not induce

 Fig. 1: Period mutants exhibit greater tolerance than isogenic controls during infection with B. cepacia. Per mutants survived longer than wild type during A) a long infection (Per, n=78; WT, n=77) and B) a short infection (Per, n=57; WT, n=64) with B. cepacia. Per mutants and wild type flies had similar bacterial loads over time following a C) long infection (n≥4 flies/time

n.s. n.s.

0 5 6

n.s.

C D

Bacte

ria

l lo

ad

/fly

B. cepacia, 29°C

Hours post-infectionHours post-infection

A B

0

25

50

75

100

Hours post-infection

0

25

50

75

100

p < 0.0001

Days post-infection

p < 0.0001

WT

Per

Rela

tive e

xpre

ssio

n

50

0

100

Hours post-infection

0 6 12

n.s. n.s.n.s.

Toll activation:

DrosomycinE Imd activation:

Diptericin

Hours post-infection

1000

0

2000

F

B. cepacia, 18°C B. cepacia, 29°C

0 6 12

n.s. n.s.n.s.

Pe

rce

nt

surv

iva

l

n.s. n.s.

0 4 14

n.s.

102

104

106

108P

erc

en

t surv

iva

lB

acte

rial lo

ad

/fly

Re

lative e

xpre

ssio

n

Figure 1: Allen et al., 2014

G

0

50

100

wound systemic

Melanization assay:

wound site vs. systemic

n.s. n.s.

Perc

ent

of flie

s

H Bead-inhibition

of phagocytosis

Hours post-infection

Perc

en

t su

rviv

al p < 0.0001

0

25

50

75

100

20 25 30

Pre-injection with:

beads, Per

beads, WT PBS, WT

PBS, Per

WT

Per

0 5 10 15 0 5 10 15 20 25

B. cepacia, 18°C

102

104

106

Page 93: Time to eat

79

point) and D) short infection (n=6 flies/time point) with B. cepacia; p-values were obtained by unpaired t-test (0h) and non-parametric Mann-Whitney test (other time points). Consistent with a tolerance phenotype, antimicrobial peptide (AMP) induction was not different between Per mutants and wild type flies after B. cepacia infection as shown by: E) Drosomycin and F) Diptericin. Other AMPs are shown in Fig. S1. G) Per mutants and wild type flies did not exhibit differences in systemic and injection site melanization after B. cepacia infection (3 trials, n=17-22 flies/trial/genotype). H) Inhibition of phagocytosis by bead pre-injection did not remove Per mutant survival advantage over wild type after B. cepacia infection (Per, n=76 with beads, n=81 with buffer; wild type, n=81 with beads, n=80 with buffer; p<0.0001 for all pair-wise curve comparisons). p-values for survival curve comparisons were obtained by log-rank analysis; all others were obtained by unpaired t-test unless otherwise noted. n.s.=p>0.05.  

systemic melanization [70]. Inhibition of phagocytosis by bead pre-injection decreased

survival of both Per and wild type controls, yet Per mutants still survived significantly

longer than wild type flies (Fig. 1H). This result suggests that phagocytosis is not

responsible for the increased tolerance of Per mutants. Taken together, these results

indicate that tolerance, and not resistance, leads to increased survival of Per mutants after

B. cepacia infection.  

Per mutants have decreased energy storage.

We next wanted to determine the mechanism(s) leading to increased tolerance in Per

mutants during B. cepacia infection. It has been shown that decreased survival of other

bacterial infections, M. marinum and L. monocytogenes, is associated with loss of

metabolic stores [86,183]. Moreover, metabolic gene expression is known to be

circadian-regulated [186,187] and regulated by immunity genes [199,200]. These

findings raise the possibility that the increased tolerance of Per mutants during infection

is due to increased metabolic stores. If so, uninfected Per mutants would be more

resistant to starvation than wild-type controls. In fact, we found the opposite: Per

mutants are more sensitive to starvation than wild-type controls (Fig. 2A, p<0.0001),

Page 94: Time to eat

80

suggesting that Per

 Fig. 2. Per mutants have lower metabolic resources and eat more than wild type flies. A) Uninfected Per mutants were more sensitive to starvation than uninfected wild type flies (Per n=15; WT n=12). B) Quantification of metabolic storage levels comparing uninfected Per mutants and wild type flies (n=12 for both) revealed that Per mutants had lower levels of triglycerides (p=0.0004) and glycogen (p=0.0007) and similar levels of primary circulating sugars. C) 16 hours after infection with B. cepacia, Per mutants relative to wild type (n=12 for both) had lower levels of triglycerides and similar levels of glycogen and primary circulating sugars. D) Schematic of the radioactive food consumption assay. E) In the radioactive food assay, Per mutants ate ~14% more than wild type (Per, n=9; WT, n=9, p=0.016). F) Schematic of the CApillary FEeder (CAFE) assay. G) Using the CAFE assay, Per mutants ate ~23% more than wild type (Per, n=24; WT, n=21; p=0.0344). n.s.=p>0.05; p-values were obtained by unpaired t-test.  

AP

erc

en

t su

rviv

al

Hours post-starvation

p < 0.0001

0

25

50

75

100

20 30 40

B

Me

tab

olic s

tore

s (

au

)

TG GLY TH +

GLU

Uninfected

0.00

0.25

0.50

0.75

1.00

n.s.*** ***

Figure 2: Allen et al., 2014

Starvation assay C

Me

tab

olic s

tore

s (

au

)

Infected with B. cepacia

***n.s.n.s.

0.00

0.25

0.50

0.75

1.00

TG GLY TH +

GLU

D

FCAFE assay

32P-labeled

food

Radioactive food assay

Measure 32P with

scintillation counter

32P

Day 1

Day 2

Measure

change

in liquid

E

0.0

0.5

1.0

Fe

ed

ing

l/fly/d

ay)

*

WT Per

Fe

ed

ing

l/fly/d

ay)

G

0.0

0.2

0.4

0.6 *

WT Per

WT

Per GLY = Glycogen

TG = Trigylcerides

p < 0.001***

n.s. ≥ 0.05p TH = Trehalose

GLU = Glucoseagar onlyPer starved

WT starved

Page 95: Time to eat

81

mutants have less energy storage than wild type. To directly test this, we measured three

major types of energy storage: fat (triglycerides), glycogen, and primary circulating

sugars (trehalose and glucose). Consistent with their increased sensitivity to starvation,

uninfected Per mutants exhibited significantly lower levels of triglycerides (p=0.0004)

and glycogen (p=0.0007) and similar levels of trehalose and glucose (p=0.7065) when

compared to wild type (Fig. 2B).  

An alternative metabolic explanation for the increased tolerance in Per mutants is

that these mutants may lose energy less quickly than wild type and have greater energy

stores late in infection. Both Per mutants and wild type lost energy stores during

infection, but Per mutants still had lower energy stores than wild type (Fig. 2C). At 16

hours post-infection, triglyceride levels in Per mutants were reduced relative to wild type

(~70% of wild type, p=0.0001), with levels of circulating sugars and glycogen similar to

wild type (p=0.9314 and 0.484, respectively). These data suggest that the increased

tolerance of Per mutants are not due to increased energy stores or decreased energy usage

during infection.

Per mutants exhibit increased feeding behavior.

We reasoned that Per mutants, because they have low energy stores, might eat more than

wild type animals and that this increased feeding behavior itself could enhance infection

tolerance. To directly investigate the feeding behavior of Per mutants, we compared the

nutrient intake of Per mutants and wild type flies by measuring consumption of 32P-

labeled, standard molasses food over 24 hours (Fig. 2D, 2E) [192,193]. In addition, we

measured the consumption of liquid food containing sugar and yeast extract using the

Page 96: Time to eat

82

Capillary Feeder (CAFE) assay (Fig. 2F, 2G) [191,193]. In the 32P-labeled food assay,

Per mutants ate 14% more than wild type controls; in the CAFE assay, Per mutants ate

23% more than wild type controls (Fig. 2F, p = 0.016; Fig. 2G, p = 0.034). Thus, despite

having decreased energy storage before and during infection, Per mutants exhibit

significantly greater food intake than wild type.

Nutrient availability enhances infection tolerance of B. cepacia.

If the increased survival of Per mutants is due to increased feeding, then decreasing

nutrient intake by dietary restriction should abolish the enhanced survival time of Per

mutants in response to B. cepacia infection. To induce dietary restriction, we used a low

sugar, protein-free diet containing only water, agar, and 1% glucose. We found that even

short-term dietary restriction (~24 hours before infection) decreased survival time for

both wild type (Fig. 3A, p<0.0001) and Per mutants (Fig. 3B, p<0.0001). Per mutants

consistently survived an average of 22% longer than wild type flies when fed standard

food (20/20 experiments, comparison of median survival times). In contrast, diet-

restricted Per mutants either had no survival advantage over wild type (4/12

experiments), survived significantly less well than wild type (2/12 experiments), or

survived an average of only 7% longer than wild type (6/12) (Fig. 3C, 3G). Bacterial

loads remained unchanged under these feeding conditions (Fig. 3D, 3E, 3F; p>0.05 for all

time points). Together, these results show that short-term dietary restriction reduces

infection tolerance and support the hypothesis that the increased feeding behavior of Per

mutants on the day of infection contributes to their increased tolerance of infection with

B. cepacia.

Page 97: Time to eat

83

Fig. 3. Dietary restriction does not increase infection tolerance of either Per mutants or wild type A-C) Dietary restriction decreased survival time after infection for both A) wild type (standard food n=66, DR n=63) and B) Per mutants (standard food n=59, DR n=62). C) Dietary restriction eliminated the consistent survival advantage of Per mutants over wild-type flies (Per n=62, WT n=63). Dietary restriction did not alter bacterial load for D) wild type (n≥5 flies/time point) or E) Per mutants (n≥5 flies/timepoint, n>0.05, all time points); moreover, F) diet-restricted wild type and Per mutants had similar bacterial loads (n≥5 flies/time point). p-values were obtained by unpaired t-test (0h) and non-parametric Mann-Whitney test (other time points). G) Multiple experimental trials showed that, on standard food, Per mutants always survived B. cepacia infection longer than wild type but not after dietary restriction; survival outcome was judged significant when p<0.05. p-values for survival curve comparisons were obtained by log-rank analysis; all others were obtained by unpaired t-test unless otherwise noted. n.s.=p>0.05; fed=standard diet, DR=diet-restricted (1% glucose).

Figure 3: Allen et al., 2014

0 6 12

A

D

Bac

teria

l loa

d/fly

Per

cent

sur

viva

l

Hours post-infection

WT: Fed vs. DRp < 0.0001

0

25

50

75

100

15 20 25 30

Hours post-infection

2

4

6

10

10

10

108

0 126

n.s. n.s.n.s. E

Hours post-infection

Per

cent

sur

viva

l

B Per mutant: Fed vs. DR

15 20 25 300

25

50

75

100 p < 0.0001

102

104

106

Hours post-infection0 6 12

Bac

teria

l loa

d/fly

2

4

6n.s. n.s.n.s.

Per

cent

sur

viva

lB

acte

rial l

oad/

fly

F

Hours post-infection

WT vs. Per: DRn.s.

C

15 20 25 300

25

50

75

100

10

102

104

6

Hours post-infection

G

n.s. n.s.n.s.

Standard diet:WT fed Per fed

Diet-restricted:WT DRPer DR

Standard diet(n = 20)

Outcomes of B. cepacia infection: WT vs. Per (number of trials)

DR(n = 12)

Per and WT same (p>0.05)WT survives longer (p<0.05)

Per survives longer (p<0.05)

Per vs. WT survival

Page 98: Time to eat

84

Dietary glucose and amino acids enhance infection tolerance

To evaluate the specific dietary components that contribute to tolerance of infection, we

supplemented a restricted diet with defined nutrients (Fig. 4A). Because Per mutants

display pleiotropic defects in metabolism and other circadian-regulated physiologies

[177], we focused on wild type flies. We first tested if increasing glucose content in the

food could fully complement the restricted diet, which contains 1% glucose. We

compared the effects of standard food with protein-free, agar-based food containing 1%,

5%, 10%, or 15% glucose (Fig. 4B). Wild-type flies survived longest on standard food

and shortest when switched to 1% glucose 24 hours before infection (Fig. 4B, p<0.0001

comparing standard food or 1% glucose with any other condition). Increasing glucose

concentration from 1% to 5% increased survival time (Fig. 4B, p<0.0001) without

changing bacterial loads (Fig. 4D, p>0.05 for all time points), but glucose concentrations

higher than 5% did not further increase survival time (4B, p>0.05 for any pair-wise

comparison of 5%, 10%, and 15% glucose). Moreover, none of these glucose-only diets

was sufficient to increase infection survival time to that observed on standard food

(p<0.0001 comparing any glucose-only diet with standard food). Thus, glucose enhances

infection tolerance but glucose alone is not sufficient for optimal survival of infection.

This result suggests that other components in standard food also contribute to survival of

infection.

In addition to sugar, standard food also contains a complex mixture of lipids,

proteins, vitamins, and other nutrients derived from yeast and cornmeal ingredients. We

tested whether 5% glucose supplemented with amino acids was sufficient to substitute for

standard food (Fig. 4A). We compared survival of flies fed three different diets: 5%

Page 99: Time to eat

85

glucose, 5% glucose plus amino acids, and standard food. A diet of 5% glucose plus

 Fig. 4: Dietary glucose and amino acids increase tolerance of B. cepacia infection A) Schematic of dietary conditions: wild-type flies were raised on standard food and switched 24 hours before B. cepacia infection to fresh standard food, glucose diets (B,D) or glucose diet plus amino acids (C,E). B) Increasing glucose concentration (5%, 10%, 15%) increased survival time relative to 1% glucose diet (p<0.0001 in all cases) and caused similar survival kinetics compared

Page 100: Time to eat

86

to each other (p>0.05 in all cases). C) Supplementing 5% glucose with amino acids increased survival time significantly longer than the glucose-only diets (p<0.0001 in all cases) and was sufficient for survival kinetics similar to standard food (p>0.05). There was no difference in bacterial load comparing flies fed D) 1% vs. 5% glucose (n=6 flies/time point) or E) 5% glucose vs. 5% glucose plus amino acids (n=6 flies/time point); p-values were obtained by unpaired t-test (0h) and non-parametric Mann-Whitney test (other time points). F) Wild-type flies survived longer when injected 1.5 hours before infection with 50 nL of 5% glucose (n=21) than with PBS control (n=18, p=0.0007). G) Multiple trials showed that 5% glucose injection enhanced survival relative to buffer injection when administered within 2 hours before or after the time of infection. Survival outcome was judged significant when p<0.05. p-values for survival curve comparisons were obtained by log-rank analysis; all others were obtained by unpaired t-test unless otherwise noted. n.s.=p>0.05.    amino acids 24 hours prior to infection resulted in significantly longer survival time

compared to 5% glucose alone (Fig. 4C, p<0.0001), with no change in bacterial load (Fig.

4E, all time points p>0.05). In fact, 5% glucose plus amino acids was sufficient to

increase survival time as long as standard food (Fig. 4C, p=0.3039). Taken together, these

results suggest that glucose and dietary amino acids alone are sufficient to promote

tolerance of infection, and that acute exposure to both within ~24 hours of B. cepacia

infection is necessary for optimal survival.

Glucose is required around the time of infection for increased host tolerance.

We set out to precisely characterize the required timing of the glucose contribution to

infection tolerance. To identify the time and place where glucose acts to increase

tolerance, we developed a method to mimic acute glucose intake. We found that a 50 nL

injection of 5% glucose administered directly into the circulatory system of diet-restricted

flies significantly increased infection survival time relative to injection of buffer alone

(Fig. 4F, p <0.0001). This amount of glucose is approximately equivalent to the quantity

ingested by a single fly in a 1-hour period (as calculated from prior feeding experiments;

Fig. 2D-G).

Page 101: Time to eat

87

We found that glucose injection was only effective if administered within a

specific time window centered around the time of infection with B. cepacia. Glucose

injections within 2 hours before or after the time of infection led to a significant increase

in survival (Fig. 4G). In contrast, injections performed 2-4 hours before or after infection

had no effect or a negative effect on survival (Fig. 4G). This time window is

unexpectedly narrow, consistent with an acute rather than chronic effect of diet upon

infection tolerance. These results suggest that acute glucose intake acts to stimulate a

specific signaling pathway that increases immune tolerance when administered at or

around the time of infection.

FOXO activity mediates a glucose-activated increase in infection tolerance.

We set out to identify the signaling pathways underlying the nutrient contributions to

tolerance. Because insulin-like signaling (ILS) is important for glucose homeostasis

[201], we hypothesized that ILS mediates the glucose contribution to infection tolerance.

Phosphorylation of dAkt, a downstream component of the ILS pathway, is indicative of

glucose-activated insulin-like signaling (Fig. 5D) [202]. Consistent with their increased

feeding, we found that Per mutants exhibit increased levels of phospho-dAkt relative to

wild type flies by Western blot analysis (Fig. 5C). Thus activation of insulin-like

signaling is correlated with increased tolerance of B. cepacia.

To directly test whether insulin signaling mediates glucose-stimulated tolerance,

we infected mutant flies lacking dFOXO, a critical downstream component of the ILS

pathway (Fig. 5D). Because dFOXO activation results in negative feedback to inhibit the

ILS pathway, loss of dFOXO function causes constitutive activation of insulin-like

Page 102: Time to eat

88

signaling. If dietary glucose alters survival through activation of ILS, then dFOXO

 Fig. 5: Insulin-like signaling mediates the glucose contribution to infection tolerance and the TORC2 complex mediates the effects of dietary amino acids. A) dFOXO mutants showed no significant difference in survival kinetics when fed 1% or 5% glucose 24 hours before infection (1%, n=20; 5%, n=23). B) dFOXO mutants survived

Page 103: Time to eat

89

significantly longer when fed 5% glucose supplemented with amino acids (a.a.) compared to 5% glucose alone (glucose alone, n=23; with a.a., n=11). C) Per mutants exhibit 1.7x higher levels of phospho-Akt than wild-type flies as determined by Western blot analysis. D) Schematic of the insulin-like signaling pathway. E) Flies co-injected with buffer at the time of infection survived longer when fed 5% glucose plus amino acids (n=52) than 5% glucose alone (n=70). F) Inhibition of TORC1 by co-injection of rapamycin at the time of infection failed to abolish the survival benefit conferred by dietary amino acids (glucose alone, n=59; with a.a., n=67). G) Gal80-ts; tub-Gal4>UAS-Tsc1/Tsc2 flies raised at 18ºC (not over-expressing Tsc1/2, see Fig. S3) exhibited increased survival time in response to dietary amino acids (glucose alone, n=71; with a.a, n=65). H) Gal80-ts; tub-Gal4>UAS-Tsc1/Tsc2 flies raised at 29ºC (over-expressing Tsc1/2, see Fig. S3) exhibited increased survival time in response to dietary amino acids (glucose alone, n=63; with a.a., n=59). I) RicTOR01 control flies exhibited increased survival time in response to dietary amino acids (both, n=71). J) RicTOR01 mutants did not exhibit increased survival time in response to dietary amino acids (glucose alone, n=50; with a.a., n=51). K) Schematic of the TOR pathway. n.s.=p>0.05.; a.a.=amino acids.  

mutants will exhibit no difference in survival whether fed 1% or 5% glucose. Consistent

with this, dFOXO mutants fed 1% glucose or 5% glucose showed no difference in

survival after B. cepacia infection (Fig. 5A, p=0.2974).

In contrast, defective insulin-like signaling had no effect on increased survival

induced by dietary amino acids. Infected dFOXO mutants given dietary amino acids

survived longer when infected with B. cepacia than those deprived of amino acids (Fig.

5B, p<0.001), similar to wild-type flies. Together, these results indicate that the insulin-

like signaling pathway mediates the effects of dietary glucose, but not dietary amino

acids, in increasing tolerance of B. cepacia infection.

TORC2 mediates the amino acid contribution to infection tolerance.

We next asked which molecular pathways might mediate the effects of amino acids in

increasing tolerance of B. cepacia infection. The canonical sensor of amino acid

availability is the TOR signaling pathway, specifically the TORC1 pathway [203]. The

TOR protein forms two distinct complexes: TOR Complex 1 (TORC1) is inhibited by the

Page 104: Time to eat

90

small molecule rapamycin and monitors amino acid availability; TOR Complex 2

(TORC2) is rapamycin-insensitive and is better characterized as a regulator of the actin

cytoskeleton (Fig. 5K) [204,205].

To determine whether TORC1 signaling plays a role in amino-acid stimulated

tolerance, we tested whether tolerance is inhibited by rapamycin [204,205]. We found

that injecting 9.6 ng of rapamycin per fly (approximately equivalent to the mammalian

dose of 16 mg/kg, [206]) significantly decreased survival of infection (Fig. S2, rapamycin

vs buffer, p<0.0001 on food with or without amino acids). This effect is likely due to

TORC1-dependent resistance mechanisms, as seen during P. aeruginosa infection

(personal communication, S. Pletcher). We found that, when injected with either

rapamycin or vehicle, flies survived infection significantly longer when fed dietary amino

acids (Fig. 5E, 5F, p<0.0001 in both cases). This result suggests that the rapamycin-

sensitive, canonical amino acid sensor TORC1 is not involved in amino acid-stimulated

tolerance of B. cepacia infection.

To confirm that TORC1 signaling is not the key mediator of amino acid-

stimulated infection tolerance, we compared the effects of perturbing another molecular

component of this pathway, Tsc1/2. Tsc1/2 activity inhibits Rheb, thus negatively

regulating TORC1 signaling (Fig. 5K) [207]. We found that mutants overexpressing

Tsc1/2 (Fig. S3A,B) still responded to dietary amino acids with increased survival similar

to controls following B. cepacia infection (Fig. 5G,H p<0.0001). These results further

confirm that the effect of amino acids on survival of infections is independent of TORC1

signaling.

We next tested whether dietary amino acids enhance infection tolerance through

Page 105: Time to eat

91

TORC2 signaling. In TORC2, TOR complexes with RicTOR (the Rapamycin-insensitive

companion of TOR) and Sin1 (Fig. 5K) [208]. We compared mutants carrying a

transposon-mediated deletion in the RicTOR gene with control flies in which this

transposon has been precisely excised from the RicTOR gene. While control flies

responded to dietary amino acids with increased survival of infection (Fig. 5I, p<0.0001),

RicTOR mutants exhibited the same survival time with or without dietary amino acids

(Fig. 5J, p>0.05). These results therefore suggest that TORC2, but not TORC1, is

required for amino acid-mediated tolerance of infection.

Amino acid contribution to survival depends on the gut microbiota.

To characterize the mechanism underlying amino acid-stimulated tolerance, we asked

whether direct injection of amino acids can increase survival time as well as ingestion of

amino acids, similar to glucose (Fig. 6A, schematic). We found that injection of amino

acids at two different concentrations at different times points before or during infection

did not improve survival time (Fig. 6B, p=0.0031 with injection of a high dose of amino

acids decreasing survival time; Fig. 6C, low amino acid dose, p>0.05; and Fig. S4A and

S4B, showing amino acid injections 2h before and after infection, respectively, p>0.05

for both). Flies injected with buffer were still able to respond to ingested amino acids

(Fig. 6C, p<0.0001). One possibility is that amino acids need to be administered in a

more sustained fashion than glucose in order to enhance tolerance. Another possibility is

that amino acids need to be ingested by the animal through the gut. If the latter is true,

we hypothesized that the gut microbiota may be required for amino acid-stimulated

infection tolerance.

Page 106: Time to eat

92

To test this, we modified a previously published method [195] to raise germ-free

Fig. 6. The effect of amino acids on survival of infection is facilitated by gut microbiota. A) Schematic of experimental set-up for B and C. Conventionally raised flies are either injected with amino acids or have amino acids included in their diet. Injection of amino acids in a high dose (B, n=69) or low dose (C, n=43) does not improve survival when compared to buffer alone (B, n=64, p=0.0031; C, n=47, p>0.05). As a control, buffer-injected flies are still able to respond to dietary amino acids with enhanced survival time (C, n=25). D) Experimental set-up for E and F. Conventionally-raised wild-type flies’ and germ-free wild-type flies’ survival of infection are compared on diets with and without amino acids. E) Germ-free flies survive infection similarly to

Page 107: Time to eat

93

conventionally-raised flies with normal microbiota when amino acids are not present in the diet (germ-free, n=38; conventional, n=54, p>0.05). F) When exposed to dietary amino acids, flies with intact microbiota exhibit enhanced survival significantly greater than germ-free flies (germ-free, n=39; conventional, n=61). p-values obtained by log-rank analysis.      flies lacking their microbiota. We then compared the effects of dietary amino acids on

survival to the effects on conventional flies with intact microbiota (Fig. 6D, schematic).

We confirmed germ-free status by testing for the presence of bacterial 16S ribosomal

genes using PCR analysis (Fig. S4C). On the protein-free 5% glucose diet, germ-free and

conventional flies had identical survival times after infection with B. cepacia, showing

that germ-free flies are not generally immunocompromised against this infection (Fig.

6E, p<0.05). However, dietary amino acids enhanced the survival of conventional flies

significantly more than germ-free flies (Fig. 6F, p<0.0001). Dietary amino acids still

increased the survival of germ-free flies relative to glucose-only diet (Fig. S4D),

suggesting that the microbiota are not exclusively required for amino acid-mediated

survival. Taken together, these results suggest that amino acids are optimally

administered by feeding to increase infection tolerance; the gut microbiota may help to

mediate absorption of these amino acids or may increase survival of infection by a

different mechanism.  

Page 108: Time to eat

94

DISCUSSION

By examining a circadian mutant with increased infection tolerance against B. cepacia,

we first identified increased feeding as a behavior that contributes to increased tolerance.

We found that two specific nutrients, glucose and amino acids, fully substitute for

standard food in promoting optimal tolerance of rapid infection with B. cepacia. We

found that the glucose contribution to survival is mediated by insulin-like signaling;

consistent with this, two previous studies have implicated ILS in Drosophila tolerance of

M. marinum and resistance against viral infections [86,209] and two other studies have

found that induction of immunity gene expression leads to decreased insulin signaling in

wild-type flies [199,200]. Our work strengthens the hypothesis that insulin-like signaling

may be a general feature of Drosophila immunity, irrespective of the specific type of

infection. Our data further suggest a time and location for this insulin activation, in that

an acute increase in circulating glucose around the time of infection can increase overall

survival time. Thus, what and how much a fly ingests near the time of infection has a

significant effect on its survival of that infection.

Our results also show that dietary amino acids can increase a host’s tolerance.

This effect does not appear to be mediated by the canonical amino acid sensing TORC1

pathway but through the less well-understood TORC2 pathway. From our rapamycin

experiments, TORC1 appears to play a role in immunity against B. cepacia independent

of amino acid-stimulated tolerance. We hypothesize that this is a TORC1-mediated

resistance mechanism, similar to an effect recently observed with the related bacterial

pathogen, P. aeruginosa (S. Pletcher, personal communication). In vertebrates,

rapamycin is a well-charactarized immunosuppressant due to its inhibition of mTORC1.

Page 109: Time to eat

95

A role for TORC2 in host tolerance has not been previously described. TORC2 is

thought to be stimulated by growth factors and PI3K [205]. Interestingly, in both yeast

and mammalian cells, TORC2 is activated by association with the ribosome but not by

translation [210,211]; and evidence exists that TORC2 may be stimulated by amino acids

[212]. In other systems, downstream targets of TORC2 include components of the

cytoskeletal system, as well as Akt and SGK1, and are thought to play a role in tissue-

specific morphology [204,213]. In our system, dietary amino acids appear to activate

TORC2 and enhance host tolerance, an intriguingly novel finding. While in most cases,

immune effects of TOR are thought to act through TORC1, recent evidence suggests that,

in mouse embryonic fibroblasts, RicTOR inhibits TLR-stimulated cytokine expression

[214]. Thus RicTOR may contribute to host tolerance by suppressing harmful effects of

the host immune response. The direct targets of TORC2 that are relevant for infection

tolerance remain unknown and their identification will be an important goal of future

studies.

Our results also show that in Drosophila, amino-acid mediated tolerance is

facilitated by the gut microbiota. Recent work has shown that the gut microbiota can

affect the pathogenesis of diabetes [215], obesity, and atherosclerosis [216]. Here we

show that the fly's microbiota also play a role in tolerance of a systemic infection. This

role could include a contribution to the fly's development so that it can respond properly

to amino acids [217-219] or the gut microbiota may be required for absorption of specific

amino acids by the gut epithelia, as is postulated for branched-chain amino acids [219].

Alternatively, the fly’s microbiome itself may be altered or stimulated by amino acids to

produce molecules or interact with the gut epithelium in a way that enhances infection

Page 110: Time to eat

96

tolerance. The Drosophila system described here will be useful in distinguishing

between these models.

The cellular and molecular mechanisms that promote host tolerance of infection

are not well-understood [82]. B. cepacia is a significant opportunistic bacterial pathogen,

particularly in hospital settings. This hospital-acquired infection is associated with high

rates of mortality, up to 50% for severe strains, and is often antibiotic-resistant [220,221].

Understanding the mechanisms that are stimulated by acute glucose and dietary amino

acids and lead to increased tolerance will help to identify targets for pharmacological

treatments. The Drosophila model of infection may therefore prove useful in further

dissection of acute, glucose and amino acid-stimulated host tolerance of this human

pathogen.

Page 111: Time to eat

97

Fig. S1. Antimicrobial peptides (AMPs) are induced by B. cepacia infection to similar expression levels in Per mutants and wild type. There is no difference in AMP expression between Per mutants and wild type (WT) observed before or after B. cepacia infection. Shown here are expression levels measured by qPCR for five antimicrobial peptides (AMPs): Attacin (A), Cecropin (B), Defensin (C), Drosocin (D), and Metchnikowin (E). p-values obtained by unpaired t-test; n.s.=p>0.05.

   

0

400

800

0

100

200

300

0

40

80

A BAttacin

Hours post-infection0 6 12

n.s. n.s.n.s.800

400

0

E

Hours post-infection0 6 12

n.s. n.s.n.s.Metchnikowan

Cecropin

Hours post-infection0 6 12

n.s.n.s.n.s.

D Drosocin

Hours post-infection0 6 12

n.s. n.s.n.s.

1000

500

0

C

Hours post-infection0 6 12

n.s. n.s.n.s.Defensin

WTPer

Figure S1: Allen et al., 2014 (Related to Fig. 1)

Rel

ativ

e ex

pres

sion

Rel

ativ

e ex

pres

sion

Page 112: Time to eat

98

Fig. S2. Rapamycin reduces survival of B. cepacia infection. Flies injected with 9.6 ng rapamycin at the time of infection (n=59) died more quickly following B. cepacia infection than flies injected with control buffer (n=72, p<0.0001).

 

Page 113: Time to eat

99

Fig. S3. Expression of Tsc1 and Tsc2 transcripts are induced by heat- shock. After heat-shock treatment at 29°C for 0, 24, 48, or 72 hours (left to right), adult Gal80-ts; tub-Gal4>UAS-Tsc1/Tsc2 flies showed increased expression of A) Tsc1 and B) Tsc2. Transcript levels for each sample were normalized to total RNA level by comparison to ribosomal protein 1 (rpl4). C) Schematic of the TOR pathway.

Page 114: Time to eat

100

Fig. S4. Amino acid-stimulated tolerance is facilitated by the gut microbiota. Amino acids must be ingested (enter through the gut) to stimulate host tolerance. Amino acids injected 2 hours prior to infection (A) or post- infection (B) did not improve survival outcome (p>0.05 for both). C) Bacterial 16S ribosomal DNA is not detected after PCR in germ-free flies, but is present in conventionally-raised flies. D) Survival curves for all 4 conditions for germ-free fly experiments: conventionally-raised flies fed glucose, conventionally-raised flies fed glucose and amino acids, germ-free flies fed glucose, and germ-free flies fed glucose and amino acids. Dietary amino acids improve survival of germ-free flies (p<0.0001) to a lesser extent than they improve survival of conventionally- raised flies (also p<0.0001).

Page 115: Time to eat

101

Table S1. Primer sequences.

Gene Left primer Right primer

AttA CACAATGTGGTGGGTCAGG GGCACCATGACCAGCATT

CecA1 TCTTCGTTTTCGTCGCTCTC CTTGTTGAGCGATTCCCAGT

Def TTCTCGTGGCTATCGCTTTT GGAGAGTAGGTCGCATGTGG

Dipt ACCGCAGTACCCACTCAATC CCCAAGTGCTGTCCATATCC

Dro CCATCGAGGATCACCTGACT CTTTAGGCGGGCAGAATG

Drs GTACTTGTTCGCCCTCTTCG

CTTGCACACACGACGACAG

Mtk TCTTGGAGCGATTTTTCTGG TCTGCCAGCACTGATGTAGC

Rpl1 TCCACCTTGAAGAAGGGCTA TCCACCTTGAAGAAGGGCTA

Tsc1 GTAAACACACCTTGTCCAAGCAGC TGACAGATGGATAGACGGAACCAC

Tsc2 ACACATAGACAACAACGAGAGG

AAGAGATATCATGGCAGGATGC

   

Page 116: Time to eat

102

AUTHOR CONTRIBUTIONS

Experiments were conceived by MSH with input from VWA and RMO. VWA, RMO,

and CGZ performed all survival and bacterial load assays after B. cepacia infection;

VWA and RMO analyzed these experiments. In addition, VWA performed and analyzed

the AMP induction experiments and starvation assay; RMO performed and analyzed the

metabolic storage assays. EFS performed and analyzed the bead inhibition experiments

and Western blot analysis. VMH performed and analyzed the melanization experiments.

RBP performed the axenic (germ-free) fly experiments. KRM and WWJ performed the

radioactive and CAFE feeding experiments. MSH, JCC, VWA, and RMO wrote the

manuscript and constructed the figures.

Page 117: Time to eat

103

ACKNOWLEDGMENTS

We thank all members of the Shirasu-Hiza lab for support and advice; in particular, we

thank Emily Marcinkevicius for insightful comments on experiments and this manuscript.

We also thank Shawn Jordan and Tim Davies for critical reading of this manuscript. We

are grateful to Joel Shirasu-Hiza for assistance in writing our custom analysis software,

Count the Dead; Marc Dionne and David Schneider for support and productive

discussions; and Adam Ryan, Albert Kim, and Paul Kim for technical support. We thank

Gerard Karsenty for use of his qRT-PCR machine, microplate reader, and film developer;

and Rodney Rothstein for use of his Nanodrop. This work was supported by: NIH

1F31NS080673 (EFS); NIH DP2 OD008773 (JCC); NIH R21 DK092735 (WWJ); a

Hirschl Career Award from the Irma T. Hirschl Foundation (MSH); and NIH R01

GM105775 (MSH).

Page 118: Time to eat

104

CHAPTER 4:

Phagocytic immune cell defects in the Drosophila model of Fragile X syndrome

Work in progress:

Elizabeth F. Stone, Reed M. O’Connor, Emily Marcinkevicius, Vanessa M. Hill, Jennifer S. Ziegenfuss, Clarice G. Zhou, Wesley B. Grueber, and Mimi M. Shirasu-Hiza

Page 119: Time to eat

105

SUMMARY

Having found significant differences in immune system function in the absence of

circadian regulation, I next wanted to know how the immune system functions in the

Drosophila model of Fragile X syndrome, a disease that is known to be associated with

circadian dysregulation. Fragile X syndrome is the most common single locus genetic

cause of intellectual disability and autism and results from the loss of the RNA-binding

protein FMRP that regulates the translation of a whole host of mRNA targets. In addition

to circadian dysregulation, patients and animal models of the disease exhibit defects in

learning and memory, altered synaptic plasticity, and increased dendritic arborization.

Altered immune system parameters observed in patients with Fragile X syndrome and

autism suggests an altered immune system function in these diseases, but immunity has

not been well characterized, either in humans or in animal models. Here I ask, how does

the loss of dFMR1, the Drosophila homolog of FMRP, affect immune system function

both in the body and in the brain?

Dr. Mimi Shirasu-Hiza and I conceived of all of the experiments enclosed in this

chapter and I have written the manuscript with input from Dr. Shirasu-Hiza. I initially

conducted all the systemic immunity experiments described in this chapter using dFMR1

transheterozygous hypomorphs, made by crossing flies with the dFMR1B55 allele to flies

with the null dFMR1 Δ50M allele and compared them to their isogenic wild-type control

(see Appendix II). At the time I was unable to generate dFMR1 null animals in lab. I

have since optimized husbandry conditions to obtain dFMR1 null animals by raising the

dFMR1 mutants and wild-type controls on food containing low amounts of glutamate

[222]. To avoid recessive background mutations, I used a transheterozygous null for all

Page 120: Time to eat

106

of the experiments listed in this chapter. The two dFMR1 mutant alleles used for these

experiments were dFMR1Δ50M and dFMR13 and their isogenic wild-type controls.

Others in the lab have also contributed to this work by optimizing experiments,

repeating experiments, and assisting in data analysis. I conducted all of the systemic

immunity experiments in this chapter with dFMR1 hypomorphs (see Appendix II) as well

as null mutants (Fig 1A-G). Vanessa Hill conducted the L. monocytogenes infection and

the melanization experiment on the dFMR1 null animals (Fig 1C, 1G, 2A-B). Reed

O’Connor and I performed the AMP experiment (Fig 2E-F). I performed all of the initial

hemocyte phagocytosis experiments, with subsequent trials performed in collaboration

with Dr. Emily Marcinkevicius (Fig 2H-I). I performed the genetic crosses for necessary

fly lines and initiated the glial phagocytosis assays in Figs 3, 4, and 5, with some of the

subsequent trials performed in collaboration with Reed O’Connor, who was invaluable in

optimizing the immunostaining and confocal microscopy conditions.

Page 121: Time to eat

107

ABSTRACT

Autism spectrum disorder has become a major national health crisis. Most research on

autism spectrum disorder has focused on neuronal defects; however, recent evidence

suggests that immune system dysfunction correlates with this neurological disease. The

role of glia, or non-neuronal immune cells in the brain, remains unclear. Here we

examine a Drosophila model of Fragile X syndrome, the most common genetic cause of

autism, for immune cell defects. Fragile X syndrome is caused by loss of FMR1 function

and flies lacking the Drosophila homolog dFMR1 display neurological phenotypes

similar to patients with Fragile X syndrome, including dendritic over-arborization,

defects in learning and memory, and loss of circadian regulation. We found that dFMR1

mutants have defective immune responses to bacterial pathogens, and that these defects

are the result of aberrant phagocytic activity in the flies’ hemocytes. Additionally, we

show that these mutants exhibit defective phagocytosis by ensheathing glia in the brain in

response to neuronal injury: glia in dFMR1 mutants show significant delays in clearance

of axonal debris and in their expression of the critical phagocytic receptor Draper. These

results represent the first in vivo study demonstrating a defect in a specific glial function

in a model of Fragile X syndrome. Thus our work establishes a new model of

neuroimmune dysregulation in an autism spectrum disorder model, demonstrating a

possible role for glia in pathogenesis, with potential future implications for therapeutic

approaches to the disease.

Page 122: Time to eat

108

INTRODUCTION

The prevalence of autism spectrum disorders has reached a modern crisis, occurring at a

rate of 1 in 68 children [223]. The vast majority of autism cases occur sporadically, but

the most common known monogenic cause of intellectual disability and autism is Fragile

X syndrome. Thus Fragile X syndrome is a highly utilized model for studying autism and

intellectual disability. In Fragile X syndrome, transcriptional silencing of the FMR1 gene

leads to absence of the protein FMRP, an mRNA binding protein and translational

regulator. FMRP is ubiquitously expressed but is particularly abundant in neurons [91].

Loss of FMR1 in humans and animal models causes neurons to display excessive growth

of dendritic spines and defects in synaptic plasticity, but the mechanism underlying the

neuronal dysfunction is unknown [132].

One factor that may contribute to the neurological symptoms of autism is immune

system dysfunction. Autism is strikingly correlated with both maternal infection and

autoimmune diseases during pregnancy; this suggests that prenatal immune challenge in

the mother may cause immune misregulation in the fetus, contributing to the

development of autism. In support of this hypothesis, numerous studies have

demonstrated that children with autism express auto-antibodies with reactivity against

brain-specific antigens [224], and others have suggested an association between specific

human leukocyte antigen alleles and autism [225]. Fragile X syndrome is also associated

with altered immune system function, including elevated proinflammatory cytokine

levels in the blood [226] and gastrointestinal disorders [95]. However, whether altered

immune system function in the body is a cause or consequence of neuronal dysfunction

remains unclear.

Page 123: Time to eat

109

Recent studies have also suggested a role for glia, which are the resident immune

cells in the brain, in the etiology of Fragile X syndrome and autism. FMR1 is expressed

in mouse glia during embryonic development as well as postnatally [227]. FMR1 mouse

mutants exhibit significant and often subtle defects in different subtypes of glia, including

a decreased number of radial glial cells during neocortical neurogenesis [228]; atypical

activated reactive astrocytes in specific brain regions [229]; decreased glutamate

transporter expression in astroglial cells [230]; and delayed myelination due to impaired

oligodendrocyte precursor cells [231]. In an in vitro co-culture assay, FMRP knockout

neurons grown with wild-type glia developed a wild-type architecture, while wild-type

neurons grown with FMRP knockout glia developed a morphology characteristic of

FMRP knockout neurons. These results demonstrate that glia are disordered in the

absence of FMRP and may play an important role in regulating neuronal structure. One

of the main cellular functions of glia is phagocytosis, or the engulfment of extracellular

material, which is mediated by phagocytic receptors and cytoskeletal rearrangement [22-

24]. Still, the role of FMR1 in glial phagocytosis has not yet been examined.

In both vertebrates and invertebrates, it has become increasingly clear that

phagocytosis by glia is important for neuronal structure and function [232-234]. Rather

than being passive structural cells or vacuum cleaners that merely clear the extracellular

spaceof debris, glia are now understood to play an active role in shaping their local

environment. Glia contribute to the restructuring of neuronal circuits by phagocytosing

synaptic elements and newborn cells [235]. Recently, phagocytic properties of microglia

have been implicated in a mouse model of Rett syndrome, which in humans causes

autism spectrum disorder [236]. This suggests that neuronal disease may be caused by a

Page 124: Time to eat

110

defect in glial phagocytosis. Glial phagocytosis is additionally known to be required for

dendritic pruning. Because FMR1 mutants exhibit defects in dendritic pruning leading to

an over-arborization of dendritic branching, glial phagocytosis in these mutants could be

dysregulated or absent. Glial phagocytosis has not yet been examined in any model of

Fragile X syndrome.

The Drosophila dFMR1 mutant is a well-established model for studying specific

aspects of Fragile X syndrome. Drosophila has a single FMR1 family member (dFMR1),

which is most similar to vertebrate FMRP with a high degree of amino acid sequence

similarity [237]. In many key parameters, loss of dFMR1 in Drosophila phenocopies

mammalian models of Fragile X syndrome. Similar to vertebrate models, Drosophila

dFMR1 mutants display neurological symptoms: defects in learning, memory, and

circadian regulation [120], as well as similar neuroanatomical defects and synaptic

dysfunction [238]. dFMR1 is also functionally conserved on the molecular level: dFMR1

regulates the translation of Drosophila homologs of putative targets regulated by

mammalian FMRP [121,122]. However, while much research has been undertaken to

examine the neuronal defects in animal models of Fragile X syndrome, immune system

function in other cell types—particularly in non-neuronal, glial cell types in the brain, of

dFMR1 mutants has not been investigated.

This is the first study to investigate specific immune mechanisms both in the body

and in the brain using the Drosophila model of Fragile X syndrome. Here we identify

several pathogens to which dFMR1 mutants are more susceptible, and we show that

dFMR1 mutants exhibit elevated basal expression of anti-microbial peptides but reduced

phagocytic activity. These results are the first demonstration of a cellular immune system

Page 125: Time to eat

111

defect in the Drosophila model for Fragile X syndrome. After observing a hemocyte

phagocytosis defect in the body, we next investigated glial phagocytosis in dFMR1

mutants using an axonal injury model. We found that dFMR1 mutants exhibit reduced

clearance of axonal remnants after injury. We observed a decrease in the induction of

Draper, a receptor crucial for phagocytosis, in dFMR1 mutants after injury, suggesting

that glia are less activated after injury in dFMR1 mutants. This is the first study to

demonstrate a defect in a specific glial function in vivo in dFMR1 mutants, namely glial

phagocytosis. Thus, as glial phagocytosis is crucial for proper dendritic pruning and

synaptic function, these results suggest that glia may play a critical role in the

pathogenesis of Fragile X syndrome.

Page 126: Time to eat

112

MATERIALS AND METHODS

Fly and bacterial strains

In the following experiments, we generated flies transheterozygous for two dFMR1 nulls

alleles, dFMR1Δ50M (a kind gift from D. Zarnescu [122]) and dFMR13 (a kind gift from T.

Jongens [239]), and their isogenic controls. Both of these alleles were outcrossed to their

isogenic controls (Oregon R and iso31b, respectively) for at least 6 generations. The

control flies were transheterozygous for OreR and iso31b. To label hemocytes with GFP,

we recombined a hemocyte-specific Gal4 driver (hmlΔ-Gal4) with UAS-GFP(2x) and

crossed this to dFMR1 and control lines. To label a specific population of olfactory

neurons with GFP, we crossed the OR85e>mCD8::GFP reporter (a kind gift from M.

Freeman [240]) into dFMR1 and control lines.

Infections were performed with the following bacteria as previously described

[63]: Streptococcus pneumoniae strain SP1, a streptomycin-resistant variant of D39 and

gift from Elizabeth Joyce in Stan Falkow’s laboratory at Stanford University [156];

Listeria monocytogenes strain 10403S, gift from Julie Theriot at Stanford University

[157]; Serratia marcescens strain DB1140, gift from Man Wah Tan at Stanford

University [158]; Escheria coli strain DH5-alpha; and Micrococcus luteus ATCC strain

4698. S. pneumoniae was frozen at OD600 from 0.15-0.4 in 1 ml aliquots with 10%

glycerol, pelleted upon thawing, rediluted in BHI (Brain Heart Infusion media, Difco) to

an OD600 0.12 for injection into flies, which were incubated at 29˚C after injection.

L. monocytogenes was grown in standing BHI, overnight at 37˚C, and diluted to OD600 of

0.1 for injection into flies, which were incubated at 25˚C. S. marcescens was grown in

shaking BHI overnight at 37˚C and diluted to OD600 ranging from 0.1 to 0.6 for injection

Page 127: Time to eat

113

into flies, which were incubated at 29˚C. E. coli was grown in shaking LB, overnight at

37˚C, and diluted to OD600 0.1 for injection into flies, which were incubated at 25˚C. M.

luteus was grown shaking in BHI, overnight at 37˚C, and diluted to OD600 0.1 for

injection into flies, which were incubated at 29˚C.

Fly rearing conditions

Flies were raised at 25°C, 55-65% humidity onyeasted low glutamate food in a 12h

light/dark cycle. The recipe for the low glutamate food was from Chang, et al, 2008

[222]. Briefly, the low glutamate food contains equal parts 6.25% v/v cornmeal, 6.25%

v/v molasses, 25.75 g/L Baker’s yeast, and 9.25 g/L agar. Flies collected for survival and

hemocyte phagocytosis experiments were maintained on standard dextrose food. The

recipe for standard dextrose food is as follows: 38 g/L cornmeal, 20.5 g/L yeast, 85.6 g/L

dextrose, 7.1 g/L agar. All infection experiments were performed with male flies, 5-7

days post-eclosion. Flies collected for glial phagocytosis assays were maintained on low

glutamate food and were kept on low glutamate food after maxillary palp excisions.

Injections

Flies were anesthetized on CO2 pads. Injections were performed using a custom Tritech

microinjector (Tritech) and pulled glass capillary needles. 50 nL of liquid were injected

into each fly, as calibrated by measuring the diameter of the expelled drop under oil.

Survival assays

Page 128: Time to eat

114

As described previously [63], 60-85 flies per genotype per condition were assayed for

each survival curve and placed in 3 vials of standard dextrose food with approximately 20

flies each. In each experiment, 20-40 flies of each line were also injected with media as a

wounding control. Death was recorded hourly or daily depending on the virulence of the

pathogen. Data was converted to Kaplan-Meier format using custom Excel-based

software called Count the Dead. Survival curves are plotted as Kaplan-Meier plots and

statistical significance is tested using log-rank analysis using GraphPad Prism software.

All experiments were performed at least three times and yielded similar results.

Bacterial load quantitation

Following microbial challenge, six individual flies were collected at each time point.

These flies were homogenized, diluted serially and plated on appropriate media (tryptic

soy blood agar for S.pneumoniae, LB agar for all others). Statistical significance was

determined using non-parametric Mann-Whitney tests. All experiments were performed

at least three times and yielded similar results.

Melanization assay

The melanization assay was performed as described previously [63]. Briefly, flies were

injected with L. monocytogenes (OD600 0.1) and incubated at 29˚C. 5 days after infection,

flies were visually inspected for melanin deposits. Each deposit was assigned a point

value based on its estimated size. The point scale ranged form 0.5 to 4, where 4

rpresented the largest deposits. The point values for each fly were totaled to give a

melanization index. Separate melanization indices were calculated for melanization at

Page 129: Time to eat

115

the injection site, which results from wounding, and disseminated melanization, which

results from the systemic response to infection.

Antimicrobial peptide gene expression

Flies were injected with 50 nL of bacterial culture at OD600 0.10 (M. luteus and E. coli).

Following injection, flies were placed in vials containing dextrose food and incubated at

25˚C for 6, 24, and 48 hours. Three groups of six flies after each time point were

homogenized in Trizol and stored at -80˚C until processed. RNA was isolated using a

standard Trizol preparation and the samples treated with DNAse (Invitrogen). Fermentas

RevertAid First Strand cDNA synthesis kit was used to produce the cDNA following the

manufacturer’s protocol. Quantitative RT-PCR was performed using a Bio-Rad CFX

Connect system, with Roche FastStart Universal SYBR Green Master (Rox) mix and the

following primer sets: Drosomycin, Diptericin, Rpl1 [160]. Total mRNA concentration

was normalized using Rpl1 expression. Differences in the infection-induced gene

expression were calculated by normalizing to basal gene expression at the same time

point; p-values were obtained by Mann-Whitney test.

Hemocyte phagocytosis assay

5-7 day old male flies were injected with 50 nL of 20 mg/ml pHrodo-labeled S. aureus in

PBS (Molecular Probes, cat# A10010). The flies were allowed to phagocytose the

particles for 35-45 minutes. The dorsal surface of the thorax was glued onto cover glass

(VWR, cat# 3406)with Loctite brand super glue. Images were taken using epifluorescent

illumination with a Nikon Eclipse E800 microscope fitted with a Photometrics Cool

Page 130: Time to eat

116

Snaps HQ2 camera. Images were captured with Nikon Elements software. The

experiment was repeated three times with 5-6 flies for per genotype. Images were

thresholded to generate ROIs corresponding to fluorescent phagocytic compartments.

The same threshold was used for all images within a given experiment. ROI pixel areas

and intensities were calculated using Nikon Elements software and summed to generate a

single intensity value for each fly, called the sum intensity. Significance was calculated

between genotypes using the student t-test. To elucidate whether dFMR1 mutants had

fewer hemocytes or hemocytes that were less active, we performed a phagocytosis assay

on flies expressing a UAS-GFP under the hemocyte-specific promoter hmlΔ-Gal4,

driving hemocyte-specific GFP expression.

Glial phagocytosis assay and immunohistochemistry

Glial phagocytosis assays were conducted as previously described [241] with some

adjustments. The maxillary palps of 7 to 10-day-old OR85e>mCD8::GFP; dFMR1

mutants and wild-type controls were excised to sever the olfactory neurons. Flies were

collected and decapitated at various time points after wounding, and the fly heads were

fixed for 40 minutes at room temperature in 4% paraformaldehyde in PBS + 0.1% Triton

X-100 (PTX). Fly heads were washed 5 times in PTX and brains were dissected in ice

cold PTX. Brains were blocked in 4% normal donkey serum in PTX and incubated in

primary antibodies at 4˚C for 22-24 hours. Primary antibody was diluted in PTX with 2%

normal donkey serum and 0.02% sodium azide; the primary antibody was reused up to

three times and stored at 4˚C for up to 4 weeks. The following primary antibodies and

dilutions were used: rabbit anti-Draper (a kind gift from M. Freeman, 1:500 [242]);

Page 131: Time to eat

117

chicken anti-GFP (Abcam, #ab13970, 1:1000). Brains were then washed 5 times over the

course of 1 hour at room termperature in PTX and incubated at 4˚C for 22-24 hours.

Secondary antibodies were diluted in PTX with 2% normal donkey serum and 0.02%

sodium azide and made fresh for each experiment. The following secondary antibodies

and dilutions were used in this study: Rhodamine-Red X donkey anti-rabbit IgG (Jackson

Immunoresearch, # 711-295-152, 1:200) and AlexaFluor 488 donkey anti-chicken IgG

(Jackson, # 703-545-155 1:200). Brains were mounted onto coverslips that had been

coated with poly-L-lysine and Kodak Photoflow (10 mL poly-L-lysine and 36 uL

Photoflow), dehydrated by incubating in increasing concentrations of alcohol for 5

minutes (50%, 75%, 95%, 100%, 100%), and incubated in two different solutions of

100% xylenes for 10 minutes each. Coverslips were mounted onto slides with DPX

(Electron Microscopy Service # 13510) and slides dried overnight in the dark at room

temperature. Slides were stored at 4˚C before imaging.

Confocal microscopy was performed using a Zeiss LSM 510 Meta with 488nm

and 533nm lasers and an AxioImager Z1 upright microscope with a PlanNeoFL 40X/1.3

oil objective. Slices with a depth of 8 bit were taken from the anterior to posterior

extremities of the bilateral glomeruli, as indicated by GFP staining. The sum of the mean

fluorescent intensity, using both the GFP and rhodamine red channels, of the middle three

slices of each glomerulus were quantified and normalized to background separately using

ImageJ.

Page 132: Time to eat

118

RESULTS

dFMR1 mutants are sensitive to infections with specific pathogens

The innate immune system of Drosophila is composed of three main immune

mechanisms that help to kill pathogens: melanization, antimicrobial peptide (AMP)

production, and phagocytosis. Drosophila use specific immune mechanisms to defend

against different bacterial pathogens. To assess whether the loss of dFMR1 causes

specific defects in immunity, we first compared the survival kinetics of dFMR1 mutants

and wild-type isogenic controls after infection with four different pathogens known to

elicit immune system activation. We found that dFMR1 mutants died faster than wild-

type controls after infection with S. pneumoniae (Fig 1A, p < 0.0001), S. marcescens (Fig

1B, p < 0.0001), and P. aeruginosa (Fig 1D, p < 0.0001). Survival rates of dFMR1

mutants and wild-type controls did not differ after infection with L. monocytogenes (Fig

1C, p = 0.1702). These results suggest that dFMR1 mutants have defects in specific

immune mechanisms, which cause sensitivity to particular infections.

Decreased survival could occur through defects in resistance, defined as the

ability to kill and clear bacteria, or through defects in tolerance, defined as the ability to

withstand the pathogenic affects of infection. To determine if dFMR1 mutants were less

resistant to infection, we next examined bacterial loads in dFMR1 mutants and wild-type

controls at different time points after infection. We found that dFMR1 mutants have more

bacteria per fly compared to wild-type controls at specific time points after infection with

S. pneumoniae (Fig 1E) and S. marcenscens (Fig 1F), indicating dFMR1 mutants were

less able to kill and clear bacteria. That is, dFMR1 mutants are less resistant to these

pathogens. As expected, dFMR1 mutants and wild-type controls had no consistent

Page 133: Time to eat

119

Fig. 1. dFMR1 mutants are less resistant to infections with specific pathogens. (A-D) Shown here are Kaplan-Meier survival curves of wild-type controls (WT, black circles) and dFMR1 mutants (dFMR1, purple circles) after injection with select pathogens. dFMR1 mutants died more quickly than wild-type controls after infection with (A) S. pneumoniae (p < 0.0001), (B) S. marcescens (p < 0.0001), and (D) P. aeruginosa (p < 0.0001). dFMR1 had the same survival kinetics after infection with (C) L. monocytogenes (p = 0.1702). (E-H) dFMR1 mutants have higher bacterial loads, in other words are less resistant, to infections with (E) S. pneumoniae and (F) S. marcescens. dFMR1 mutants have no consistent difference in bacterial loads after infections with (G) L. monocytogenes and (H) P. aeruginosa. p-values for survival

101

102

103

104

105

6.0 6.5 7.0 7.5 8.00

25

50

75

100

103

105

107

0 5 10 15

A B

0

25

50

75

100

Days post-infection

0

25

50

75

100

p < 0.0001

Days post-infection

p < 0.0001

WT

dFMR1

S. pneumoniae S. marcescens

Perc

ent surv

ival

Perc

ent surv

ival

Figure 1: Stone et al., 2014

n.s. n.s.

0 3 18

**

E F

Bacte

rial lo

ad/fly

S. marcescens

Hours post-infectionHours post-infection

n.s.

0 24 48

n.s.

Bacte

rial lo

ad

/fly

S. pneumoniae

0 2 4 6

0 3 6 9

n.s. n.s.

0 24 48

n.s.

G H

Bacte

rial lo

ad/fly

P. aeruginosa

Hours post-infectionHours post-infection

n.s. n.s.

0 2 4

n.s.

105

106

107

108

Bacte

rial lo

ad

/fly

L. monocytogenes

C D

0

25

50

75

100

Hours post-infection

p < 0.0001

Days post-infection

n.s.

L. monocytogenes P. aeruginosa

Perc

ent surv

ival

Perc

ent surv

ival

102

103

104

p = 0.0043

**

p = 0.005

104

106

Page 134: Time to eat

120

curves were obtained by log-rank analysis. p-values for bacterial load analyses were obtained by Mann-Whitney nonparametric t-tests.    difference in bacterial loads after infection with L. monocytogenes (Fig 1G), a pathogen

that produced no survival difference between these two groups. In contrast, dFMR1

mutants and wild-type controls had equivalent bacterial loads after P. aeruginosa

infection (Fig 1H), even though dFMR1 mutants were more sensitive to this infection.

Similar bacterial loads with reduced survival after infection with P. aeruginosa suggest

that dFMR1 mutants are less tolerant of P. aeruginosa compared to wild-type controls; in

other words, dFMR1 mutants are less able to withstand the pathogenic effects of this

infection. Because the fly uses different immune mechanisms to fight these specific

pathogens and dFMR1 mutants have different survival kinetics depending on the

pathogen, these data suggest that only specific immune mechanisms are regulated by

dFMR1.  

dFMR1 mutants exhibit a reduced melanization response

While it is difficult to investigate a defect in tolerance, a phenomenon that is not well

understood, there are many established assays for measuring particular resistance

mechanisms. Thus, we next sought to identify which specific resistance mechanisms are

regulated by dFMR1. The three best-characterized immune system mechanisms that flies

use to combat infections are antimicrobial peptide synthesis, melanization, and

phagocytosis by hemocytes, which are analogous to primitive macrophages. To

determine if any of the three immune mechanisms are regulated by the dFMR1, we next

Page 135: Time to eat

121

examined each of these three immune mechanisms in dFMR1 mutants and wild-type

controls.

We first tested whether the loss of dFMR1 leads to changes in the antimicrobial

peptide response, the most well-characterized branch of the Drosophila innate immune

system. Antimicrobial peptides (AMPs) are small peptides that have bacteriocidal

properties and are produced by the Drosophila fat body in response to bacterial and

fungal infections. The well-conserved Toll and imd signaling pathways culminate in

AMP induction and are mainly induced by Gram positive and Gram negative bacterial

infections, respectively. In order to determine if dFMR1 mutants have altered AMP

induction, we examined mRNA levels of two different AMPs, Drosomycin (Drs) and

Diptericin (Dpt), which are induced after activation of the toll and imd signaling

pathways, respectively. Although AMP induction is known to occur following P.

aeruginosa infection [243], the rapid time course of this infection is not suitable for

assaying long-term AMP induction. Instead, we examined basal levels of Drs and Dpt in

uninfected flies and in flies injected with M. luteus and E. coli, used most often to induce

the toll and imd pathways, respectively. Basal levels of Drs and Dpt were increased in

dFMR1 mutants (Fig 2A, p = 0.0370 and Fig 2B, p = 0.0010, respectively), but there

were no significant differences in Drs and Dpt expression at 5 and 24 hours after

infection in dFMR1 mutants and wild-type controls (Figs 2A and 2B). These results

suggest that loss of dFMR1 causes an increase in basal levels of AMP expression but

does not alter AMP response after infection in Drosophila.

Page 136: Time to eat

122

Fig. 2. dFMR1 mutants exhibit altered immune system parameters.

(A-B) dFMR1 mutants have increased basal levels of antimicrobial peptides. dFMR1 mutants exhibited increased basal levels of (A) Drosomycin (p = 0.0370) and (B) Diptericin (p = 0.0010) compared to wild-type controls, but infection-induced levels of Drosomycin and Diptericin were equivalent between dFMR1 mutants and wild-type controls. (C-E) dFMR1 mutants have reduced systemic melanization after infection with L. monocytogenes compared to wild-type controls. (C) Schematic of the melanization assay: 5 days after infection with L. monocytogenes, systemic melanization and melanin deposited at the wounding site were calculated. Briefly, for each fly, areas of melanin deposition were given a score from smallest to largest of 0.5 to 4. These scored were summed for each fly to generate the melanization index for each fly. (D) dFMR1 mutants have reduced systemic melanization when the melanization index was calculated (p < 0.0001) (E) but equivalent melanin deposition at wounding sites (p = 0.1949) after infection with L. monocytogenes. (F-H) dFMR1 mutants exhibit reduced phagocytic activity compared to wild-type controls. (F) Schematic of the phagocytosis assay: S. aureus labeled with the pH sensitive pHrodo fluorophore was injected into wild-type and dFMR1 mutants flies. pHrodo only fluoresces in acidic environments; only hemocytes that have phagocytosed the pHrodo-labeled bacteria fluoresces. The dashed line represents the

Page 137: Time to eat

123

thorax where fluorescence was quantified. (G) Representative micrographs showing thoraces from a wild-type fly and a dFMR1 mutant. (H) Phagocytic activity in wild-type and dFMR1 mutants was quantified by measuring the sum of the intensity, which is a measurement generated from both the area and the intensity of each fluorescent puncta in the thorax of each fly. dFMR1 mutants exhibit significantly less phagocytic activity compared to wild-type controls (p = 0.0159). p-values for melanization and antimicrobial peptides were calculated by the Mann-Whitney nonparametric t-test. p-values for antimicrobial peptide were obtained by the Mann-Whitney nonparametric t-test. p-values for phagocytic activity were calculated using an unpaired t-test. dFMR1 mutants exhibit reduced systemic melanization after infection with L.

monocytogenes compared to wild-type controls

We next tested whether the loss of dFMR1 affects the melanization response. The process

of melanization produces reactive oxygen species that are toxic to pathogens and

parasites and leads to the deposition of melanin both systemically and at local sites of

wounding and infection. Melanization is the main defense used by Drosophila to fight

infections with L. monocytogenes [70]. To measure melanization, dFMR1 mutants and

wild-type controls were infected with L. monocytogenes and inspected for melanin

deposit five days after infection (Fig 2C). A melanization index was calculated for each

fly based on the number and size of melanin deposits present. Deposits at the injection

site, which primarily develop in response to wounding rather than infection, were scored

separately. Although we had previously found there to be no survival difference between

dFMR1 mutants and wild-type controls after infection with L. monocytogenes, we found

a significant decrease in systemic melanin deposition in dFMR1 mutants (Fig 2D, p <

0.0001). However, melanization at the injection site was comparable between the two

groups (Fig 2E, p = 0.1949), indicating that dFMR1 mutants are capable of producing

melanin but are specifically defective in depositing melanin systemically in response to a

pathogen. These results suggest that the loss of dFMR1 may decrease the systemic

melanization response after infection but not locally after wounding. Additionally, this

Page 138: Time to eat

124

altered systemic melanization response does not affect survival after L. monocytogenes

infection. dFMR1 mutants may exhibit wild-type survival rates and bacterial loads after

infection with L. monocytogenes despite reduced rates of melanization in part because of

their increased basal rate of AMP induction, as Toll signaling has been found to protect

flies synergistically with melanization against L. monocytogenes [69].

 

dFMR1 mutants exhibit decreased phagocytosis of bacteria by hemocytes

Drosophila is particularly dependent on phagocytosis by hemocytes to combat certain

infections, including S. pneumoniae, S. marcescens, and S. aureus [138,155]. To

determine if dFMR1 regulates phagocytosis, we then measured phagocytosis of

fluorescently-labeled bacterial particles in vivo in dFMR1 mutants and wild-type controls.

Flies were injected S. aureus bioparticles conjugated to a pH-sensitive rhodamine

fluorophore. Therefore, only particles that are engulfed and processed in acidic

phagolysosomes generate fluorescent signal (Fig 2F). We imaged the dorsal surface of

the thorax, where fluorescent puncta, representing phagocytosed bacteria, were visible

through the cuticle (Fig 2F). We found a decrease in the intensity of fluorescence in

dFMR1 mutants when compared to wild-type controls (Fig 2F and Fig 2G, p = 0.0159).

This result show that the loss of dFMR1 causes decreased phagocytic activity by

hemocytes, suggesting that dFMR1 promotes phagocytosis.

Loss of dFMR1 results in a decrease in axonal clearance after axotomy

Having found a significant defect in hemocyte phagocytosis, we hypothesized that the

loss of dFMR1 may impact phagocytosis by glia, the phagocytic immune cells in the

Page 139: Time to eat

125

brain. Glia and hemocytes use the same receptors and phagocytic pathways for

phagocytosis [244]. Glial phagocytosis is crucial for proper pruning during development

and for the removal of dead neurons [242,245]. To determine if dFMR1 mutants have a

defect in glial phagocytosis, we used a well-characterized assay developed by Marc

Freeman’s group to examine axonal clearance by glia after axotomy in dFMR1 mutants

[242]. Briefly, we genetically-labeled olfactory neurons that innervate the maxillary

palps with membrane-bound GFP. The cell bodies of these neurons reside in the

maxillary palp and the neurons project to glomeruli in the antennal lobes. After

surgically excising the maxillary palps, these olfactory neurons cannot survive without

their cell bodies, and the axonal remnants are phagocytosed by glia. With successful glial

phagocytosis, we see a decrease in GFP intensity in the glomeruli after maxillary palp

excision. GFP intensity is identical in unwounded dFMR1 mutants and wild-type

controls. We found that, after axotomy, dFMR1 mutants have less of a reduction of GFP

intensity in the glomeruli 18 hours after maxillary palp excision when compared to wild-

type controls (Fig 3A-B, p = 0.0078), suggesting that dFMR1 mutants have reduced

axonal clearance after axotomy. These results suggest that dFMR1 mutants may have a

defect in glial phagocytosis.

 

There is a decrease in the induction of Draper, the phagocytic receptor, after

axotomy in dFMR1 mutants

Because we saw a significant decrease in the clearance of axonal remnants in the dFMR1

Page 140: Time to eat

126

 Fig. 3. Clearance of severed axons is delayed in dFMR1 mutants relative to wild-type controls. (A) Shown here are representative micrographs of GFP-labeled glomeruli at 0 and 18 hours after axonal severing of wild-type (top) and dFMR1 mutants (bottom). (B) With quantification, dFMR1 mutants exhibit significantly higher levels of GFP-labeled axonal remnants at 18 hours after injury (p = 0.0078). GFP levels are similar in uninjured dFMR1 mutants and wild-type controls (p > 0.05 ). p-values were obtained by unpaired t-test.    mutants, we next asked if there is a defect in ensheathing glia, the specific glial

population that phagocytoses axonal remnants after injury. Both hemocytes and glia

express Draper, a phagocytic receptor required for phagocytosis [246]. Draper

expression is induced early after injury and is associated with the degree of glial

phagocytosis [241]. To investigate if dFMR1 mutants have reduced glial activity in

addition to reduced clearance of axonal remnants, we tested Draper induction after injury

as a measure of glial activation. We found that there is less Draper induction 18 hours

after injury in dFMR1 mutants compared to wild-type controls (Fig 4A and B, p <

0.0001). These results suggest that glial activation after axonal wounding is impaired in

dFMR1 mutants compared to wild-type controls.

Figure 3: Stone et al., 2014

Uninjured Post-injury

WT

dFMR1

A

WT dFMR1 WT dFMR10

100

200

300

400

Mea

n in

tens

ity o

f G

FPim

mun

oflu

ores

cenc

e (a

u)

uninjured post-injury

**p = 0.0078

n.s.p > 0.05

B

Page 141: Time to eat

127

 

Fig. 4. Fewer activated glia are seen in dFMR1 mutants than in wild-type controls after axonal severing. (A) Shown here are representative micrographs of GFP-labeled axons and Draper-expressing glia for uninjured wild-type and dFMR1 mutants and18 hours after injury. (B) Quantification shows that Draper induction is the same in uninjured wild-type and dFMR1 mutants but significantly lower in dFMR1 mutants after injury (p < 0.0001). p-values were obtained by unpaired t-test.    Clearance of axonal remnants is delayed in dFMR1 mutants while Draper induction

is reduced at numerous time points after axotomy

Having found a defect in axonal clearance and Draper induction at 18 hours after

wounding, we next sought to determine whether glial phagocytosis in dFMR1 mutants is

inhibited or delayed after axotomy. To test this, we performed a time course, examining  

Figure 4: Stone et al., 2014

A Uninjured Post-injury

WT

dFMR1

GFP GFPDRAPER DRAPER

B

WT dFMR1 WT dFMR10

255075

100125

uninjured post-injury

****p < 0.0001

n.s.p > 0.05

Mea

n in

tens

ity o

f D

rape

rim

mun

oflu

ores

cenc

e (a

u)

Page 142: Time to eat

128

Fig. 5. Time course of Draper expression after axonal severing shows that glial phagocytosis is delayed but not abolished in dFMR1 mutants. (A) Shown here are representative micrographs of GFP-labeled axons for wild-type (top) and dFMR1 mutants (bottom) in uninjured flies, 12 hours, 18 hours, 24 hours, and 48 hours after injury. (B) Shown here are representative micrographs of Draper-expressing glia for wild-type (top) and dFMR1 mutants (bottom) in uninjured flies, 12 hours, 18 hours, 24 hours, and 48 hours after injury. (C) Quantification of GFP expression in wild-type and dFMR1 mutants shows significantly higher levels of GFP-labeled axonal remnants at 12 (p = 0.0026), 18 (p = 0.0062), and 48 hours (p = 0.0126) after wounding. Levels of GFP-labeled remnants were the same in uninjured flies (p = 0.1527), and

Figure 5: Stone et al., 2014

0 12 18 24 480

2500

5000

7500

10000

12500

Tota

l GFP

inte

nsity

C

Hours post-injury

**n.s. n.s.** *

0 12 18 24 480

2500

5000

7500

10000

12500

Tota

l DR

APER

inte

nsity

D

Hours post-injury

n.s. n.s. n.s.* *

A

WT

dFMR1

Uninjured 12 hours 18 hours 24 hours 48 hours

GFP

WT

dFMR1

Uninjured 12 hours 18 hours 24 hours 48 hours

DR

AP

ER

B

WTdFMR1

Page 143: Time to eat

129

levels were not consistently higher at 24 hours (p = 0.3265 for this experiment). (D) Quantificiation shows that Draper induction is the same in uninjured wild-type and dFMR1 mutants (p = 0.8069) and in wild-type and dFMR1 mutants 12 (p = 0.4663) and 48 hours (p = 0.8844) after wounding. Draper induction is significantly lower in dFMR1 mutants 18 (p = 0.0256) and 24 hours (p = 0.0261) after wounding. p-values were obtained by unpaired t-test.    GFP intensity of the olfactory neurons and Draper induction at different times after

maxillary palp excision. We found that dFMR1 mutants have significantly more GFP

intensity at 12 (p = 0.0026), 18 (p = 0.0062) and 48 hours (p = 0.0126) after maxillary

palp excision (Fig 5A). There was no consistent difference in GFP intensity at 24 hours

after wounding (p = 0.3265 for this particular experiment). These results show that

axonal remnants are cleared after axotomy in dFMR1 mutants, although the axonal

remnants persist longer in the mutants compared to wild-type controls.

We next investigated Draper induction at various times after axotomy to

determine if Draper induction is also delayed. We found that Draper induction is reduced

at 18 (p = 0.0256) and 24 hours (p = 0.0261) but is no different from wild-type levels by

48 hours post-axotomy (p = 0.8844). These results demonstrate that Draper induction is

not delayed but is moderately inhibited in dFMR1 mutants. Taken together, these

findings suggest that the loss of dFMR1 not only causes general immune system

dysregulation, but also results in a defect in phagocytosis by glia, the resident immune

cells in the brain.

Page 144: Time to eat

130

DISCUSSION

In addition to neurological symptoms, patients with Fragile X syndrome and autism

spectrum disorders often present with other symptoms, including recurrent otitis media

[247] and gastrointestinal disease [248,249]. Patients with Fragile X syndrome also have

elevated cytokines levels [250], suggesting that they have increased inflammation.

Additionally, when the mothers of patients with Fragile X syndrome had aggravations of

an autoimmune disease or asthma during pregnancy, the children are more likely to have

seizures and tics, suggesting that immune activation affects the brain during development

[251]. The non-neurological symptoms in the absence of FMRP, the contribution of the

immune system on the neurological symptoms, and the etiology of these symptoms are

not well understood. Using the Drosophila model for Fragile X syndrome, we probed the

innate immune system function in the absence of dFMR1. We found the dFMR1 mutants

are more susceptible to infection with particular pathogens. When we examined the

three main branches of the fly’s innate immune system in dFMR1 mutants, we found that

the dFMR1 mutants have increased basal levels of antimicrobial peptides, reduced

reactive oxygen species production after infection, and decreased phagocytosis of

bacteria. Just as human patients with Fragile X syndrome and autism have altered

immune system function, these results demonstrate that dFMR1 mutants have altered

immune system function in all three branches of the fly’s innate immune system.

Glia are the phagocytic immune cells in the brain, and glial phagocytosis is

crucial for proper synaptic and dendritic pruning and for the clearance of axonal debris

[242,252]. Patients with Fragile X syndrome and many forms of autism have defects in

synaptic pruning and synaptic plasticity, but how the absence of FMRP affects glial

Page 145: Time to eat

131

function on synaptic pruning and plasticity has not been previously studied. Using an

axonal injury model, we investigated ensheathing glial function in the absence of dFMR1.

We found that dFMR1 mutants exhibit reduced axonal clearance after injury. We next

examined the induction of Draper, a phagocytic receptor that is crucial for the activation

of glial phagocytosis, and we found that dFMR1 mutants have reduced Draper induction

after injury, corresponding with reduced glial activation. When we examined axonal

clearance at numerous times after injury, we found that glial phagocytosis is delayed but

is not abolished in dFMR1 mutants. These results suggest that there is a defect in glial

phagocytosis in the absence of dFMR1 mutants. This is the first in vivo evidence of

defects in a specific glial function, phagocytosis, in an animal model of Fragile X

syndrome.

To further characterize the glial phagocytosis phenotype, it would be important to

determine whether the glial phagocytosis defect stems from a defect in glial migration,

glial phagocytic activity, or a combination of these two possibilities. In order to

phagocytose the axonal remnants in the glomeruli, ensheathing glia need to migrate from

the neuropil and infiltrate the glomeruli. If glial motility is delayed, this could be due to

intracellular defects (such as defects in the molecular components of the glial

cytoskeleton or an altered expression of surface receptors) or extracellular defects (such

as defects in the extracellular matrix or defects in the secretion of glial enzymes, such as

metalloproteases, that allow motility through the extracellular matrix). This last

possibility is especially interesting because minocycline, a drug that inhibits

metalloproteases, has been shown to ameliorate some of the neuroanatomical and

neurological defects in animal models of autism spectrum disorder [136,137]. If a

Page 146: Time to eat

132

migration defect is observed, we can next characterize ensheathing glia during migration

for their cytoskeletal components (including actin, microtubules, intermediate filaments,

and integrins). If, on the other hand, glial motility is unaffected and glial expression of

phagocytic components is delayed, this result would point to a different set of

mechanisms. For example, this defect in phagocytic receptor expression could be due to

defects in glial gene expression, protein translation, or protein trafficking to the

membrane. These results would begin to suggest a potential mechanism that dFMR1 is

employing in glia.

Another aspect of the glial phagocytosis defect that is essential to investigate is

the location in which dFMR1 is required for proper glial function. Both neurons and glia

are thought to actively participate in the process of axonal clearance after severing. After

axonal severing, axonal remnants putatively secrete two types of signals: a "find me"

signal that causes ensheathing glia to migrate specifically to the glomeruli and an "eat

me" signal that activates ensheathing glia to induce Draper expression [253]. The precise

nature and order of secretion of these signals is not known; nor is it known if there are

two discrete signals or a single signal that serves both functions. We hypothesize that

dFMR1 regulates either these neuronal signals and/or the glial response to these signals.

In order to account for the seemingly contradictory increased basal levels of

antimicrobial peptides (AMPs) but decreased hemocyte phagocytosis, we speculate that

the absence of dFMR1 induces an inappropriately primed and “distracted” immune

system that is less able to combat infections. Thus, when the immune system encounters a

pathogen, the immune system may not realize the threat and would then be unable to

combat the infection effectively. How the primed immune system function in dFMR1

Page 147: Time to eat

133

mutants affects brain function is not yet know. However, recent studies have emphasized

the importance of communication between the gut microbiome, the immune system, and

brain function in mammals. Studies in wild-type mouse [254], rat [255], and rhesus

monkey [256] have indicated that maternal immune activation (MIA) during pregnancy

can cause autistic features in the offspring of these animals. These offspring have

increased inflammatory cytokine signaling and abnormalities in hematopoietic lineages in

both the serum and in different areas of the brain [257]. Furthermore, the offsprings’

microbiome was altered, consistent with the observed differences in the microbiota

between patients with autism and typically developing controls [257]. What causes the

altered microbiota composition and the consequence of this cellular defect on immune

system and neuronal functions have not been fully elucidated. It would be important to

examine whether the differences in immune system function in dFMR1 mutants also

extend to an altered gut microbiota in these mutants.

Using the power of Drosophila genetics, it would be beneficial to dissect the

mechanism behind the so-called gut-microbiota-brain axis. In mammals, the phagocytic,

antigen-presenting dendritic cells, an innate immune cell type, is the link between the

microbiota and the adaptive immune system. If these cells have a defect in phagocytosis,

proper communication between the gut microbiota and the immune system may not

occur. Additionally, proper signaling between the immune system and the developing

brain, mediated by cytokine secretion and circulating immune cells, may be altered,

resulting in the exacerbation of some of the symptoms seen in Fragile X syndrome and

autism. Our results highlight the utility of the Drosophila model of Fragile X syndrome

to examine the innate immune system, and one future question is whether the innate

Page 148: Time to eat

134

immune system function in the brain apart from glia is dysregulated in the absence of

dFMR1.

Because Fragile X syndrome and autism are considered developmental disorders,

defects in phagocytosis seen during adulthood in dFMR1 mutants may reflect other, more

general defects in phagocytosis by glia during development. In both mouse and

Drosophila, there appear to be two waves of dFMR1 expression, one during development

and the other during adulthood. The wave of dFMR1 expression during development,

specifically the pupal stage, appears to be required for neuronal function in adulthood

[257,258]. The mushroom body, the neuronal structure important for learning and

memory analogous to the hippocampus in vertebrates, has morphological defects in

dFMR1 mutants during adulthood [259]. Interestingly, neurite pruning of gamma neurons

in the mushroom body has been shown to require proper phagocytosis by glia during

neurodevelopment in wild-type flies [252]. We speculate that the mushroom body defect

observed in dFMR1 mutants stems from a glial phagocytosis defect. Taken together,

these studies highlight the need for future experiments to examine immune system

function, particularly glial phagocytosis, on Fragile X syndrome and autism, with the

potential for new therapeutics for these diseases.

Page 149: Time to eat

135

CHAPTER 5:

CONCLUSIONS

Page 150: Time to eat

136

Many complex physiologies, including immune system function, feeding, metabolism,

and learning and memory, are regulated by circadian rhythms. Circadian rhythms are

generated by the aptly named molecular clock, whose daily oscillations originating in

clock neurons in the brain drive the expression of hundreds of genes both in the neurons

and in the body. However, the manner in which the molecular clock and circadian

rhythms regulate these other complex physiologies has yet to be elucidated. In this thesis,

I have used the Drosophila model organism to investigate the interaction between the

brain and immune system function. In Chapter 2, I showed that a specific immune

mechanism, phagocytosis, is altered in the absence of circadian regulation. In Chapter 3, I

presented evidence that circadian regulation of two tolerance mechanisms, feeding

behaviors and metabolism, impacts survival after two sepsis-like infections in Drosophila.

And finally, in Chapter 4 I demonstrated that the Drosophila model of Fragile X

syndrome, a neurological disease that is characterized by intellectual disability, autistic

features, and circadian dysregulation, exhibits altered phagocytosis both by immune cells

in the body and in the brain. Here I will summarize my findings and propose future

investigations.

Phagocytosis of bacteria is circadian-regulated in Drosophila

An unfortunate consequence of modern living is circadian dysregulation and is evident in,

for instance, jet lag and shift workers. An interaction between sleep and immune system

function has long been known [260], however, the mechanism behind this interaction has

not been well elucidated. Circadian regulation and innate immunity are well conserved

between Drosophila and mammals, and the genetic tools and assays available for

Drosophila making Drosophila a useful model organism to study the relationship

Page 151: Time to eat

137

between the two physiologies. We had previously found that wild-type flies lose

circadian regulation when infected with pathogenic bacteria and that two circadian

mutants, per01 and tim01, are more sensitive to infection with particular pathogens

compared to wild-type controls [81]; however, it was not known which (if any) of the

immunity mechanisms is circadian regulated. I investigated this relationship by

examining particular immune system mechanisms using the genetically tractable

Drosophila circadian tim01 mutant.

As I reported here, we found that tim01 mutants are more sensitive to particular

infections, but not all infections tested, suggesting that only particular immune

mechanisms are circadian regulated. In order to fight against pathogens, Drosophila

innate immune system utilizes three main well-conserved immune mechanisms:

antimicrobial peptide synthesis, melanization, and phagocytosis. We found that

phagocytosis, but not antimicrobial peptide synthesis or melanization, is upregulated

during the night compared to during the day in wild-type flies and is reduced in the tim01

circadian mutant. The definition of a circadian-regulated physiology is a process that

exhibits variation of the course of the circadian cycle in wild-types, and that change is

abolished in circadian mutants; thus, our results suggested that phagocytosis is circadian

regulated in Drosophila. Additionally, we found that a defect in phagocytosis resulting

from the absence of circadian regulation affects survival after infection. These results

suggest that Tim protein regulates phagocytosis.

As the mechanism by which Tim specifically regulates phagocytosis remains

unknown, we have a significant opportunity to build on these findings. For instance,

determining if Tim is expressed in phagocytes themselves and required for wild-type

Page 152: Time to eat

138

phagocytosis would suggest whether Tim acts cell autonomously or non-cell

autonomously to regulate phagocytosis. If Tim is found to work non-cell autonomously, it

would be important to screen for potential downstream effectors between the molecular

clock in the brain and the phagocytes in the body. In addition, we do not know what is

altered in phagocytes in the absence of Tim. For instance, Tim expression at different

times of the day may induce phagocytic surface receptors or other phagocytic machinery

to be differentially expressed, enhancing phagocytosis activity at certain times of the day.

Alternatively, Tim expression may lead to an increased number of phagocytes during

development, leading to increased numbers of phagocytes during the life of the fly, thus

increasing phagocytosis rate compared to tim01 mutants after infection. In the meantime,

the results published here and in PLoS pathogens [63] demonstrate that the absence of

circadian oscillation has a direct, negative effect on phagocytosis after infection,

contributing to poor survival outcomes.

The circadian regulation of feeding and metabolism affects survival of infection in

Drosophila

Sepsis and its related syndromes, severe inflammatory response syndrome and severe

sepsis, are lethal consequences of extreme immune system activation in the presence of

overwhelming infection. The best therapy available to patients with sepsis include

supportive care, such as fluid replenishment, antibiotic treatment, and isolating and

treating the source of the infection [176]. However, patients are still dying of sepsis at an

alarming rate [174], and the optimal treatment is controversial [176]. I pioneered a sepsis-

like infection in Drosophila to further characterize potential metabolic mediators during

Page 153: Time to eat

139

infection with the aim that these data may deepen the understanding of sepsis.

Here I highlighted our findings that circadian regulation of two tolerance

mechanisms, feeding behaviors and metabolism, impacts survival after infection

in Drosophila. I show that per01 circadian mutants are more tolerant of rapidly lethal,

sepsis-like infections with P. aeruginosa and B. cepecia when compared to wild-type

controls, as they live longer but they have equivalent bacterial loads. These results

suggest that the per01 mutants are better able to tolerate the pathogenic effects of these

infections compared to wild-type flies. The improved survival in per01 mutants was found

to be due to a difference in feeding behavior, as uninfected per01 mutants consumed more

food over the course of the day and that starvation of infected per01 mutants and wild-

type controls eliminated the difference in survival. Increased feeding was demonstrated to

be beneficial for survival after infection with B. cepecia and we determined that glucose

and amino acids were necessary food components required for this survival benefit.

Interestingly, although glucose could be injected into the fly to improve survival, amino

acids needed to be delivered to the flies orally to improve survival, and the fly’s

microbiota mediated the improved survival. These results suggest a complicated

relationship between diet, metabolism and microbiota on survival after infection.

Possible future experiments include elucidating how the gut microbiota interacts

with immunity after infection, either by priming the fly to absorb amino acids during

development [261,262] or by promoting the absorption of specific amino acids that

would otherwise be unavailable to the fly [263]. Another important avenue of study

would be to compare the microbiome between wild-type and per01 mutants and clarify

which Genus or species of bacteria is required for microbiota-mediated immunity. Taken

Page 154: Time to eat

140

collectively, these results published here contribute to our understanding of feeding and

metabolism on tolerance mechanisms after rapid, sepsis-like infections.

The Drosophila model of Fragile X syndrome exhibits defects in phagocytosis

Lastly, in addition to examining the altered phagocytic immune system function in

Drosophila circadian mutants, I also examined the immune system function in the

Drosophila model of a particular developmental disease, Fragile X syndrome, known to

have circadian dysregulation. Fragile X syndrome is the most common monogenic cause

of intellectual disability and autism. While the effect of Fragile X syndrome on the

immune system has not been well-characterized, patients with autism have altered

immune system parameters, including cytokine expression patterns [264]. I investigated

immune system function both systemically and in the brain in dFMR1 mutants, the

Drosophila model of Fragile X syndrome.

dFMR1 mutants were found to be more sensitive to infections with specific

bacteria but survived like wild-type flies when infected with other bacteria. Drosophila

use different immune mechanisms to fight against specific pathogens, and interestingly, I

found that dFMR1 mutants have defects in all three immune mechanisms, antimicrobial

peptide generation, melanization, which generates reactive oxygen species, and

phagocytosis. Although dFMR1 mutants have equivalent induction of antimicrobial

peptides after infection compared to wild-type controls, they exhibit an increase in the

basal level of two antimicrobial peptides. This result suggests that in an uninfected state,

the immune system function might be upregulated in dFMR1 mutants. We next examined

melanization in dFMR1 mutants with L. monocytogenes, which Drosophila use

melanization primarily to combat this particular infection [70]. Interestingly, dFMR1

Page 155: Time to eat

141

mutants were found to exhibit decreased systemic melanization after infection with L.

monocytogenes compared to wild-type controls, even though dFMR1 mutants had

identical survival and bacterial loads compared to wild-type controls when infected with

that same pathogen. These results suggest that dFMR1 is required for proper systemic

melanization, although the altered melanization does not affect survival or bacterial load.

This melanization and survival result should be investigated further, not only with other

pathogens to explore whether the melanization phenotype is pathogen-specific or a

general defect in melanization but also to elucidate the mechanism behind this

unexpected phenotype. Because L. monocytogenes infection also leads to antimicrobial

peptide induction [69], the increased basal levels of antimicrobial peptides may keep

bacterial growth in check in lieu of proper melanization, and this relationship should be

examined. Expression levels and activities of key players in the melanization enzymatic

cascade should be also investigated. Additionally, the effects of increased basal levels of

antimicrobial peptide expression and decreased systemic melanization on neuronal

function should be explored.

When dFMR1 mutants were examined for macrophage phagocytic activity, they

were found to have decreased macrophage phagocytosis compared to wild-type controls.

These results suggest that dFMR1 promotes survival after infection and is require for

proper phagocytosis. However, there are many other questions about this phenotype that

should be explored. For instance, do dFMR1 mutants exhibit less phagocytic activity

and/or fewer total phagocytes compared to wild-type controls? Is there a developmental

defect in macrophage phagocytosis? Where is dFMR1 required for proper macrophage

phagocytosis: in the phagocytes themselves or another cell type? Do dFMR1 mutants

Page 156: Time to eat

142

exhibit different patterns of macrophage distribution compared to wild-type controls,

making our assay examining phagocytosis through the dorsal cuticle a poor measure for

macrophage phagocytosis in these mutants? Is there altered circadian regulation of

phagocytosis in dFMR1 mutants? Additionally, do mammalian models of Fragile X

syndrome also have cellular phagocytosis immune cell defects? It is clear from these

questions that the characterization of the macrophage phagocytosis phenotype needs to be

expanded.

Next, I examined the absence of dFMR1 on glial function. Glia are the resident

immune cells in the brain, and glial phagocytosis has previously been found by others to

be important for proper neuronal connections, learning, and memory [233,265]. Using an

axonal injury model, I found that there is a delay in the clearance of axonal remnants

in dFMR1 mutants and that glia are less activated compared to wild-type controls. These

results suggest that the absence of dFMR1 leads to a defect in glial phagocytosis.

In additional to the experiments suggested in the Discussion of Chapter 3, further

examination of the glial phagocytosis phenotype after axotomy needs to be conducted.

There are three basic steps for glial phagocytosis of axonal remnants after axotomy:

presumed glial activation by an unknown “eat me” signal from the axonal remnants, glial

migration to the axonal remnants, and glial phagocytosis of the axonal remnants [266].

Using Drosophila genetics and the imaging techniques described in Chapter 3, whether

the cause of the defect in glial phagocytosis is due to reduced glial activation, glial

migration, or the glial phagocytosis machinery can be explored. Another important

avenue of study is to determine where dFMR1 is required for proper glial phagocytosis: is

it required in glia, neurons, or both? The results to these experiments would suggest the

Page 157: Time to eat

143

mechanism behind the role of dFMR1 on glial phagocytosis after axotomy.

Whether this observed adult glial phagocytosis phenotype has any consequences

on neuronal structure or function during development remains unknown. Preliminary

studies suggest that, similar to the glial phagocytosis phenotype observed after axotomy

in adults, there is a delay in normal axonal pruning during larval and pupal development

in dFMR1 mutants, suggesting a defect in glial phagocytosis during development. The

developmental pruning of a population of neurons has been tested; these neurons are

located in the gamma lobe of the mushroom body, the structure in the fly brain important

for learning and memory. This population of neurons is pruned at specific times during

development [245,252]. Other preliminary studies highlight that these developmental

defects in glial phagocytosis may be ameliorated by minocycline, a tetracycline analog

that is presumed to inhibit matrixmetalloproteases that is currently in drug trials for the

treatment of Fragile X syndrome and autism [134,267]. If minocycline is found to restore

glial phagocytosis in dFMR1 mutants, then this would suggest that the defect in glial

phagocytosis may be due to matrixmetalloproteases or the extracellular matrix and would

lead to further avenues of study. These preliminary results suggest that the glial

phagocytosis phenotype extends to a defect during development, and these are the first

data suggesting that glial glial dysfunction may contribute to observed defects in neuronal

structure. Additionally, these results illustrate the need for the characterization of the role

of dFMR1 on glial function as it relates to neuronal dendritic pruning and synaptogenesis,

and thus neural function. If glial function on dendritic pruning and synaptogenesis is

found to be dysregulated in dFMR1 mutants, these results may help guide future

therapeutics for the disease.

Page 158: Time to eat

144

As I explored in detail above, future studies are required to explore the role of

circadian regulation on systemic immune system function in Fragile X syndrome and

whether the altered systemic immunity impacts neuronal function. While neuronal

function in animal models of Fragile X syndrome has been investigated thoroughly, my

results reported here highlight the need for future studies on how glial function impacts

learning and memory in Fragile X syndrome. In particular, as glial phagocytosis is crucial

for proper synaptic pruning both during development and in adulthood [252,265], the

mechanism behind the glial phagocytosis defect needs to be clarified.

In summary, the findings discussed in this dissertation have built on previous

research by us and other groups and have demonstrated novel interactions between the

brain, metabolism, and immune system function. Each of these three projects has rich

possibilities for future work and discoveries for these complex physiologies with possible

future therapeutic consequences.

Page 159: Time to eat

145

REFERENCES

1. Zylka MJ SL, Weaver DR, Reppert SM (1998) Three period homologs in mammals: Differential light responses in the suprachiasmatic circadian clock and oscillating transcripts outside of brain. Neuron 20: 1103-1110.

2. Zheng X, Sehgal A (2010) AKT and TOR signaling set the pace of the circadian pacemaker. Curr Biol 20: 1203-1208.

3. Hoffmann JA, Reichhart, J-M. (2002) Drosophila innate immunity: an evolutionary perspective. Nature Immunology 3: 121-126.

4. Wan L, Dockendorff TC, Jongens TA, Dreyfuss G (2000) Characterization of dFMR1, a Drosophila melanogaster Homolog of the Fragile X Mental Retardation Protein. Molecular and Cellular Biology 20: 8536-8547.

5. McClung CR (2006) Plant Circadian Rhythms. The Plant Cell 18: 792-803.

6. Kumar Jha P CE, Kalsbeek A. (2015) Circadian rhythms in glucose and lipid metabolism in nocturnal and diurnal mammals. Mol Cell Endocrinol.

7. Smarr BL JK, Driscoll JR, Kriegsfeld LJ (2014) A time to remember: the role of circadian clocks in learning and memory. Behav Neurosci 128: 283-303.

8. Curtis AM BM, Sassoone-Corsi P, O'Neill LA (2014) Circadian clock proteins and immunity. Immunity 40: 178-186.

9. Burke TM MR, Chinoy Ed, Snider JA, Bessman SC, Jung CM, Wright KP Jr. (2013) Combination of light and melatonin time cues for phase advancing the human circadian clock. Sleep 26: 1617-1624.

10. Muindi F ZJ, Colas D, Heller HC (2013) The acute effects of light on murine sleep during the dark phase: importance of melanopsin for maintenance of light-induced sleep. Eur J Neurosci 37: 1727-1736.

11. Bennie JJ DJ, Inger R, Gaston KJ (2014) Biogeography of time partitioning in mammals. Proc Natl Acad Sci U S A 111: 13727-13732.

12. Ensing EP CS, de Wijs FA, Lentferink DH, Ten Hoedt A, Boyce MS, Hut RA (2014) GPS based daily activity patterns in European red deer and North American elk (Cervus elaphus): indication for a weak circadian clock in ungulates. PLoS One 9: e106997.

13. Wang D OT, Mower J, Shimmin LC, Yim S, Honeycutt RL, Tsao H, Li WH. (2004) Molecular evolution of bat color vision genes. Mol Bio Evol 21: 295-302.

14. Zimmerman RH LL, Nishimura N, Galardo AK, Galardo CD, Arruda ME. (2013) Nightly biting cycles of malaria vectors in a heterogeneous transmission area of eastern Amazonian Brazil. Malar J 12: 262.

15. Rieger D FC, Popp J, Bichler D, Dittmann R, Helfrich-Förster C. (2007) The fruit fly Drosophila melanogaster favors dim light and times its activity peaks to early dawn and late dusk. J Biol Rhythms 22: 387-399.

Page 160: Time to eat

146

16. Tahara Y SS (2013) Chronobiology and nutrition. Neuroscience 253: 78-88.

17. Nikhil KL GG, Ratna K, Sharma VK (2014) Role of temperature in mediating morning and evening emergence chronotypes in fruit flies Drosophila melanogaster. J Biol Rhythms 29: 427-441.

18. Bloch G HE, Levine JD, Schwartz WJ (2013) Socially synchronized circadian oscillators. Proc Biol Sci 280: 20130035.

19. Mistiberger RE SD (2004) Social influcences on mammalian circadian rhythms: animal and human studies. Biol Rev Camb Philos Soc 79: 533-556.

20. Konopka RJ, & Benzer, S. (1971). Clock mutants of Drosophila melanogaster. Proceedings of the National Academy of Sciences of the United States of America, 68(9), 2112–2116.

21. Allada R, Chung BY (2010) Circadian organization of behavior and physiology in Drosophila. Annu Rev Physiol 72: 605-624.

22. Darlington TKCtCLC-IToIOIpatS, 280(5369), 1599–1603. doi:10.1126/science.280.5369.1599.

23. Saez L, & Young, M. W. (1996). Regulation of nuclear entry of the Drosophila clock proteins period and timeless. Neuron, 17(5), 911–920.

24. Dubruille R, Emery P (2008) A plastic clock: how circadian rhythms respond to environmental cues in Drosophila. Mol Neurobiol 38: 129-145.

25. Cryan JF, Dinan TG (2012) Mind-altering microorganisms: the impact of the gut microbiota on brain and behaviour. Nat Rev Neurosci 13: 701-712.

26. Kadener S, Stoleru, D., McDonald, M., Nawathean, P., & Rosbash, M. (2007). Clockwork Orange is a transcriptional repressor and a new Drosophila circadian pacemaker component. Genes & Development, 21(13), 1675–1686. doi:10.1101/gad.1552607.

27. Ko HW, Jiang, J., & Edery, I. (2002). Role for Slimb in the degradation of Drosophila Period protein phosphorylated by Doubletime. Nature, 420(6916), 673–678. doi:10.1038/nature01272.

28. Zheng XaAS (2008) Probing the Relative Importance of Molecular Oscillations in the Circadian Clock. Genetics 178: 1147-1155.

29. Ko CH, & Takahashi, J. S. (2006). Molecular components of the mammalian circadian clock. Human Molecular Genetics, 15 Spec No 2(Review Issue 2), R271–7. doi:10.1093/hmg/ddl207.

30. Price JL, Blau, J., Rothenfluh, A., Abodeely, M., Kloss, B., & Young, M. W. (1998). double-time is a novel Drosophila clock gene that regulates PERIOD protein accumulation. Cell, 94(1), 83–95.

31. Martinek S, Inonog, S., Manoukian, A. S., & Young, M. W. (2001). A role for the segment polarity gene shaggy/GSK-3 in the Drosophila circadian clock. Cell, 105(6), 769–779.

32. Sathyanarayanan S, Zheng, X., Xiao, R., & Sehgal, A. (2004). Posttranslational regulation of Drosophila PERIOD protein by protein phosphatase 2A. Cell, 116(4), 603–615.

Page 161: Time to eat

147

33. Fang Y, Sathyanarayanan, S., & Sehgal, A. (2007). Post-translational regulation of the Drosophila circadian clock requires protein phosphatase 1 (PP1). Genes & Development, 21(12), 1506–1518. doi:10.1101/gad.1541607.

34. Koh KJRtDCCbPL-IDoTS, 312(5781), 1809–1812. doi:10.1126/science.1124951.

35. Eide EJ, Vielhaber, E. L., Hinz, W. A., & Virshup, D. M. (2002). The circadian regulatory proteins BMAL1 and cryptochromes are substrates of casein kinase Iepsilon. Journal of Biological Chemistry, 277(19), 17248–17254. doi:10.1074/jbc.M111466200.

36. Akashi M, Tsuchiya, Y., Yoshino, T., & Nishida, E. (2002). Control of Intracellular Dynamics of Mammalian Period Proteins by Casein Kinase I (CKI ) and CKI in Cultured Cells. Molecular and Cellular Biology, 22(6), 1693–1703. doi:10.1128/MCB.22.6.1693-1703.2002.

37. Gallego M, Kang, H., & Virshup, D. M. (2006). Protein phosphatase 1 regulates the stability of the circadian protein PER2. The Biochemical Journal, 399(1), 169. doi:10.1042/BJ20060678.

38. Siepka SM, Yoo, S.-H., Park, J., Song, W., Kumar, V., Hu, Y., et al. (2007). Circadian Mutant Overtime Reveals F-box Protein FBXL3 Regulation of Cryptochrome and Period Gene Expression. Cell, 129(5), 1011–1023. doi:10.1016/j.cell.2007.04.030.

39. Nitabach MN, Taghert PH (2008) Organization of the Drosophila circadian control circuit. Curr Biol 18: R84-93.

40. Stoleru D, Peng, Y., Agosto, J., & Rosbash, M. (2004). Coupled oscillators control morning and evening locomotor behaviour of Drosophila. Nature, 431(7010), 862–868. doi:10.1038/nature02926.

41. Lear BC, Zhang L, Allada R (2009) The neuropeptide PDF acts directly on evening pacemaker neurons to regulate multiple features of circadian behavior. PLoS Biol 7: e1000154.

42. Yoshii T, Wulbeck C, Sehadova H, Veleri S, Bichler D, et al. (2009) The neuropeptide pigment-dispersing factor adjusts period and phase of Drosophila's clock. J Neurosci 29: 2597-2610.

43. Partch CL, Green, C. B., & Takahashi, J. S. (2014). Molecular architecture of the mammalian circadian clock. Trends in Cell Biology, 24(2), 90–99. doi:10.1016/j.tcb.2013.07.002.

44. Dibner C, Schibler, U., & Albrecht, U. (2010). The Mammalian Circadian Timing System: Organization and Coordination of Central and Peripheral Clocks. Annual Review of Physiology, 72(1), 517–549. doi:10.1146/annurev-physiol-021909-135821.

45. Yang X, Lamia, K. A., & Evans, R. M. (2007). Nuclear receptors, metabolism, and the circadian clock. Cold Spring Harbor Symposia on Quantitative Biology, 72(1), 387–394. doi:10.1101/sqb.2007.72.058.

46. Cailotto C, Lei, J., van der Vliet, J., van Heijningen, C., van Eden, C. G., Kalsbeek, A., et al. (2009) Effects of Nocturnal Light on (Clock) Gene Expression in Peripheral Organs: A Role for the Autonomic Innervation of the Liver. PLoS One 4: e5650.

Page 162: Time to eat

148

47. Brown SA, & Azzi, A. (2013). Peripheral Circadian Oscillators in Mammals. In Circadian Clocks (Vol. 217, pp. 45–66). Berlin, Heidelberg: Springer Berlin Heidelberg. doi:10.1007/978-3-642-25950-0_3.

48. Buhr ED, Yoo, S.-H., & Takahashi, J. S. (2010). Temperature as a universal resetting cue for mammalian circadian oscillators. Science, 330(6002), 379–385. doi:10.1126/science.1195262.

49. Liu AC, Welsh, D. K., Ko, C. H., Tran, H. G., Zhang, E. E., Priest, A. A., et al. (2007). Intercellular coupling confers robustness against mutations in the SCN circadian clock network. Cell, 129(3), 605–616. doi:10.1016/j.cell.2007.02.047.

50. Stratmann M, & Schibler, U. (2006). Properties, Entrainment, and Physiological Functions of Mammalian Peripheral Oscillators. Journal of Biological Rhythms, 21(6), 494–506. doi:10.1177/0748730406293889.

51. Alvarez JD, Chen, D., Storer, E., & Sehgal, A. (2003) Non-cyclic and developmental stage-specific expression of circadian clock proteins during murine spermatogenesis. Biology of Reproduction 69: 81-91.

52. Schmutz I, Albrecht, U., & Ripperger, J. A. (2012) The role of clock genes and rhythmicity in the liver. . Molecular and Cellular Endocrinology 349: 38-44.

53. Reddy AB, Maywood, E. S., Karp, N. A., King, V. M., Inoue, Y., Gonzalez, F. J., et al. (2007) Glucocorticoid signaling synchronizes the liver circadian transcriptome. Hepatology 45: 1478-1488.

54. Yang X, Downes, M., Yu, R. T., Bookout, A. L., He, W., Straume, M., et al. (2006) Nuclear receptor expression links the circadian clock to metabolism. Cell 126: 801-810.

55. Kornmann B, Schaad, O., Bujard, H., Takahashi, J. S., & Schibler, U. (2007) System-Driven and Oscillator-Dependent Circadian Transcription in Mice with a Conditionally Active Liver Clock. PLoS Biol 5: e34.

56. Schibler U RJ, Brown SA. (2003) Peripheral circadian oscillators in mammals: time and food. Journal of Biological Rhythms 18: 250-260.

57. Stokkan KA, Yamazaki, S., Tei, H., Sakaki, Y., & Menaker, M. (2001). Science, 291(5503), 490–493. (2001) Entrainment of the circadian clock in the liver by feeding. . Science 291: 490-493.

58. Y-J KTaK (2005) Overview of Innate Immunity in Drosophila. Journal of Biochemistry and Molecular Biology 38: 121-127.

59. Hashimoto C, Hudson, K. L., & Anderson, K. V. (1988). Cell, 52(2), 269–279. (1988) The Toll gene of Drosophila, required for dorsal-ventral embryonic polarity, appears to encode a transmembrane protein. Cell 52: 269-279.

60. Lemaitre B, Nicolas, E., Michaut, L., Reichhart, J. M., & Hoffmann, J. A. (1996) The dorsoventral regulatory gene cassette spätzle/Toll/cactus controls the potent antifungal response in Drosophila adults. Cell 86: 973-983.

Page 163: Time to eat

149

61. Ferrandon D, Imler JL, Hetru C, Hoffmann JA (2007) The Drosophila systemic immune response: sensing and signalling during bacterial and fungal infections. Nat Rev Immunol 7: 862-874.

62. Ligoxygakis P PN, Ji C, Leclerc V, Duvic B, Belvin M, Jiang H, Hoffmann JA, and Reichhart J-M. (2002) A serpin mutant links Toll activation to melanization in the host defense of Drosophila. The EMBO Journa 21: 6330-6337.

63. Stone EF, Fulton BO, Ayres JS, Pham LN, Ziauddin J, et al. (2012) The circadian clock protein timeless regulates phagocytosis of bacteria in Drosophila. PLoS Pathog 8: e1002445.

64. Medzhitov R, Preston-Hurlburt, P., & Janeway, C. A. (1997) A human homologue of the Drosophila Toll protein signals activation of adaptive immunity. Nature 388: 394-397.

65. Poltorak A, He, X., Smirnova, I., Liu, M. Y., Van Huffel, C., Du, X., et al. (1998) Defective LPS signaling in C3H/HeJ and C57BL/10ScCr mice: mutations in Tlr4 gene. Science 282: 2085-2088.

66. Hoshino K, Takeuchi, O., Kawai, T., Sanjo, H., Ogawa, T., Takeda, Y., et al. (1999) Cutting edge: Toll-like receptor 4 (TLR4)-deficient mice are hyporesponsive to lipopolysaccharide: evidence for TLR4 as the Lps gene product. . The Journal of Immunology 162: 3749-3752.

67. Jiravanichpaisal P, Lee BL, Soderhall K (2006) Cell-mediated immunity in arthropods: hematopoiesis, coagulation, melanization and opsonization. Immunobiology 211: 213-236.

68. Tang H (2009) Regulation and function of the melanization reaction in Drosophila. Fly 3: 105-111.

69. Binggeli O NC, Poidevin M, Lemaitre B. (2014) Prophenoloxidase activation is required for survival to microbial infections in Drosophila. PLoS Pathog 10: 1-15.

70. Ayres JS, Schneider DS (2008) A signaling protease required for melanization in Drosophila affects resistance and tolerance of infections. PLoS Biol 6: 2764-2773.

71. Stuart LM, Ezekowitz RA (2008) Phagocytosis and comparative innate immunity: learning on the fly. Nat Rev Immunol 8: 131-141.

72. Ulvila J, Vanha-Aho, L. M. & Ramet, M. (2011) Drosophila phagocytosis - still many unknowns under the surface. . APMIS 119: 651-662.

73. Imeri L, & Opp, M. R. (2009) How (and why) the immune system makes us sleep. Nature Reviews Neuroscience 10: 199-210.

74. Krueger JM, Obál, F. J., Fang, J., Kubota, T., & Taishi, P. (2001) The role of cytokines in physiological sleep regulation. Annals of the New York Academy of Sciences 933: 211-221.

75. Dantzer R (2001) Cytokine-Induced Sickness Behavior: Where Do We Stand? Brain Behavior and Immunity 15: 7-24.

Page 164: Time to eat

150

76. Gemma C, Imeri, L., de Simoni, M. G., & Mancia, M. (1997). The American Journal of Physiology, 272(2 Pt 2), R601–6. (1997) Interleukin-1 induces changes in sleep, brain temperature, and serotonergic metabolism. The American Journal of Physiology 272: R601-606.

77. Hurtado-Alvarado G, Pavón, L., Castillo-García, S. A., Hernández, M. E., Domínguez-Salazar, E., Velázquez-Moctezuma, J., & Gómez-González, B. (2013) Sleep Loss as a Factor to Induce Cellular and Molecular Inflammatory Variations. Clinical and Developmental Immunology 2013: 1-14.

78. Dimitrov S, Lange, T., Nohroudi, K., & Born, J. (2007) Number and function of circulating human antigen presenting cells regulated by sleep. . Sleep 30: 401-411.

79. Born J, Lange, T., Hansen, K., Mölle, M., & Fehm, H. L. (1997). The Journal of Immunology, 158(9), 4454–4464. (1997) Effects of sleep and circadian rhythm on human circulating immune cells. . The Journal of Immunology 158: 4454-4464.

80. Williams JA, Sathyanarayanan, S., Hendricks, J. C., & Sehgal, A. (2007) Interaction between sleep and the immune response in Drosophila: a role for the NFkappaB relish. Sleep 30: 389-400.

81. Shirasu-Hiza MM, Dionne MS, Pham LN, Ayres JS, Schneider DS (2007) Interactions between circadian rhythm and immunity in Drosophila melanogaster. Curr Biol 17: R353-355.

82. Medzhitov R, Schneider DS, Soares MP (2012) Disease tolerance as a defense strategy. Science 335: 936-941.

83. Schneider DS, & Ayres, J. S. (2008) Two ways to survive infection: what resistance and tolerance can teach us about treating infectious diseases. Nature Reviews Immunology 8: 889-895.

84. Brown AE, Baumbach, J., Cook, P.E., and Ligoxygakis, P. (2009). PloS one 4, e4490 (2009) Short-term starvation of immune deficient Drosophila improves survival to gram-negative bacterial infections. PLoS one 4: e4490.

85. Ayres JS, and Schneider, D.S. (2009) The role of anorexia in resistance and tolerance to infections in Drosophila. PLoS Biol 7: e1000150.

86. Dionne MS, Pham LN, Shirasu-Hiza M, Schneider DS (2006) Akt and FOXO dysregulation contribute to infection-induced wasting in Drosophila. Curr Biol 16: 1977-1985.

87. Chambers MC, Song, K.H., and Schneider, D.S. (2012). PloS one 7, e50679. (2012) Listeria monocytogenes infection causes metabolic shifts in Drosophila melanogaster. Plos one 7: 50679.

88. Green CB, Takahashi, J.S., and Bass, J. (2008). The meter of metabolism. Cell 134, 728-742. (2008) The meter of metabolism. Cell 134.

89. Xu K, Zheng, X., and Sehgal, A. (2008). Regulation of feeding and metabolism by neuronal and peripheral clocks in Drosophila. Cell metabolism 8, 289-300. (2008) Regulation of feeding and metabolism by neuronal and peripheral clocks in Drosophila. Cell metabolism 8: 289-300.

Page 165: Time to eat

151

90. Lee JE, Edery I (2008) Circadian regulation in the ability of Drosophila to combat pathogenic infections. Curr Biol 18: 195-199.

91. Jin P, & Warren, S. T. (2000) Understanding the molecular basis of fragile X syndrome. Human Molecular Genetics 9: 901-908.

92. Hernandez RN, Feinberg RL, Vaurio R, Passanante NM, Thompson RE, et al. (2009) Autism spectrum disorder in fragile X syndrome: a longitudinal evaluation. Am J Med Genet A 149A: 1125-1137.

93. Muhle R, Trentacoste, S. V., & Rapin, I. (2004). The genetics of autism. Pediatrics, 113(5), e472–86. (2004) The genetics of autism. Pediatrics 113: e472-486.

94. Schaefer GB, & Mendelsohn, N. J. (2008) Genetics evaluation for the etiologic diagnosis of autism spectrum disorders. Genetics in Medicine 10: 4-12.

95. Kidd SA, Lachiewicz A, Barbouth D, Blitz RK, Delahunty C, et al. (2014) Fragile X Syndrome: A Review of Associated Medical Problems. Pediatrics 134: 995-1005.

96. Warren ST, & Nelson, D. L. (1994) Advances in molecular analysis of fragile X syndrome. JAMA 271: 536-542.

97. Comery TA, Harris, J. B., Willems, P. J., Oostra, B. A., Irwin, S. A., Weiler, I. J., & Greenough, W. T. (1997) Abnormal dendritic spines in fragile X knockout mice: maturation and pruning deficits. Proceedings of the National Academy of Sciences of the United States of America 94: 5401-5404.

98. Hinton VJ BW, Wisniewski K, Rudelli RD. (1991) Analysis of neocortex in three males with the fragile X syndrome. 41 3.

99. Pietrobono R, Tabolacci, E., Zalfa, F., Zito, I., Terracciano, A., Moscato, U., et al. (2005) Molecular dissection of the events leading to inactivation of the FMR1 gene. Human Molecular Genetics 14: 267-277.

100. Wheeler AC, Bailey, D. B., Jr, Berry-Kravis, E., Greenberg, J., Losh, M., Mailick, M., et al. (2014). , 6(1), 1–14. (2014) Associated features in females with an FMR1 premutation. J Neurodev Disord 6: 1-14.

101. Darnell JC, Fraser CE, Mostovetsky O, Darnell RB (2009) Discrimination of common and unique RNA-binding activities among Fragile X mental retardation protein paralogs. Hum Mol Genet 18: 3164-3177.

102. Martin JP, & Bell, J. (1943) A PEDIGREE OF MENTAL DEFECT SHOWING SEX-LINKAGE. Journal of Neurology and Psychiatry 6: 154-157.

103. Verkerk AJ, Pieretti, M., Sutcliffe, J. S., Fu, Y. H., Kuhl, D. P., Pizzuti, A., et al. (1991) Identification of a gene (FMR-1) containing a CGG repeat coincident with a breakpoint cluster region exhibiting length variation in fragile X syndrome. cell 65: 905-914.

104. Ashley CT, Wilkinson, K. D., Reines, D., & Warren, S. T. (1993) FMR1 protein: conserved RNP family domains and selective RNA binding. Science 262: 563-566.

Page 166: Time to eat

152

105. Siomi MC, Higashijima K, Ishizuka A, Siomi H (2002) Casein Kinase II Phosphorylates the Fragile X Mental Retardation Protein and Modulates Its Biological Properties. Molecular and Cellular Biology 22: 8438-8447.

106. Siomi H, Siomi, M. C., Nussbaum, R. L., & Dreyfuss, G. (1993) The protein product of the fragile X gene, FMR1, has characteristics of an RNA-binding protein. Cell 74: 291-298.

107. Eberhart DE, Malter, H. E., Feng, Y., & Warren, S. T. (1996) The fragile X mental retardation protein is a ribonucleoprotein containing both nuclear localization and nuclear export signals. Human Molecular Genetics 5: 1083-1091.

108. Khandjian EW, Corbin, F., Woerly, S., & Rousseau, F. (1996) The fragile X mental retardation protein is associated with ribosomes. Nature Genetics 12: 91-93.

109. Darnell JC, Van Driesche SJ, Zhang C, Hung KY, Mele A, et al. (2011) FMRP stalls ribosomal translocation on mRNAs linked to synaptic function and autism. Cell 146: 247-261.

110. McBride SM, Choi CH, Wang Y, Liebelt D, Braunstein E, et al. (2005) Pharmacological rescue of synaptic plasticity, courtship behavior, and mushroom body defects in a Drosophila model of fragile X syndrome. Neuron 45: 753-764.

111. Darnell JC, Klann E (2013) The translation of translational control by FMRP: therapeutic targets for FXS. Nat Neurosci 16: 1530-1536.

112. Zhang Y OCJ, Siomi MC, Srinivasan S, Dutra A, Nussbaum RL, Dreyfuss G. (1995) The fragile X mental retardation syndrome protein interacts with novel homologs FXR1 and FXR2. EMBO J 14: 5358-5366.

113. Coffee RL, Jr., Tessier CR, Woodruff EA, 3rd, Broadie K (2010) Fragile X mental retardation protein has a unique, evolutionarily conserved neuronal function not shared with FXR1P or FXR2P. Dis Model Mech 3: 471-485.

114. Darnell JC, Fraser CE, Mostovetsky O, Stefani G, Jones TA, et al. (2005) Kissing complex RNAs mediate interaction between the Fragile-X mental retardation protein KH2 domain and brain polyribosomes. Genes Dev 19: 903-918.

115. Darnell JC, Jensen, K. B., Jin, P., Brown, V., Warren, S. T., & Darnell, R. B. (2001) Fragile X mental retardation protein targets G quartet mRNAs important for neuronal function. Cell 107: 489-499.

116. Zang JB, Nosyreva ED, Spencer CM, Volk LJ, Musunuru K, et al. (2009) A mouse model of the human Fragile X syndrome I304N mutation. PLoS Genet 5: e1000758.

117. De Boulle K, Verkerk, A. J., Reyniers, E., Vits, L., Hendrickx, J., Van Roy, B., et al. (1993) A point mutation in the FMR-1 gene associated with fragile X mental retardation. Nature Genetics 3: 31-35.

118. Feng Y, Absher, D., Eberhart, D. E., Brown, V., Malter, H. E., & Warren, S. T. (1997) FMRP associates with polyribosomes as an mRNP, and the I304N mutation of severe fragile X syndrome abolishes this association. Molecular Cell 1: 109-118.

Page 167: Time to eat

153

119. Qin M, Kang, J., Burlin, T. V., Jiang, C., & Smith, C. B. (2005) Postadolescent changes in regional cerebral protein synthesis: an in vivo study in the FMR1 null mouse. The Journal of Neuroscience : the Official Journal of the Society for Neuroscience 25: 5087-5095.

120. Bolduc FV, Bell K, Cox H, Broadie KS, Tully T (2008) Excess protein synthesis in Drosophila fragile X mutants impairs long-term memory. Nat Neurosci 11: 1143-1145.

121. Brown V, Jin, P., Ceman, S., Darnell, J. C., O'Donnell, W. T., Tenenbaum, S. A., Jin, X., Feng, Y., Wilkinson, K. D., Keene, J. D., Darnell, R. B., Warren, S. T. (2001) Microarray identification of FMRP-associated brain mRNAs and altered mRNA translational profiles in fragile X syndrome. Cell 107: 477-487.

122. Zhang YQ, Bailey, A. M., Matthies, H. J., Renden, R. B., Smith, M. A., Speese, S. D., Rubin, G. M., Broadie, K. (2001) Drosophila fragile X-related gene regulates the MAP1B homolog Futsch to control synaptic structure and function1-3. Cell 107: 591-603.

123. Ule J, Jensen, K., Mele, A., & Darnell, R. B. (2005) CLIP: A method for identifying protein–RNA interaction sites in living cells. . Methods 37: 376-386.

124. Irwin SA, Idupulapati M, Gilbert ME, Harris JB, Chakravarti AB, et al. (2002) Dendritic spine and dendritic field characteristics of layer V pyramidal neurons in the visual cortex of fragile-X knockout mice. Am J Med Genet 111: 140-146.

125. Malenka RC, Bear MF (2004) LTP and LTD: an embarrassment of riches. Neuron 44: 5-21.

126. Huber KM, Kayser, M. S., & Bear, M. F. (2000) Role for rapid dendritic protein synthesis in hippocampal mGluR-dependent long-term depression. Science 288: 1254-1257.

127. Li Z, Zhang, Y., Ku, L., Wilkinson, K. D., Warren, S. T., & Feng, Y. (2001) The fragile X mental retardation protein inhibits translation via interacting with mRNA. Nucleic Acids Research 29: 2276-2283.

128. Godfraind JM, Reyniers, E., De Boulle, K., D'Hooge, R., De Deyn, P. P., Bakker, C. E., et al. (1996) Long-term potentiation in the hippocampus of fragile X knockout mice. American Journal of Medical Genetics Part A, 64: 246-251.

129. Huber KM, Gallagher SM, Warren ST, Bear MF (2002) Altered synaptic plasticity in a mouse model of fragile X mental retardation. Proc Natl Acad Sci U S A 99: 7746-7750.

130. Weiler IJ, & Greenough, W. T. (1999) Synaptic synthesis of the Fragile X protein: possible involvement in synapse maturation and elimination. American Journal of Medical Genetics Part A 83: 248-252.

131. Nosyreva ED, & Huber, K. M. (2006). (2006) Metabotropic receptor-dependent long-term depression persists in the absence of protein synthesis in the mouse model of fragile X syndrome. . Journal of Neurophysiology 95: 3291–3295.

132. Bear MF, Huber KM, Warren ST (2004) The mGluR theory of fragile X mental retardation. Trends Neurosci 27: 370-377.

133. Michalon A SM, Ballard TM, Ozmen L, Spooren W, Wettstein JG, Jaeschke G, Bear MF, Lindemann L. Chronic Pharmacological mGlu5 Inhibition Corrects Fragile X in Adult Mice. Neuron 74: 49–56. .

Page 168: Time to eat

154

134. Paribello C, Tao, L., Folino, A., Berry-Kravis, E., Tranfaglia, M., Ethell, I. M., & Ethell, D. W. (2010) Open-label add-on treatment trial of minocycline in fragile X syndrome. BMC Neurology 10: 91.

135. Kim H-S, & Suh, Y.-H. (2009) Minocycline and neurodegenerative diseases. Behavioural Brain Research 196: 168-179.

136. Bilousova TV, Dansie L, Ngo M, Aye J, Charles JR, et al. (2009) Minocycline promotes dendritic spine maturation and improves behavioural performance in the fragile X mouse model. J Med Genet 46: 94-102.

137. Siller SS, Broadie K (2011) Neural circuit architecture defects in a Drosophila model of Fragile X syndrome are alleviated by minocycline treatment and genetic removal of matrix metalloproteinase. Dis Model Mech 4: 673-685.

138. Pham LN, Dionne MS, Shirasu-Hiza M, Schneider DS (2007) A specific primed immune response in Drosophila is dependent on phagocytes. PLoS Pathog 3: e26.

139. Lowrey PL, Takahashi JS (2004) Mammalian circadian biology: elucidating genome-wide levels of temporal organization. Annu Rev Genomics Hum Genet 5: 407-441.

140. Storch KF, Lipan O, Leykin I, Viswanathan N, Davis FC, et al. (2002) Extensive and divergent circadian gene expression in liver and heart. Nature 417: 78-83.

141. Keller M, Mazuch J, Abraham U, Eom GD, Herzog ED, et al. (2009) A circadian clock in macrophages controls inflammatory immune responses. Proc Natl Acad Sci U S A 106: 21407-21412.

142. Hardin PE (2005) The circadian timekeeping system of Drosophila. Curr Biol 15: R714-722.

143. Beaver LM, Giebultowicz JM (2004) Regulation of copulation duration by period and timeless in Drosophila melanogaster. Curr Biol 14: 1492-1497.

144. Krishnan N, Davis AJ, Giebultowicz JM (2008) Circadian regulation of response to oxidative stress in Drosophila melanogaster. Biochem Biophys Res Commun 374: 299-303.

145. Williams JA, Sathyanarayanan S, Hendricks JC, Sehgal A (2007) Interaction between sleep and the immune response in Drosophila: a role for the NFkappaB relish. Sleep 30: 389-400.

146. Schneider DS, Ayres JS (2008) Two ways to survive infection: what resistance and tolerance can teach us about treating infectious diseases. Nat Rev Immunol 8: 889-895.

147. McDonald MJ, Rosbash M (2001) Microarray analysis and organization of circadian gene expression in Drosophila. Cell 107: 567-578.

148. Ceriani MF, Hogenesch JB, Yanovsky M, Panda S, Straume M, et al. (2002) Genome-wide expression analysis in Drosophila reveals genes controlling circadian behavior. J Neurosci 22: 9305-9319.

149. Mendez-Ferrer S, Lucas D, Battista M, Frenette PS (2008) Haematopoietic stem cell release is regulated by circadian oscillations. Nature 452: 442-447.

Page 169: Time to eat

155

150. Kondo Y, Cahyaningsih U, Abe A, Tanabe A (1992) Presence of the diurnal rhythms of monocyte count and macrophage activities in chicks. Poult Sci 71: 296-301.

151. Kurepa Z, Rabatic S, Dekaris D (1992) Influence of circadian light-dark alternations on macrophages and lymphocytes of CBA mouse. Chronobiol Int 9: 327-340.

152. Hriscu ML (2005) Modulatory factors of circadian phagocytic activity. Ann N Y Acad Sci 1057: 403-430.

153. Roy B, Singh R, Kumar S, Rai U (2008) Diurnal variation in phagocytic activity of splenic phagocytes in freshwater teleost Channa punctatus: melatonin and its signaling mechanism. J Endocrinol 199: 471-480.

154. Nguyen-Legros J, Hicks D (2000) Renewal of photoreceptor outer segments and their phagocytosis by the retinal pigment epithelium. Int Rev Cytol 196: 245-313.

155. Schneider DS, Ayres JS, Brandt SM, Costa A, Dionne MS, et al. (2007) Drosophila eiger mutants are sensitive to extracellular pathogens. PLoS Pathog 3: e41.

156. Joyce EA, Kawale A, Censini S, Kim CC, Covacci A, et al. (2004) LuxS is required for persistent pneumococcal carriage and expression of virulence and biosynthesis genes. Infect Immun 72: 2964-2975.

157. Mansfield BE, Dionne MS, Schneider DS, Freitag NE (2003) Exploration of host-pathogen interactions using Listeria monocytogenes and Drosophila melanogaster. Cell Microbiol 5: 901-911.

158. Flyg C, Xanthopoulos K (1983) Insect Pathogenic Properties of Serratia marcescens. Passive and Active Resistance to Insect Immunity Studied with Protease-deficient and Phage-resistant Mutants. Journal of General Microbiology 129: 453-464.

159. Hoiseth SK, Stocker BA (1981) Aromatic-dependent Salmonella typhimurium are non-virulent and effective as live vaccines. Nature 291: 238-239.

160. Clark RI, Woodcock KJ, Geissmann F, Trouillet C, Dionne MS (2011) Multiple TGF-beta superfamily signals modulate the adult Drosophila immune response. Curr Biol 21: 1672-1677.

161. Elrod-Erickson M, Mishra S, Schneider D (2000) Interactions between the cellular and humoral immune responses in Drosophila. Curr Biol 10: 781-784.

162. Lee J-E, Edery I (2008) Circadian regulation in the ability of Drosophila to combat pathogenic infections. Curr Biol 18: 195-199.

163. Naidoo N, Song W, Hunter-Ensor M, Sehgal A (1999) A role for the proteasome in the light response of the timeless clock protein. Science 285: 1737-1741.

164. Konopka RJ, Pittendrigh C, Orr D (1989) Reciprocal behaviour associated with altered homeostasis and photosensitivity of Drosophila clock mutants. J Neurogenet 6: 1-10.

165. Peng Y, Stoleru D, Levine JD, Hall JC, Rosbash M (2003) Drosophila free-running rhythms require intercellular communication. PLoS Biol 1: E13.

Page 170: Time to eat

156

166. Rieger D, Stanewsky R, Helfrich-Förster C (2003) Cryptochrome, compound eyes, Hofbauer-Buchner eyelets, and ocelli play different roles in the entrainment and masking pathway of the locomotor activity rhythm in the fruit fly Drosophila melanogaster. Journal of Biological Rhythms 18: 377-391.

167. Ayres J, Schneider D (2009) The role of anorexia in resistance and tolerance to infections in Drosophila. PLoS Biol 7: e1000150.

168. Lemaitre B, Reichhart JM, Hoffmann JA (1997) Drosophila host defense: differential induction of antimicrobial peptide genes after infection by various classes of microorganisms. Proc Natl Acad Sci U S A 94: 14614-14619.

169. De Gregorio E, Spellman PT, Tzou P, Rubin GM, Lemaitre B (2002) The Toll and Imd pathways are the major regulators of the immune response in Drosophila. Embo J 21: 2568-2579.

170. Ramet M, Pearson A, Manfruelli P, Li X, Koziel H, et al. (2001) Drosophila scavenger receptor CI is a pattern recognition receptor for bacteria. Immunity 15: 1027-1038.

171. Garver LS, Wu J, Wu LP (2006) The peptidoglycan recognition protein PGRP-SC1a is essential for Toll signaling and phagocytosis of Staphylococcus aureus in Drosophila. Proc Natl Acad Sci USA 103: 660-665.

172. Castanon-Cervantes O, Wu M, Ehlen JC, Paul K, Gamble KL, et al. (2010) Dysregulation of inflammatory responses by chronic circadian disruption. J Immunol 185: 5796-5805.

173. Feigin RD, San Joaquin VH, Haymond MW, Wyatt RG (1969) Daily periodicity of susceptibility of mice to pneumococcal infection. Nature 224: 379-380.

174. Liu V, Escobar, G. J., Greene, J. D., Soule, J., Whippy, A., Angus, D. C., & Iwashyna, T. J. (2014) Hospital Deaths in Patients With Sepsis From 2 Independent Cohorts. JAMA 312: 90.

175. Anne Elixhauser BFES (2011) Statistical Brief #122: Septicemia in U.S. Hospitals, 2009,. 1–13.

176. Marik PE (2014) Early Management of Severe Sepsis. . Chest 145: 1407-1418.

177. Lin Y, Han M, Shimada B, Wang L, Gibler TM, et al. (2002) Influence of the period-dependent circadian clock on diurnal, circadian, and aperiodic gene expression in Drosophila melanogaster. Proc Natl Acad Sci U S A 99: 9562-9567.

178. Brennan CA, Anderson KV (2004) Drosophila: the genetics of innate immune recognition and response. Annu Rev Immunol 22: 457-483.

179. Lemaitre B, Nicolas E, Michaut L, Reichhart JM, Hoffmann JA (1996) The dorsoventral regulatory gene cassette spatzle/Toll/cactus controls the potent antifungal response in Drosophila adults. Cell 86: 973-983.

180. Repicky S, Broadie K (2009) Metabotropic glutamate receptor-mediated use-dependent down-regulation of synaptic excitability involves the fragile X mental retardation protein. J Neurophysiol 101: 672-687.

Page 171: Time to eat

157

181. Raberg L, Sim D, Read AF (2007) Disentangling genetic variation for resistance and tolerance to infectious diseases in animals. Science 318: 812-814.

182. Ayres JS, Freitag N, Schneider DS (2008) Identification of Drosophila mutants altering defense of and endurance to Listeria monocytogenes infection. Genetics 178: 1807-1815.

183. Chambers MC, Song KH, Schneider DS (2012) Listeria monocytogenes infection causes metabolic shifts in Drosophila melanogaster. PLoS One 7: e50679.

184. Brown AE, Baumbach J, Cook PE, Ligoxygakis P (2009) Short-term starvation of immune deficient Drosophila improves survival to gram-negative bacterial infections. PLoS One 4: e4490.

185. Ayres JS, Schneider DS (2009) The role of anorexia in resistance and tolerance to infections in Drosophila. PLoS Biol 7: e1000150.

186. Xu K, Zheng X, Sehgal A (2008) Regulation of feeding and metabolism by neuronal and peripheral clocks in Drosophila. Cell Metab 8: 289-300.

187. Green CB, Takahashi JS, Bass J (2008) The meter of metabolism. Cell 134: 728-742.

188. Tapon N, Ito N, Dickson BJ, Treisman JE, Hariharan IK (2001) The Drosophila tuberous sclerosis complex gene homologs restrict cell growth and cell proliferation. Cell 105: 345-355.

189. Hietakangas V, Cohen SM (2007) Re-evaluating AKT regulation: role of TOR complex 2 in tissue growth. Genes & Development 21: 632-637.

190. McGowan MW, Artiss JD, Strandbergh DR, Zak B (1983) A peroxidase-coupled method for the colorimetric determination of serum triglycerides. Clin Chem 29: 538-542.

191. Ja WW, Carvalho GB, Mak EM, de la Rosa NN, Fang AY, et al. (2007) Prandiology of Drosophila and the CAFE assay. Proc Natl Acad Sci U S A 104: 8253-8256.

192. Ja WW, Carvalho GB, Zid BM, Mak EM, Brummel T, et al. (2009) Water- and nutrient-dependent effects of dietary restriction on Drosophila lifespan. Proc Natl Acad Sci U S A 106: 18633-18637.

193. Deshpande SA, Carvalho GB, Amador A, Phillips AM, Hoxha S, et al. (2014) Quantifying Drosophila food intake: comparative analysis of current methodology. Nat Methods 11: 535-540.

194. Emery P (2007) Protein extraction from Drosophila heads. Methods Mol Biol 362: 375-377.

195. Brummel T, Ching A, Seroude L, Simon AF, Benzer S (2004) Drosophila lifespan enhancement by exogenous bacteria. Proceedings of the National Academy of Sciences of the United States of America 101: 12974-12979.

196. Castonguay-Vanier J, Vial L, Tremblay J, Deziel E (2010) Drosophila melanogaster as a model host for the Burkholderia cepacia complex. PLoS One 5: e11467.

197. D'Argenio DA, Gallagher LA, Berg CA, Manoil C (2001) Drosophila as a model host for Pseudomonas aeruginosa infection. J Bacteriol 183: 1466-1471.

Page 172: Time to eat

158

198. Schwager S, Agnoli K, Kothe M, Feldmann F, Givskov M, et al. (2013) Identification of Burkholderia cenocepacia Strain H111 Virulence Factors Using Nonmammalian Infection Hosts. Infection and Immunity 81: 143-153.

199. Clark RI, Tan SWS, Pean CB, Roostalu U, Vivancos V, et al. (2013) MEF2 Is an In Vivo Immune-Metabolic Switch. Cell 155: 435-447.

200. DiAngelo JR, Bland ML, Bambina S, Cherry S, Birnbaum MJ (2009) The immune response attenuates growth and nutrient storage in Drosophila by reducing insulin signaling. Proceedings of the National Academy of Sciences of the United States of America 106: 20853-20858.

201. Grewal SS (2009) Insulin/TOR signaling in growth and homeostasis: A view from the fly world. International Journal of Biochemistry & Cell Biology 41: 1006-1010.

202. Kockel L, Kerr KS, Melnick M, Bruckner K, Hebrok M, et al. (2010) Dynamic switch of negative feedback regulation in Drosophila Akt-TOR signaling. PLoS Genet 6: e1000990.

203. Staschke KA, Dey S, Zaborske JM, Palam LR, McClintick JN, et al. (2010) Integration of general amino acid control and target of rapamycin (TOR) regulatory pathways in nitrogen assimilation in yeast. J Biol Chem 285: 16893-16911.

204. Loewith R, Jacinto E, Wullschleger S, Lorberg A, Crespo JL, et al. (2002) Two TOR complexes, only one of which is rapamycin sensitive, have distinct roles in cell growth control. Molecular Cell 10: 457-468.

205. Wullschleger S, Loewith R, Hall MN (2006) TOR signaling in growth and metabolism. Cell 124: 471-484.

206. Lee N, Woodrum Cl Fau - Nobil AM, Nobil Am Fau - Rauktys AE, Rauktys Ae Fau - Messina MP, Messina Mp Fau - Dabora SL, et al. Rapamycin weekly maintenance dosing and the potential efficacy of combination sorafenib plus rapamycin but not atorvastatin or doxycycline in tuberous sclerosis preclinical models.

207. Huang J, Manning BD (2008) The TSC1-TSC2 complex: a molecular switchboard controlling cell growth. Biochemical Journal 412: 179-190.

208. Sarbassov DD, Ali SM, Kim DH, Guertin DA, Latek RR, et al. (2004) Rictor, a novel binding partner of mTOR, defines a rapamycin-insensitive and raptor-independent pathway that regulates the cytoskeleton. Current Biology 14: 1296-1302.

209. Xu J, Hopkins K, Sabin L, Yasunaga A, Subramanian H, et al. (2013) ERK signaling couples nutrient status to antiviral defense in the insect gut. Proc Natl Acad Sci U S A 110: 15025-15030.

210. Oh WJ, Wu CC, Kim SJ, Facchinetti V, Julien LA, et al. (2010) mTORC2 can associate with ribosomes to promote cotranslational phosphorylation and stability of nascent Akt polypeptide. Embo Journal 29: 3939-3951.

211. Zinzalla V, Stracka D, Oppliger W, Hall MN (2011) Activation of mTORC2 by Association with the Ribosome. Cell 144: 757-768.

Page 173: Time to eat

159

212. Tato I, Bartrons R, Ventura F, Rosa JL (2011) Amino Acids Activate Mammalian Target of Rapamycin Complex 2 (mTORC2) via PI3K/Akt Signaling. Journal of Biological Chemistry 286: 6128-6142.

213. Dimitroff B, Howe K, Watson A, Campion B, Lee HG, et al. (2012) Diet and Energy-Sensing Inputs Affect TorC1-Mediated Axon Misrouting but Not TorC2-Directed Synapse Growth in a Drosophila Model of Tuberous Sclerosis. Plos One 7: 16.

214. Brown J, Wang HZ, Suttles J, Graves DT, Martin M (2011) Mammalian Target of Rapamycin Complex 2 (mTORC2) Negatively Regulates Toll-like Receptor 4-mediated Inflammatory Response via FoxO1. Journal of Biological Chemistry 286: 44295-44305.

215. Kootte RS, Vrieze A, Holleman F, Dallinga-Thie GM, Zoetendal EG, et al. (2012) The therapeutic potential of manipulating gut microbiota in obesity and type 2 diabetes mellitus. Diabetes Obesity & Metabolism 14: 112-120.

216. Caesar R, Fak F, Backhed F (2010) Effects of gut microbiota on obesity and atherosclerosis via modulation of inflammation and lipid metabolism. Journal of Internal Medicine 268: 320-328.

217. Lee WJ, Brey PT (2013) How Microbiomes Influence Metazoan Development: Insights from History and Drosophila Modeling of Gut-Microbe Interactions. In: Schekman R, editor. Annual Review of Cell and Developmental Biology, Vol 29. Palo Alto: Annual Reviews. pp. 571-592.

218. Shin SC, Kim SH, You H, Kim B, Kim AC, et al. (2011) Drosophila Microbiome Modulates Host Developmental and Metabolic Homeostasis via Insulin Signaling. Science 334: 670-674.

219. Storelli G, Defaye A, Erkosar B, Hols P, Royet J, et al. (2011) Lactobacillus plantarum Promotes Drosophila Systemic Growth by Modulating Hormonal Signals through TOR-Dependent Nutrient Sensing. Cell Metabolism 14: 403-414.

220. Lipuma JJ (2005) Update on the Burkholderia cepacia complex. Curr Opin Pulm Med 11: 528-533.

221. Kunz AN, Brook I (2010) Emerging resistant Gram-negative aerobic bacilli in hospital-acquired infections. Chemotherapy 56: 492-500.

222. Chang S, Bray SM, Li Z, Zarnescu DC, He C, et al. (2008) Identification of small molecules rescuing fragile X syndrome phenotypes in Drosophila. Nat Chem Biol 4: 256-263.

223. Developmental Disabilities Monitoring Network Surveillance Year 2010 Principal Investigators CfDCaPC (2014) Prevalence of autism spectrum disorder among children aged 8 years - autism and developmental disabilities monitoring network, 11 sites, United States, 2010. Morbidity and Mortality Weekly Report Surveillance Summaries (Washington, DC : 2002) 63: 1-21.

224. Wills S, Cabanlit, M., Bennett, J., Ashwood, P., Amaral, D., & van de Water, J. (2007) Autoantibodies in Autism Spectrum Disorders (ASD). Annals of the New York Academy of Sciences 1107: 79-91.

Page 174: Time to eat

160

225. Mostafa GA, Shehab, A. A., & Al-Ayadhi, L. Y. (2013) The link between some alleles on human leukocyte antigen system and autism in children. Journal of Neuroimmunology 255: 70-74.

226. Ashwood P, Nguyen DV, Hessl D, Hagerman RJ, Tassone F (2010) Plasma cytokine profiles in Fragile X subjects: is there a role for cytokines in the pathogenesis? Brain Behav Immun 24: 898-902.

227. Pacey LK, Doering LC (2007) Developmental expression of FMRP in the astrocyte lineage: implications for fragile X syndrome. Glia 55: 1601-1609.

228. Saffary R, Xie Z (2011) FMRP regulates the transition from radial glial cells to intermediate progenitor cells during neocortical development. J Neurosci 31: 1427-1439.

229. Yuskaitis CJ, Beurel E, Jope RS (2010) Evidence of reactive astrocytes but not peripheral immune system activation in a mouse model of Fragile X syndrome. Biochim Biophys Acta 1802: 1006-1012.

230. Higashimori H, Morel L, Huth J, Lindemann L, Dulla C, et al. (2013) Astroglial FMRP-dependent translational down-regulation of mGluR5 underlies glutamate transporter GLT1 dysregulation in the fragile X mouse. Hum Mol Genet 22: 2041-2054.

231. Pacey LK, Xuan IC, Guan S, Sussman D, Henkelman RM, et al. (2013) Delayed myelination in a mouse model of fragile X syndrome. Hum Mol Genet 22: 3920-3930.

232. Logan MA, Hackett R, Doherty J, Sheehan A, Speese SD, et al. (2012) Negative regulation of glial engulfment activity by Draper terminates glial responses to axon injury. Nat Neurosci 15: 722-730.

233. Chung WS, Barres BA (2012) The role of glial cells in synapse elimination. Curr Opin Neurobiol 22: 438-445.

234. Blank T, Prinz M (2012) Microglia as modulators of cognition and neuropsychiatric disorders. Glia.

235. Tremblay ME, Stevens, B., Sierra, A., Wake, H., Bessis, A., Nimmerjahn, A. (2011) The role of microglia in the healthy brain. J Neurosci 31: 16064-16069.

236. Derecki NC, Cronk, J. C., Lu, Z., Xu, E., Abbott, S. B. G., Guyenet, P. G., & Kipnis, J. (2012). Nature, 483(7392), 105–109. doi:10.1038/nature10907 (2012) Wild-type microglia arrest pathology in a mouse model of Rett syndrome. Nature 483: 105-109.

237. Gao FB (2002) Understanding fragile X syndrome: insights from retarded flies. . Neuron 34: 859-862.

238. Morales J, Hiesinger, P. R., Schroeder, A. J., Kume, K., Verstreken, P., Jackson, F. R., Nelson, D. L., Hassan, B. A. (2002) Drosophila fragile X protein, DFXR, regulates neuronal morphology and function in the brain. Neuron 34: 961-972.

239. Dockendorff TC, Su, H. S., McBride, S. M. J., Yang, Z., Choi, C. H., Siwicki, K. K., Sehgal A., Jongens T. A. (2002). Drosophila lacking dfmr1 activity show defects in circadian output and fail to maintain courtship interest. Neuron, 34(6), 973–984. (2002) Drosophila lacking dfmr1 activity show defects in circadian output and fail to maintain courtship interest. Neuron 34: 973-984.

Page 175: Time to eat

161

240. Ziegenfuss JS, Biswas R, Avery MA, Hong K, Sheehan AE, et al. (2008) Draper-dependent glial phagocytic activity is mediated by Src and Syk family kinase signalling. Nature 453: 935-939.

241. Ziegenfuss JS, Doherty J, Freeman MR (2012) Distinct molecular pathways mediate glial activation and engulfment of axonal debris after axotomy. Nat Neurosci 15: 979-987.

242. MacDonald JM, Beach MG, Porpiglia E, Sheehan AE, Watts RJ, et al. (2006) The Drosophila cell corpse engulfment receptor Draper mediates glial clearance of severed axons. Neuron 50: 869-881.

243. Lau GW, Goumnerov BC, Walendziewicz CL, Hewitson J, Xiao W, et al. (2003) The Drosophila melanogaster Toll Pathway Participates in Resistance to Infection by the Gram-Negative Human Pathogen Pseudomonas aeruginosa. Infection and Immunity 71: 4059-4066.

244. Kurant E (2011) Keeping the CNS clear: glial phagocytic functions in Drosophila. Glia 59: 1304-1311.

245. Awasaki T, Ito K (2004) Engulfing action of glial cells is required for programmed axon pruning during Drosophila metamorphosis. Curr Biol 14: 668-677.

246. Hashimoto Y TY, Sakurai K, Kutsuna M, Kurokawa K, Awasaki T, Sekimizu K, Nakanishi Y, Shiratsuchi A. (2009) Identification of lipoteichoic acid as a ligand for draper in the phagocytosis of Staphylococcus aureus by Drosophila hemocytes. J Immunol 183: 7451-7460.

247. Gallagher A, Hallahan B (2012) Fragile X-associated disorders: a clinical overview. J Neurol 259: 401-413.

248. Adams JB, Johansen LJ, Powell LD, Quig D, Rubin RA (2011) Gastrointestinal flora and gastrointestinal status in children with autism--comparisons to typical children and correlation with autism severity. BMC Gastroenterol 11: 22.

249. Utari A, Adams E, Berry-Kravis E, Chavez A, Scaggs F, et al. (2010) Aging in fragile X syndrome. J Neurodev Disord 2: 70-76.

250. Molloy CA, Morrow AL, Meinzen-Derr J, Schleifer K, Dienger K, et al. (2006) Elevated cytokine levels in children with autism spectrum disorder. J Neuroimmunol 172: 198-205.

251. Chonchaiya W, Tassone F, Ashwood P, Hessl D, Schneider A, et al. (2010) Autoimmune disease in mothers with the FMR1 premutation is associated with seizures in their children with fragile X syndrome. Hum Genet 128: 539-548.

252. Tasdemir-Yilmaz OE, Freeman MR (2014) Astrocytes engage unique molecular programs to engulf pruned neuronal debris from distinct subsets of neurons. Genes Dev 28: 20-33.

253. Doherty J, Logan MA, Tasdemir OE, Freeman MR (2009) Ensheathing glia function as phagocytes in the adult Drosophila brain. J Neurosci 29: 4768-4781.

254. Malkova NV, Yu CZ, Hsiao EY, Moore MJ, Patterson PH (2012) Maternal immune activation yields offspring displaying mouse versions of the three core symptoms of autism. Brain Behav Immun 26: 607-616.

Page 176: Time to eat

162

255. Baharnoori M, Bhardwaj SK, Srivastava LK (2012) Neonatal behavioral changes in rats with gestational exposure to lipopolysaccharide: a prenatal infection model for developmental neuropsychiatric disorders. Schizophr Bull 38: 444-456.

256. Bauman MD, Iosif AM, Smith SE, Bregere C, Amaral DG, et al. (2014) Activation of the maternal immune system during pregnancy alters behavioral development of rhesus monkey offspring. Biol Psychiatry 75: 332-341.

257. Hsiao EY, McBride SW, Hsien S, Sharon G, Hyde ER, et al. (2013) Microbiota modulate behavioral and physiological abnormalities associated with neurodevelopmental disorders. Cell 155: 1451-1463.

258. Gatto CL, Broadie K (2009) Temporal requirements of the fragile x mental retardation protein in modulating circadian clock circuit synaptic architecture. Front Neural Circuits 3: 8.

259. Michel CI, Kraft R, Restifo LL (2004) Defective neuronal development in the mushroom bodies of Drosophila fragile X mental retardation 1 mutants. J Neurosci 24: 5798-5809.

260. Majde JA KJ (2005) Links between the innate immune system and sleep. J Allergy Clin Immunol 116: 1188-1198.

261. Shin SC, Kim, S.H., You, H., Kim, B., Kim, A.C., Lee, K.A., Yoon, J.H., Ryu, J.H., and Lee, W.J. (2011) Drosophila Microbiome Modulates Host Developmental and Metabolic Homeostasis via Insulin Signaling. . Science 334: 670-674.

262. Lee WJ, and Brey, P.T. (2013). How Microbiomes Influence Metazoan Development: Insights from History and Drosophila Modeling of Gut-Microbe Interactions. In Annual Review of Cell and Developmental Biology, Vol 29, Volume 29, R. Schekman, ed. (Palo Alto: Annual Reviews), pp. 571-592. (2013) How Microbiomes Influence Metazoan Development: Insights from History and Drosophila Modeling of Gut-Microbe Interactions.; R. Schekman e, editor: Palo Alto: Annual Reviews. 571-592 p.

263. Storelli G, Defaye, A., Erkosar, B., Hols, P., Royet, J., and Leulier, F. (2011). Lactobacillus plantarum Promotes Drosophila Systemic Growth by Modulating Hormonal Signals through TOR-Dependent Nutrient Sensing. Cell metabolism 14, 403-414. (2011) Lactobacillus plantarum Promotes Drosophila Systemic Growth by Modulating Hormonal Signals through TOR-Dependent Nutrient Sensing. Cell metabolism 14: 403-414.

264. Ashwood P, Wills S, Van de Water J (2006) The immune response in autism: a new frontier for autism research. J Leukoc Biol 80: 1-15.

265. Chung WS, Clarke LE, Wang GX, Stafford BK, Sher A, et al. (2013) Astrocytes mediate synapse elimination through MEGF10 and MERTK pathways. Nature 504: 394-400.

266. Corty MM, Freeman MR (2013) Cell biology in neuroscience: Architects in neural circuit design: glia control neuron numbers and connectivity. J Cell Biol 203: 395-405.

267. Pardo CA BA, Thurm A, Lee LC, Azhagiri A, Neville DM, Swedo SE. (2013) A pilot open-label trial of minocycline in patients with autism and regressive features. J Neurodev Disord 5: 9.

   

Page 177: Time to eat

163

APPENDIX I  

Fig. 1. Per mutants are more tolerant of infection with P. aeruginosa, and starvation eliminates this survival advantage. (A) Here is a Kaplan-Meier survival curve of wild-type and Per mutants that were infected with P. aeruginosa at an OD of 1.0. Per mutants survived significantly longer than wild-type controls (p > 0.0001). (B) Per mutants have equivalent bacterial loads compared to wild-type controls 3 and 6 hours after infection with P. aeruginosa. These results demonstrate that Per mutants are more tolerant of infection with P. aeruginosa. (C) Under starvation conditions, Per mutants lose their survival advantage after infection with P. aeruginosa (n.s., p = 0.1575). (D) Under starvation conditions, Per mutants still have equivalent bacterial loads compared to wid-type controls. These data suggest that the tolerance phenotype that Per mutants exhibit is a result of feeding or metabolic state.  

5 6 7 8 9

0

25

50

75

100

Time (hours)

Perc

ent surv

ival

p = 0.1575

n.s.

6 7 8 9

0

25

50

75

100

Time (hours)

Perc

ent surv

ival

p < 0.0001

WT fed

Per fed

AAA

B

102

103

104

105

106

Bacte

rial lo

ad p

er

fly

0 3 6

Hours post infection

n.s. n.s. n.s.

B

103

104

105

106

107

Bacte

rial lo

ad p

er

fly

n.s. n.s. n.s.

0 3 6

Hours post infection

DC

WT starved

Per starved

Page 178: Time to eat

164

APPENDIX II                                  

Phagocytosis is upregulated in dFMR1 hypomorphs

Elizabeth Stone and Mimi Shirasu-Hiza

Data and observations unpublished.

Page 179: Time to eat

165

ABSTRACT

Loss of function of FMR1 causes Fragile X syndrome, the most common cause of

intellectual disability and autism. Autism and Fragile X syndrome are associated with a

loss of circadian rhythms, or physiological oscillations with a period of approximately 24

hours, and altered immune system function; however, the molecular mechanisms driving

these symptoms are poorly understood. Using the established Drosophila model of

Fragile X syndrome, we can examine immune system function. We found that a dFRM1

hypomorph, the Drosophila homolog of FMR1, has increased phagocytic activity by

immune blood cells. Surprisingly, the phagocytic phenotype in dFMR1 hypomorphs is

the opposite of the phenotype observed in dFMR1 null flies, which exhibit reduced

phagocytic activity by immune blood cells. Previously, we showed that phagocytosis by

macrophages is circadian-regulated [44]. We report here that dFMR1 hypomorph mutants

have lost circadian regulation of phagocytosis by macrophages. These results are the first

demonstration that loss of circadian regulation in the context a neurological disease leads

to immune system dysfunction. We anticipate that our research will introduce a new

model of Fragile X syndrome with possible future implications for therapeutic

approaches to this and other neurological diseases.

Page 180: Time to eat

166

Materials and Methods

Fly and bacterial strains

A transheterozygous Drosophila dFMR1 mutant was used in these experiments,

made from crossing the hypomorph dFMR1B55 mutant to the null dFMR1Δ50M. Infections

were performed with the following bacteria as described in Chapters 2, 3, and 4 [44].

 Fly rearing conditions

Flies were raised as described in Chapter 4.

 Injections

Flies were anesthetized on CO2 pads. Injections were performed using a custom

Tritech microinjector (Tritech) and pulled glass capillary needles. 50 nL of liquid were

injected into each fly, and the volume of each injection was calibrated by measuring the

diameter of the expelled drop under oil.

 Survival assays

As described previously in Chapters 2, 3, and 4 [40], 60-85 flies per genotype per

condition were assayed for each survival curve and placed in 3 vials of either low

glutamate food or standard dextrose food with approximately 20 flies each. In each

experiment, 20-40 flies of each line were also injected with media as a wounding control.

Death was recorded daily. Data was converted to Kaplan-Meier format using custom

Excel-based software called Count the Dead. Survival curves are plotted as Kaplan-

Meier plots and statistical significance is tested using log-rank analysis using GraphPad

Prism software. All experiments were performed at least three times and yielded similar

results.

Page 181: Time to eat

167

Bacterial load quantitation

Following microbial challenge, six individual flies were collected at each time

point. These flies were homogenized, diluted serially and plated on appropriate media

(tryptic soy blood agar for S.pneumoniae, LB agar for all others). Statistical significance

was determined using non-parametric Mann-Whitney tests. All experiments were

performed at least three times and yielded similar results.

Antimicrobial peptide gene expression

Flies were injected with 50 nL of bacterial culture at OD600 0.10 (M. luteus or E.

coli), or media alone (LB, Difco). Following injection, flies were placed in vials

containing dextrose food and incubated at 25˚C for 6, 24, and 48 hours. Three groups of

six flies after each time point were homogenized in Trizol and stored at -80˚C until

processed. RNA was isolated using a standard Trizol preparation and the samples treated

with DNAse (Invitrogen). Fermentas RevertAid First Strand cDNA synthesis kit was

used to produce the cDNA following the manufacturer’s protocol. Quantitative RT-PCR

was performed using a Stratagene Mx3000 qPCR machine, with Roche FastStart

Universal SYBR Green Master (Rox) mix and the following primer sets: Drosomycin,

Diptericin, and Rpl1. Total mRNA concentration was normalized using Rpl1

expression. Differences in the infection-induced gene expression were calculated by

normalizing to basal gene expression in uninfected wild-type sample taken at the same

time point; p-values were obtained by Mann-Whitney test.

Phagocytosis assay

Page 182: Time to eat

168

5-7 day old male flies were injected with 50 nL of 20 mg/ml pHrodo-labeled S.

aureus in PBS (Molecular Probes, cat# A10010). The flies were allowed to phagocytose

the particles for 35-45 minutes. The wings of the flies were removed and flies were

pinned with a minutien pin onto a silicon pad. Fluorescence images were taken of the

dorsal surface using epifluorescent illumination with a Nikon Eclipse E800 microscope

fitted with an Photometrics Cool Snaps HQ2 camera. Images were captured with Nikon

Elements software. The exposure was set such that the brightest images had a very small

number of saturated pixels. The experiment was repeated three times with 8-14 flies for

each treatment.

Bead inhibition of phagocytosis

Fluorescent 0.2 um polystyrene beads (Molecular Probes, cat# F13080) were

injected into hemolymph to inhibit phagocytosis as previously described [138]. Briefly,

200 ul of beads in solution were washed three times in sterile water and resuspended in

20 ul final volume. 5-7 day old male flies were injected with 50 nL of bead solution. To

confirm that phagocytosis was inhibited with this protocol, the in vivo phagocytosis assay

was performed as described above. Phagocytosis was completely inhibited within 2 days

of bead injection. Flies were then injected with S. pneumoniae at OD600 ranging from 0.1

(see Survival Assays, above).

Page 183: Time to eat

169

Results

dFMR1 hypomorph mutants are more resistant to infections with specific pathogens

Drosophila use specific immune mechanisms to defend against different bacterial

pathogens. By infecting dFMR1 hypmorph mutants with different pathogens, I can assess

whether the loss of dFMR1 causes specific defects in immunity. I first compared the

survival kinetics of dFMR1 mutants and wild-type isogenic controls after infection with

four different pathogens known to elicit immune system activation. I found that dFMR1

mutants survived significantly longer than wild-type controls after infection with S.

pneumoniae (Fig 1A, p < 0.0001), S. marcescens (Fig 1B, p < 0.0001), and L.

monocytogenes (Fig 1C, p < 0.0001). Survival rates of dFMR1 mutants and wild-type

controls did not differ after infection with P. aeruginosa (Fig 1C, p = 0.6898). These

results show that dFMR1 mutants survive longer than wild-type flies after particular

infections.

Figure 1. dFMR1 mutants survive longer than wild-type controls after infection with certain pathogens. The graphs shown here are Kaplan-Meier survival curves of wild-type (black squares) and dFMR1 mutant (dFMR1Δ50M/dFMR1B55, green triangles) flies. dFMR1 mutants survive significantly longer than wild-type flies after infection with (A) S. pneumonia (p < 0.0001) or (B) S. marcenscens (p < 0.0001). dFMR1 mutants have identical survival kinetics when compared to wild-type controls after infection with P. aeruginosa (C, p = 0.6898). dFMR1 mutants are more resistant to L. monocytogenes compared to wild-type controls (D, p < 0.0001). p-values are determined using log-rank analysis.

0 2 4 6 8 10 12

0

25

50

75

100

Time (days)

L. monocytogenes:

dFMR1 is resistant

D

p < 0.0001***

0 2 4 6 8 100

25

50

75

100

Time (days)

S. marcenscens:

dFMR1 is resistant

B

p < 0.0001***

S. pneumoniae:

dFMR1 is resistant

Infection with S. pneumoniae:

dFMR1 mutants are more resistant

0 5 10 150

25

50

75

100

Time (days)

Pe

rce

nt su

rviv

al

A

p < 0.0001***

7.0 7.5 8.0 8.5 9.0

0

25

50

75

100

Time (hours)

P. aeruginosa:

dFMR1 is like wild-type

C

p = 0.6898n.s.

n.s. not significant

*** p < 0.0001dFMR1 mutant

Wild-type

Page 184: Time to eat

170

To determine if dFMR1 mutants were more resistant, meaning they were able to

kill and clear bacteria more efficiently, or if dFMR1 mutants were more tolerant, namely

if they were better able to withstand the pathogenic effects of infection, I next examined

bacterial loads in dFMR1 mutants and wild-type controls at different time points after

infection. I found that dFMR1 mutants have fewer bacterial colonies per fly compared to

wild-type controls at specific time points after infection with S. pneumoniae (Fig 2A) and

S. marcenscens (Fig 2B), indicating dFMR1 mutants were more resistant to these two

bacteria following infection when compared to wild-type controls. dFMR1 mutant flies

had similar bacterial loads after infection with L. monocytogenes (Fig 2C), dFMR1

mutants and wild-type controls had similar bacterial loads after P. aeruginosa infection

(Fig 2C). As the fly uses different immune mechanisms to fight these individual

pathogens and dFMR1 mutants have different survival kinetics compared to wild-type

flies depending on the pathogens, these data suggest that perhaps only specific immune

mechanisms are regulated by dFMR1.

Figure 2. dFMR1 mutants are more resistant to infection with S. pneumoniae and S. marcenscens compared to wild-type controls. After infection, dFMR1 mutants (green triangles) are better able to kill and clear (A) S. pneumonia and (B) S. marcenscens when compared to wild-type controls (black squares). dFMR1 mutants have similar bacterial loads after infection with (C) P. aeruginosa and (D) L. monocytogenes, regardless of nutritional state (data not shown). p-values were determined using Mann-Whitney non-parametric test.

n.s. not significant * p < 0.05 ** p < 0.005dFMR1 mutant

Wild-type

L. monocytogenes:dFMR1 mutants are like wild-type

n.s.

102103104105106107

n.s. n.s.

0 1 2Days post infection

DC

n.s. n.s. n.s.

103104105106107108

0 3 6Hours post infection

P. aeruginosa:dFMR1 mutants are like wild-type

S. pneumoniae:dFMR1 is resistant

100101102103104105106

Col

ony-

Form

ing

Uni

ts

per f

ly (C

FUs)

Hours post infection0 5.5 20

** *

A

n.s.

0 84

S. marcescens:dFMR1 is resistant

103

104

105

106 ** *

B

n.s.

Hours post infection

Page 185: Time to eat

171

Antimicrobial peptide induction is not regulated by dFMR1

I next tested whether the loss of dFMR1 leads to changes in the antimicrobial peptide

response, the most well characterized branch of the Drosophila innate immune system.

Antimicrobial peptides (AMPs) are small peptides that have bacteriocidal properties and

are produced by the Drosophila fat body in response to bacterial and fungal infections.

The well conserved Toll and imd signaling pathways culminate in AMP induction and are

mainly induced by Gram positive and Gram negative bacterial infections, respectively. In

order to determine if dFMR1 mutants have altered AMP induction, we examined mRNA

levels of two different AMPs, Drosomycin (Drs) and Diptericin (Dpt), which are induced

after activation of the toll and imd signaling pathways, respectively. Although AMP

induction is known to occur following P. aeruginosa infection [222], the rapid time

course of this infection is not suitable for assaying long-term AMP induction. Instead, I

examined basal levels of Drs and Dpt in uninfected flies and in flies injected with M.

luteus and E. coli, used most often to induce the toll and imd pathways, respectively.

There were no significant differences in Drs and Dpt expression either basally or at 6, 24,

and 48 hours after infection in dFMR1 mutants and wild-type controls (Fig 3). These

results suggest that loss of dFMR1 does not alter AMP response in Drosophila.

dFMR1 hypomorph mutants exhibit increased phagocytosis of bacteria by

hemocytes

I measured phagocytosis in dFMR1 mutants and wild-type controls. Drosophila

hemocytes use phagocytosis to combat particular infections, including S. pneumoniae, S.

marcescens, and S. aureus. To determine if dFMR1 regulates phagocytosis, I measured

Page 186: Time to eat

172

 

Figure 3. Antimicrobial peptide synthesis is not regulated by dFMR1 Here are representative graphs of the relative expression of (A) Drosomycin and (B) Diptericin assayed by qRT-PCR. dFMR1 hypomorph mutants were found in three trials not to have a statistically-significant difference at all time points assayed after infection in either (A) Drosomycin (6 hours: p = 1.0000; 24 hours: p = 0.1000; 48 hours: p = 0.4000) or (B) Diptericin (6 hours: p = 1.0000; 24 hours: p = 0.2000; 48 hours: p = 0.2000) compared to wild-type controls.

phagocytosis of fluorescently-labeled bacteria by hemocytes in live dFMR1 mutants and

wild-type controls. Briefly, to assay phagocytosis, I injected dFMR1 mutants and wild-

type controls with FITC-labeled, dead S. aureus. I allowed the hemocytes to phagocytose

the dead bacteria, and then we quenched the extracellular fluorescence; thus, the only

fluorescence observed would be due to intracellular, phagocytosed bacteria. I imaged the

dorsal surface of the flies, and fluorescent puncta, representing phagocytosed bacteria,

were visible through the cuticle (Fig 4A). I found an increased number of fluorescent

puncta and an increase in the area of fluorescence in dFMR1 mutants when compared to

wild-type controls (Fig 4B-C). These results show that the loss of dFMR1 causes an

increase in phagocytosis by hemocytes, suggesting that the presence of dFMR1 inhibits

phagocytosis.

Time course of relative Drosomycin expression

0

200

400

600

800

1000

6 hr 24 hr 48 hrGenotype/time point

Rel

ativ

e Ex

pres

sion

WTdFMR1

Time course of relative Diptericin expression

0

2000

4000

6000

8000

6 hr 24 hr 48 hrGenotype/time point

Rel

ativ

e Ex

pres

sion

A B

dFMR1 mutantWild-type

Page 187: Time to eat

173

Figure 4. dFMR1 mutants exhibit increased phagocytosis of bacteria by hemocytes. (A) dFMR1 mutants and wild-type controls were injected with fluorescein-labeled, dead S. aureus, and, after an incubation during which hemocytes phagocytosed the fluorescent bacteria, extracellular fluorescence was quenched. Hemocytes with intracellular bacteria become fluorescent. The dashed line illustrates the dorsal surface of the flies imaged in the adjacent micrographs. (B) Shown here are representative images of phagocytosis in wild-type controls and dFMR1 mutants. The insets are enlarged regions represented by dashed lines. (C) Phagocytosis in wild-type and dFMR1 mutant flies was quantitated by counting the number of fluorescent puncta. dFMR1 mutant flies exhibit significantly more phagocytosis when compared to wild-type controls (p = 0.0047, obtained by unpaired, two-tailed t-test).

dFMR1 mutants have lost circadian regulation of phagocytosis.

One of the most striking phenotypes of dFMR1 mutants is loss of circadian

regulation, a phenotype also seen in human patients with Fragile X syndrome. I showed

in a previous publication that phagocytosis is circadian-regulated [40]. When I found that

dFMR1 mutants have altered phagocytic activity by their macrophages, I set out to test

whether this is because they have lost circadian regulation of phagocytosis. I tested this

hypothesis by performing phagocytosis assays as described above at different times of

day and night (Fig 5). As shown previously, wild-type flies exhibited less phagocytosis

during the day and more at night. In contrast, dFMR1 mutants exhibited similarly high

levels of phagocytosis during the day and night. This result suggests that dFMR1

mutants have macrophages stuck in the “night state”—that is, rather than oscillating

between low and high activity, they have constant high phagocytic activity. Thus dFMR1

appears to inhibit phagocytosis by macrophages.

Page 188: Time to eat

174

Figure 5. dFMR1 mutants do no exhibit circadian changes in phagocytosis. Shown here are analyses of phagocytic activity. For each genotype, two groups were entrained to antiphase circadian rhythms before being assayed. Thus day and night conditions could be assayed simultaneously. (A) Wild-type exhibited circadian differences in phagocytosis, with upregulated activity at night as previously published. (B) In contrast, dFMR1 mutants did not exhibit these circadian differences.

Inhibiting phagocytosis prior to infection with S. pneumoniae abolishes the survival

benefit observed in dFMR1 mutants

I next asked whether the increase in phagocytosis observed in dFMR1 mutants

was responsible for their increase in survival time after S. pneumoniae infection. To

address this question, we inhibited phagocytosis in vivo and then assayed the survival

kinetics in dFMR1 mutants and wild-type controls. Briefly, we inhibited phagocytosis by

injecting flies with 0.2 mm polystyrene beads; hemocytes engulf these beads and

subsequently are unable to phagocytose bacteria 3 days post bead injection (Fig 6A).

Bead-inhibited flies were then infected with S. pneumoniae, and their survival rates were

compared to flies that were not bead-inhibited. Consistent with previous results showing

that bead-inhibiting flies decreases survival time after infection with S. pneumoniae [40],

we found that bead-inhibited wild-type and dFMR1 mutant flies died significantly more

quickly than uninhibited wild-type and dFMR1 mutant flies. We also found that bead-

inhibited dFMR1 mutants and wild-type controls had identical survival kinetics after

day night0

20

40

60

80 n.s.

dFMR1

# flu

ores

cent

pun

cta

day night0

20

40

60

80 *

WT

# flu

ores

cent

pun

cta

A B

Page 189: Time to eat

175

infection with S. pneumoniae (Fig 6B). That is, bead inhibition caused dFMR1 mutants

to lose their survival advantage after S. pneumoniae infection. This result suggests that

the increased phagocytosis observed in dFMR1 mutants is responsible for the increase in

survival time after S. pneumoniae infection when compared to wild-type controls.

Figure 6. Inhibiting phagocytosis abolishes the survival benefit observed in dFMR1 mutants with S. pneumonia infection. (A) Phagocytosis was inhibited by injecting wild-type and dFMR1 mutants flies with polystyrene beads prior to infection with S. pneumoniae (“+ beads”); wild-type and dFMR1 mutant controls were injected with an equal volume of PBS (“no beads”). (B) The graph shown here is a Kaplan-Meier survival curve comparing wild-type and dFMR1 mutant flies with and without inhibited phagocytes. Uninhibited dFMR1 mutants survived significantly longer than wild-type uninhibited flies (p = 0.0144). Bead-inhibited wild-type flies were more sensitive to infection with S. pneumoniae when compared to uninhibited wild-type controls (p < 0.0001), and bead-inhibited dFMR1 mutants were also more sensitive when compared to dFMR1 uninhibited controls. However, bead-inhibited dFMR1 mutant flies had similar survival kinetics as bead-inhibited wild-type flies (p = 0.1779). p-values are determined using log-rank analysis.

Together, these results demonstrate that reduced amounts of dFMR1 improves

phagocytic activity compared to wild-type flies while the complete absence of dFMR1

worsens phagocytic activity compared to wild-type controls, suggesting that there is a

dosage effect of the amount of dFMR1 present in Drosophila for proper immune cell

function.

Page 190: Time to eat

176