development of novel chemical biology tools to probe malaria

117
Development of Novel Chemical Biology Tools to Probe Malaria Parasite Physiology and Aid in Antimalarial Drug Discovery by James R. Abshire B.S., University of Maryland – College Park (2008) Submitted to the Department of Biological Engineering in Partial Fulfillment of the Requirements for the Degree of DOCTOR OF PHILOSOPHY IN BIOLOGICAL ENGINEERING at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY June 2015 © 2015 Massachusetts Institute of Technology. All rights reserved. Signature of Author ........................................................................................................................... James R. Abshire Department of Biological Engineering Certified by ....................................................................................................................................... Jacquin C. Niles Associate Professor, Department of Biological Engineering Thesis Supervisor Accepted by....................................................................................................................................... Forest M. White Associate Professor, Department of Biological Engineering Co-Chair, Course XX Graduate Program Committee

Upload: lekiet

Post on 13-Feb-2017

216 views

Category:

Documents


1 download

TRANSCRIPT

Development of Novel Chemical Biology Tools to Probe Malaria Parasite Physiology and Aid in Antimalarial Drug Discovery

by

James R. Abshire

B.S., University of Maryland – College Park (2008)

Submitted to the Department of Biological Engineering

in Partial Fulfillment of the Requirements for the Degree of

DOCTOR OF PHILOSOPHY IN BIOLOGICAL ENGINEERING

at the

MASSACHUSETTS INSTITUTE OF TECHNOLOGY

June 2015

© 2015 Massachusetts Institute of Technology. All rights reserved. Signature of Author ...........................................................................................................................

James R. Abshire Department of Biological Engineering

Certified by .......................................................................................................................................

Jacquin C. Niles Associate Professor, Department of Biological Engineering

Thesis Supervisor Accepted by.......................................................................................................................................

Forest M. White Associate Professor, Department of Biological Engineering

Co-Chair, Course XX Graduate Program Committee

2

This doctoral thesis has been examined by a committee of the Department of Biological Engineering as follows: Certified by .......................................................................................................................................

K. Dane Wittrup Professor, Departments of Chemical Engineering and Biological Engineering

Thesis Committee Chair Certified by .......................................................................................................................................

Jacquin C. Niles Associate Professor, Department of Biological Engineering

Thesis Supervisor Certified by .......................................................................................................................................

Peter C. Dedon Professor, Department of Biological Engineering

Thesis Committee Member

3

Development of Novel Chemical Biology Tools to Probe Malaria Parasite Physiology and Aid in Antimalarial Drug Discovery

by

James R. Abshire

Submitted to the Department of Biological Engineering on April 14, 2015 in partial fulfillment

of the requirements for the degree of Doctor of Philosophy in Biological Engineering ABSTRACT

Malaria remains a major burden to global public health. Antimalarial drugs are a mainstay in efforts to control and eventually eradicate this disease. However, increasing drug resistance threatens to reverse recent gains in malaria control, making the discovery of new antimalarials critical. Antimalarial discovery is especially challenging due to the unique biology of malaria parasites, the scarcity of tools for identifying new drug targets, and the poorly understood mechanisms of action of existing antimalarials. Therefore, this work describes the development of two chemical biology tools to address unmet needs in antimalarial drug discovery.

A particular challenge in antimalarial development is a shortage of validated parasite drug targets. Potent antimalarials with demonstrated clinical efficacy, like the aminoquinolines and artemisinins, represent a promising basis for rational drug development. Unfortunately, the molecular targets of these drugs have not been identified. While both are thought to interact with parasite heme, linking in vitro heme binding with drug potency remains challenging because labile heme is difficult to quantify in live cells. This work presents a novel genetically-encoded heme biosensor and describes its application to quantify labile heme in live malaria parasites and test mechanisms of antimalarial action.

Another challenge is posed by the widespread malaria parasite Plasmodium vivax, which, unlike P. falciparum, cannot be propagated in vitro, hindering research into parasite biology and drug target identification. P. vivax preferentially invades reticulocytes, which are impractical to obtain in continuous supply. The basis for this invasion tropism remains incompletely understood, mainly because current tools cannot directly link molecular binding events to invasion outcomes. This work presents novel methods for immobilizing synthetic receptors on the red blood cell surface. These receptors are used in proof-of-concept experiments to investigate requirements for efficient invasion via a well-characterized P. falciparum invasion pathway, suggesting this method can be used to elucidate molecular mechanisms underlying parasite invasion tropisms. Future receptor designs could promote the invasion of P. vivax into mature red blood cells and potentially facilitate practical in vitro culture. Taken together, these tools present new opportunities for drug discovery to aid efforts in malaria control and eventual eradication. Thesis Supervisor: Jacquin C. Niles Title: Associate Professor of Biological Engineering

4

ACKNOWLEDGEMENTS This thesis represents roughly six years of focused effort, during which I have benefited from the guidance and support of mentors, colleagues, friends, and family. I would first like to thank Doug Lauffenburger and the Department of Biological Engineering for the opportunity to pursue my graduate education at MIT. The BE department has a unique blend of exciting, diverse research and a collegial, supportive environment fostered by both the faculty and the students. I am grateful to have started my scientific career as a part of this community. I would especially like to thank my advisor, Jacquin Niles, for his support and mentorship over the past six years. During my time in his lab, Jacquin has been consistently attentive, encouraging, and thoughtful in helping me approach challenging scientific questions, design and execute research plans, and think critically about results. I would also like to thank my thesis committee, Dane Wittrup, Manoj Duraisingh, and Pete Dedon for their feedback and suggestions along the way. Additionally, I would like to acknowledge several other students, postdocs, and core facility personnel for their assistance. In particular, Ceth Parker, Helena de Puig Guixé, Prabhani Atukorale, Matthew Wohlever, Charlie Knutson, Koli Taghizadeh, Wendy Salmon, and Glenn Paradis have provided technical help in various aspects of my thesis work. Matthew Edwards and Hunter Elliott also provided valuable technical feedback. In addition, I would like to thank Professors Steven Tannenbaum, John Essigmann, Leona Samson, Darrell Irvine, Robert Sauer, Kim Hamad-Schifferli, and Lee Gehrke for use of their equipment and facilities. I would also like to extend my gratitude to the other members and alumni of the Niles Lab. Brian Belmont, Steve Goldfless, and Jeff Wagner were instrumental in much of my day-to-day training as I started working in the lab. Erika Bechtold, Chris Birch, Sumanta Dey, Suresh Ganesan, Sebastian Nasamu, Bridget Wall, and Daiying Xu all provided helpful discussions and feedback. In addition to the technical help, everyone in the lab contributed to a working environment that was both unique (with our diverse musical tastes) and enjoyable. I am also grateful to this group for their friendship both inside and outside the lab. Additionally, I would like to thank Denise MacPhail for her enthusiastic work behind the scenes helping our lab run smoothly. I would also like to thank my BE classmates and as well as my Boston-area friends for helping make my time here so enjoyable. I will always be grateful for their camaraderie and friendship during the challenging and celebratory moments of graduate school. I was fortunate to receive several fellowships that funded my education and research, including the DuPont Presidential Fellowship and the NIGMS Biotechnology Training Program Fellowship. I would especially like to acknowledge the opportunity provided by the BTP fellowship to pursue an industrial internship during my graduate studies. To that end, I would like to thank Eugene Antipov of Amyris, Inc. for his guidance as my mentor during my internship.

5

I would also like to acknowledge some of my early mentors for fostering my interests in science and engineering. I am especially grateful to Professors Daniel Stein, Ann Smith, and William Bentley for the opportunity to conduct research as an undergraduate student at the University of Maryland. Colin Hebert, a graduate student at the time, was instrumental in my early training as my mentor in the Bentley Lab. I would also like to extend my gratitude to Professor Anne Simon and Dr. Bonnie Dixon, whose engaging Biology and Organic Chemistry lectures inspired me to continue my education and explore a career in these fields. I am also grateful to Dr. Xufeng Wu and Dr. John Hammer at the National Institutes of Health for introducing me to molecular biology research and for the opportunity to learn in the Hammer Lab. I am also incredibly grateful to my family, especially to my parents, for their love and support throughout all my endeavors. Having their advice, encouragement, and perspective “just a phone call away” has been truly indispensible. Finally, I would like to thank my wonderful girlfriend Cheryl for her strength, optimism, and good humor, all of which have helped immensely.

6

ATTRIBUTIONS In addition to the general acknowledgements described above, I would like to detail several specific contributions to the work described in this thesis. The fluorescence lifetime measurements described in Chapter 2 were performed in collaboration with Professor Peter So and his postdoctoral fellow Christopher Rowlands. In addition, Suresh Ganesan built and tested several P. falciparum strains that helped inform my strain construction efforts. The text of Chapter 2 represents a collaborative writing effort between Jacquin Niles, Christopher Rowlands, and me, with input from Peter So. Finally, Professor Carolyn Bertozzi and her graduate student Jason Hudak synthesized and provided the aminooxy-functionalized reagents for several of the experiments described in Chapter 3.

7

TABLE OF CONTENTS CHAPTER 1: INTRODUCTION . . . . . . . . . . . . . . . . 11

1.1 Malaria burden and pathogenesis . . . . . . . . . . . . 11

1.2 Malaria chemotherapy and resistance . . . . . . . . . . . 13

1.3 Challenges and opportunities in antimalarial drug development . . . . 15

1.4 Heme metabolism in malaria parasites . . . . . . . . . . 16

1.4.1 Degradation of host cell hemoglobin . . . . . . . . 16

1.4.2 Heme biosynthesis and utilization . . . . . . . . 18

1.4.3 Other potential sources of parasite heme . . . . . . . 19

1.4.4 Role of heme in antimalarial potency . . . . . . . . 20

1.5 In vitro culture of P. vivax . . . . . . . . . . . . . 23

1.6 Invasion of red blood cells by malaria parasites . . . . . . . . 24

1.6.1 Overview of the invasion process . . . . . . . . . 24

1.6.2 Ligand-receptor interactions governing invasion . . . . . 26

1.6.3 Linking ligand-receptor interactions to invasion outcomes . . 27

1.7 Summary of rationale and work presented . . . . . . . . . 28

1.8 References . . . . . . . . . . . . . . . . . . 30

CHAPTER 2: DEVELOPMENT OF A NOVEL GENETICALLY-ENCODED FRET BIOSENSOR AND

QUANTIFICATION OF A LABILE CYTOSOLIC HEME POOL IN LIVE MALARIA PARASITES . . 38

2.0 Note . . . . . . . . . . . . . . . . . . . 38

2.1 Abstract . . . . . . . . . . . . . . . . . . 38

2.2 Introduction . . . . . . . . . . . . . . . . . 39

8

2.3 Methods . . . . . . . . . . . . . . . . . . 42

2.3.1 Molecular cloning . . . . . . . . . . . . . 42

2.3.2 Protein expression and purification . . . . . . . . 42

2.3.3 Absorbance titrations . . . . . . . . . . . . 43

2.3.4 Fluorescence titrations and FRET efficiency calculations . . 44

2.3.5 Fluorescence lifetime spectroscopy . . . . . . . . 44

2.3.6 Malaria parasite culture . . . . . . . . . . . 46

2.3.7 Preparation of giant multilamellar vesicles . . . . . . 46

2.3.8 In situ FRET analysis . . . . . . . . . . . . 47

2.3.9 Western immunoblotting . . . . . . . . . . . 48

2.4 Results . . . . . . . . . . . . . . . . . . . 49

2.4.1 Design and characterization of initial FRET-based heme biosensor 49

2.4.2 Optimization of initial heme biosensor design . . . . . 54

2.4.3 Correlating FRET efficiencies determined by imaging microscopy

and fluorimetry for calibrating heme concentrations . . . . 60

2.4.4 Measuring labile heme in live malaria parasites . . . . . 61

2.4.5 Quantitative analysis of perturbed heme homeostasis by

heme-interacting antimalarials . . . . . . . . . 65

2.5 Discussion . . . . . . . . . . . . . . . . . . 68

2.6 References . . . . . . . . . . . . . . . . . . 72

2.7 Appendix: MATLAB Scripts . . . . . . . . . . . . . 75

2.7.1 Calculate heme concentration and confidence intervals

from image data . . . . . . . . . . . . . 75

9

2.7.2 Calculate average heme concentration and confidence intervals

from bootstrapping data obtained from multiple experiments . 77

CHAPTER 3: USING SYNTHETIC RECEPTORS TO ELUCIDATE HOST CELL REQUIREMENTS

FOR PARASITE INVASION . . . . . . . . . . . . . . . . . 78

3.1 Abstract . . . . . . . . . . . . . . . . . . 78

3.2 Introduction . . . . . . . . . . . . . . . . . 79

3.3 Methods . . . . . . . . . . . . . . . . . . 83

3.3.1 Malaria parasite culture . . . . . . . . . . . 83

3.3.2 Neuraminidase treatment . . . . . . . . . . . 83

3.3.3 Sialyltransferase treatment . . . . . . . . . . 83

3.3.4 Oxime ligation . . . . . . . . . . . . . 84

3.3.5 Flow cytometry . . . . . . . . . . . . . 84

3.3.6 Fluorimetric sialic acid quantitation . . . . . . . . 85

3.3.7 HPLC sialic acid quantitation . . . . . . . . . . 86

3.3.8 Glycophorin extraction and biotinylation . . . . . . . 86

3.3.9 Glycophorin immobilization . . . . . . . . . . 87

3.3.10 Invasion assay . . . . . . . . . . . . . . 88

3.4 Results . . . . . . . . . . . . . . . . . . . 89

3.4.1 Effect of surface receptor density on parasite invasion rates . . 89

3.4.2 Enzymatic restoration of sialic acid receptors . . . . . 91

3.4.3 Enzymatic attachment of sialic acid receptors with an

alternate terminal linkage . . . . . . . . . . . 93

10

3.4.4 Synthetic glycan receptor construction using

aminooxy-functionalized reagents . . . . . . . . 95

3.4.5 Synthetic glycoprotein receptor construction using

biotin-NeutrAvidin interactions . . . . . . . . . 97

3.5 Discussion . . . . . . . . . . . . . . . . . . 101

3.6 References . . . . . . . . . . . . . . . . . . 104

CHAPTER 4: CONCLUSIONS AND FUTURE WORK . . . . . . . . . . . 107

4.1 Parasite heme biology . . . . . . . . . . . . . . . 107

4.2 Antimalarial drug action . . . . . . . . . . . . . . 108

4.3 Antimalarial drug discovery . . . . . . . . . . . . . 109

4.4 Heme sensing in other biological systems . . . . . . . . . 111

4.5 Synthetic receptor development . . . . . . . . . . . . 112

4.6 Towards synthetic receptor use for in vitro culture of P. vivax . . . . 114

4.7 Conclusions . . . . . . . . . . . . . . . . . 115

4.8 References . . . . . . . . . . . . . . . . . . 116

11

CHAPTER 1: INTRODUCTION

1.1 Malaria burden and pathogenesis

Malaria is an ancient parasitic disease that remains a major burden to global public health. In

2013, there were an estimated 198 million cases of malaria worldwide, which led to an estimated

584,000 deaths, mostly in young children living in sub-Saharan Africa1. Nearly half the world’s

population is at risk for malaria infection, with active disease transmission occurring in 97

countries1. While a single malaria infection can be effectively treated and cured with modern

antimalarial drugs, the widespread distribution of the disease, the possibility for repeat infections,

the limited infrastructure in the most severely affected countries, and the lack of an effective

vaccine currently preclude malaria eradication.

Malaria in humans is caused by five species of the eukaryotic parasite genus Plasmodium – P.

falciparum, P. vivax, P. knowlesi, P. malariae, and P. ovale. Of these, P. falciparum and P. vivax

are responsible for the vast majority of malaria morbidity, while P. falciparum infection accounts

for most malaria-associated deaths1. These parasites are transmitted by the bite of an infected

female Anopheles mosquito, where haploid sporozoites are injected from the mosquito’s salivary

glands, and travel to the liver of the host. Sporozoites then multiply within hepatocytes and are

released into the bloodstream as merozoites, which invade and replicate inside red blood cells. P.

vivax infections retain a population of quiescent parasites in the liver, termed hypnozoites, which

can reactivate and cause disease relapse after the blood-stage infection has been cleared2.

The blood stage of infection is solely responsible for the symptoms of malaria, which include

a recurring high fever and anemia9. During this stage, the infective merozoites bind to and invade

the red blood cell, and begin digesting the contents of the red blood cell cytosol. Parasites of this

stage, termed trophozoites, consume more than 75% of the hemoglobin from the host red blood

12

cell3 before undergoing schizogeny to produce daughter merozoites. Rupture of the schizont

releases the daughter merozoites, which can then infect other red blood cells. Once released,

merozoites are only viable for a few minutes, and typically reinvade new host red blood cells

within 90 seconds4. Haploid blood-stage parasites can also differentiate into sexual-stage

gametocytes, through a process that is not fully understood, but appears to involve epigenetic

regulation5 of gametocyte-specific transcription factors6,7. These gametocytes can then be taken

up by a mosquito during a blood meal, where fertilization and oogenesis lead to the production

of new sporozoites8 (Fig. 1-1).

In addition to fever and anemia, P. falciparum infections often lead to further complications

due to the sequestration of infected red blood cells in the host microvasculature. Blood-stage P.

falciparum parasites extensively remodel the surface of their red blood cell hosts, expressing

proteins that adhere the infected red blood cell to endothelial cells9. Presumably, this enables

infected red blood cells to avoid clearance by the host spleen, but can lead to coagulation,

breakdown in blood vessel structure, and inflammation in the host10-12, with further complications

in individual organs13. Cerebral malaria has a high mortality rate in children and can lead to

permanent neurological impairment14. Sequestration can also occur in the placenta during

pregnancy, leading to anemia in the mother and reducing fetal birth weight15, thereby increasing

the risk of infant mortality16.

13

Figure 1-1. Malaria parasite life cycle in both the human host and the mosquito vector. After inoculation by an infected mosquito, sporozoites invade and replicate inside liver cells (A). Rupture of infected liver cells releases merozoites into the bloodstream, where parasites infect red blood cells (B). Blood-stage parasites can differentiate into gametocytes, which are acquired by the mosquito during a blood meal. Gametocyte fertilization occurs in the mosquito and produces new sporozoites (C). Figure from [17].

1.2 Malaria chemotherapy and resistance

Global efforts to control malaria rely on a combination of approaches. Vector control

methods, through both physical barriers (e.g. bed nets) and insecticide spraying, aim to reduce

disease transmission by preventing mosquito bites. While these methods can be effective, bed

nets must be replaced regularly, and mosquito populations can develop resistance to

insecticides18. Development is also ongoing on a variety of malaria vaccines, with over 40

candidates reaching clinical trials19. However, the most advanced vaccine candidate, which

14

targets the circumsporozoite protein20, has achieved only partial protection with approximately

31% efficacy in Phase III trials21. Therefore, chemotherapy remains a mainstay in combating

malaria.

Two of the most important classes of antimalarial drugs, the aminoquinolines and the

artemisinins, have formed the backbone of modern efforts against malaria. The potent

antimalarial activity of chloroquine was first highlighted by clinical trials in the United States

during World War II22, and chloroquine quickly became the most extensively-used antimalarial

drug23. In addition to its rapid activity against blood-stage malaria parasites, chloroquine was

easily administered, safe, and inexpensive24. The extraordinary success of chloroquine, along

with the insecticide DDT, generated optimism in the 1950s and 1960s that malaria would soon

be eradicated25. However, after extensive use as a monotherapy26, widespread resistance to

chloroquine emerged in the 1960s and 1970s, leading to a devastating resurgence of morbidity

and mortality, especially in sub-Saharan Africa27,28. Today, chloroquine is no longer

recommended to treat P. falciparum malaria due to the high rates of resistance in endemic areas29.

Interest in chloroquine remains, however, due to its unmatched combination of safety, and

affordability, and historical efficacy30.

Artemisinins, in combination with other drugs, have become the standard-of-care in treating

chloroquine-resistant malaria. Artemisinins rapidly kill all blood stages of the parasite (including

gametocytes, making them active against transmission), and exhibit the most rapid clearing of

malaria-induced fever of any antimalarial drug class31. However, artemisinin resistance, first

noted as increased parasite clearance times among patients in Cambodia32, now appears to be

spreading across Southeast Asia33. While artemisinin combination therapies are largely still

15

effective in these regions, likely due to action of their partner drugs, rates of treatment failure are

increasing33. Therefore, new antimalarial drugs are urgently needed34.

1.3 Challenges and opportunities in antimalarial drug development

Antimalarial drug discovery is especially daunting due a unique combination of scientific

and public health challenges. While an in vitro culture system for P. falciparum malaria

developed in the 1970s35 has revolutionized our understanding of parasite biology, such a system

does not exist for P. vivax. This has hindered efforts to measure efficacy of current antimalarial

drugs, and identify new drug targets34. Attempts at P. vivax culture have met with only limited

success (discussed below). Therefore, developing a practical method for in vitro culture of P.

vivax is a major priority for malaria research34,36,37.

Even with a practical in vitro culture system, drug development in P. falciparum remains

challenging. Although sequenced in 200238, the genome of P. falciparum remains poorly

understood, as the functions of many predicted gene sequences have not been determined34. In

addition, its extreme A-T richness and sparse toolkit for gene manipulation have hindered drug

development efforts. While exciting new technologies promise to accelerate this process39-41,

identifying promising drug targets remains a top priority42. In addition, public health challenges

in endemic areas place additional constraints on drug development. Future drugs must be well

tolerated when given in combination with other drugs to minimize the need for follow-up care,

which is often limited, and delay the development of resistance. Additionally, drugs must also be

orally bioavailable and rapidly cure the underlying disease to enable practical mass

administration and maximize patient compliance. Finally, drugs must be especially inexpensive

to be broadly accessible to populations in endemic areas43,44.

16

Taken together, the scarcity of validated drug targets and the stringent requirements for

successful drug candidates suggest that understanding the mechanisms of action of existing

antimalarial drugs is critical. Antimalarials with demonstrated clinical efficacy like the

aminoquinolines and artemisinins represent a promising basis for rational drug development45.

However, the molecular targets of aminoquinoline and artemisinin antimalarials remain

controversial, which precludes broader efforts to exploit these targets. Both classes of drugs have

been shown to interact with heme in vitro, but connecting this in vitro interaction to a mechanism

of parasite toxicity has proven difficult, partly due to a limited understanding of heme

metabolism in the malaria parasite.

1.4 Heme metabolism in malaria parasites 1.4.1 Degradation of host cell hemoglobin

Blood-stage malaria parasites ingest roughly 75% of the hemoglobin from the host red blood

cell into the lysosome-like digestive vacuole3 (Fig. 1-2). Here, the polypeptide chains of

hemoglobin are cleaved into short peptides and individual amino acids by the concerted action of

multiple classes of proteases3. Proteolytic degradation products are then transported into the

parasite cytoplasm, where the individual amino acids are used by the parasite in protein

translation46. In addition to liberating peptides and amino acids, hemoglobin proteolysis releases

large amounts of heme. Given that the digestive vacuole represents only 3-5% of the parasite’s

total volume47, heme liberated from hemoglobin digestion could reach concentrations up to 500

mM in this compartment absent sequestration or destruction of the excess heme48. High

concentrations of free heme are cytotoxic, due to its affinity for lipids in cellular membranes and

its ability to generate reactive oxygen species49.

17

To prevent vacuolar damage from free heme accumulation, the parasite sequesters liberated

heme into inert crystals of heme dimers termed hemozoin50,51. However, the mechanism of

hemozoin crystallization in the parasite is not completely understood. Crystallization of β-

hematin, a synthetic analogue of hemozoin, propagates readily from seed crystals in vitro,

suggesting hemozoin crystallization may be autocatalytic52. In vitro, several parasite proteins

localized to the digestive vacuole, namely histidine-rich proteins 2 and 3 (PfHRP2, 3) and heme

detoxification protein (PfHDP) have been shown to expedite the formation of hemozoin53,54.

Interestingly, while knockouts of PfHRP2 and PfHRP3 still form hemozoin55, PfHDP appears to

be essential54. Crystallization can also be nucleated by parasite-derived lipids in vitro56, which

corroborates electron microscopy data showing hemozoin crystals localized near membrane

structures in the digestive vacuole57.

Importantly, heme sequestration into hemozoin is the only known method by which parasites

can detoxify surplus heme. Recent studies in P. falciparum showed that parasites lack heme

oxygenase activity, and that the heme oxygenase-like enzyme encoded in the parasite genome

appears not to degrade heme58. Others have proposed non-enzymatic degradation pathways for

heme in the food vacuole59 and cytosol60 based on in vitro experiments, but it remains unknown

whether these reactions contribute appreciably to heme degradation in the parasite61. Finally, it is

not known whether heme liberated from hemoglobin degradation is able to escape the digestive

vacuole. The acidic pH of the digestive vacuole would tend to protonate the propionate groups of

free heme molecules, perhaps allowing them to diffuse across the vacuolar membrane48,61. Heme

may also exit the digestive vacuole via specific transporters, although none have been

definitively identified61.

18

1.4.2 Heme biosynthesis and utilization

In addition to large-scale hemoglobin degradation and crystallization of the liberated heme,

malaria parasites also contain a complete pathway for heme biosynthesis. Heme biosynthesis in

the parasite spans three organelles – the cytosol, mitochondrion, and apicoplast62,63 (Fig. 1-2).

This is likely a result of endosymbiotic events that left the parasite with two complete heme

biosynthesis pathways, where redundant functions were eliminated over time64. Heme

biosynthesis begins in the parasite mitochondrion, where a condensation reaction combines

succinyl-CoA and glycine to form δ-aminolevulinic acid (ALA). Next, ALA is converted in a

series of steps to coproporphyrinogen III in the parasite apicoplast65,66, and is then oxidized to

protoporphyrinogen IX in the cytosol67. The final conversion steps involving further oxidation

and loading with iron occur in the mitochondrion68,69. Presumably, trafficking of heme

biosynthesis intermediates between the mitochondrion, apicoplast, and cytosol relies on

transporters or specific binding proteins, as cellular membranes are generally impermeable to

these compounds70. How these intermediates are trafficked between parasite organelles remains

to be elucidated62.

The parasite genome encodes only a small number of known hemoproteins38,63,71. Multiple

cytochromes are present in the parasite mitochondrion and function in the electron transport

chain, which appears to be essential for parasite survival. Atovaquone, which binds to

cytochrome b and inhibits electron transport72 is toxic to the parasite73, while certain mutations in

cytochrome b can render parasites atovaquone-resistant74. The electron transport chain is

required, however, for regenerating ubiquinone, the electron acceptor for dihydroorotate

dehydrogenase (DHOD) during pyrimidine biosynthesis. Expressing a yeast DHOD, which is

cytosolic and operates independently of ubiquinone, renders parasites insensitive to atovaquone75,

19

suggesting that other functions of mitochondrial electron transport (such as ATP generation) are

dispensable. While the parasite genome encodes orthologues of cytochrome b5, the functions of

these proteins have not been determined63. Binding to heme has also been demonstrated with

other recombinantly-expressed parasite proteins76, but the physiological relevance of these

interactions has not been specifically addressed61.

1.4.3 Other potential sources of parasite heme

Recent evidence suggests that blood-stage parasites can meet their metabolic needs without

synthesizing heme de novo. First, the penultimate enzyme of the heme biosynthesis pathway,

protoporphyrinogen IX oxidase, requires an electron acceptor coupled to the mitochondrial

electron transport chain68. Given that parasites expressing yeast DHOD survive electron transport

inhibition with atovoquone75, protoporphyrinogen IX oxidase activity appears not to be required

for growth. Other steps in the heme biosynthesis pathway appear dispensable, as well. Double-

crossover knockouts of the first and last enzymes in the heme biosynthesis pathway (δ-

aminolevulinic acid synthase and ferrochelatase, respectively) have been successfully generated

in blood-stage P. berghei77 and P. falciparum78 parasites, which grew normally but were unable

to progress to the mosquito stages, suggesting that heme biosynthesis is only required for the

exoerythrocytic stages of the parasite life cycle.

Therefore, blood-stage parasites are likely able to obtain heme from other sources. In a recent

study, radiolabeled hemoglobin-heme (obtained by incubating mouse reticulocytes with 14C-

ALA) was found in the mitochondrial cytochromes of ferrochelatase-null P. berghei parasites77,

suggesting that hemoglobin-heme can be trafficked outside the digestive vacuole. Additionally, a

micromolar pool of “free” heme has been indirectly measured in the cytosol of erythrocytes79,

20

which may be accessible to the parasite63. Studies with zinc protoporphyrin IX suggest that

parasites in culture can accumulate protoporphyrins added to the extracellular media80, which

may represent another pathway for scavenging. Finally, heme escape from the digestive vacuole

has been suggested, based on the ability of several cytosolic parasite proteins (especially

glyceraldehyde-3-phosphate dehydrogenase [GAPDH] and thioredoxin reductase [TrxR]) to bind

and be regulated by heme76. However, mechanistic details regarding these proposed trafficking

pathways remain to be elucidated.

1.4.4 Role of heme in antimalarial potency

Multiple classes of antimalarial drugs are known to interact with heme. Artemisinin and its

derivatives are potent drugs extensively used for treating P. falciparum malaria. Artemisinin

activity requires an endoperoxide moiety81 which is thought to undergo iron-assisted reductive

cleavage in the parasite to form damaging radicals82. Recent experiments have shown that

inhibiting hemoglobin degradation attenuates artemisinin toxicity83, as does iron chelation84,

implicating heme and ferrous iron as potential activators in the parasite. However, the

mechanism of artemisinin activation remains incompletely understood, partly due to an inability

to quantify pools of labile heme or labile iron in the parasite. Elucidating this mechanism is

particularly critical given the potency of the artemisinins, which is not well understood, and the

recent emergence of artemisinin resistance33. Mechanistic study of artemisinin action could aid in

developing additional artemisinin derivatives and identify validated molecular targets for new

antimalarials.

Chloroquine, in particular, has been one of the most potent and successful drugs ever

developed against an infectious disease25, despite the devastating spread of resistance in the

21

1960s and 1970s85. However, the mechanism(s) of chloroquine action remain controversial.

Chloroquine has been shown to accumulate significantly in the digestive vacuoles of treated

parasites86. In vitro, chloroquine binds both free87 and exposed heme on growing hemozoin

crystals88 and inhibits crystal growth. Additionally, chloroquine-treated parasites contain less

detectable hemozoin89-91, and higher amounts of heme unassociated with hemozoin or

hemoglobin91. Recently, increased extravacuolar iron density has also been observed in

chloroquine-treated parasites91. Taken together, these results support a model where the heme-

chloroquine complex blocks heme detoxification to hemozoin in the digestive vacuole and

causes free heme to accumulate in the parasite. However, a direct link between chloroquine and

toxicity through accumulation of unbound heme has not been demonstrated, partly due to the

inability to reliably quantify unbound heme in live cells.

Other effects of chloroquine treatment have been observed in a series of in vitro and in situ

experiments. Chloroquine has been proposed to target polyamine biosynthesis based on its

activity against ornithine decarboxylase in cultured parasites92. Chloroquine has also shown

inhibitory activity against protein synthesis both in cell-free extracts and in cultured parasites93.

In vitro, chloroquine has been shown to inhibit proteases involved in hemoglobin degradation94,

and to inhibit proposed heme degradation pathways involving hydrogen peroxide59 and

glutathione95. However, the extent to which these interactions contribute to parasite toxicity are

not known. Furthermore, recent studies with chloroquine analogues found that the inhibition of

hemozoin formation was correlated with cytostatic but not cytocidal activity96, suggesting that

chloroquine toxicity may be the result of multiple mechanisms. Further studies are needed to

dissect the mechanism of chloroquine cytotoxicity and elucidate the role(s) played by heme.

22

Resistance to chloroquine has been mapped to mutations in the P. falciparum chloroquine

resistance transporter (PfCRT)97,98, a transmembrane protein associated with the digestive

vacuolar membrane99. While the native role of PfCRT is not known, bioinformatic studies have

suggested a possible function as a transporter of small molecules100,101. Similarly, the role of

PfCRT in determining chloroquine resistance has not been fully defined. Less chloroquine

appears to accumulate in chloroquine-resistant parasites expressing mutant PfCRT than in

sensitive strains102, and similar results have been obtained in experiments using isolated digestive

vacuoles103. This suggests that resistance may be mediated primarily by reducing parasite

exposure to chloroquine, and therefore chloroquine’s mechanism(s) of action may hold promise

for future antimalarial development.

Figure 1-2. Summary of heme metabolism in blood-stage parasites, depicting hemoglobin digestion, hemozoin formation, interactions with antimalarial drugs, and heme biosynthesis. Abbreviations used: amino acids (AA), heme detoxification protein (HDP), chloroquine (CQ), artemisinin (ART), and activated artemisinin (ART**). Figure from [61].

23

1.5 In vitro culture of P. vivax

A landmark 1976 study established a method for continuously propagating P. falciparum

cultures using human red blood cells35, which are readily available. This discovery was critical in

subsequent research into P. falciparum biology, and as of early 2015, had been cited in over

5,000 articles (statistic from Web of Science). However, such a system does not exist for P. vivax,

which is currently impractical to culture in vitro because it preferentially invades reticulocytes

(immature red blood cells)104, which constitute between 0.5% and 1.5% of the circulating cells in

human peripheral blood105. In contrast, P. falciparum invades both reticulocytes and mature red

blood cells efficiently106. Furthermore, reticulocytes mature rapidly into normocytes in vitro, with

a half-life of approximately 30 hours107. Therefore, propagating P. vivax relies on a continuous

supply of a rare and transient blood component to maintain an adequate population of invadable

cells. Studies reporting P. vivax propagation without enriched reticulocytes were either of very

short duration108-112, could not maintain high parasitemia113, or could not be reproduced114,115.

Obtaining enriched reticulocytes for P. vivax culture is difficult and costly. In one study,

reticulocytes were supplied from the blood of a hemochromatosis patient being treated by

therapeutic phlebotomy and enriched by centrifuging the blood cells in homologous plasma.

While this technique resulted in stable P. vivax propagation over a two-week period, total

reticulocyte yields were low (< 20%)105, and these results have not been replicated in other

groups115. Human umbilical cord blood, which is also naturally enriched in reticulocytes, has also

been used for continuous P. vivax culture up to two months116-118. However, these techniques

were unable to maintain high parasitemia. In both cases, these techniques required continuous

access to patient-derived samples that are not readily available. Reticulocytes can also be

generated by differentiating hematopoietic stem cells (HSCs) derived from cord blood, and HSC-

24

derived reticulocytes have been shown to be usable in P. vivax culture119,120. However, this

method is labor-intensive, as HSCs require two weeks of culture to mature into reticulocytes, and

expensive, due to the mixture of growth factors and cytokines required121. Parasites can be

obtained for brief ex vivo studies from infected research animals, as P. vivax also infects New

World monkeys of the Saimiri and Aotus genera105,122. However, maintaining infected research

animals (especially primates) is often cost-prohibitive and raises ethical concerns. In contrast to

these existing methods, an ideal culturing system would propagate this parasite continuously

using human normocytes and other reagents that are readily and inexpensively available.

However, this would depend on overcoming the P. vivax preference for invading reticulocytes,

the basis of which is only partially understood.

1.6 Invasion of red blood cells by malaria parasites

1.6.1 Overview of the invasion process

During the blood stage of malaria infection, parasites bind to and invade red blood cells in a

multi-step process (Fig. 1-3)123. Merozoites released from a bursting schizont quickly associate

with erythrocytes, averaging less than 40 seconds between schizont rupture and contact with a

potential host cell4. This initial contact is mediated by long distance and relatively low-affinity

interactions and can occur with the merozoite in any orientation124. Invasion proteins are then

released to the merozoite surface from secretory organelles termed rhoptries and micronemes,

which are located at the apical end of the merozoite125. Secretion of invasion proteins occurs

shortly before the proteins are needed, which minimizes exposure to the host’s immune system

and slows the development of an immune response126.

25

The parasite then repositions to place its apical end in contact with the red blood cell127, and

through a series of ligand-receptor interactions, attaches irreversibly to the red cell membrane126.

By electron microscopy, the interface between parasite and host cell shows the electron density

and close contact between the two cells typical of a tight junction128. Finally, the parasite enters

the red blood cell by moving the junction along its length, effectively pushing itself into the host

cell and sealing the membrane closed behind it. Shortly after invasion, the red cell undergoes

morphological changes induced by ion fluxes but then quickly returns to its previous shape4.

Figure 1-3. Overview of the red blood cell invasion process. Merozoites (Mrz) initially attach to the red blood cell (RBC) surface in any orientation through low-affinity interactions. After attachment, the merozoite reorients to place its apical end in close proximity to the RBC surface, where a series of ligand-receptor interactions (a) stabilize the formation of a tight junction (b). The parasite then moves the junction along its length and sheds its protein coat (c-d), creating the parasitophorous vacuole (PVM) and sealing the RBC membrane closed (e). Figure from [129], copyright © the authors.

26

1.6.2 Ligand-receptor interactions governing invasion

Ligand-receptor interactions that precede tight junction formation define the cell types that

can be invaded by the merozoite and irreversibly begin the invasion process126,130,131 (Fig. 1-3).

Two families of parasite ligands have been identified: the Duffy-binding like (DBL) and

reticulocyte-binding like homologues (Rh), which originate in micronemes132,133 and rhoptries134,

respectively, before release to the merozoite surface. DBL-family proteins are homologous to P.

vivax proteins that bind to the Duffy antigen and mediate invasion135,136. The Rh proteins are

homologous to a family of P. vivax proteins that bind to reticulocytes137,138, and are believed to

underlie the parasite’s reticulocyte-specific invasion tropism, although their cognate receptors

have not yet been identified.

Like P. vivax, P. falciparum expresses ligands from both the DBL and Rh families. Ligand

expression varies by strain, allowing strain-specific differences in host cell preference for

invasion139. The DBL-family proteins EBA-175 (for Erythrocyte Binding Antigen, 175 kDa),

EBA-140, and EBA-181 mediate interactions with sialylated receptors on the red blood cell

surface140. EBA-175 binding to glycophorin A is likely the dominant interaction, as deleting

EBA-175 from a sialic acid-dependent strain results in a switch to sialic acid-independent

pathways for invasion139. Structural data showing EBA-175 co-crystallized with sialyllactose

demonstrates glycan contacts with two DBL domains, suggesting that dimerization of EBA-175

is also important for receptor binding141. Additionally, inhibition studies with glycophorin A

peptides demonstrate that the glycophorin A protein backbone participates in binding, either by

direct contacts to EBA-175 or by maintaining a specific conformation of sialic acid residues140.

27

P. falciparum can also invade host red blood cells through sialic acid-independent pathways

mediated by Rh-family proteins. Of these, only the receptors for PfRh4 (complement receptor

1)142 and PfRh5 (basigin)143 have been identified. PfRh5 appears unique in that it cannot be

disrupted and has limited homology to other proteins in the Rh family, suggesting that it may

have an unrelated function144,145

After engaging receptors on the red cell surface through DBL and Rh-family proteins,

parasites then secrete additional rhoptry proteins into the membrane of the host red blood cell146.

These proteins (termed RONs, for their apparent origin in the rhoptry neck) form a complex that

provides a high-affinity anchor for the merozoite147. The merozoite protein AMA-1 (for Apical

Membrane Antigen-1) then binds to RON2 to form the tight junction and initiate invasion129.

1.6.3 Linking ligand-receptor interactions to invasion outcomes

While multiple ligand-receptor interactions have been identified, it has remained challenging

to link individual binding events to invasion outcomes. Although P. vivax expresses ligands that

preferentially bind reticulocytes, it is unclear whether this preferential binding is solely

responsible for the inability of P. vivax to efficiently invade mature red blood cells. Current

techniques to link ligand-receptor interactions with invasion outcomes rely on protease or

glycosylase treatments that cleave necessary receptors from the host cell surface, or inhibit

binding interactions with soluble competitors. These interventions are limited in their specificity:

protease and glycosylase treatments remove broad classes of receptors from the host cell surface,

and soluble competitors are often added at high concentrations in order to inhibit invasion.

Synthetic receptors represent a promising potential strategy for linking binding events to

invasion outcomes. Providing a specific receptor in trans and promoting invasion into an

28

otherwise-refractory cell type would provide conclusive evidence that particular ligand-receptor

interaction(s) are necessary and/or sufficient for a given invasion tropism. In the case of P.

falciparum, synthetic receptors could demonstrate causal links between engaging various DBL or

Rh proteins and strain-specific invasion preferences. In the case of P. vivax, synthetic receptors

could demonstrate whether engaging the reticulocyte-binding proteins is sufficient to promote

invasion into Duffy-positive mature red blood cells, and potentially facilitate culture system

development.

1.7 Summary of rationale and work presented

This thesis documents the development of novel chemical biology tools to address critical

needs in antimalarial drug discovery. Among these are validated molecular targets to guide drug

discovery efforts. Some of the most potent and successful antimalarial drugs are thought to

interact with parasite heme, although their mechanisms of action remain controversial.

Elucidating the mechanisms of action of these drugs, in the context of a broader understanding of

parasite heme metabolism, would identify validated targets for future drug development efforts.

Currently, heme metabolism and the action of heme-binding drugs are poorly understood

because heme is difficult to quantify in situ. Chapter 2 describes the development of a novel

genetically-encoded biosensor for quantifying labile heme in live cells, and applications in P.

falciparum to derive new insights about parasite heme metabolism and antimalarial drug action.

Another critical need is a method for in vitro propagation of blood-stage P. vivax, which is

currently impractical due to its preference for invading reticulocytes. The molecular basis for this

preference is incompletely understood, precluding culture system development. Chapter 3

describes a chemoenzymatic toolkit for displaying synthetic receptors on the surface of the red

blood cell, which can be used to link molecular interactions with invasion outcomes. Proof-of-

29

concept experiments in P. falciparum demonstrate the utility of this approach for elucidating

structural requirements in a well-understood invasion pathway, and define rules for developing

future synthetic receptor technologies geared towards facilitating in vitro culture of P. vivax

parasites in mature erythrocytes.

30

1.8 References

1. WHO Global Malaria Programme. World Malaria Report 2014. (World Health Organization).

2. Krotoski, W. A. et al. Demonstration of hypnozoites in sporozoite-transmitted Plasmodium vivax infection. Am J Trop Med Hyg 31, 1291–1293 (1982).

3. Goldberg, D. E. Hemoglobin degradation. Curr Top Microbiol Immunol 295, 275–291 (2005).

4. Gilson, P. R. & Crabb, B. S. Morphology and kinetics of the three distinct phases of red blood cell invasion by Plasmodium falciparum merozoites. Int J Parasitol 39, 91–96 (2009).

5. Brancucci, N. M. B. et al. Heterochromatin protein 1 secures survival and transmission of malaria parasites. Cell Host Microbe 16, 165–176 (2014).

6. Kafsack, B. F. C. et al. A transcriptional switch underlies commitment to sexual development in malaria parasites. Nature 507, 248–252 (2014).

7. Sinha, A. et al. A cascade of DNA-binding proteins for sexual commitment and development in Plasmodium. Nature 507, 253–257 (2014).

8. Ghosh, A., Edwards, M. J. & Jacobs-Lorena, M. The journey of the malaria parasite in the mosquito: hopes for the new century. Parasitol. Today (Regul. Ed.) 16, 196–201 (2000).

9. Miller, L. H., Baruch, D. I., Marsh, K. & Doumbo, O. K. The pathogenic basis of malaria. Nature 415, 673–679 (2002).

10. Francischetti, I. M. B., Seydel, K. B. & Monteiro, R. Q. Blood coagulation, inflammation, and malaria. Microcirculation 15, 81–107 (2008).

11. Moxon, C. A., Heyderman, R. S. & Wassmer, S. C. Dysregulation of coagulation in cerebral malaria. Mol Biochem Parasitol 166, 99–108 (2009).

12. Miller, L. H., Ackerman, H. C., Su, X.-Z. & Wellems, T. E. Malaria biology and disease pathogenesis: insights for new treatments. Nature medicine 19, 156–167 (2013).

13. Smith, J. D., Rowe, J. A., Higgins, M. K. & Lavstsen, T. Malaria's deadly grip: cytoadhesion of Plasmodium falciparum-infected erythrocytes. Cell Microbiol 15, 1976–1983 (2013).

14. Idro, R., Marsh, K., John, C. C. & Newton, C. R. J. Cerebral malaria: mechanisms of brain injury and strategies for improved neurocognitive outcome. Pediatr. Res. 68, 267–274 (2010).

15. Steketee, R. W., Nahlen, B. L., Parise, M. E. & Menendez, C. The burden of malaria in pregnancy in malaria-endemic areas. Am J Trop Med Hyg 64, 28–35 (2001).

16. McCormick, M. C. The contribution of low birth weight to infant mortality and childhood morbidity. N. Engl. J. Med. 312, 82–90 (1985).

17. da Silva, A. J. & Moser, M. Malaria. (2002). ID #3405 at <http://phil.cdc.gov/> 18. Liu, N. Insecticide resistance in mosquitoes: impact, mechanisms, and research

directions. Annu. Rev. Entomol. 60, 537–559 (2015). 19. Schwartz, L., Brown, G. V., Genton, B. & Moorthy, V. S. A review of malaria vaccine

clinical projects based on the WHO rainbow table. Malar J 11, 11 (2012). 20. Vekemans, J., Leach, A. & Cohen, J. Development of the RTS,S/AS malaria candidate

vaccine. Vaccine 27 Suppl 6, G67–71 (2009). 21. RTS,S Clinical Trials Partnership et al. A phase 3 trial of RTS,S/AS01 malaria vaccine

31

in African infants. N. Engl. J. Med. 367, 2284–2295 (2012). 22. Loeb, F. et al. ACTIVITY OF A NEW ANTIMALARIAL AGENT, CHLOROQUINE

(SN 7618): Statement Approved by the Board for Coordination of Malarial Studies. JAMA 130, 1069–1070 (1946).

23. Slater, A. F. Chloroquine: mechanism of drug action and resistance in Plasmodium falciparum. Pharmacol. Ther. 57, 203–235 (1993).

24. Sá, J. M., Chong, J. L. & Wellems, T. E. Malaria drug resistance: new observations and developments. Essays Biochem 51, 137–160 (2011).

25. Wellems, T. E. & Plowe, C. V. Chloroquine-resistant malaria. J Infect Dis 184, 770–776 (2001).

26. D'Alessandro, U. & Buttiëns, H. History and importance of antimalarial drug resistance. Trop. Med. Int. Health 6, 845–848 (2001).

27. Trape, J. F. et al. Impact of chloroquine resistance on malaria mortality. C. R. Acad. Sci. III, Sci. Vie 321, 689–697 (1998).

28. Marsh, K. Malaria disaster in Africa. Lancet 352, 924 (1998). 29. Wells, T. N. C., Alonso, P. L. & Gutteridge, W. E. New medicines to improve control

and contribute to the eradication of malaria. Nat Rev Drug Discov 8, 879–891 (2009). 30. Ecker, A., Lehane, A. M., Clain, J. & Fidock, D. A. PfCRT and its role in antimalarial

drug resistance. Trends Parasitol 28, 504–514 (2012). 31. Meshnick, S. R., Taylor, T. E. & Kamchonwongpaisan, S. Artemisinin and the

antimalarial endoperoxides: from herbal remedy to targeted chemotherapy. Microbiol. Rev. 60, 301–315 (1996).

32. Noedl, H. et al. Evidence of artemisinin-resistant malaria in western Cambodia. N. Engl. J. Med. 359, 2619–2620 (2008).

33. Ashley, E. A. et al. Spread of artemisinin resistance in Plasmodium falciparum malaria. N. Engl. J. Med. 371, 411–423 (2014).

34. malERA Consultative Group on Basic Science and Enabling Technologies. A research agenda for malaria eradication: basic science and enabling technologies. PLoS Medicine 8, e1000399 (2011).

35. Trager, W. & Jensen, J. B. Human malaria parasites in continuous culture. Science 193, 673–675 (1976).

36. Mueller, I. et al. Key gaps in the knowledge of Plasmodium vivax, a neglected human malaria parasite. The Lancet Infectious Diseases 9, 555–566 (2009).

37. Price, R. et al. Vivax Malaria: Neglected and Not Benign. American Journal of Tropical Medicine and Hygiene 77, 79 (2007).

38. Gardner, M. J. et al. Genome sequence of the human malaria parasite Plasmodium falciparum. Nature 419, 498–511 (2002).

39. Armstrong, C. M. & Goldberg, D. E. An FKBP destabilization domain modulates protein levels in Plasmodium falciparum. Nature Methods 4, 1007–1009 (2007).

40. Goldfless, S. J., Wagner, J. C. & Niles, J. C. Versatile control of Plasmodium falciparum gene expression with an inducible protein-RNA interaction. Nat Commun 5, 5329 (2014).

41. Wagner, J. C., Platt, R. J., Goldfless, S. J., Zhang, F. & Niles, J. C. Efficient CRISPR-Cas9-mediated genome editing in Plasmodium falciparum. Nature Methods 11, 915–918 (2014).

42. Wells, T. N. & Poll, E. M. When is enough enough? The need for a robust pipeline of high-quality antimalarials. Discov Med 9, 389–398 (2010).

32

43. Nwaka, S. & Hudson, A. Innovative lead discovery strategies for tropical diseases. Nat Rev Drug Discov 5, 941–955 (2006).

44. Gelb, M. H. Drug discovery for malaria: a very challenging and timely endeavor. Current Opinion in Chemical Biology 11, 440–445 (2007).

45. O'Neill, P. M. et al. Isoquine and related amodiaquine analogues: a new generation of improved 4-aminoquinoline antimalarials. J. Med. Chem. 46, 4933–4945 (2003).

46. Rosenthal, P. J. & Meshnick, S. R. Hemoglobin catabolism and iron utilization by malaria parasites. Mol Biochem Parasitol 83, 131–139 (1996).

47. Yayon, A., Timberg, R., Friedman, S. & Ginsburg, H. Effects of chloroquine on the feeding mechanism of the intraerythrocytic human malarial parasite Plasmodium falciparum. J. Protozool. 31, 367–372 (1984).

48. Becker, K. et al. Oxidative stress in malaria parasite-infected erythrocytes: host-parasite interactions. Int J Parasitol 34, 163–189 (2004).

49. Jeney, V. et al. Pro-oxidant and cytotoxic effects of circulating heme. Blood 100, 879–887 (2002).

50. Pagola, S., Stephens, P. W., Bohle, D. S., Kosar, A. D. & Madsen, S. K. The structure of malaria pigment beta-haematin. Nature 404, 307–310 (2000).

51. Egan, T. J. et al. Fate of haem iron in the malaria parasite Plasmodium falciparum. Biochem J 365, 343–347 (2002).

52. Dorn, A., Stoffel, R., Matile, H., Bubendorf, A. & Ridley, R. G. Malarial haemozoin/beta-haematin supports haem polymerization in the absence of protein. Nature 374, 269–271 (1995).

53. Sullivan, D. J., Gluzman, I. Y. & Goldberg, D. E. Plasmodium hemozoin formation mediated by histidine-rich proteins. Science 271, 219–222 (1996).

54. Jani, D. et al. HDP-a novel heme detoxification protein from the malaria parasite. PLoS Pathog 4, e1000053 (2008).

55. Sullivan, D. J. Theories on malarial pigment formation and quinoline action. Int J Parasitol 32, 1645–1653 (2002).

56. Bendrat, K., Berger, B. J. & Cerami, A. Haem polymerization in malaria. Nature 378, 138–139 (1995).

57. Hempelmann, E., Motta, C., Hughes, R., Ward, S. A. & Bray, P. G. Plasmodium falciparum: sacrificing membrane to grow crystals? Trends Parasitol 19, 23–26 (2003).

58. Sigala, P. A., Crowley, J. R., Hsieh, S., Henderson, J. P. & Goldberg, D. E. Direct Tests of Enzymatic Heme Degradation by the Malaria Parasite Plasmodium falciparum. J Biol Chem 287, 37793–37807 (2012).

59. Loria, P., Miller, S., Foley, M. & Tilley, L. Inhibition of the peroxidative degradation of haem as the basis of action of chloroquine and other quinoline antimalarials. Biochem J 339 ( Pt 2), 363–370 (1999).

60. Atamna, H. & Ginsburg, H. Heme degradation in the presence of glutathione. A proposed mechanism to account for the high levels of non-heme iron found in the membranes of hemoglobinopathic red blood cells. J Biol Chem 270, 24876–24883 (1995).

61. Sigala, P. A. & Goldberg, D. E. The peculiarities and paradoxes of Plasmodium heme metabolism. Annu. Rev. Microbiol. 68, 259–278 (2014).

62. Lim, L. & McFadden, G. I. The evolution, metabolism and functions of the apicoplast. Philos. Trans. R. Soc. Lond., B, Biol. Sci. 365, 749–763 (2010).

33

63. van Dooren, G. G., Kennedy, A. T. & Mcfadden, G. I. The use and abuse of heme in apicomplexan parasites. Antioxid Redox Signal 17, 634–656 (2012).

64. Ralph, S. A. et al. Tropical infectious diseases: metabolic maps and functions of the Plasmodium falciparum apicoplast. Nat. Rev. Microbiol. 2, 203–216 (2004).

65. Sato, S., Clough, B., Coates, L. & Wilson, R. J. M. Enzymes for heme biosynthesis are found in both the mitochondrion and plastid of the malaria parasite Plasmodium falciparum. Protist 155, 117–125 (2004).

66. Nagaraj, V. A. et al. Localisation of Plasmodium falciparum uroporphyrinogen III decarboxylase of the heme-biosynthetic pathway in the apicoplast and characterisation of its catalytic properties. Int J Parasitol 39, 559–568 (2009).

67. Nagaraj, V. A., Prasad, D., Arumugam, R., Rangarajan, P. N. & Padmanaban, G. Characterization of coproporphyrinogen III oxidase in Plasmodium falciparum cytosol. Parasitology International 59, 121–127 (2010).

68. Nagaraj, V. A., Arumugam, R., Prasad, D., Rangarajan, P. N. & Padmanaban, G. Protoporphyrinogen IX oxidase from Plasmodium falciparum is anaerobic and is localized to the mitochondrion. Mol Biochem Parasitol 174, 44–52 (2010).

69. Nagaraj, V. A., Prasad, D., Rangarajan, P. N. & Padmanaban, G. Mitochondrial localization of functional ferrochelatase from Plasmodium falciparum. Mol Biochem Parasitol 168, 109–112 (2009).

70. Severance, S. & Hamza, I. Trafficking of heme and porphyrins in metazoa. Chem Rev 109, 4596–4616 (2009).

71. Kořený, L., Oborník, M. & Lukeš, J. Make it, take it, or leave it: heme metabolism of parasites. PLoS Pathog 9, e1003088 (2013).

72. Kessl, J. J. et al. Molecular basis for atovaquone binding to the cytochrome bc1 complex. J Biol Chem 278, 31312–31318 (2003).

73. Srivastava, I. K., Rottenberg, H. & Vaidya, A. B. Atovaquone, a broad spectrum antiparasitic drug, collapses mitochondrial membrane potential in a malarial parasite. J Biol Chem 272, 3961–3966 (1997).

74. Srivastava, I. K., Morrisey, J. M., Darrouzet, E., Daldal, F. & Vaidya, A. B. Resistance mutations reveal the atovaquone-binding domain of cytochrome b in malaria parasites. Mol Microbiol 33, 704–711 (1999).

75. Painter, H. J., Morrisey, J. M., Mather, M. W. & Vaidya, A. B. Specific role of mitochondrial electron transport in blood-stage Plasmodium falciparum. Nature 446, 88–91 (2007).

76. Campanale, N. et al. Identification and characterization of heme-interacting proteins in the malaria parasite, Plasmodium falciparum. J Biol Chem 278, 27354–27361 (2003).

77. Nagaraj, V. A. et al. Malaria parasite-synthesized heme is essential in the mosquito and liver stages and complements host heme in the blood stages of infection. PLoS Pathog 9, e1003522 (2013).

78. Ke, H. et al. The heme biosynthesis pathway is essential for Plasmodium falciparum development in mosquito stage but not in blood stages. Journal of Biological Chemistry 289, 34827–34837 (2014).

79. Liu, S. C., Zhai, S. & Palek, J. Detection of hemin release during hemoglobin S denaturation. Blood 71, 1755–1758 (1988).

80. Sartorello, R. et al. In vivo uptake of a haem analogue Zn protoporphyrin IX by the human malaria parasite P. falciparum-infected red blood cells. Cell. Biol. Int. 34, 859–

34

865 (2010). 81. Brossi, A. et al. Arteether, a new antimalarial drug: synthesis and antimalarial properties.

J. Med. Chem. 31, 645–650 (1988). 82. Meshnick, S. R., Thomas, A., Ranz, A., Xu, C. M. & Pan, H. Z. Artemisinin

(qinghaosu): the role of intracellular hemin in its mechanism of antimalarial action. Mol Biochem Parasitol 49, 181–189 (1991).

83. Klonis, N. et al. Artemisinin activity against Plasmodium falciparum requires hemoglobin uptake and digestion. Proc Natl Acad Sci USA 108, 11405–11410 (2011).

84. Stocks, P. A. et al. Evidence for a common non-heme chelatable-iron-dependent activation mechanism for semisynthetic and synthetic endoperoxide antimalarial drugs. Angew Chem Int Ed Engl 46, 6278–6283 (2007).

85. Payne, D. Spread of chloroquine resistance in Plasmodium falciparum. Parasitol. Today (Regul. Ed.) 3, 241–246 (1987).

86. Ginsburg, H. & Geary, T. G. Current concepts and new ideas on the mechanism of action of quinoline-containing antimalarials. Biochem. Pharmacol. 36, 1567–1576 (1987).

87. Bachhawat, K., Thomas, C. J., Surolia, N. & Surolia, A. Interaction of chloroquine and its analogues with heme: An isothermal titration calorimetric study. Biochem Biophys Res Commun 276, 1075–1079 (2000).

88. Sullivan, D. J., Matile, H., Ridley, R. G. & Goldberg, D. E. A common mechanism for blockade of heme polymerization by antimalarial quinolines. J Biol Chem 273, 31103–31107 (1998).

89. Slater, A. F. & Cerami, A. Inhibition by chloroquine of a novel haem polymerase enzyme activity in malaria trophozoites. Nature 355, 167–169 (1992).

90. Zhang, J., Krugliak, M. & Ginsburg, H. The fate of ferriprotorphyrin IX in malaria infected erythrocytes in conjunction with the mode of action of antimalarial drugs. Mol Biochem Parasitol 99, 129–141 (1999).

91. Combrinck, J. M. et al. Insights into the role of heme in the mechanism of action of antimalarials. ACS Chem Biol 8, 133–137 (2013).

92. Königk, E. & Putfarken, B. Inhibition of ornithine decarboxylase of in vitro cultured Plasmodium falciparum by chloroquine. Tropenmed Parasitol 34, 1–3 (1983).

93. Surolia, N. & Padmanaban, G. Chloroquine inhibits heme-dependent protein synthesis in Plasmodium falciparum. Proc Natl Acad Sci USA 88, 4786–4790 (1991).

94. Gluzman, I. Y. et al. Order and specificity of the Plasmodium falciparum hemoglobin degradation pathway. J Clin Invest 93, 1602–1608 (1994).

95. Ginsburg, H., Famin, O., Zhang, J. & Krugliak, M. Inhibition of glutathione-dependent degradation of heme by chloroquine and amodiaquine as a possible basis for their antimalarial mode of action. Biochem. Pharmacol. 56, 1305–1313 (1998).

96. Gorka, A. P. et al. Cytostatic versus cytocidal activities of chloroquine analogues and inhibition of hemozoin crystal growth. Antimicrob. Agents Chemother. 57, 356–364 (2013).

97. Fidock, D. A. et al. Mutations in the P. falciparum digestive vacuole transmembrane protein PfCRT and evidence for their role in chloroquine resistance. Mol. Cell 6, 861–871 (2000).

98. Wellems, T. E., Walker-Jonah, A. & Panton, L. J. Genetic mapping of the chloroquine-resistance locus on Plasmodium falciparum chromosome 7. Proc Natl Acad Sci USA 88,

35

3382–3386 (1991). 99. Cooper, R. A. et al. Alternative mutations at position 76 of the vacuolar transmembrane

protein PfCRT are associated with chloroquine resistance and unique stereospecific quinine and quinidine responses in Plasmodium falciparum. Mol Pharmacol 61, 35–42 (2002).

100. Martin, R. E. & Kirk, K. The malaria parasite's chloroquine resistance transporter is a member of the drug/metabolite transporter superfamily. Mol. Biol. Evol. 21, 1938–1949 (2004).

101. Tran, C. V. & Saier, M. H. The principal chloroquine resistance protein of Plasmodium falciparum is a member of the drug/metabolite transporter superfamily. Microbiology (Reading, Engl.) 150, 1–3 (2004).

102. Chinappi, M., Via, A., Marcatili, P. & Tramontano, A. On the mechanism of chloroquine resistance in Plasmodium falciparum. PLoS ONE 5, e14064 (2010).

103. Saliba, K. J., Folb, P. I. & Smith, P. J. Role for the plasmodium falciparum digestive vacuole in chloroquine resistance. Biochem. Pharmacol. 56, 313–320 (1998).

104. Kitchen, S. The infection of reticulocytes by Plasmodium vivax. Am J Trop Med Hyg 1, 347 (1938).

105. Golenda, C. F., Li, J. & Rosenberg, R. Continuous in vitro propagation of the malaria parasite Plasmodium vivax. Proc Natl Acad Sci USA 94, 6786–6791 (1997).

106. Wilson, R. J., Pasvol, G. & Weatherall, D. J. Invasion and growth of Plasmodium falciparum in different types of human erythrocyte. Bull World Health Organ 55, 179–186 (1977).

107. Skadberg, O., Brun, A. & Sandberg, S. Human reticulocytes isolated from peripheral blood: maturation time and hemoglobin synthesis. Lab Hematol 9, 198–206 (2003).

108. Bass, C. C. & Johns, F. M. THE CULTIVATION OF MALARIAL PLASMODIA (PLASMODIUM VIVAX AND PLASMODIUM FALCIPARUM) IN VITRO. J Exp Med 16, 567–579 (1912).

109. Brockelman, C. R., Tan-Ariya, P. & Laovanitch, R. Observation on complete schizogony of Plasmodium vivax in vitro. J. Protozool. 32, 76–80 (1985).

110. Sutar, N. K. & Renapurkar, D. M. Effect of liver extract on growth of Plasmodium vivax in vitro. Indian J. Exp. Biol. 29, 286–287 (1991).

111. Zhou, W. Z. & Hu, L. Q. [Erythrocytic schizogony of Plasmodium vivax under various conditions of in vitro cultivation]. Zhongguo Ji Sheng Chong Xue Yu Ji Sheng Chong Bing Za Zhi 9, 258–260 (1991).

112. Chotivanich, K. et al. Ex-vivo short-term culture and developmental assessment of Plasmodium vivax. Trans. R. Soc. Trop. Med. Hyg. 95, 677–680 (2001).

113. Larrouy, G., Magnaval, J. F. & Moro, F. [Obtaining intraerythrocytic forms of Plasmodium vivax by in vitro culture]. Comptes rendus des séances de l'Académie des sciences. Série III, Sciences de la vie 292, 929–930 (1981).

114. Renapurkar, D. M., Pradhan, V. R. & Sutar, N. K. The continuous in vitro cultivation of Plasmodium vivax. IRCS Medical Science 11, 7–8 (1983).

115. Noulin, F., Borlon, C., Van Den Abbeele, J., D’Alessandro, U. & Erhart, A. 1912-2012: a century of research on Plasmodium vivax in vitro culture. Trends Parasitol 29, 286–294 (2013).

116. Borlon, C. et al. Cryopreserved Plasmodium vivax and cord blood reticulocytes can be used for invasion and short term culture. Int J Parasitol 42, 155–160 (2012).

36

117. Devi, C. U., Pillai, C. R., Subbarao, S. K. & Dwivedi, S. C. Short term in vitro cultivation of erythrocytic stages of Plasmodium vivax. Journal of Parasitic Diseases 24, 61–66 (2000).

118. Udomsangpetch, R. et al. Short-term in vitro culture of field isolates of Plasmodium vivax using umbilical cord blood. Parasitology International 56, 65–69 (2007).

119. Panichakul, T. et al. Production of erythropoietic cells in vitro for continuous culture of Plasmodium vivax. Int J Parasitol 37, 1551–1557 (2007).

120. Noulin, F. et al. Cryopreserved reticulocytes derived from hematopoietic stem cells can be invaded by cryopreserved Plasmodium vivax isolates. PLoS ONE 7, e40798 (2012).

121. Giarratana, M.-C. et al. Ex vivo generation of fully mature human red blood cells from hematopoietic stem cells. Nature Biotechnology 23, 69–74 (2005).

122. Lanners, H. N. Prolonged in vitro cultivation of Plasmodium vivax using Trager's continuous-flow method. Parasitol Res 78, 699–701 (1992).

123. Dvorak, J. A., Miller, L. H., Whitehouse, W. C. & Shiroishi, T. Invasion of erythrocytes by malaria merozoites. Science 187, 748–750 (1975).

124. Bannister, L. H. & Dluzewski, A. R. The ultrastructure of red cell invasion in malaria infections: a review. Blood Cells 16, 257–92– discussion 293–7 (1990).

125. Singh, S., Alam, M. M., Pal-Bhowmick, I., Brzostowski, J. A. & Chitnis, C. E. Distinct external signals trigger sequential release of apical organelles during erythrocyte invasion by malaria parasites. PLoS Pathog 6, e1000746 (2010).

126. Cowman, A. F., Berry, D. & Baum, J. The cellular and molecular basis for malaria parasite invasion of the human red blood cell. J Cell Biol 198, 961–971 (2012).

127. Mitchell, G. H. & Bannister, L. H. Malaria parasite invasion: interactions with the red cell membrane. Crit. Rev. Oncol. Hematol. 8, 225–310 (1988).

128. Aikawa, M., Miller, L. H., Johnson, J. & Rabbege, J. Erythrocyte entry by malarial parasites. A moving junction between erythrocyte and parasite. J Cell Biol 77, 72–82 (1978).

129. Srinivasan, P. et al. Binding of Plasmodium merozoite proteins RON2 and AMA1 triggers commitment to invasion. Proc Natl Acad Sci USA (2011).

130. Cowman, A. F. & Crabb, B. S. Invasion of red blood cells by malaria parasites. 124, 755–766 (2006).

131. Tham, W.-H., Healer, J. & Cowman, A. F. Erythrocyte and reticulocyte binding-like proteins of Plasmodium falciparum. Trends Parasitol 28, 23–30 (2012).

132. Adams, J. H. et al. The Duffy receptor family of Plasmodium knowlesi is located within the micronemes of invasive malaria merozoites. 63, 141–153 (1990).

133. Sim, B. K., Toyoshima, T., Haynes, J. D. & Aikawa, M. Localization of the 175-kilodalton erythrocyte binding antigen in micronemes of Plasmodium falciparum merozoites. Mol Biochem Parasitol 51, 157–159 (1992).

134. Rayner, J. C., Galinski, M. R., Ingravallo, P. & Barnwell, J. W. Two Plasmodium falciparum genes express merozoite proteins that are related to Plasmodium vivax and Plasmodium yoelii adhesive proteins involved in host cell selection and invasion. Proc Natl Acad Sci USA 97, 9648–9653 (2000).

135. Adams, J. H. et al. A family of erythrocyte binding proteins of malaria parasites. Proc Natl Acad Sci USA 89, 7085–7089 (1992).

136. Horuk, R. et al. A receptor for the malarial parasite Plasmodium vivax: the erythrocyte chemokine receptor. Science 261, 1182–1184 (1993).

37

137. Galinski, M. R., Medina, C. C., Ingravallo, P. & Barnwell, J. W. A reticulocyte-binding protein complex of Plasmodium vivax merozoites. 69, 1213–1226 (1992).

138. Triglia, T. et al. Identification of proteins from Plasmodium falciparum that are homologous to reticulocyte binding proteins in Plasmodium vivax. Infect Immun 69, 1084–1092 (2001).

139. Duraisingh, M. T., Maier, A. G., Triglia, T. & Cowman, A. F. Erythrocyte-binding antigen 175 mediates invasion in Plasmodium falciparum utilizing sialic acid-dependent and -independent pathways. Proc Natl Acad Sci USA 100, 4796–4801 (2003).

140. Sim, B. K., Chitnis, C. E., Wasniowska, K., Hadley, T. J. & Miller, L. H. Receptor and ligand domains for invasion of erythrocytes by Plasmodium falciparum. Science 264, 1941–1944 (1994).

141. Tolia, N. H., Enemark, E. J., Sim, B. K. L. & Joshua-Tor, L. Structural basis for the EBA-175 erythrocyte invasion pathway of the malaria parasite Plasmodium falciparum. 122, 183–193 (2005).

142. Tham, W.-H. et al. Complement receptor 1 is the host erythrocyte receptor for Plasmodium falciparum PfRh4 invasion ligand. Proc Natl Acad Sci USA 107, 17327–17332 (2010).

143. Crosnier, C. et al. Basigin is a receptor essential for erythrocyte invasion by Plasmodium falciparum. Nature 480, 534–537 (2011).

144. Hayton, K. et al. Erythrocyte binding protein PfRH5 polymorphisms determine species-specific pathways of Plasmodium falciparum invasion. Cell Host Microbe 4, 40–51 (2008).

145. Baum, J. et al. Reticulocyte-binding protein homologue 5 - an essential adhesin involved in invasion of human erythrocytes by Plasmodium falciparum. Int J Parasitol 39, 371–380 (2009).

146. Besteiro, S., Michelin, A., Poncet, J., Dubremetz, J.-F. & Lebrun, M. Export of a Toxoplasma gondii rhoptry neck protein complex at the host cell membrane to form the moving junction during invasion. PLoS Pathog 5, e1000309 (2009).

147. Alexander, D. L., Mital, J., Ward, G. E., Bradley, P. & Boothroyd, J. C. Identification of the moving junction complex of Toxoplasma gondii: a collaboration between distinct secretory organelles. PLoS Pathog 1, e17 (2005).

38

CHAPTER 2: DEVELOPMENT OF A NOVEL GENETICALLY-ENCODED

FRET BIOSENSOR AND QUANTIFICATION OF A LABILE CYTOSOLIC

HEME POOL IN LIVE MALARIA PARASITES

2.0 Note

This chapter is adapted from a manuscript submitted for publication, with Christopher J.

Rowlands, Suresh M. Ganesan, Peter T. C. So, and Jacquin C. Niles as co-authors.

2.1 Abstract

Heme is ubiquitous, yet little is known about the maintenance of labile pools of this cofactor

that ensure its timely bioavailability for proper cell function. Quantitative analysis of labile heme

is of broad fundamental importance to understanding how nature preserves access to the diverse

chemistry heme enables, while minimizing cellular damage caused by its redox-activity. Here,

we have developed a novel, genetically-encoded FRET sensor for quantifying labile heme in

intact cells, and measured the physiologic cytosolic heme pool in the malarial parasite,

Plasmodium falciparum. Our findings indicate that a labile heme pool (~1.1 µM) is stably

maintained throughout parasite development within red blood cells, even during a period

coincident with extensive hemoglobin degradation by the parasite. We also find that the heme-

binding antimalarial drug chloroquine specifically increases labile cytosolic heme, indicative of

homeostatic dysregulation of this pool that may directly relate to the antimalarial activity of this

drug class. We propose that application of this technology in other organisms could similarly

yield new, quantitative insights into fundamental heme biology.

39

2.2 Introduction

Heme is a cofactor of central importance across biology, and plays vital roles in diverse

processes including energy production, oxygen transport, gas sensing and signaling 1 and

catalysis 2. Its inherently high and tunable redox potential together with its diverse ligand-

binding properties make it an extremely versatile cofactor suited to a broad range of chemistries.

Free heme redox cycles in the aerobic and reducing cellular environment, which can induce

potentially cytotoxic oxidative stress. To minimize this, both heme levels and reactivity are

restricted in several ways, including sequestering it into protein scaffolds that determine the

selectivity and specificity of its chemistry, degradation, export and inactivation by physical

processes such as polymerization 2-4. Cells maintain labile pools of critical cofactors to meet

rapidly changing metabolic demands. Such pools for transition metal cofactors including iron

and zinc, which can also be cytotoxic, have been quantitatively defined using an extensive toolkit

5,6. However, similar and generally accessible tools for studying labile heme pools in live cells

have not previously been established. This has precluded achieving a detailed and quantitative

understanding of cellular heme pool composition and dynamics under both physiologic and

perturbed states.

We have been particularly interested in characterizing labile heme pools in the human

malarial parasite, Plasmodium falciparum. This pathogen is a major cause of the 198 million

cases and 584,000 deaths per year due to malaria 7. Several aspects of heme metabolism in P.

falciparum are counterintuitive, and its exquisite sensitivity to heme-interacting antimalarial

drugs suggests a critical and finely balanced role for heme in its biology. During development

within red blood cells (RBCs), P. falciparum takes up and digests between 44-80% of the

hemoglobin in a specialized subcellular digestive vacuole (DV) to release peptides and heme 8-11.

40

The majority of this heme is converted into the relatively redox inert crystalline hemozoin

polymer within the DV 8,10. While the extent of hemoglobin digestion and heme polymerization

is minimal in early stage parasites (rings), this progressively increases as parasites develop

through mid- (trophozoite) and late- (schizont) stages.

It is presently unknown whether hemoglobin-derived heme is quantitatively converted into

hemozoin and exclusively confined to the DV, or whether it escapes the DV to accumulate in

other compartments such as the parasite cytoplasm during normal development. Such a heme

pool may be important for meeting metabolic needs, signaling to coordinate DV biochemistry

with cytosolic and nuclear processes, or simply a consequence the parasite must endure due to its

obligate degradation of hemoglobin. Along these lines, despite liberating large quantities of

heme from hemoglobin that should be more than adequate to meet the parasite’s needs, the P.

falciparum genome encodes a complete heme biosynthetic pathway that appears to be active in

blood stage parasites 12-14. Nevertheless, de novo heme biosynthesis is dispensable during the

blood stage infection, as the genes encoding δ-aminolevulinic acid synthase (ALAS) and

ferrochelatase that are required for de novo heme biosynthesis can be deleted without observable

defects in parasite growth 13,15. Based on these studies, it has been suggested that hemoglobin-

derived heme may escape the DV to completely meet the parasite’s heme requirement. However,

the physiologic levels of bioavailable heme, irrespective of its source, are yet to be defined.

Further highlighting the importance of heme biochemistry in the parasite is the potent

antimalarial activity of chloroquine, an exemplar of the heme-binding 4-aminoquinoline drug

class. These compounds accumulate within the parasite’s DV to disrupt hemozoin formation, and

the unpolymerized heme is proposed to escape the DV to cause toxicity 10. Consistent with this,

electron spectroscopic imaging of fixed, chloroquine-treated parasites revealed a qualitative

41

increase in cytosolic iron content, suggestive of increased heme content in the cytoplasm 16.

However, heme can be degraded in a glutathione-dependent manner to release iron 17, the extent

of which cannot be inferred from the data. Fractionation studies on chloroquine-treated parasites

also support an increase in labile heme, but its precise subcellular distribution cannot be inferred

16. Thus, direct and quantitative evidence of cytosolic heme accumulation in chloroquine-treated

parasites is still lacking, despite the central importance of this knowledge to understanding the

mechanism of action of arguably the most successful antimalarial drug class used to date.

Here, we have addressed the fundamental challenge of directly quantifying labile heme in

live cells by systematically developing, validating, and optimizing a genetically-encoded FRET-

based heme biosensor. Using the optimized biosensor, we demonstrate for the first time that P.

falciparum maintains a labile cytosolic heme pool throughout its blood-stage development.

Furthermore, we directly show that disrupting heme sequestration in the digestive vacuole using

a heme-binding antimalarial drug causes a significant increase in the concentration of cytosolic

labile heme, thus directly linking chloroquine to cell-wide heme perturbation for the first time.

We believe that this novel biosensor will be broadly useful for directly interrogating heme

biology in P. falciparum, and in other organisms.

42

2.3 Methods

2.3.1 Molecular cloning

A fragment of the P. falciparum HRP2 gene (109-916 bp) lacking the N-terminal signal

peptide was amplified by PCR from plasmid MRA-67 (ATCC/MR4) and cloned into pET28b

(Novagen) vectors containing ECFP and EYFP using standard restriction and ligation techniques.

To generate sensors based on truncated PfHRP2, a forward oligonucleotide primer was designed

to anneal to a repeated PfHRP2 sequence motif identified using MEME 42 was used with a fixed

reverse primer to PCR amplify fragments of varying sizes using the PfHRP2 gene as a template.

These were separated by agarose gel electrophoresis, and then cloned into ECFP and EYFP-

containing vectors as above. Fragments mapped to full-length PfHRP2 except for some minor in-

frame insertion/deletion mutations in the histidine-rich repeats. For CH18Y, the oligonucleotide

encoding the heme-binding motif (HHAHHAADA)2 was generated in a Klenow reaction, and

cloned as above. For expression in P. falciparum, coding sequences from pET28b-based vectors

were PCR amplified and cloned using the Gibson Assembly Master Mix (New England Biolabs)

to replace the ENR-GFP fusion protein in plasmid MRA-846 (ATCC/MR4).

2.3.2 Protein expression and purification

Plasmids encoding FRET sensor and control proteins were transformed into E. coli strain

BL21(DE3). Cultures were grown overnight at 37˚C in ZYM-505 media supplemented with

kanamycin. Saturated cultures were diluted 1:200 in kanamycin-containing ZYM-505 media and

grown at 37˚C until OD600 = 0.6-0.8. Protein expression was induced by adding 0.1 mM

isopropyl β-D-1-thiogalactopyranoside (IPTG) for 24 hours at room temperature. Cells were

harvested by centrifugation and pellets were frozen and stored at -80˚C. For protein purification,

43

cell pellets were thawed at room temperature and lysed with B-PER II bacterial protein

extraction reagent (Thermo Scientific) supplemented with lysozyme, Benzonase (Novagen), and

protease inhibitor cocktail (Sigma-Aldrich). Lysates were clarified by centrifugation, and applied

to purification beads that had been previously washed with Equilibration/Wash Buffer (50 mM

Tris-HCl pH 8.0, 200 mM NaCl, 5% glycerol). Hexahistidine-tagged protein constructs were

bound to HisPur Cobalt Resin (Thermo Scientific), while Strep-Tactin Superflow Plus (Qiagen)

was used to bind Strep-tagged proteins. In both cases, lysates were incubated with purification

beads for 1 hour at 4˚C with gentle agitation. Purification beads were then washed extensively

with ice-cold Equilibration/Wash buffer, before loading on to a gravity-flow column.

Hexahistidine tagged proteins were eluted from the column with Elution Buffer H (50 mM Tris-

HCl pH 8.0, 200 mM NaCl, 500 mM imidazole, 5% glycerol). Strep-tagged proteins were eluted

from the column with Elution Buffer S (50 mM Tris-HCl pH 8.0, 200 mM NaCl, 2.5 mM

desthiobiotin, 5% glycerol). Fractions were spectroscopically monitored using a NanoDrop

spectrophotometer (Thermo Scientific), and those containing ECFP and/or EYFP were pooled

and concentrated using Amicon Ultra-15 centrifugal filters (Thermo Scientific). Concentrated

protein solutions were then dialyzed against 2x PBS pH 7.4. Glycerol was then added to 50%

and protein solutions were stored at -20˚C.

2.3.3 Absorbance titrations

Titrations to measure heme binding affinity and stoichiometry were performed as previously

described 18,21. Stock solutions of recombinant protein were prepared in 100 mM HEPES-KOH

pH 7.0. Stock solutions of 1 mM hemin were (Sigma-Aldrich) prepared in DMSO. All

concentrations were verified spectrophotometrically. Heme binding titrations were performed in

44

3 ml-capacity quartz cuvettes at 37˚C with stirring using a Cary 100 Bio Spectrophotometer

(Varian). Heme titrations were performed with 2 ml of 0.5 µM protein solution, using 2 ml of

HEPES-KOH as a reference. For each concentration, heme was added to both the sample and

reference cells and stirred for 5 min before difference spectra were measured. Heme binding was

quantified based on a differential absorption peak at 416 nm. The ∆A416nm versus heme

concentration data were plotted and analyzed using Prism (GraphPad Software, La Jolla, CA).

2.3.4 Fluorescence titrations and FRET efficiency calculations

Hemin stock solution was serially diluted in HEPES-KOH in a 96-well plate. Protein stock

solution was added to 0.5 µM final concentration using a multichannel pipetter, and the plate was

incubated at 37˚C for 5 min. Fluorescence spectra were measured using a Fluoromax-3

fluorimeter (Horiba Jobin Yvon). ECFP was excited at 420 nm, with emission scanning from

440-600 nm. EYFP was excited at 500 nm, with emission scanning from 505-600 nm. FRET

efficiencies were calculated using the ratioA method as previously described 43,44.

2.3.5 Fluorescence lifetime spectroscopy

The multiphoton FLIM microscope consisted of a Ti:Sapphire laser (Tsunami HP, Spectra

Physics, Santa Clara CA, USA), tunable between 780 nm and 880 nm (Supplemental Figure S2).

Power control was achieved using a half waveplate and Glan-Laser polarizer. The beam first

struck a tilted glass coverslip beamsplitter where part of the beam was focused onto a photodiode

in order to create a reference pulse. The majority of the pulse intensity passed through the

beamsplitter to the rest of the microscope. The beam was subsequently reflected off a dichroic

beamsplitter (675DCSX, Chroma Technology Inc., Brattleboro VT, USA) and two

45

galvanometric scanning mirrors (6350, Cambridge Technology, Watertown MA, USA) before

entering the scan lens, which produced a moving focal spot at the image plane of the microscope

(Axiovert 100TV, Zeiss, Göttingen, Germany). The scanning mirrors were controlled using

software written in-house, and the synchronization signals were created using an FPGA (Spartan

XCS30, Xilinx, San Jose CA, USA).

The spot in the image plane was imaged onto the sample by the microscope objective (C-

Apochromat 40× water immersion 44-00-52, Zeiss, Göttingen, Germany), and the resulting

fluorescence emission was captured by the objective, descanned by the scanning mirrors and

passed through the dichroic mirror and filter (ESP650, Chroma Technology, Brattleboro VT,

USA) where it was focused onto a photomultiplier tube (R7400P, Hamamatsu, Bridgewater NY,

USA). The signal from the photomultiplier and the signal from the reference photodiode were

measured by a Time-Correlated Single Photon Counting card (SPC-730, Becker and Hickl,

Berlin, Germany) and the resulting image displayed using the Becker and Hickl software.

The instrument response was compensated for by measuring a sample (Fluorescein in pH 9

DMSO solution, from R14782 reference sample kit, Life Technologies, Grand Island NY, USA,

known lifetime of 4.1 ns) and deconvolving the known decay curve from the measured curve to

yield the instrument response. This instrument response was then deconvolved from every

measured curve to yield a corrected decay curve. After this correction, the exponentially-

decaying section of the curve was taken from the data, and a ‘single exponential decay with

unknown offset’ function was fitted to the data to recover the fluorescence lifetime while

compensating for the small background in the measurement. All calculations were performed

using MATLAB 2011b (MathWorks, Natick MA).

46

2.3.6 Malaria parasite culture

Blood-stage malaria parasites were cultured at 2% hematocrit in 5% O2 and 5% CO2 in RPMI

1640 Medium supplemented with 5 g/l AlbuMAX II (Life Technologies), 2 g/l NaHCO3, 25 mM

HEPES-KOH pH 7.4, 1 mM hypoxanthine, and 50 mg/l gentamicin. Strains were synchronized

using a solution of 0.3 M alanine supplemented with 25 mM HEPES-KOH pH 7.4. Parasite

transfections were performed by pre-loading red blood cells with plasmid DNA by

electroporation. All expression constructs were integrated at the cg6 locus in NF54attB parasites 26

by co-transfecting with the pINT plasmid (MRA-847) 45. For each transfection, 100 µl of washed

red blood cells was mixed with 25-50 µg plasmid DNA and then electroporated with 8 x 1 ms

square pulses at 365 V. Late-stage parasites were then split to 0.1% parasitemia using half of the

loaded RBCs. After 48 hours, transfected cultures were split 1:2 using the remainder of the

loaded RBCs. After another 48 hours, cultures were split again and drug selection was initiated.

Drug-resistant parasites were then cloned via limiting dilution, and clones were screened for

expression of ECFP and/or EYFP using flow cytometry.

2.3.7 Preparation of giant multilamellar vesicles

Stock solutions of 1,2-dioleoyl-sn-glycero-3-phospho-(1'-rac-glycerol) (DOPG) and 1,2-

dioleoyl-sn-glycero-3-phosphocholine (DOPC, Avanti Polar Lipids) in chloroform were mixed

in a 1:1 ratio, and 1 µmol of total lipid was added to a glass scintillation vial. Lipid films were

deposited by overnight evaporation of the chloroform at room temperature. Lipid films were then

hydrated by incubating scintillation vials in a humidified 70˚C oven for 6 hours. Vesicles were

prepared by gently adding protein solutions (0.5 µM protein in 100 mM HEPES-KOH pH 7 plus

47

50 mM sucrose) and incubating overnight at room temperature in the dark. After incubation,

vesicles were washed extensively in HEPES-KOH plus 50 mM glucose and imaged.

2.3.8 In situ FRET analysis

Synchronized late-stage parasite cultures were washed and resuspended in Opti-Klear media

(Marker Gene Technologies) at 0.05% hematocrit for imaging. Culture suspensions were added

to glass-bottom 24-well plates (In vitro Scientific, Sunnyvale, CA) pre-treated with 0.1%

polyethyleneimine. Cultures were imaged using a Nikon Ti microscope using the following

filtersets: ECFP (436nm/20nm EX, 455nm LP, 480nm/40nm EM), FRET (436nm/20nm EX,

455nm LP, 535nm/30nm EM), and YFP (500nm/20nm EX, 515nm LP, 535nm/30nm EM).

Images were acquired using an Andor iXon+ 897 EMCCD camera and MetaMorph acquisition

software (Molecular Devices). Images were processed using Biosensor Processing Software v2.1

46. Briefly, images were shade-corrected using averaged reference images, and segmentation

values were chosen for each channel by manual inspection in order to draw background masks,

which were then grown using a 5-pixel radius. Background subtraction was performed according

to software defaults. The FRET channel was corrected for bleedthrough from CFP and crosstalk

from YFP using correction factors of 0.4 and 0.1, respectively. FRET images were obtained by

calculating the ratio between the corrected FRET channel image and the YFP image. The

resulting FRET images were inspected manually to identify individual cells. Average per-cell

FRET efficiencies for CH49Y and CSY strains were calculated and tabulated using the FIJI

distribution of ImageJ 47. A bootstrapping algorithm with 10,000 iterations was used to estimate

the ratio of averages between CH49Y and CSY FRET distributions, correct for the offset

measured between microscopy and fluorimetry data, and compute a 95% confidence interval.

48

The ratio of averages and the confidence interval bounds were used to calculate heme

concentrations according to the following empirical relationship between normalized EFRETSensor

and heme (Figs. 2-14a and 2-14b):

Normalized EFRETSensor =  

!FRET,    CH49Y!FRET,  CSY

= !.!"!!!!.!"# !"#"

!.!"#!!.!"#$ !"#"

Calculations were performed using MATLAB R2013a software (MathWorks, Natick MA).

2.3.9 Western immunoblotting

Parasites were obtained by lysing 180 µl infected red blood cells with an ice-cold solution of

0.1% saponin (Fluka) in PBS and incubating on ice until solutions cleared. Parasite pellets were

washed extensively with saponin-PBS solution to remove residual hemoglobin and then lysed

with the addition of 200 µl 1X SDS-Urea sample buffer (40 mM Tris base, 80 mM Gly-Gly, 40

mM dithiothreitol, 1.6% SDS and 6.4 M urea adjusted to pH 6.8 with HCl). Parasite lysates were

diluted 1:10 before gel electrophoresis and transfer. Blots were probed with a mouse monoclonal

anti-GFP (B-2) primary antibody (Catalog # SC-9996; Santa Cruz Biotech) diluted 1:2000, and a

goat anti-mouse secondary antibody conjugated to horseradish peroxidase (H+L) (Catalog #

71045; Novagen), and visualized using SuperSignal West Femto chemiluminescent substrate

(Thermo Scientific).

49

2.4 Results

2.4.1 Design and characterization of initial FRET-based heme biosensor

We selected the enhanced cyan and yellow fluorescent proteins (ECFP and EYFP) as the

FRET donor-acceptor pair, and P. falciparum histidine rich protein 2 (PfHRP2) as the heme-

binding domain to use for our FRET sensor design. For this initial sensor design, we created a

construct (CHY) in which PfHRP2 was flanked by ECFP and EYFP. We also made a control

construct (CSY) that should exhibit heme-independent and constitutive FRET by substituting the

PfHRP2 for a (Gly4Ser)3 peptide spacer (Fig. 2-1).

Figure 2-1. Schematic of CHY heme sensor and CSY non-sensing control.

We chose PfHRP2 as our heme sensor domain. This protein has previously been shown to

bind ~15-18 heme molecules/monomer with modest (~0.3 µM) apparent affinity 18,19. This

minimizes the potential for a sensor based on this protein to function as a heme sink in cells.

Additionally, PfHRP2 is non-essential to the parasite, and is predominantly trafficked to the

infected red blood cell cytoplasm and DV to a lesser extent 19,20. This suggests that PfHRP2 does

not play an integral role in the parasite’s cytosolic or other subcellular compartments, and that

FRET Donor (ECFP)

FRET Acceptor(EYFP)

His6

FRET Donor (ECFP)

FRET Acceptor(EYFP)ST2

CHY

CSY

Heme-binding domain

(PfHRP2)

Non-binding linker

(Gly4Ser)3

a

-0.2

0

0.2

0.4

0.6

350 400 450 500 550 600 650 700

Wavelength (nm)

hemeequivalentsadded

b c

d

∆Abs

fe

CHYCSYECFPEYFP

[Heme] (µM)

0 2 4 6 8 10 120.0

0.2

0.4

0.6

∆Abs

41

6

1.0

1.5

2.0

2.5

3.0

CHY

ECFP

CSY

CH

[Heme] (µM)

0 2 4 6 8 10 12

τ (

ns)

hemeequivalentsadded

[Heme] (µM)

No

rma

lize

d E

FR

ET

440

Norm

aliz

ed F

luore

scence

Wavelength (nm)

480 520 560 600

1.2

1.0

0.8

0.6

0.4

0.2

0.00 2 4 6 8 10 12

1.2

1.0

0.8

0.6

0.4

0.2

0.0

CHY

CSY

ECFP

50

expression of PfHRP2 fusion proteins in the cytosol should minimally impinge on its physiologic

role.

We recombinantly expressed and purified both CHY and CSY to characterize their

biochemical and spectroscopic properties in vitro (Fig. 2-2).

Figure 2-2. SDS-PAGE purity analysis of recombinant CHY, CSY, ECFP, EYFP, and CH used for fluorescence lifetime spectroscopy.

We first tested whether flanking PfHRP2 by ECFP and EYFP interfered with its heme

binding properties. Heme binding to PfHRP2 occurs via bis-histidyl ligation, which causes a

shift in the heme Soret peak absorbance from ~396 nm to 416 nm. Monitoring this change by

electronic absorption spectroscopy while titrating heme levels facilitates determination of heme

binding stoichiometry and apparent heme binding affinity to PfHRP2 18,21. Heme titrations with

CHY produced the expected increase in absorbance at 416 nm, consistent with heme binding to

PfHRP2 (Fig. 2-3a). Analysis of these binding data by breakpoint detection and fitting to a

ligand binding with depletion model 21 (Fig. 2-3b), respectively, revealed that CHY bound ~15

heme equivalents/monomer with an apparent KD ~0.25 µM. These data are in good agreement

with previously published data obtained for recombinant PfHRP2 18,19, indicating that fusion to

ECFP and EYFP does not significantly alter its heme binding properties. Furthermore, no heme

S1

CHY

CSY

ECFP

EYFP

CH

51

binding to ECFP, EYFP and CSY could be detected in similar heme titration experiments (Fig.

2-3b), indicating that PfHRP2 accounts for all detectable heme binding to CHY.

Figure 2-3. (a) Difference absorption spectrum of CHY during titration with heme. (b) Heme binding isotherms based on ∆A416nm Soret peak absorbance for CHY (blue circles), CSY (red triangles), ECFP (cyan inverted triangles) and EYFP (yellow diamonds). The solid line indicates the best fit for the CHY data using a single-site binding model and accounting for ligand depletion.

Next, we assessed the FRET properties of CHY and CSY. Based on earlier studies with

model FRET constructs of the design CFP-linker-YFP 22, we expected that CHY and CSY should

produce efficient FRET in the absence of heme. Indeed, for both CHY and CSY, direct ECFP

excitation at 420 nm produced emission spectra with maxima at 475 nm (ECFP emission) and

525 nm (EYFP emission) (Fig. 2-4a). The latter emission peak is indicative of FRET between

the ECFP donor and EYFP acceptor. Further supporting this interpretation, when a 1:1 mixture

of ECFP and EYFP was excited at 420 nm, only the characteristic ECFP emission spectrum with

a maximum at 475 nm but no emission maximum at 525 nm was detected. This indicates that

interaction of the donor and acceptor pair in trans is insufficient to produce the observed FRET.

We then titrated CHY and CSY with heme while exciting at 420 nm. For CHY, detected FRET

sharply decreased upon adding heme, and at 10 µM heme, the FRET signal was almost

FRET Donor (ECFP)

FRET Acceptor(EYFP)

His6

FRET Donor (ECFP)

FRET Acceptor(EYFP)ST2

CHY

CSY

Heme-binding domain

(PfHRP2)

Non-binding linker

(Gly4Ser)3 -0.2

0

0.2

0.4

0.6

350 400 450 500 550 600 650 700

Wavelength (nm)

hemeequivalentsadded

a b

a

∆Abs

b

CHYCSYECFPEYFP

[Heme] (µM)

0 2 4 6 8 10 120.0

0.2

0.4

0.6

∆Abs

41

6

1.0

1.5

2.0

2.5

3.0

CHY

ECFP

CSY

CH

[Heme] (µM)

0 2 4 6 8 10 12

τ (

ns)

hemeequivalentsadded

[Heme] (µM)

Norm

aliz

ed E

FR

ET

440

No

rma

lize

d F

luo

resc

en

ce

Wavelength (nm)

480 520 560 600

1.2

1.0

0.8

0.6

0.4

0.2

0.00 2 4 6 8 10 12

1.2

1.0

0.8

0.6

0.4

0.2

0.0

CHY

CSY

ECFP

52

completely abrogated (Fig. 2-4a,b). For CSY, only a modest decrease in FRET was detected

upon titrating heme (Fig. 2-4a,b). These data are consistent with CHY functioning as a ‘turn off’

sensor that responds to heme binding to the PfHRP2 domain.

Figure 2-4. (a) Normalized fluorescence intensity spectra for CHY titrated with heme. (b) Normalized FRET efficiency for CHY (blue circles) and CSY (red triangles) fitted to a single-exponential decay model (blue solid line) and a line (red dashed line), respectively.

To gain some insight into why CHY functioned as a ‘turn off’ heme sensor, we compared

heme-dependent ECFP fluorescence lifetimes for ECFP, CSY, CHY, and ECFP-PfHRP2 (CH)

(Figs. 2-2, 2-5, and 2-6).

FRET Donor (ECFP)

FRET Acceptor(EYFP)

His6

FRET Donor (ECFP)

FRET Acceptor(EYFP)ST2

CHY

CSY

Heme-binding domain

(PfHRP2)

Non-binding linker

(Gly4Ser)3 -0.2

0

0.2

0.4

0.6

350 400 450 500 550 600 650 700

Wavelength (nm)

hemeequivalentsadded

a b

a

∆Abs

b

CHYCSYECFPEYFP

[Heme] (µM)

0 2 4 6 8 10 120.0

0.2

0.4

0.6

∆Abs

41

6

1.0

1.5

2.0

2.5

3.0

CHY

ECFP

CSY

CH

[Heme] (µM)

0 2 4 6 8 10 12

τ (

ns)

hemeequivalentsadded

[Heme] (µM)N

orm

aliz

ed E

FR

ET

440

No

rma

lize

d F

luo

resc

en

ce

Wavelength (nm)

480 520 560 600

1.2

1.0

0.8

0.6

0.4

0.2

0.00 2 4 6 8 10 12

1.2

1.0

0.8

0.6

0.4

0.2

0.0

CHY

CSY

ECFP

53

Figure 2-5. Diagram of the fluorescence lifetime instrumentation setup.

ECFP fluorescence lifetime in both CHY and CH decreased upon titrating heme, but was

unchanged for ECFP and CSY (Fig. 2-6). With no heme bound, PfHRP2 exists as a random coil,

but undergoes a conformational change to adopt 310-helical structure upon heme binding 18. A

transition from random coil to a more rigid helical structure that physically separates the donor-

acceptor pair could account for a decrease in FRET efficiency. However, this should be

accompanied by an increase in ECFP lifetime upon titration with heme, rather than the observed

decrease. Therefore, this cannot be the dominant mechanism underlying the observed heme-

dependent change in FRET. Alternatively, heme-dependent changes in both CH and CHY FRET

could be due to either dynamic or static quenching of the emitted ECFP or EYFP fluorescence by

heme bound to PfHRP2. This mechanism is consistent with the decrease in ECFP lifetime

observed in our experiments upon heme binding, and is supported by an earlier report showing

that irreversible heme binding to a cytochrome b562-GFP fusion strongly quenched GFP

fluorescence 23.

54

Figure 2-6. Dependence of ECFP fluorescence lifetime on heme concentration for the CHY (blue circles), CH (green squares), ECFP (cyan inverted triangles) and CSY (red triangles) constructs.

2.4.2 Optimization of initial heme biosensor design

Having established that inserting PfHRP2 between ECFP and EYFP produces a heme-

dependent FRET sensor, we sought to improve this design by defining a minimal PfHRP2

fragment that preserves heme binding while maximizing heme-dependent changes in FRET.

Based on model studies, these parameters may be simultaneously optimized by selecting shorter

linkers between the donor and acceptor fluorophores 22. We also reasoned that reducing the heme

binding capacity of our sensor together with its inherently modest binding affinity would

minimize the potential for the sensor to act as a heme sink, which could potentially interfere with

both heme physiology and the ability of the sensor to respond to changes in labile heme

concentration.

We made a mini-library of PfHRP2 fragments containing variable numbers of heme-binding

motifs 24,25, and cloned these between ECFP and EYFP to create several sensors. These are

annotated as CHxY, where x is the number of PfHRP2-derived amino acids making up the heme

binding domain (Figs. 2-7 and 2-8).

FRET Donor (ECFP)

FRET Acceptor(EYFP)

His6

FRET Donor (ECFP)

FRET Acceptor(EYFP)ST2

CHY

CSY

Heme-binding domain

(PfHRP2)

Non-binding linker

(Gly4Ser)3 -0.2

0

0.2

0.4

0.6

350 400 450 500 550 600 650 700

Wavelength (nm)

hemeequivalentsadded

a b

a

∆Abs

b

CHYCSYECFPEYFP

[Heme] (µM)

0 2 4 6 8 10 120.0

0.2

0.4

0.6

∆Abs

41

6

1.0

1.5

2.0

2.5

3.0

CHY

ECFP

CSY

CH

[Heme] (µM)

0 2 4 6 8 10 12 τ

(n

s)

hemeequivalentsadded

[Heme] (µM)

No

rma

lize

d E

FR

ET

440

Norm

aliz

ed F

luore

scence

Wavelength (nm)

480 520 560 600

1.2

1.0

0.8

0.6

0.4

0.2

0.00 2 4 6 8 10 12

1.2

1.0

0.8

0.6

0.4

0.2

0.0

CHY

CSY

ECFP

55

Figure 2-7. Schematic of the mini-library of PfHRP2 fragments (shown in blue) evaluated for improved FRET properties. The fragment length (in amino acids) and the amino acid coordinates for mapping each fragment onto full length PfHRP2 are indicated.

Figure 2-8. SDS-PAGE purity analysis of the recombinant mini-library of truncated biosensors.

We recombinantly expressed and purified these constructs, and determined that all bound heme

with apparent affinities similar to full-length CHY and stoichiometries that were directly

proportional to the length of the heme-binding domain (Figs. 2-9 and 2-10).

b

d e

Kd,app (µM)

0.25

0.082

0.050

0.13

0.12

0.067

CHY37 309

273 aaECFP EYFP

CH115Y309195

115 aaECFP EYFP

CH91Y309219

91 aaECFP EYFP

CH65Y309243

65 aaECFP EYFP

CH49Y309261

49 aaECFP EYFPMEME (no SSC)7.2.2012 07:42

0

1

2

bits

1

G2

C3C

AT

4

C5A 6

CT

7

C8A 9T 10

G11

TC

12

TA x 34

PfHRP2

Construct

CH18YECFP EYFP Heme binding domain (amino acids)0 100 200 300

0

5

10

15

20

Hem

e eq

uiva

lent

s bo

und

c

∆EFR

ET

Heme binding domain (amino acids)0 100 200 300

0.2

0.4

0.6

0.00 2 4 6 8 10 12

[Heme] (µM)

0.1

0.2

0.3

0.4

0.5

EFR

ET

[Heme] (µM)

Nor

mal

ized

EFR

ETS

enso

r

0.0

0.2

0.4

0.6

0.8

1.0

0 2 4 6 8 10 12

CHY

0.8

0.0

CHYCSYCH49Y

S3

CH115Y

CH91Y

CH65Y

CH49Y

CH18Y

56

Figure 2-9. Heme binding titrations of truncated sensor mini-library showing heme binding stoichiometry and apparent heme binding affinity.

57

Figure 2-10. Heme-binding stoichiometry varies linearly with length of the PfHRP2-derived heme-binding domain.

In addition, the heme sensing dynamic range was improved with shorter heme-binding domains

(Fig. 2-11).

Figure 2-11. Dependence of ∆EFRET [= EFRET (no heme) – EFRET (saturating heme)] on length of the PfHRP2-derived heme binding domain.

All sensors exhibited heme-dependent decreases in FRET efficiency, with half-maximal FRET

values between 2-5 µM heme (Figs. 2-12 and 2-13). At a given heme concentration, the shorter

the PfHRP2 fragment the greater the detected FRET signal (Figs. 2-11 and 2-12).

b

Kd,app (µM)

0.25

0.082

0.050

0.13

0.12

0.067

CHY37 309

273 aaECFP EYFP

CH115Y309195

115 aaECFP EYFP

CH91Y309219

91 aaECFP EYFP

CH65Y309243

65 aaECFP EYFP

CH49Y309261

49 aaECFP EYFPMEME (no SSC)7.2.2012 07:42

0

1

2

bits

1

G

2

C

3CAT

4

C

5A 6

CT

7

C

8A 9T 10

G

11

TC

12

TA x 34

PfHRP2

Construct

CH18YECFP EYFP Heme binding domain (amino acids)0 100 200 300

0

5

10

15

20

Hem

e eq

uiva

lent

s bo

und

a

∆EFR

ET

Heme binding domain (amino acids)0 100 200 300

0.2

0.4

0.6

0.00 2 4 6 8 10 12

[Heme] (µM)

0.1

0.2

0.3

0.4

0.5

EFR

ET

[Heme] (µM)

Nor

mal

ized

EFR

ETS

enso

r

0.0

0.2

0.4

0.6

0.8

1.0

0 2 4 6 8 10 12

CHY

0.8

0.0

CHYCSYCH49Y

b

Kd,app (µM)

0.25

0.082

0.050

0.13

0.12

0.067

CHY37 309

273 aaECFP EYFP

CH115Y309195

115 aaECFP EYFP

CH91Y309219

91 aaECFP EYFP

CH65Y309243

65 aaECFP EYFP

CH49Y309261

49 aaECFP EYFPMEME (no SSC)7.2.2012 07:42

0

1

2

bits

1

G

2

C

3CAT

4

C5A 6

CT

7

C8A 9T 10

G11

TC

12

TA x 34

PfHRP2

Construct

CH18YECFP EYFP Heme binding domain (amino acids)0 100 200 300

0

5

10

15

20

Hem

e eq

uiva

lent

s bo

und

a

∆EFR

ET

Heme binding domain (amino acids)0 100 200 300

0.2

0.4

0.6

0.00 2 4 6 8 10 12

[Heme] (µM)

0.1

0.2

0.3

0.4

0.5

EFR

ET

[Heme] (µM)

Nor

mal

ized

EFR

ETS

enso

r

0.0

0.2

0.4

0.6

0.8

1.0

0 2 4 6 8 10 12

CHY

0.8

0.0

CHYCSYCH49Y

58

Figure 2-12. Heme-dependent changes in FRET efficiency for the mini-library of truncated sensors.

S5

59

Figure 2-13. Effect of truncating the heme-binding domain on sensitivity (IC50) of the resulting biosensor. Generally, shorter heme-binding domains result in sensors with lower sensitivity (higher IC50). CH49Y was the most sensitive of the truncated biosensor family with the second-largest dynamic range (after CH18Y), and was therefore chosen as the optimal biosensor design.

We selected CH49Y for further development due to its improved dynamic range over CHY

and retained sensitivity to heme. We reasoned that for quantitative analyses, the CSY construct

would serve as a normalization reference to account for factors unrelated to heme binding that

may affect ECFP and EYFP fluorescence intensities in situ. Based on curve fits to our in vitro

data (Fig. 2-14a), we derived an empirical relationship between the normalized CH49Y FRET,

(EFRETSensor), and heme (Fig. 2-14b, see Online Methods).

Figure 2-14. Quantifying heme using CH49Y sensor and CSY control. (a) EFRET dependence on heme concentration for optimized CH49Y (blue hollow circles) relative to the original sensor CHY (filled blue circles) and CSY (red triangles). (b) Calibration curve relating normalized EFRET

Sensor to heme concentration (see Online Methods).

S6

b

Kd,app (µM)

0.25

0.082

0.050

0.13

0.12

0.067

CHY37 309

273 aaECFP EYFP

CH115Y309195

115 aaECFP EYFP

CH91Y309219

91 aaECFP EYFP

CH65Y309243

65 aaECFP EYFP

CH49Y309261

49 aaECFP EYFPMEME (no SSC)7.2.2012 07:42

0

1

2

bits

1

G

2

C

3CAT

4

C

5A 6

CT

7

C

8A 9T 10

G

11

TC

12

TA x 34

PfHRP2

Construct

CH18YECFP EYFP Heme binding domain (amino acids)0 100 200 300

0

5

10

15

20H

eme

equi

vale

nts

boun

d

a

∆EFR

ET

Heme binding domain (amino acids)0 100 200 300

0.2

0.4

0.6

0.00 2 4 6 8 10 12

[Heme] (µM)

0.1

0.2

0.3

0.4

0.5

EFR

ET

[Heme] (µM)

Nor

mal

ized

EFR

ETS

enso

r

0.0

0.2

0.4

0.6

0.8

1.0

0 2 4 6 8 10 12

CHY

0.8

0.0

CHYCSYCH49Y

60

2.4.3 Correlating FRET efficiencies determined by imaging microscopy and fluorimetry for

calibrating heme concentrations

To accurately measure labile heme pools in cells using FRET imaging microscopy, we next

defined the relationship between microscopy and fluorimetric data. This facilitates converting

FRET efficiency data collected by microscopy into heme concentrations based on fluorimetric

calibration data (Fig. 2-14b). To achieve this, we encapsulated solutions of purified proteins into

giant multilamellar vesicles (GMVs) and measured FRET efficiencies by microscopy. In parallel,

we determined FRET efficiencies for the same solutions by fluorimetry. While GMVs emitted

fluorescence consistent with the loaded proteins, only those containing CH49Y or CSY exhibited

significant FRET (Fig. 2-15).

Figure 2-15. Representative images of CH49Y, CSY, ECFP, EYFP and ECFP+EYFP encapsulated in GMVs illustrate relative brightness in each fluorescence channel. Only GMVs containing CH49Y or CSY exhibited significant FRET.

bECFP EYFP

ECFPEYFP CH49Y CSY

0.8

0

CFPexCFPem

CFPexYFPem

YFPexYFPem

EFRET

BF

a

c

61

These data showed that the mean apparent FRET efficiencies determined by imaging and

fluorimetry agreed closely (Fig. 2-16a,b). Furthermore, by encapsulating the apo-sensor mini-

library having a range of inherently different FRET efficiencies (Figs. 2-7, 2-9, and 2-11), we

determined that imaging and fluorimetry yield highly correlated FRET efficiencies, albeit with a

small offset (Fig. 2-16b). Altogether, these data validate that FRET efficiencies measured by

microscopy can be accurately translated into heme concentrations based on fluorimetric

calibration data.

Figure 2-16. Relating FRET efficiencies measured by microscopy to fluorimetric calibration data. (a) Distributions of FRET efficiency for GMVs containing apo-CH49Y (blue solid line) and CSY (red dashed line). (b) FRET efficiencies of sensor library described in Fig. 2-7 determined by microscopy are plotted against those determined by fluorimetry. Error bars represent 95% confidence intervals.

2.4.4 Measuring labile heme in live malaria parasites

Having developed, characterized and optimized our heme sensor in vitro, we focused next on

implementing it to quantify cytosolic labile heme concentrations in live P. falciparum parasites.

Using NF54attB [26] as a parental strain, we created clonal parasite lines expressing our optimized

CH49Y sensor alongside the CSY control. We also created several spectral control lines

expressing ECFP, EYFP, and both ECFP and EYFP as individual proteins, in order to calibrate

0 0.1 0.2 0.3 0.4 0.5 0.60

0.2

0.4

0.6

0.8

1.0

EFRET

Rel

ativ

e Fr

eque

ncy

a CH49YCSY

b

0

0.2

0.4

0.6

0.8

0 0.2 0.4 0.6 0.8

GM

V E

FRE

T (M

icro

scop

e)

Solution EFRET (Fluorimeter)

Slope = 1.04Offset = 0.049R2 = 0.98

62

our FRET efficiency calculations. All lines homogenously expressed the respective proteins

within the parasite’s cytosol as determined by epifluorescence microscopy (Fig. 2-17a) and flow

cytometry (Fig. 2-18). We also confirmed by Western blot that both CH49Y and CSY were

expressed as full-length proteins, thus eliminating proteolysis or premature translational

termination as potential confounding factors in our subsequent data analyses (Fig. 2-17b).

Figure 2-17. Expression of genetically-encoded heme sensor and controls in P. falciparum. (a) Representative images of CH49Y, CSY, ECFP, EYFP and ECFP+EYFP in trophozoite-stage P. falciparum parasites. (b) Anti-GFP Western blot of lysates obtained from trophozoite stage parasites expressing CH49Y, CSY and YFP.

a

0.8

0

ECFP EYFPECFPEYFP CH49Y CSY

CFPexCFPem

CFPexYFPem

YFPexYFPem

EFRET

BF60 kD

50 kD

40 kD

30 kD

25 kD

CH

49Y

CS

Y

EY

FP

Ladd

erb

63

Figure 2-18. Analysis of CH49Y sensor and CSY control expression in trophozoite-stage P. falciparum by flow cytometry. The transgenic parasite pools (blue line) showed positive fluorescence over background (black) in the Pacific Blue (CFP only), AmCyan (CFP and FRET), and FITC (YFP) channels. Clonal parasites used in subsequent experiments (red) exhibited homogenous expression of the respective fluorophores.

During initial studies, we observed that normalized EFRETSensor for normally developing late

stage parasites (0.69; 95% CI = 0.65-0.72) was significantly lower than in heme-free GMVs

(0.93; 95% CI = 0.91-0.95) (Figs. 2-19a and 2-16a). As expected, parasites expressing ECFP and

EYFP as individual proteins exhibited a mean EFRET close to zero (-0.0060; 95% CI = -0.013-

0.0022). As both CH49Y and CSY control are expressed as full-length proteins (Fig. 2-17b),

degradation that preferentially reduces CH49Y FRET relative to that of CSY does not account for

the observation. Based on our calibration data (Figs. 2-14b and 2-16b) and 15 independent

experiments, we observed an average normalized EFRETSensor = 0.714 (95% CI = 0.711-0.718) in

late stage parasites, consistent with a 1.14 µM cytosolic labile heme pool (95% CI = 1.13-1.16

µM).

0 103 104 105AmCyan-A

0

20

40

60

80

100

0 103 104 105FITC-A

0

20

40

60

80

100

% o

f Max

0 103 104 105Pacific Blue-A

0

20

40

60

80

100

% o

f Max

0 103 104 105AmCyan-A

0

20

40

60

80

100

% o

f Max

0 103 104 105FITC-A

0

20

40

60

80

100

% o

f Max

0 103 104 105Pacific Blue-A

0

20

40

60

80

100

CH49Y

CSY

CFP(Pacific Blue)

FRET(AmCyan)

YFP(FITC)

% o

f Max

% o

f Max

64

We next sought to understand the dynamics of this labile heme pool over the course of the

48-hour P. falciparum intra-erythrocytic developmental cycle (IDC). As previously discussed,

hemoglobin degradation is a potential source of a labile cytosolic heme pool in the parasite 13,15.

As the quantity of hemoglobin degraded significantly increases during progression through the

IDC, we wished to determine whether labile cytosolic heme levels would similarly rise or if

these would be maintained at relatively constant levels throughout. Using highly synchronous

parasite cultures, we monitored normalized EFRETSensor in ring-stage (4 hours post-invasion (hpi)),

trophozoite-stage (16 hpi and 28 hpi), and schizont-stage (38 hpi) parasites. Intriguingly, the

metabolic changes associated with progression through the IDC appear to have minimal effects

on the labile cytosolic heme pool (Fig. 2-19b). This suggests that, despite the large-scale

hemoglobin degradation in the DV of trophozoite and schizont-stage parasites, labile heme levels

in the cytosol remain tightly controlled during parasite development.

Figure 2-19. Using genetically-encoded heme sensor to measure cytosolic labile heme concentrations in P. falciparum. (a) Representative FRET efficiency distributions for P. falciparum trophozoites expressing CH49Y (solid line) and CSY (dashed line). Using the calibration curve in Fig. 2-14b yields an average cytosolic labile heme concentration of 1.14 µM (95% CI = 1.13-1.16 µM) across 15 independent experiments. (b) Quantitation of cytosolic labile heme over the P. falciparum intra-erythrocytic developmental cycle. Representative images of Giemsa-stained parasites to confirm the parasite stage being analyzed are shown for each time point. Measurements of CH49Y and CSY lines were made in triplicate (9 calculations, see Methods). Error bars represent 95% confidence intervals.

0 0.1 0.2 0.3 0.4 0.5 0.60

0.2

0.4

0.6

0.8

1

EFRET

Rel

ativ

e Fr

eque

ncy

CH49YCSY 4 h.p.i.

16 h.p.i.

28 h.p.i.

38 h.p.i.

CH49Y CSY

0

0.5

1.0

1.5

Labi

le h

eme

(µM

)

Time post-invasion (hours)4 16 28 38

2.0a b

65

2.4.5 Quantitative analysis of perturbed heme homeostasis by heme-interacting antimalarials

We next used our heme sensor to gain direct insight into a proposed mechanism of

antimalarial drug action. Heme-interacting antimalarial drugs, such as chloroquine, comprise one

of the most clinically successful classes discovered to date, yet fundamental aspects of their

mechanism of action remain unknown. Indeed, a more detailed understanding of these

mechanisms could potentially be exploited for developing the next generation of antimalarial

drugs that are not immediately susceptible to described resistance mechanisms 27. Heme-binding

antimalarial drugs have been proposed to interfere with detoxification of hemoglobin-derived

heme by inhibiting its polymerization into hemozoin, thereby increasing labile heme

concentrations to toxic levels 10. However, the effects of inhibiting hemozoin formation in the

DV on heme levels in other parasite compartments has been a challenging question to directly

and quantitatively address. Therefore, we sought to use our heme sensor to specifically quantify

changes in the concentration of cytosolic labile heme upon parasite exposure to chloroquine.

We first established that parasites expressing CH49Y and CSY did not show altered

sensitivity to chloroquine by determining IC50 values. We determined that both lines had similar

chloroquine IC50 values of 7.8 nM and 8.4 nM, respectively (Fig. 2-20). These data indicate that

CH49Y does not directly interfere with chloroquine action by sequestering heme and preventing

potential toxic effects associated with any increase in labile heme levels induced by chloroquine.

66

Figure 2-20. Growth inhibition data comparing the sensitivity of parasites expressing CH49Y (blue) and CSY (red) to chloroquine.

We then exposed highly synchronous parasites expressing either CH49Y or CSY to various

concentrations of chloroquine, and determined normalized EFRETSensor and cytosolic labile heme

concentrations. In five independent experiments, we detected a significant increase in labile

heme concentration with 60 nM chloroquine treatment compared to the untreated control

(average ∆[heme] = +0.67 µM; p = 0.018) (Figs. 2-21a and 2-21b). As a negative control, we

treated highly synchronous parasites with the antifolate pyrimethamine, which exerts its

antimalarial activity by inhibiting DNA synthesis rather than interfering with heme metabolism.

No increase in labile heme concentration was detected between the treated and untreated

parasites (Fig. 2-21b), consistent with pyrimethamine’s mode of action. Altogether, these data

demonstrate that a model heme-interacting antimalarial compound specifically induces a

significant increase in cytosolic labile heme that can be quantified using our FRET-based heme

sensor technology. These data also add new biological insight by demonstrating that the increase

in cytosolic labile heme is unlikely to be a generalized cytotoxic response, but rather one that is

directly and specifically linked to chloroquine-mediated dysregulation of parasite heme

homeostasis.

log [Cq] (nM)

1.5

1.0

0.5

0

Rel

ativ

e ex

pans

ion

-1 0 1 2 3

CH49YCSY

67

Figure 2-21. Quantifying the impact of chloroquine on cytosolic labile heme pool in P. falciparum. (a) Change in labile heme concentration at various chloroquine concentrations after 24-hour exposures. Data is representative of five independent experiments. Error bars show 95% confidence interval. (b) Cytosolic labile heme concentrations in parasites that were untreated or exposed to chloroquine (60 nM) or pyrimethamine (125 nM and 500 nM) for 32 hours. Data is representative of three independent experiments. Error bars represent 95% confidence intervals derived from bootstrapping calculations (see Online Methods).

0

0.5

1.5

2.5

[Cq] (nM)0 20 40 60

ba

2.0

1.0

[Hem

e] (µ

M)

[Hem

e] (µ

M)

Untreated Cq(60 nM)

Pyr(125 nM)

Pyr(500 nM)

0

0.5

1.5

2.5

2.0

1.0

68

2.5 Discussion

Here we report the development, characterization, and application of a novel, FRET-based

heme biosensor to measure cytosolic labile heme levels in live, blood stage P. falciparum

malarial parasites. Our approach is distinct from previous strategies used to measure intracellular

labile heme, in providing the important advantages of obtaining quantitative information using

intact cells and with subcellular resolution. Previous methods have relied on measuring the

heme-dependent activities of L-tryptophan-2,3-dioxygenase and ∂-aminolevulinic acid synthase

(ALAS) 28,29 natively present in the cell type of interest, or recombinantly expressed horseradish

peroxidase (HRP) 30. Fractionation methods have also been reported 16,31. These all require cell

lysis, which makes it difficult to confidently infer subcellular localization of detected labile heme.

Our genetically-encoded sensor facilitates direct visualization of heme in intact cells through

imaging, and can be specifically targeted to different cell compartments using appropriate signal

sequences. Therefore, in principle, it should broadly enable determination of the distribution and

levels of labile heme with subcellular resolution that exceeds previous standards that have relied

on fractionation or bulk cell analyses.

In applying our system to study blood stage P. falciparum, we show for the first time that the

parasite maintains a ~1.1 µM labile cytosolic heme pool throughout its intraerythrocytic

development. Based on earlier work 32, labile heme has been thought to be inherently highly

toxic to the parasite. The observation that heme-binding antimalarial drugs induce accumulation

of membrane-associated heme in treated parasites 33 has further contributed to widespread

acceptance of this hypothesis. This proposal has been based largely on qualitative data, however.

Our findings suggest that malaria parasites do not stringently restrict labile heme. Rather, readily

measurable levels are maintained, and these are consistent with those reported in other

69

eukaryotic cells 34. Thus, it appears that the physiologic requirement for maintaining labile heme

in P. falciparum and other previously studied organisms might be more conserved than

previously thought.

The source of heme used by the parasite to maintain this cytosolic pool is unknown, but

scavenging from the DV, de novo biosynthesis, and uptake from the extracellular compartment

are formal possibilities. Previous qualitative evidence suggests that blood stage parasites

synthesize heme de novo 13,14, but the extent to which this contributes to the heme pool we

measure is unknown. Moreover, recent studies have shown that the parasite’s biosynthetic

pathway is dispensable during blood stage development 13,15, suggesting that biosynthesis may

not be a critical source of heme for the parasite during this stage. Developing parasites are also

known to become increasingly permeable to extracellular low molecular weight solutes via the

new permeability pathway 35, so labile heme could potentially be acquired via this route.

However, the largest obvious flux of heme in blood stage parasites is through hemoglobin

degradation in the DV. Therefore, this seems to be a reasonable candidate source of heme for the

parasite. Additional studies will be needed to definitively address this possibility.

While the exact source and physiologic role(s) of this labile cytosolic heme pool remain to be

defined, our data show that the heme-binding antimalarial drug chloroquine specifically

dysregulates this pool. While chloroquine accumulates to high levels within the DV to interfere

with heme polymerization 10, our data suggest that some of the heme that is not polymerized

escapes the DV to reach the parasite’s cytosol. It is important to emphasize that our sensor likely

responds only to the soluble fraction of released heme, and not the parasite cell membrane-

associated fraction 33. Thus, our measurements do not reflect the total heme flux induced by

chloroquine treatment. This distinction is important, as these two heme pools can potentially

70

induce different outcomes. Membrane-associated heme will likely contribute to cell membrane

damage, while increased soluble heme could either induce cytosolic oxidative stress and/or

directly interfere with critical protein function(s) 36. In this model, both heme pools may directly

contribute to cytotoxicity. Alternatively, as resting labile cytosolic heme levels are already

reasonably high, moderate increases may be tolerated without significantly increased toxicity.

Instead, increases in soluble heme levels may be directly sensed by the parasite to initiate critical

heme-dependent responses, such as changes in transcription 1,37,38, translation and/or proteasome-

mediated protein turnover 39,40, as in other organisms. These may serve to coordinate changes in

DV metabolism with nuclear and cytosolic events that either counteract or exacerbate any

adverse effects induced by increased efflux of labile heme from the DV into the cytoplasm.

The responses to increased heme flux induced by antimalarial compounds likely overlap with

the mechanisms for maintaining heme homeostasis. Guided by our new quantitative

understanding of labile heme levels in P. falciparum, elucidating these mechanisms can stimulate

new therapeutic strategies that recapitulate important aspects of chloroquine’s antimalarial

mode(s) of action, while circumventing resistance mechanisms that have made it increasingly

ineffective. Novel compounds that induce dysregulation of parasite labile heme pools and/or key

heme-regulated processes may be especially promising leads given the extraordinary success of

the 4-aminoquinoline antimalarial drug class.

Here, we have used our technology to examine a longstanding gap in our knowledge of labile

heme pools in the malaria parasite. However, this approach is broadly applicable to studying

other cellular systems where a quantitative understanding of intracellular labile heme pools is

still noticeably absent. For instance, many pathogenic bacteria of human health relevance, such

as S. aureus and N. meningitidis, delicately balance heme synthesis, acquisition, sequestration

71

and degradation/extrusion to minimize toxicity due to excess intracellular heme accumulation 41.

Various molecular mechanisms mediating each of these outcomes and their heme-dependent

responses have been described, and these are clearly linked to pathogenicity 41. However, critical

insights into the actual intracellular heme concentrations that define a toxic threshold or that

integrates these mechanisms to ensure proper heme homeostasis are still lacking. Application of

our technology here could yield new insights into heme homeostasis, and establish a stronger

quantitative basis for the fundamental link between this central cofactor and infectious disease

caused by very distinct pathogens.

72

2.6 References 1. Girvan, H. M. & Munro, A. W. Heme sensor proteins. J Biol Chem 288, 13194-13203,

doi:10.1074/jbc.R112.422642 (2013). 2. Poulos, T. L. Heme enzyme structure and function. Chem Rev 114, 3919-3962,

doi:10.1021/cr400415k (2014). 3. Khan, A. A. & Quigley, J. G. Control of intracellular heme levels: heme transporters and

heme oxygenases. Biochim Biophys Acta 1813, 668-682, doi:10.1016/j.bbamcr.2011.01.008 (2011).

4. Ziegler, J., Linck, R. & Wright, D. W. Heme Aggregation inhibitors: antimalarial drugs targeting an essential biomineralization process. Curr Med Chem 8, 171-189 (2001).

5. Pluth, M. D., Tomat, E. & Lippard, S. J. Biochemistry of mobile zinc and nitric oxide revealed by fluorescent sensors. Annu Rev Biochem 80, 333-355, doi:10.1146/annurev-biochem-061009-091643 (2011).

6. Que, E. L., Domaille, D. W. & Chang, C. J. Metals in neurobiology: probing their chemistry and biology with molecular imaging. Chem Rev 108, 1517-1549, doi:10.1021/cr078203u (2008).

7. WHO Global Malaria Programme. World Malaria Report 2014. (World Health Organization).

8. Egan, T. J. et al. Fate of haem iron in the malaria parasite Plasmodium falciparum. Biochem J 365, 343-347, doi:10.1042/BJ20020793 (2002).

9. Esposito, A. et al. FRET imaging of hemoglobin concentration in Plasmodium falciparum-infected red cells. PLoS ONE 3, e3780, doi:10.1371/journal.pone.0003780 (2008).

10. Francis, S. E., Sullivan, D. J. & Goldberg, D. E. Hemoglobin metabolism in the malaria parasite Plasmodium falciparum. Annu Rev Microbiol 51, 97-123, doi:10.1146/annurev.micro.51.1.97 (1997).

11. Hanssen, E. et al. Soft X-ray microscopy analysis of cell volume and hemoglobin content in erythrocytes infected with asexual and sexual stages of Plasmodium falciparum. J Struct Biol 177, 224-232, doi:10.1016/j.jsb.2011.09.003 (2012).

12. Gardner, M. J. et al. Genome sequence of the human malaria parasite Plasmodium falciparum. Nature 419, 498-511, doi:10.1038/nature01097 (2002).

13. Nagaraj, V. A. et al. Malaria parasite-synthesized heme is essential in the mosquito and liver stages and complements host heme in the blood stages of infection. PLoS Pathog 9, e1003522, doi:10.1371/journal.ppat.1003522 (2013).

14. Surolia, N. & Padmanaban, G. de novo biosynthesis of heme offers a new chemotherapeutic target in the human malarial parasite. Biochem Biophys Res Commun 187, 744-750 (1992).

15. Ke, H. et al. The Heme Biosynthesis Pathway is Essential for Plasmodium falciparum Development in Mosquito Stage but not in Blood Stages. J Biol Chem, doi:10.1074/jbc.M114.615831 (2014).

16. Combrinck, J. M. et al. Insights into the role of heme in the mechanism of action of antimalarials. ACS Chem Biol 8, 133-137, doi:10.1021/cb300454t (2013).

17. Atamna, H. & Ginsburg, H. Heme degradation in the presence of glutathione. A proposed mechanism to account for the high levels of non-heme iron found in the membranes of hemoglobinopathic red blood cells. J Biol Chem 270, 24876-24883 (1995).

73

18. Schneider, E. L. & Marletta, M. A. Heme binding to the histidine-rich protein II from Plasmodium falciparum. Biochemistry 44, 979-986, doi:10.1021/bi048570p (2005).

19. Sullivan, D. J., Gluzman, I. Y. & Goldberg, D. E. Plasmodium hemozoin formation mediated by histidine-rich proteins. Science 271, 219-222 (1996).

20. Akompong, T. et al. Trans expression of a Plasmodium falciparum histidine-rich protein II (HRPII) reveals sorting of soluble proteins in the periphery of the host erythrocyte and disrupts transport to the malarial food vacuole. J Biol Chem 277, 28923-28933, doi:10.1074/jbc.M201968200 (2002).

21. Choi, C. Y., Cerda, J. F., Chu, H. A., Babcock, G. T. & Marletta, M. A. Spectroscopic characterization of the heme-binding sites in Plasmodium falciparum histidine-rich protein 2. Biochemistry 38, 16916-16924 (1999).

22. Evers, T. H., van Dongen, E. M. W. M., Faesen, A. C., Meijer, E. W. & Merkx, M. Quantitative understanding of the energy transfer between fluorescent proteins connected via flexible peptide linkers. Biochemistry 45, 13183-13192, doi:10.1021/bi061288t (2006).

23. Takeda, S., Kamiya, N., Arai, R. & Nagamune, T. Design of an artificial light-harvesting unit by protein engineering: cytochrome b(562)-green fluorescent Protein chimera. Biochem Biophys Res Commun 289, 299-304, doi:10.1006/bbrc.2001.5966 (2001).

24. Pandey, A. V., Joshi, R., Tekwani, B. L., Singh, R. L. & Chauhan, V. S. Synthetic peptides corresponding to a repetitive sequence of malarial histidine rich protein bind haem and inhibit haemozoin formation in vitro. Mol Biochem Parasitol 90, 281-287 (1997).

25. Ziegler, J., Chang, R. T. & Wright, D. W. Multiple-antigenic peptides of histidine-rich protein II of Plasmodium falciparum: Dendrimeric biomineralization templates. J Am Chem Soc {121}, {2395-2400}, doi:{10.1021/ja983220+} (1999).

26. Adjalley, S. H. et al. Quantitative assessment of Plasmodium falciparum sexual development reveals potent transmission-blocking activity by methylene blue. Proc Natl Acad Sci USA 108, E1214-1223, doi:10.1073/pnas.1112037108 (2011).

27. Sidhu, A. B. S., Verdier-Pinard, D. & Fidock, D. A. Chloroquine resistance in Plasmodium falciparum malaria parasites conferred by pfcrt mutations. Science 298, 210-213, doi:10.1126/science.1074045 (2002).

28. Granick, S., Sinclair, P., Sassa, S. & Grieninger, G. Effects by heme, insulin, and serum albumin on heme and protein synthesis in chick embryo liver cells cultured in a chemically defined medium, and a spectrofluorometric assay for porphyrin composition. J Biol Chem 250, 9215-9225 (1975).

29. Lincoln, B. C., Alvarado, A., Setty, N. & Bonkovsky, H. L. Tryptophan 2,3-Dioxygenase in Chick-Embryo Hepatocytes - Studies Inovo and in Culture. Proceedings of the Society for Experimental Biology and Medicine 188, 308-315 (1988).

30. White, C. et al. HRG1 is essential for heme transport from the phagolysosome of macrophages during erythrophagocytosis. Cell Metab 17, 261-270, doi:10.1016/j.cmet.2013.01.005 (2013).

31. Liu, S. C., Zhai, S. & Palek, J. Detection of hemin release during hemoglobin S denaturation. Blood 71, 1755-1758 (1988).

32. Orjih, A. U., Banyal, H. S., Chevli, R. & Fitch, C. D. Hemin lyses malaria parasites. Science 214, 667-669 (1981).

74

33. Ginsburg, H., Famin, O., Zhang, J. & Krugliak, M. Inhibition of glutathione-dependent degradation of heme by chloroquine and amodiaquine as a possible basis for their antimalarial mode of action. Biochem Pharmacol 56, 1305-1313 (1998).

34. Garrick, M. D. et al. Evidence for and consequences of chronic heme deficiency in Belgrade rat reticulocytes. Biochim Biophys Acta 1449, 125-136 (1999).

35. Martin, R. E., Ginsburg, H. & Kirk, K. Membrane transport proteins of the malaria parasite. Mol Microbiol 74, 519-528, doi:10.1111/j.1365-2958.2009.06863.x (2009).

36. Radfar, A., Diez, A. & Bautista, J. M. Chloroquine mediates specific proteome oxidative damage across the erythrocytic cycle of resistant Plasmodium falciparum. Free Radic Biol Med 44, 2034-2042, doi:10.1016/j.freeradbiomed.2008.03.010 (2008).

37. Ogawa, K. et al. Heme mediates derepression of Maf recognition element through direct binding to transcription repressor Bach1. EMBO J 20, 2835-2843, doi:10.1093/emboj/20.11.2835 (2001).

38. Smart, J. L. & Bauer, C. E. Tetrapyrrole biosynthesis in Rhodobacter capsulatus is transcriptionally regulated by the heme-binding regulatory protein, HbrL. J Bacteriol 188, 1567-1576, doi:10.1128/JB.188.4.1567-1576.2006 (2006).

39. Hu, R.-G., Wang, H., Xia, Z. & Varshavsky, A. The N-end rule pathway is a sensor of heme. Proc Natl Acad Sci USA 105, 76-81, doi:10.1073/pnas.0710568105 (2008).

40. Yang, F., Xia, X., Lei, H.-Y. & Wang, E.-D. Hemin Binds to Human Cytoplasmic Arginyl-tRNA Synthetase and Inhibits Its Catalytic Activity. J Biol Chem 285, 39437-39446, doi:10.1074/jbc.M110.159913 (2010).

41. Anzaldi, L. L. & Skaar, E. P. Overcoming the heme paradox: heme toxicity and tolerance in bacterial pathogens. Infect Immun 78, 4977-4989, doi:10.1128/IAI.00613-10 (2010).

42. Bailey, T. L. & Elkan, C. Fitting a mixture model by expectation maximization to discover motifs in biopolymers. Proc Int Conf Intell Syst Mol Biol 2, 28-36 (1994).

43. Clegg, R. M. Fluorescence resonance energy transfer and nucleic acids. Meth Enzymol 211, 353-388 (1992).

44. Kolossov, V. L. et al. Engineering redox-sensitive linkers for genetically encoded FRET-based biosensors. Exp Biol Med (Maywood) 233, 238-248, doi:10.3181/0707-RM-192 (2008).

45. Nkrumah, L. J. et al. Efficient site-specific integration in Plasmodium falciparum chromosomes mediated by mycobacteriophage Bxb1 integrase. Nat Methods 3, 615-621, doi:10.1038/nmeth904 (2006).

46. Hodgson, L., Shen, F. & Hahn, K. Biosensors for characterizing the dynamics of rho family GTPases in living cells. Current protocols in cell biology / editorial board, Juan S Bonifacino [et al] Chapter 14, Unit 14.11.11-26, doi:10.1002/0471143030.cb1411s46 (2010).

47. Schindelin, J. et al. Fiji: an open-source platform for biological-image analysis. Nat Methods 9, 676-682, doi:10.1038/nmeth.2019 (2012).

75

2.7 Appendix: MATLAB Scripts

2.7.1. Calculate heme concentration and confidence intervals from image data

% Takes two column vectors as input: % X = CH49Y FRET Sensor, S = CSY FRET+ Control % Each vector consists of cell-wide averages of FRET efficiency % Vectors of means and variances for each bootstrap sample size means = []; vars = []; bootstrap_ratios = []; resample_number = 10000; % number of times calculation is performed test_resample_size = 0; % test sensitivity to size of bootstrap dataset? if test_resample_size % test how mean & variance are affected by size % of subsamples max_resample_size = 100; % set maximum size of subsample to test for resample_size = 1:max_resample_size bootstrap_ratios = []; % vector of ratios of bootstrapped means for(i = 1:resample_number) % loop for resampling & ratio calculation % resample FRET sensor dataset bootstrap_x = datasample(X, resample_size);

% resample FRET+ control dataset bootstrap_s = datasample(S, resample_size); % calculate ratio of means, converting microscopy to % fluorimetry EFRET values by subtracting y-intercept % (0.04897)... remember that image values are ratio*1000 avg_subset_ratio = (mean(bootstrap_x) - 48.97) /(mean(bootstrap_s) – 48.97); % create output dataset bootstrap_ratios = [bootstrap_ratios; avg_subset_ratio]; end means = [means; mean(bootstrap_ratios)] % mean of output dataset vars = [vars; var(bootstrap_ratios)]; % variance of output datset end xvalues = 1:max_resample_size; plot(xvalues, means); % plot mean vs. subsample size xlabel('Bootstrap Dataset Size') ylabel('FRET Ratio (Sensor/Control)') figure plot(xvalues, vars); % plot variance vs. subsample size xlabel('Bootstrap Dataset Size') ylabel('Variance in FRET Ratio') else % perform bootstrapping on full-size datasets for(j = 1:resample_number) % loop to do resampling & ratio calculation % resample FRET sensor datset bootstrap_x = datasample(X, length(X)); % resample FRET+ control dataset bootstrap_s = datasample(S, length(S));

76

% calculate ratio of means, converting microscopy to % fluorimetry EFRET values by subtracting y-intercept % (0.04897)... remember that image values are ratio*1000 avg_subset_ratio = (mean(bootstrap_x) - 48.97) / (mean(bootstrap_s) - 48.97); % create output dataset bootstrap_ratios = [bootstrap_ratios; avg_subset_ratio]; end hist(bootstrap_ratios, 100); % display histogram of bootstrapped ratios xlabel('FRET Ratio (Sensor/Control)') avgRatio = mean(bootstrap_ratios) % mean of bootstrapped ratios % sort dataset to calculate confidence intervals sorted_ratios = sort(bootstrap_ratios); % 2.5th percentile for two-tailed 95% CI bootstrap_lowCI = sorted_ratios(round(0.025*length(sorted_ratios))) % 97.5th percentile for two-tailed 95% CI bootstrap_highCI = sorted_ratios(round(0.975*length(sorted_ratios))) if ~adtest(bootstrap_ratios) % Anderson-Darling test: is distribution of calculated ratios normal? sigma = sqrt(var(bootstrap_ratios))

% if so, calculate 95% CI based on +/- 2 standard devs from mean normal_lowCI = avgRatio - 2*sigma normal_highCI = avgRatio + 2*sigma end end % Solve for heme concentrations for ratios calculated by bootstrapping syms heme; span = 0.3830; k = 0.2835; plateau = 0; slope = -0.01282; yint = 0.4023; Heme_Avg = double(vpasolve(avgRatio == ((span*exp(-k*heme) + plateau)/(slope*heme + yint)), heme)); Heme_High = double(vpasolve(bootstrap_lowCI == ((span*exp(-k*heme) +

plateau)/(slope*heme + yint)), heme)); Heme_Low = double(vpasolve(bootstrap_highCI == ((span*exp(-k*heme) +

plateau)/(slope*heme + yint)), heme));

77

2.7.2. Calculate average heme concentration and confidence intervals from bootstrapping data

obtained from multiple experiments

% Takes a matrix (master_bootstrap) as input, where each column consists of bootstrap values from a single experiment, and the number of rows is the number of iterations performed for each bootstrapping calculation. The 16 independent experiments in this study yielded a master_bootstrap matrix with dimensions 10000 x 16. for j = 1:length(master_bootstrap) grand_mean(j) = mean(master_bootstrap(j,:)); end hist(grand_mean, 100); % display histogram of bootstrapped ratios xlabel('FRET Ratio (Sensor/Control)') avgRatio = mean(grand_mean) % mean of bootstrapped ratios % sort dataset to calculate confidence intervals sorted_ratios = sort(grand_mean); % 2.5th percentile for two-tailed 95% CI bootstrap_lowCI = sorted_ratios(round(0.025*length(sorted_ratios))) % 97.5th percentile for two-tailed 95% CI bootstrap_highCI = sorted_ratios(round(0.975*length(sorted_ratios))) % Anderson-Darling test: is distribution of calculated ratios normal? if ~adtest(grand_mean) % if so, calculate 95% CI based on +/- 2 standard devs from mean sigma = sqrt(var(grand_mean)) normal_lowCI = avgRatio - 2*sigma normal_highCI = avgRatio + 2*sigma end % Solve for heme concentrations for ratios calculated by bootstrapping syms heme; span = 0.3830; k = 0.2835; plateau = 0; slope = -0.01282; yint = 0.4023; Heme_Avg = double(vpasolve(avgRatio == ((span*exp(-k*heme) +

plateau)/(slope*heme + yint)), heme)); Heme_High = double(vpasolve(bootstrap_lowCI == ((span*exp(-k*heme) +

plateau)/(slope*heme + yint)), heme)); Heme_Low = double(vpasolve(bootstrap_highCI == ((span*exp(-k*heme) +

plateau)/(slope*heme + yint)), heme)

78

CHAPTER 3: USING SYNTHETIC RECEPTORS TO ELUCIDATE HOST CELL

REQUIREMENTS FOR PARASITE INVASION

3.1 Abstract

Research on the parasite P. vivax, a major cause of malaria-associated morbidity, lags behind

that of P. falciparum. Unlike P. falciparum, P. vivax preferentially invades reticulocytes, which

are relatively rare and transient in whole blood, precluding in vitro culture. While P. vivax

merozoites express proteins that preferentially bind reticulocytes, the basis of the parasite’s

invasion tropism remains incompletely understood, mainly due to the inability to link ligand-

receptor binding to invasion. We hypothesize that critical receptors present on the reticulocyte

surface are lost during red blood cell maturation, preventing efficient invasion by the parasite

and precluding propagation in abundant mature red blood cells. Therefore, we have developed a

chemical biology toolkit to immobilize synthetic receptors on the red blood cell surface and

probe the relationship between ligand-receptor interactions and invasion outcomes. Proof-of-

concept experiments using the sialic acid-dependent invasion pathway in P. falciparum

demonstrate that synthetic sialic acid receptors can restore invasion into desialylated red blood

cells, as long as the synthetic glycan exactly recapitulates the native glycan structure. Alternative

receptor structures, which have been shown to inhibit invasion in vitro, appear unable to

facilitate sialic acid-dependent invasion. This suggests that synthetic receptor-based strategies

can provide crucial information about the biochemical context necessary for surface receptors to

support parasite invasion, and can guide future efforts to develop synthetic receptors for P. vivax

culture.

79

3.2 Introduction

The vast majority of human malaria cases are caused by two parasites, Plasmodium

falciparum and P. vivax1. Although P. falciparum causes the majority of malaria-related

mortality, vivax malaria is more widespread and significantly hinders the health, longevity, and

prosperity of the population in endemic areas2. Despite the significant economic and public

health burden imposed by this disease, basic research on P. vivax lags behind that of P.

falciparum2-4. A major advance in P. falciparum research was the development of a practical

method for continuously propagating P. falciparum cultures using mature human red blood cells5

which can be easily obtained. In contrast, P. vivax is extremely difficult to culture in vitro

because it preferentially invades reticulocytes6 which constitute only 1% of the circulating cells

in human whole blood7. Further complicating culture system design is reticulocyte maturation, as

reticulocytes mature rapidly into normocytes in vitro, with a half-life of roughly 30 hours8,

similar to the generation time of blood-stage P. vivax parasites9. Therefore, in vitro culture of P.

vivax relies on a continuous supply of a cell type that is relatively rare and transient in whole

blood.

As a result, long-term in vitro culture of P. vivax is currently impractical because of the labor

and expense involved in obtaining reticulocytes. Reticulocytes can be enriched from human

whole blood using gradient centrifugation, although this process is inefficient and results in low

reticulocyte yields7. Human umbilical cord blood, which is naturally enriched in reticulocytes,

has also been used for P. vivax culture10. However, this method can maintain cultures for only

one month, and requires uninterrupted access to discarded umbilical cords. Infected research

animals can be used to supply P. vivax isolates, as P. vivax also infects Aotus and Saimiri

monkeys11. However, maintaining continuously infected animals for parasite supply is extremely

80

expensive and raises ethical concerns. Recently, CD34+ hematopoietic stem cells have been

differentiated to generate continuous supplies of red blood cells12, and this technique has been

applied to P. vivax culture13. However, this method produces limited yields and requires

expensive growth factors. Because of these challenges, developing a practical method for in vitro

culture of P. vivax has been identified as a major priority in malaria research4,14. An ideal

culturing system would propagate this parasite continuously using human normocytes and other

reagents that are readily and inexpensively available. However, this would depend on

overcoming the P. vivax preference for invading and developing within reticulocytes.

Considering that P. vivax expresses ligands that preferentially bind reticulocytes15, and that

loss of surface proteins is a major feature of reticulocyte maturation16, it is reasonable to

hypothesize that normocytes lack receptors critical for efficient P. vivax invasion. As P. vivax

still invades normocytes at a lower rate6, these receptors are not absolutely required for invasion

to occur, but increase the efficiency of invasion. The related parasite P. knowlesi expresses

ligands that bind to the Duffy antigen17, and requires this antigen to efficiently invade human red

blood cells18. In addition, sialic acid-dependent P. falciparum strains invade through interactions

with terminal sialic acid residues presented on the glycophorins, a family of abundant

erythrocyte glycoproteins19. Treating red blood cells with neuraminidase removes these sialic

acids and greatly decreases P. falciparum invasion rates20, but invasion can be rescued by

incubating parasites and neuraminidase-treated cells with sialylated α1-acid glycoprotein20, which

is structurally dissimilar to the glycophorins21.

81

Therefore, we hypothesize that P. vivax is unable to invade normocytes with high efficiency

because critical receptors are lost upon reticulocyte maturation, and that invasion could be

restored by attaching suitable receptors to the normocyte surface (Fig. 3-1).

Figure 3-1. We hypothesize that normocytes (1) are refractory to P. vivax invasion because necessary receptors are missing from the cell surface, and that restoring these interactions in trans (2) should allow invasion to proceed successfully (3).

Due to the aforementioned difficulty of culturing P. vivax in vitro, we turned to the well-

characterized sialic acid-dependent invasion pathway in P. falciparum as a model system for

proof-of-concept studies. The dominant interaction in this pathway is that of the parasite protein

EBA-175 binding to glycophorin A22, one of the most abundant proteins on the red blood cell

surface23. After generating refractory host cells by removing the critical sialic acid residues from

otherwise unaltered human red blood cells, we develop and evaluate multiple methods for

displaying the critical sialic acid receptor in trans. We then use these results to derive parameters

governing productive ligand-receptor interactions that drive sialic acid-dependent parasite

invasion to inform future synthetic receptor designs.

We first quantify the relationship between sialic acid content and invasion efficiency to

establish the theoretical needs for synthetic receptor density on the treated red blood cell surface.

We then use sialyltransferase enzymes to assess the importance of the chemical linkage type

(1) (2) (3)

82

between the terminal sialic acid and the underlying glycan moiety, demonstrating that the

parasite discriminates between native and non-native terminal sialic acid linkages. We finally

demonstrate two broadly applicable biochemical strategies for immobilizing alternate receptor

structures on the cell surface at high density. We show that while our immobilization strategies

are effective, the resulting receptors cannot facilitate parasite invasion, suggesting that

productive ligand-receptor interactions are highly specific to receptor structure.

83

3.3 Methods 3.3.1 Malaria parasite culture

Blood-stage P. falciparum parasites were cultured at 2% hematocrit in 5% O2 and 5% CO2 in

RPMI-complete media (RPMI 1640 Medium supplemented with 5 g/l AlbuMAX II [Life

Technologies], 2 g/l NaHCO3, 25 mM HEPES-KOH pH 7.4, 1 mM hypoxanthine, and 50 mg/l

gentamicin). P. falciparum strains 3D7, Dd2, and W2mef were obtained from the Malaria

Reference and Reagent Resource Center (MR4). Parasite cultures were synchronized by

centrifugation, resuspension in 0.3M L-alanine (Research Products International) supplemented

with 25 mM HEPES-KOH pH 7.4, and incubation for 8 minutes at 37˚C. Cultures were then

centrifuged and resuspended in culture media.

3.3.2 Neuraminidase treatment

Human red blood cells (Research Blood Components, Brighton, MA) were washed in wash

media (RPMI 1640 supplemented with 25 mM HEPES-KOH pH 7.4) and resuspended to 50%

hematocrit. Neuraminidase from C. perfringens (New England Biolabs, Ipswich, MA) was added

and the red blood cell suspension was incubated at 37˚C with gentle mixing for 30 minutes. The

neuraminidase-treated red blood cells were then washed extensively with RPMI-Complete

culture media before use.

3.3.3 Sialyltransferase treatment

Neuraminidase-treated red blood cells were resuspended to 50% hematocrit in α2,3 or α2,6

sialyltransferase reaction solution, consisting of 25 mM buffer, 75 mM NaCl, 100 mM glucose,

10 mg/mL bovine serum albumin, and 4.4 mM CMP-sialic acid (Calbiochem, catalog #233264),

84

and sialyltransferase enzyme. Treatment with α2,3 sialyltransferase was performed at 37˚C in 2-

(N-morpholino)ethanesulfonic acid (MES) pH 6.5 with 20 mU recombinant α2,3(O)-

sialyltransferase from rat (Calbiochem, catalog #566227). Treatment with α2,6 sialyltransferase

was performed at 30˚C in bis-Tris pH 6.0 with 50 mU recombinant β-galactoside-α2,6-

sialyltransferase from Photobacterium damselae JT0160 (Catalog #GEJ-001, Cosmo Bio Ltd.,

Tokyo, Japan). Reactions were incubated for 4 hours with gentle mixing. The red blood cell

suspension was then centrifuged gently. Red blood cells treated with α2,3-sialyltransferase were

washed extensively with RPMI-complete media supplemented with 7% glycerol in order to

prevent osmotic lysis. Glycerol was then removed from the red blood cell suspension by a series

of centrifugation and 1:1 dilution steps with RPMI-complete media. Red blood cells treated with

α2,6 sialyltransferase were washed extensively with hypertonic resealing solution (280 mM

NaCl, 40 mM KCl, 11 mM glucose), then RPMI-HEPES, before resuspending in RPMI-

complete media.

3.3.4 Oxime ligation

Oxime ligations were performed as described previously24, but with several modifications.

Packed red blood cells (40µL per sample) were washed in PBS pH 6.0, then simultaneously

treated with neuraminidase and 5U galactose oxidase (Worthington Biochemical, Lakewood, NJ)

for 30 minutes at 37˚C with gentle mixing. Treated cells were washed in RPMI-complete media,

and then with oxime ligation buffer, containing either 10 mM aniline or 100 mM p-anisidine in

PBS pH 6.8. Aminooxy CF488A dye (Biotium, Hayward, CA) or aminooxy glycans

(synthesized as described25) were added to the indicated final concentrations and the reactions

85

were incubated for 1 hour at room temperature. Treated cells were washed extensively with

RPMI-complete media before analysis or use in an invasion assay.

3.3.5 Flow cytometry

Parasite cultures were briefly centrifuged and resuspended in a solution of 1% formaldehyde

in Alsever’s ACD media (114 mM glucose, 27 mM sodium citrate pH 6.1, and 72 mM NaCl) for

fixing and incubated for one hour at room temperature. For incubations longer than one hour

were performed at 4˚C. Suspensions of fixed cells were diluted 1:10 in staining solution,

containing 10 mM Tris-HCl pH 8.8, 138 mM NaCl, and a 1:5000 dilution of SYBR Green I (Life

Technologies) added before each experiment. Fixed cells were incubated in staining solution

before parasites were counted using the FITC channel of an Accuri C6 flow cytometer (BD

Biosciences).

3.3.6 Fluorimetric sialic acid quantitation

Aliquots of 10µL packed red blood cells were washed and resuspended to 250µL in PBS pH

6.0. Cells were treated with 250U neuraminidase for 30 minutes at 37˚C with gentle mixing.

After neuraminidase treatment, cells were pelleted and 200 µL of each supernatant was harvested

and stored at -80˚C. Standard curves were prepared by serial dilution of N-acetylneuraminic acid

(Sigma-Aldrich) in PBS. Fluorimetric quantitation of sialic acids was performed as described

previously26.

86

3.3.7 HPLC sialic acid quantitation

Samples and standard curves were prepared as described above, except neuraminidase

digestions were performed in 40µL total volume. Supernatants were harvested and sialic acid

was quantified as described previously27. Briefly, cell supernatants were treated with 20µL of a

freshly prepared solution of 0.2 M sodium periodate in 48% phosphoric acid for 20 minutes at

room temperature. To terminate the reaction, 100µL of freshly prepared 10% sodium arsenite in

0.1N sulfuric acid was added slowly, then vortexed until clear and incubated 5 minutes at room

temperature. Finally, 600 µL of thiobarbituric acid (6 mg/mL) was added, and the reaction was

incubated at 100˚C for 15 minutes, then chilled on ice. Prior to injection, each sample was

centrifuged briefly. Samples were analyzed using an 1100 series HPLC system (Agilent

Technologies) equipped with a variable wavelength detector and a 250 x 4.6mm ZORBAX

Eclipse C18 column (Agilent Technologies). Elution was performed isocratically using a running

buffer consisting of 115 mM sodium perchlorate, 30% methanol, and 1% phosphoric acid, and

absorbance was monitored at 549nm.

3.3.8 Glycophorin extraction and biotinylation

Crude glycophorins were extracted from human red blood cells as previously described28.

Briefly, human red blood cells from 450mL blood were divided into 10 ml aliquots and washed

with 40 mL PBS. Washed red blood cells were then lysed via repeated washing with 40 ml ice-

cold 5 mM sodium phosphate pH 8.0 supplemented with 1 mM phenylmethylsulfonyl fluoride

(PMSF). Each tube was incubated on ice for 5 min before centrifuging at 4500g in swinging

buckets for 60 minutes at 2˚C with no braking. Red cell ghosts were washed 6-8 times until the

ghost pellet appeared white with a clear supernatant. Ghosts were then resuspended to 50% in the

87

same buffer supplemented with 1.2 M NaCl. Nine volumes of a 2:1 v:v chloroform methanol

mixture were added to the red cell ghost suspension, and stirred vigorously (with occasional

shaking) for 30 minutes at room temperature. The mixture was then incubated at 4˚C overnight.

The aqueous phase was then recovered and centrifuged at 40,000g for 30 minutes, then dialyzed

extensively into 5 mM ammonium bicarbonate pH 8.3. The dialyzed solution was then

concentrated 10X using centrifugal filters (Amicon Ultra 30,000 MWCO, EMD Millipore). The

concentrated protein solutions were then lyophilized using a SpeedVac vacuum concentrator

(Thermo Scientific) with no heating. Before use, aliquots of extracted glycophorins were

reconstituted in 400 µL PBS pH 8.0, analyzed for purity by SDS-PAGE, and analyzed for

protein concentration using the BCA Protein Assay Kit (Thermo Scientific). Extracted

glycophorins were then biotinylated using 1 mg sulfo-NHS-biotin (Thermo Scientific) per

protein aliquot. The biotinylation reaction was incubated at room temperature for 30 minutes

before extensive dialysis in PBS pH 8.0. Biotinylation was verified using the Fluorescence

Biotin Quantitation Kit (Thermo Scientific), and the biotinylated glycophorins were stored at 4˚C

for up to two weeks.

3.3.9 Glycophorin immobilization

Red blood cells were washed in ice-cold PBS pH 8.0 and resuspended to 10% hematocrit.

From this suspension, approximately 109 red blood cells (100 µL packed) were incubated with

0.5 mg sulfo-NHS biotin for 30 minutes with gentle mixing at room temperature. Biotinylated

red blood cells were then washed and treated with neuraminidase as described above, with 5U

enzyme per µL of packed red blood cells, except that PBS pH 6.0 supplemented with 2 mg/mL

bovine serum albumin was substituted for RPMI-HEPES in the washing step. After

88

neuraminidase treatment, biotinylated red blood cells were resuspended to 10% hematocrit, split

into aliquots containing 2 x 108 cells each, and set aside. Biotinylated glycophorins were pre-

complexed with NeutrAvidin (Thermo Scientific) by adding the glycophorin solution to the side

of a microcentrifuge tube containing 60 µL of 1 mg/mL NeutrAvidin at the bottom and

immediately vortexing. Mixing continued using a bead-beater for 15 minutes at room

temperature. The glycophorin-NeutrAvidin complex was then added to the biotinylated red blood

cell suspension, mixed immediately by inversion, then incubated for 1 hour at room temperature

with gentle mixing. The binding reaction was terminated by adding 100 µL of saturated biotin in

PBS, then incubating for 15 minutes at room temperature with gentle mixing.

3.3.10 Invasion assay

Parasite invasion into treated red blood cells was measured using flow cytometry and SYBR

Green staining as described previously29, with several modifications. Tightly-synchronized

schizont-stage parasites at ~10% parasitemia were treated with 250U neuraminidase to reduce

reinvasion, washed, and then mixed with target cells. Invasion assays were seeded in triplicate

200µL samples at ~1% parasitemia at 0.5% hematocrit in a 96-well plate. To quantify inoculum

parasitemia, a subset of samples was fixed immediately with formaldehyde as described above,

and stored at 4˚C. The remaining samples were incubated for 48 hours, then fixed for 1 hour.

Inoculum and post-invasion samples were then stained with SYBR Green I and analyzed by flow

cytometry as described above. Expansion rates were calculated by dividing the post-invasion

parasitemia by the parasitemia of the inoculum sample. Relative expansion rates were calculated

by normalizing the expansion rate to that of the untreated control.

89

3.4 Results

3.4.1 Effect of surface receptor density on parasite invasion rates

We first sought to determine the relationship between density of the terminal sialic acid

receptors and parasite invasion. Given the high degree of avidity between carbohydrates and

carbohydrate-binding molecules like EBA-175, we hypothesized that a critical density of surface

sialic acid receptors would be necessary to support efficient invasion into the host red blood cell.

To test this hypothesis, we treated human red blood cells with various amounts of neuraminidase,

which cleaves terminal sialic acid residues from surface glycans, and quantified the sialic acid

released. Treating with neuraminidase resulted in a dose-dependent release of surface sialic acid,

where 50 U enzyme released nearly all accessible sialic acid (Fig. 3-2).

Figure 3-2. Neuraminidase treatment liberates terminal sialic acids from the red blood cell surface in a dose-dependent manner. Sialic acid release is saturable with approx. 50U neuraminidase. Unless otherwise noted, sialic acid release was quantified using the TBA-HPLC sialic acid assay (see Methods).

Sialic acid removal by neuraminidase resulted in reduced invasion efficiency by the sialic acid-

dependent P. falciparum strains Dd2 and W2mef (Fig. 3-3). While extensive neuraminidase

treatment could essentially eliminate parasite invasion into treated cells, treatment with more

0 50 100 150 200 2500.0

0.2

0.4

0.6

0.8

1.0

1.2

Neuraminidase (U)

Rel

ativ

e si

alic

aci

d re

leas

e

90

modest amounts of enzyme appeared to remove substantial sialic acid from the red cell surface

without affecting invasion rates (Figs. 3-3 and 3-4).

Figure 3-3. Release of terminal sialic acids with neuraminidase inhibits invasion by sialic acid-dependent P. falciparum strains Dd2 and W2mef.

Comparing sialic acid content to relative invasion rates revealed three distinct regimes. At

low levels (≤ 25% of native sialic acid content), parasite invasion was almost completely

inhibited, and changes in sialic acid content caused only minimal changes in invasion rates.

Between 25-50%, parasite invasion and sialic acid content appeared directly proportional.

Maximal parasite invasion occurred when surface sialic acid content was ≥ 50% that of untreated

red blood cells (Fig. 3-4). These results suggested that neuraminidase treatment represented a

tunable platform for comparing the ability of candidate synthetic receptors to restore parasite

invasion rates into sialic acid-deficient host cells.

Invasion correlates with sialic acid content

0 U

3.125

U6.2

5 U12

.5 U

25 U

50 U

100 U

0

25

50

75

100

125

150

Sialic acid contentDd2 expansionW2mef expansion

Neuraminidase

Rel

ativ

e am

ount

(%)

91

Figure 3-4. Efficient invasion of sialic acid-dependent strains into neuraminidase-treated host cells requires only 50% of native sialic acid content. Roughly 1/3 of native sialic acid content can support half-maximal invasion rates.

3.4.2 Enzymatic restoration of sialic acid receptors

Having established that parasite invasion rates are sensitive to sialic acid removal, we next

assessed whether the neuraminidase-induced invasion defect could be reversed by restoring sialic

acid to glycoproteins on the red blood cell surface. We first attempted to restore sialic acid

content by treating sialic acid-deficient red blood cells with a mammalian sialyltransferase that

attaches sialic acid to galactose via an α2,3(O) linkage, the predominant linkage type on human

red blood cells30,31. Incubating neuraminidase-treated cells with α2,3(O)-sialyltransferase and

excess CMP-sialic acid substrate resulted in significant restoration of surface sialic acid content

(Fig. 3-5). Greater restoration occurred on red cells that had been treated more extensively with

neuraminidase.

Invasion correlates with sialic acid content

0 25 50 75 1000

25

50

75

100

125

150

W2mefDd2EC50 = 32%

Sialic acid content (%)

Expa

nsio

n (%

)

92

Figure 3-5. Incubating neuraminidase-treated red blood cells with CMP-sialic acid and α2,3(O)-sialyltransferase (ST) restores sialic acid content, with greatest restoration (+50%) occurring in cells treated with 50 U neuraminidase.

Sialic acid restoration with α2,3(O)-sialyltransferase resulted in a significant restoration of

parasite invasion rates in both Dd2 and W2mef parasite strains (Fig. 3-6). Treatment with 50 U

neuraminidase resulted in nearly complete inhibition of parasite invasion, which was almost

completely restored by sialyltransferase treatment. Taken together, these results suggest that

neuraminidase treatment removes necessary invasion receptors from the red blood cell surface,

and that the necessary receptors presented in the correct biochemical context could restore the

ligand-receptor interactions required for efficient invasion.

0 U12

.5 U

25 U

50 U

0

25

50

75

100

125– ST+ α2,3(O)ST

Neuraminidase

Sial

ic a

cid

(%)

93

Figure 3-6. Sialic acid restoration by α2,3(O)-sialyltransferase (ST) rescues the invasion defects introduced by neuraminidase treatment.

3.4.3 Enzymatic attachment of sialic acid receptors with an alternate terminal linkage

Next, we assessed whether the neuraminidase-induced invasion defect could be rescued by

sialic acid receptors presented in non-native biochemical contexts. Using a recombinant α2,6

sialyltransferase from Photobacterium damselae, we were able to restore measureable amounts

of sialic acid to the neuraminidase-treated red blood cell surface, although restoration was less

complete than with the α2,3 sialyltransferase (Fig. 3-7). However, sialic acid restoration with

α2,6 sialyltransferase did not enhance parasite invasion into treated cells (Fig. 3-8), suggesting

that sialic acid-dependent invasion is sensitive to the linkage between the terminal sialic acid and

the underlying glycan structure.

0 U12

.5 U

25 U

50 U

0

25

50

75

100

125

Neuraminidase

Rel

ativ

e ex

pans

ion

(%) – ST (Dd2)

+ α2,3(O)ST (Dd2)– ST (W2mef)

+ α2,3(O)ST (W2mef)

94

Figure 3-7. Incubating neuraminidase-treated red blood cells with CMP-sialic acid and α2,6 sialyltransferase (ST) restores surface sialic acid content, but to a much more limited extent than with the α2,3(O)-sialyltransferase enzyme. The greatest amount of restoration (+19%) occurred in cells treated with 50 U neuraminidase.

Figure 3-8. Sialic acid restoration by α2,6-sialyltransferase (ST) does not rescue the invasion defect introduced by neuraminidase treatment.

0 U12

.5 U

25 U

50 U

0

25

50

75

100

125

– ST+ α2,6ST

Neuraminidase

Sial

ic a

cid

(%)

0 U12

.5 U

25 U

50 U

0

25

50

75

100

125

Neuraminidase

Rel

ativ

e ex

pans

ion

(%)

– ST

+ α2,6ST

95

3.4.4 Synthetic glycan receptor construction using aminooxy-functionalized reagents

Finally, we investigated whether synthetic receptors containing the required α2,3-linked

sialic acid presented in non-native biochemical contexts could facilitate sialic acid-dependent

invasion into neuraminidase-treated red blood cells. By treating red blood cells with

neuraminidase and galactose oxidase, we were able to chemically conjugate aminooxy-

functionalized 2,3-sialyllactose25 to the red blood cell surface, restoring sialic acid to nearly

native levels (Figs. 3-9 and 3-10).

Figure 3-9. Treatment of the native glycan structure (1) with neuraminidase removes the terminal sialic acid reside, exposing the penultimate galactose (2). Treatment with galactose oxidase results in oxidation of the C6 carbon into an aldehyde (4). By combining this oxidized galactose with an aminooxy-functionalized sialyllactose moiety (3), the sialyllactose can be covalently attached to the native glycan with a bioorthogonal oxime linkage (5).

Gal

GalNAc Neu5Ac

β1-3

α2-6 Ser/Thr

Neu5Ac

Gal

Glc

α2-3

β1-4

N

O

Gal

GalNAc Neu5Ac

β1-3

α2-6 Ser/Thr

O

HCH2OH

Gal

GalNAc Neu5Ac

β1-3

α2-6 Ser/Thr

Gal

GalNAc Neu5Ac

β1-3

α2-6 Ser/Thr

Neu5Ac

α2-3

Neu5Ac

Gal

Glc

α2-3

β1-4

ONH2 (1)

(2)

(3)

(4)

(5)

96

Figure 3-10. Incubating neuraminidase-treated red blood cells with galactose oxidase, p-anisidine, and aminooxy-2,3-sialyllactose (2,3SLac-ONH2) effectively restores sialic acid to the red blood cell surface.

However, this structure was unable to rescue the neuraminidase-induced invasion defect in Dd2

and W2mef strains, while invasion by sialic acid-independent P. falciparum strain 3D7 was

unaffected (Fig. 3-11). This suggested that sialic acid-dependent invasion is sensitive to glycan

structure, and that glycans on synthetic receptors have specific structural requirements to

productively engage parasite ligands and facilitate invasion.

0 U 25 U

50 U

100 U

0

25

50

75

100

125– SLac-ONH2

+ 2,3SLac-ONH2

Neuraminidase

Sial

ic a

cid

(%)

97

Figure 3-11. Sialic acid restored by incubating neuraminidase-treated cells with galactose oxidase, p-anisidine, and aminooxy-2,3-sialyllactose (2,3SLac-ONH2) is not able to facilitate invasion by sialic acid-dependent P. falciparum strains Dd2 and W2mef. Invasion by the sialic acid-independent strain 3D7 was unaffected by these treatments.

3.4.5 Synthetic glycoprotein receptor construction using biotin-NeutrAvidin interactions

Finally, we hypothesized that the complete native glycophorin structure (including glycans

and peptide backbone) was required by the parasite, and that immobilizing the extracted

glycophorin fraction from invasion-competent red blood cells could rescue the defect induced by

neuraminidase treatment. To immobilize extracted glycophorins on the red blood cell surface, we

developed a method combining chemical biotinylation with sulfo-NHS biotin and binding to

NeutrAvidin 32 to create a stable linkage between cell surface and protein (Fig. 3-12).

0 U 25 U

50 U

100 U

0

25

50

75

100

125 – SLac-ONH2 (Dd2)

+ 2,3SLac-ONH2 (Dd2)

– SLac-ONH2 (W2mef)

+ 2,3SLac-ONH2 (W2mef)

Neuraminidase

Rel

ativ

e ex

pans

ion

(%)

– SLac-ONH2 (3D7)

+ 2,3SLac-ONH2 (3D7)

98

Figure 3-12. Treatment of human red blood cells (1) with neuraminidase and sulfo-NHS biotin yielded biotinylated red blood cells lacking surface sialic acid receptors (5). Extracted glycophorins, also biotinylated with sulfo-NHS-biotin (2), were mixed with NeutrAvidin (3) to yield a pre-bound receptor-NeutrAvidin complex (4), which was then bound to the red blood cell surface (6). Incubation with excess biotin (7) capped the remaining biotin-binding sites on the immobilized NeutrAvidin complex (8).

We first prepared an extract of the crude glycophorin fraction of red blood cell ghosts by

chloroform-methanol extraction28 (Fig. 3-13), and biotinylated aliquots of the extracted protein

using sulfo-NHS-biotin. Total yields of extracted glycophorins were 2.2 mg by BCA assay from

120 mL of packed red blood cells. We then pre-complexed the biotinylated glycophorin fraction

with NeutrAvidin. Finally, we combined this glycophorin-NeutrAvidin complex with

neuraminidase-treated and biotinylated red blood cells. Y"

Y"

Y"Y"

Y"Y"

Y"Y"Y"Y" Y"

Y"

Y"Y" Y"

Y"Y"

Y" Y"

Y"

Y"

Y"

Y"

Y"

(1) (5)

(2)

(3)

(4)

(6)

Y"

Y" Y"

Y"

Y"

Y"

Y"

Y"

(8)

Y" Y"

(7)

99

Figure 3-13. SDS-PAGE purity analysis of the glycophorin fraction extracted from red blood cell ghosts using chloroform:methanol.

This biotinylation and immobilization technique effectively restored sialic acid content to the

surface of neuraminidase-treated red blood cells by a fluorimetric assay26, with nearly complete

restoration at 0.5:1 and 1:1 stoichiometries of glycophorins : NeutrAvidin. However, despite the

high levels of terminal sialic acid restored, none of these treatments enhanced invasion into

neuraminidase-treated red blood cells (Fig. 3-14).

MW

66.4

55.6

42.7

34.6

27.0

20.0

97.2 116 158

4.6 µg 2.3 µg 1.2 µg kDa

100

Figure 3-14. Glycophorin immobilization restores sialic acid content to the neuraminidase-treated RBC surface (by fluorimetric assay), although this restoration does not enhance sialic acid-dependent invasion in strains W2mef or Dd2.

Untrea

ted

Neuram

inida

se

0.25 G

PA : NA

0.5 G

PA : NA

1.0 G

PA : NA

0

25

50

75

100

125

Rel

ativ

e am

ount

(%)

Sialic acid contentInvasion (W2mef)Invasion (Dd2)

101

3.5 Discussion

Here we demonstrate a proof-of-concept to restore parasite invasion into otherwise refractory

red blood cells by providing the necessary invasion receptors in trans. In a model P. falciparum

system, we were able to quantify the relationship between receptor density and invasion rates,

showing for the first time that parasites require only a fraction of the sialic acid content of human

erythrocytes for efficient sialic acid-dependent invasion. Additionally, by replacing sialic acid

liberated by neuraminidase using a mammalian sialyltransferase that recreates the native α2-3

linkage type, we showed that supplying the requisite invasion receptor could rescue parasite

invasion into otherwise refractory cells.

Our results also demonstrate for the first time that the interaction between sialic acid-binding

parasite ligands and sialylated red cell receptors is highly specific to receptor glycan structure.

Modest sialic acid restoration with a bacterial α2-6 sialyltransferase did not rescue invasion into

neuraminidase-treated cells. While the total amount of sialic acid restored with the α2-6

sialyltransferase was lower (+19% rather than +50%), cells treated with α2-6 sialyltransferase

still contained up to 31% of the sialic acid content of untreated cells (Fig. 3-7), close to our

measured EC50. Therefore, these results suggest that α2-3- and α2-6-linked terminal sialic acids

are not interchangeable as parasite receptors.

Previous studies have indirectly assessed the relationship between sialic acid linkage type

and utility as a receptor for parasite invasion. Soluble sialyllactose has been used to inhibit

binding of EBA-175 to erythrocytes, where it was shown that α2-3-sialyllactose was a more

potent inhibitor of binding than its α2-6-linked counterpart33. However, the IC50 values for each

compound were each above 1 mM, suggesting that these interactions were limited in their

specificity. In addition, the sialic acid analogue 3’-N-acetyl neuraminyl-N-acetyl lactosamine,

102

which contains sialic acid linked α2-3- to a penultimate galactose, has been shown to disrupt

binding of EBA-175 to glycophorin A at micromolar concentrations34. However, this structure is

not found on the surface of the red blood cell. Finally, the crystal structure of EBA-175 was

solved with α2-3-sialyllactose bound35, but to our knowledge similar attempts have not been

made with α2-6-sialyllactose.

Additionally, we demonstrate biocompatible immobilization of aminooxy-functionalized

sugars onto the surface of neuraminidase-treated red blood cells. By conjugating aminooxy-

functionalized α2-3-sialyllactose onto the neuraminidase-treated red blood cell surface through

an oxime linkage, we restored sialic acid content nearly to native levels, and yet this restored

sialic acid was unable to facilitate invasion by the parasite. Importantly, the oxime ligation and

α2-3-sialyllactose immobilization were not toxic to parasite invasion, because the sialic acid-

independent strain 3D7 was still able to invade successfully. These results are consistent with

previous studies showing multiple contacts between EBA-175 and glycans on glycophorin A35,

as well as binding to the glycophorin A peptide backbone 36,37, although the relationships

between these individual contacts and overall invasion rates have not been elucidated. It is

possible that the geometry of the oxime-linked α2-3-sialyllactose synthetic receptor does not

support formation of intermolecular contacts necessary to support invasion.

Finally, we demonstrate a technique for immobilizing biotinylated proteins on the red blood

cell surface. Immobilizing extracted glycophorins through biotin-NeutrAvidin interactions

restored sialic acid on neuraminidase-treated cells to near-native levels, but did not support

parasite invasion. Similar glycophorin preparations have been shown to bind EBA-175

domains33,34 and peptides36 in vitro, suggesting that the immobilized glycophorins could

recapitulate the EBA-175-glycophorin A interaction. However, invasion by the parasite involves

103

multiple ligand-receptor binding events38, and therefore these synthetic receptors may have been

unable to recapitulate the other contacts necessary for productive invasion. However, this would

not explain the restoration of invasion observed when α1-acid glycoprotein was added to

neuraminidase-treated cells. Further investigation will be required to explain these results.

We believe these proof-of-concept experiments can inform future synthetic receptor designs

towards the goal of facilitating P. vivax invasion into mature red blood cells. We observe that

invasion can be restored into target cells lacking cognate receptors by supplying the necessary

interactions in trans, although invasion was only restored when the synthetic receptor structure

matched the native receptor exactly. Given the multiple contacts that must form between parasite

and host cell both within and across ligand-receptor pairs, finding a synthetic receptor design that

can facilitate all the necessary binding events may be challenging. However, we believe the

generalizable immobilization strategies developed here will be useful in future screening efforts

dedicated to finding compatible synthetic receptor structures.

104

3.6 References

1. Miller, L. H., Baruch, D. I., Marsh, K. & Doumbo, O. K. The pathogenic basis of malaria. Nature 415, 673–679 (2002).

2. Mendis, K., Sina, B. J., Marchesini, P. & Carter, R. The neglected burden of Plasmodium vivax malaria. Am J Trop Med Hyg 64, 97–106 (2001).

3. Price, R. et al. Vivax Malaria: Neglected and Not Benign. American Journal of Tropical Medicine and Hygiene 77, 79 (2007).

4. Mueller, I. et al. Key gaps in the knowledge of Plasmodium vivax, a neglected human malaria parasite. The Lancet Infectious Diseases 9, 555–566 (2009).

5. Trager, W. & Jensen, J. B. Human malaria parasites in continuous culture. Science 193, 673–675 (1976).

6. Kitchen, S. The infection of reticulocytes by Plasmodium vivax. Am J Trop Med Hyg 1, 347 (1938).

7. Golenda, C. F., Li, J. & Rosenberg, R. Continuous in vitro propagation of the malaria parasite Plasmodium vivax. Proc Natl Acad Sci USA 94, 6786–6791 (1997).

8. Skadberg, O., Brun, A. & Sandberg, S. Human reticulocytes isolated from peripheral blood: maturation time and hemoglobin synthesis. Lab Hematol 9, 198–206 (2003).

9. Bozdech, Z. et al. The transcriptome of Plasmodium vivax reveals divergence and diversity of transcriptional regulation in malaria parasites. Proceedings of the National Academy of Sciences 105, 16290–16295 (2008).

10. Udomsangpetch, R. et al. Short-term in vitro culture of field isolates of Plasmodium vivax using umbilical cord blood. Parasitology International 56, 65–69 (2007).

11. Nayar, J. K. et al. Studies on a primaquine-tolerant strain of Plasmodium vivax from Brazil in Aotus and Saimiri monkeys. J Parasitol 83, 739–745 (1997).

12. Giarratana, M.-C. et al. Ex vivo generation of fully mature human red blood cells from hematopoietic stem cells. Nature Biotechnology 23, 69–74 (2005).

13. Panichakul, T. et al. Production of erythropoietic cells in vitro for continuous culture of Plasmodium vivax. Int J Parasitol 37, 1551–1557 (2007).

14. NIAID Malaria Working Group. NIAID Research Agenda for Malaria. niaid.nih.gov (2008). at <http://www.niaid.nih.gov/topics/Malaria/Documents/researchagenda.pdf>

15. Galinski, M. R., Medina, C. C., Ingravallo, P. & Barnwell, J. W. A reticulocyte-binding protein complex of Plasmodium vivax merozoites. 69, 1213–1226 (1992).

16. Vidal, M., Mangeat, P. & Hoekstra, D. Aggregation reroutes molecules from a recycling to a vesicle-mediated secretion pathway during reticulocyte maturation. Journal of Cell Science 110 ( Pt 16), 1867–1877 (1997).

17. Haynes, J. D. et al. Receptor-like specificity of a Plasmodium knowlesi malarial protein that binds to Duffy antigen ligands on erythrocytes. J Exp Med 167, 1873–1881 (1988).

18. Miller, L. H., Mason, S. J., Dvorak, J. A., McGinniss, M. H. & Rothman, I. K. Erythrocyte receptors for (Plasmodium knowlesi) malaria: Duffy blood group determinants. Science 189, 561–563 (1975).

19. Camus, D. & Hadley, T. A Plasmodium falciparum antigen that binds to host erythrocytes and merozoites. Science 230, 553 (1985).

20. Friedman, M. J., Blankenberg, T., Sensabaugh, G. & Tenforde, T. S. Recognition and invasion of human erythrocytes by malarial parasites: contribution of sialoglycoproteins to attachment and host specificity. J Cell Biol 98, 1672–1677 (1984).

105

21. Fournier, T., Medjoubi-N, N. & Porquet, D. Alpha-1-acid glycoprotein. Biochim Biophys Acta 1482, 157–171 (2000).

22. Duraisingh, M. T., Maier, A. G., Triglia, T. & Cowman, A. F. Erythrocyte-binding antigen 175 mediates invasion in Plasmodium falciparum utilizing sialic acid-dependent and -independent pathways. Proc Natl Acad Sci USA 100, 4796–4801 (2003).

23. Anstee, D. J. The nature and abundance of human red cell surface glycoproteins. J. Immunogenet. 17, 219–225 (1990).

24. Zeng, Y., Ramya, T. N. C., Dirksen, A., Dawson, P. E. & Paulson, J. C. High-efficiency labeling of sialylated glycoproteins on living cells. Nature Methods 6, 207–209 (2009).

25. Hudak, J. E., Yu, H. H. & Bertozzi, C. R. Protein glycoengineering enabled by the versatile synthesis of aminooxy glycans and the genetically encoded aldehyde tag. J Am Chem Soc 133, 16127–16135 (2011).

26. Matsuno, K. & Suzuki, S. Simple fluorimetric method for quantification of sialic acids in glycoproteins. Anal Biochem 375, 53–59 (2008).

27. Powell, L. D. & Hart, G. W. Quantitation of picomole levels of N-acetyl- and N-glycolylneuraminic acids by a HPLC-adaptation of the thiobarbituric acid assay. Anal Biochem 157, 179–185 (1986).

28. Cochet, S., Volet, G., Cartron, J. P. & Bertrand, O. New procedures for glycophorin A purification with high yield and high purity. J Chromatogr B Biomed Sci Appl 750, 109–119 (2001).

29. Bei, A. K. et al. A flow cytometry-based assay for measuring invasion of red blood cells by Plasmodium falciparum. American Journal of Hematology 85, 234–237 (2010).

30. Fukuda, M., Lauffenburger, M., Sasaki, H., Rogers, M. E. & Dell, A. Structures of novel sialylated O-linked oligosaccharides isolated from human erythrocyte glycophorins. J Biol Chem 262, 11952–11957 (1987).

31. Reid, M. E. & Mohandas, N. Red blood cell blood group antigens: structure and function. Semin. Hematol. 41, 93–117 (2004).

32. Marttila, A. T. et al. Recombinant NeutraLite avidin: a non-glycosylated, acidic mutant of chicken avidin that exhibits high affinity for biotin and low non-specific binding properties. FEBS Lett 467, 31–36 (2000).

33. Orlandi, P. A., Klotz, F. W. & Haynes, J. D. A malaria invasion receptor, the 175-kilodalton erythrocyte binding antigen of Plasmodium falciparum recognizes the terminal Neu5Ac(alpha 2-3)Gal- sequences of glycophorin A. J Cell Biol 116, 901–909 (1992).

34. Bharara, R., Singh, S., Pattnaik, P., Chitnis, C. E. & Sharma, A. Structural analogs of sialic acid interfere with the binding of erythrocyte binding antigen-175 to glycophorin A, an interaction crucial for erythrocyte invasion by Plasmodium falciparum. Mol Biochem Parasitol 138, 123–129 (2004).

35. Tolia, N. H., Enemark, E. J., Sim, B. K. L. & Joshua-Tor, L. Structural basis for the EBA-175 erythrocyte invasion pathway of the malaria parasite Plasmodium falciparum. 122, 183–193 (2005).

36. Jakobsen, P. H. et al. Identification of an erythrocyte binding peptide from the erythrocyte binding antigen, EBA-175, which blocks parasite multiplication and induces peptide-blocking antibodies. Infect Immun 66, 4203–4207 (1998).

37. Sim, B. K., Chitnis, C. E., Wasniowska, K., Hadley, T. J. & Miller, L. H. Receptor and ligand domains for invasion of erythrocytes by Plasmodium falciparum. Science 264, 1941–1944 (1994).

106

38. Baum, J., Maier, A. G., Good, R. T., Simpson, K. M. & Cowman, A. F. Invasion by P. falciparum merozoites suggests a hierarchy of molecular interactions. PLoS Pathog 1, e37 (2005).

107

CHAPTER 4: CONCLUSIONS AND FUTURE WORK 4.1 Parasite heme biology In developing our heme biosensor, we were able to quantify for the first time a labile heme

pool in the cytosol of live malaria parasites, both under normal physiological conditions and

under stresses imposed by aminoquinoline treatment. In its current state of development, this

technology has the potential to address many unanswered questions regarding parasite heme

metabolism, and the mechanisms of action of current antimalarial drugs. In addition, the system

could be further optimized for use in screens to identify potentially promising heme-perturbing

antimalarial compounds.

The heme biosensor described in Chapter 2 could be used to quantify the labile heme pools in

other parasite organelles. In particular, the apicoplast and mitochondria contain many of the

enzymes involved in parasite heme biosynthesis, yet many key details remain to be elucidated.

For example, the parasite heme biosynthesis pathway has its final stages in the mitochondrion,

which is also the location for where most parasite hemoproteins reside. However, a labile heme

pool in the mitochondria has not been identified. Additionally, the heme biosynthesis pathway

involves enzymes localized in the mitochondria, apicoplast, and cytosol. However, it is not

known whether the apicoplast maintains a labile heme pool as well, or whether hypothesized

transporters for heme biosynthesis intermediates between these three compartments could also

transport heme. These questions could be addressed by expressing the heme biosensor in these

organelles, for which N-terminal protein targeting sequences have been identified5,6.

In addition, the source of the parasite’s cytosolic labile heme pool remains to be elucidated.

Although malaria parasites have been shown to biosynthesize heme during the blood stages7, this

is evidently not essential, as pathway enzymes can be chemically inhibited or knocked out with

108

no apparent effect. It is not currently known whether inhibiting heme biosynthesis affects the

cytosolic labile heme pool, and future experiments could compare the concentration of labile

heme between wild-type parasites and those unable to biosynthesize heme. If a deficiency in

heme biosynthesis resulted in cytosolic labile heme being depleted, that would suggest that a

pool of labile heme in the cytosol is not critical to parasite growth. However, if parasites

deficient in heme biosynthesis were still able to maintain a cytosolic labile heme pool, this would

suggest that parasites could obtain necessary heme from other sources. Potential alternate

sources include scavenging from the food vacuole, and import from the red blood cell cytosol.

As protein sequences have been elucidated that mediate trafficking to both of these

compartments, the heme biosensor could in principle be localized to either, although the low pH8

and proteolytic activity in the food vacuole may preclude accurate heme quantitation.

The role of the cytosolic labile heme pool could be further investigated by manipulating

heme concentrations directly. While chloroquine appears to cause labile heme to accumulate in

the parasite cytosol, the effects of depleting cytosolic labile heme are not yet known. Malaria

parasites have recently been shown to lack the ability to degrade heme due to a non-functional

heme oxygenase enzyme9. In conjunction with the heme biosensor, inducibly degrading heme by

expressing a heme oxygenase under translational10 or post-translational11 control could assess the

degree to which a cytosolic labile heme pool is required for parasite growth.

4.2 Antimalarial drug action

In this work, we have demonstrated that chloroquine treatment causes labile heme to

accumulate in the parasite cytosol, directly linking heme dysregulation to chloroquine toxicity.

However, important mechanistic details remain to be elucidated. For instance, the kinetics of

109

heme accumulation following chloroquine treatment could give insight into the underlying

mechanisms of chloroquine-induced heme transport. In addition, coupling these measurements to

assays of parasite viability12 could establish whether heme accumulation in the cytosol correlates

with parasite health, substantiating the role of heme as the toxic effector of chloroquine action.

Additionally, the heme biosensor could be used to gain insight into the mechanisms of action

of other successful antimalarials. For instance, artemisinin and its derivatives are first-line

therapies when used in combination with other drugs. However, the mechanism of artemisinin

action remains controversial. Artemisinin is hypothesized to be activated by heme to form a

reactive radical that can cause oxidative damage, although this has not been directly

demonstrated. Correlating artemisinin toxicity with labile heme concentration could substantiate

this proposed mechanism.

4.3 Antimalarial drug discovery

The potent antimalarial activity of the 4-aminoquinolines, and the parasite’s apparent

inability to develop resistance to their mechanism of action, represent a promising starting point

for antimalarial drug discovery. In principle, the heme biosensor could be used to assess the

heme-perturbing activity of promising antimalarial compounds. For example, the Malaria Box

consists of 400 compounds shown to potently inhibit parasite growth, although their mechanisms

of action have not been fully elucidated13. In principle, our heme biosensor could be used to

identify compounds within this set that cause heme dysregulation. In order to expedite this

screening process and select candidates for further screening, compounds could first be screened

for their ability to inhibit β-hematin formation in vitro. Potent β-hematin inhibitors could then be

tested for their ability to disrupt heme homeostasis using the heme biosensor, identifying the

most promising compounds for future development.

110

However, screening large numbers of compounds using the approaches detailed in Chapter 2

is not likely to be practical, as microscopy-based heme quantitation is time-consuming and labor-

intensive. Flow cytometry presents an attractive alternative to microscopy for FRET quantitation,

based on its ability to rapidly quantify multiple fluorescence parameters with single-cell

resolution. In addition, flow cytometry has been used to detect FRET between ECFP and EYFP14,

and this assay should in principle be adaptable to high-throughput FRET quantitation. One issue

that may arise in flow cytometric FRET quantitation is CFP signal limitation. Although ECFP

signal could be maximized using a laser-filterset combination optimized for CFP, using a

brighter donor fluorophore could significantly improve detection and signal-to-noise ratio in high

throughput experiments. The CFP mutant Cerulean has been widely used as a FRET donor and is

roughly 2.5x brighter15. A newer cyan fluorescent protein, mTurquoise2, is roughly 3.7x brighter

than ECFP, has a longer fluorescence lifetime16, and has been successfully applied to improve

dynamic range of multiple FRET sensors17. In addition, newer variants of EYFP are available

that display greater brightness, robustness to environmental changes, and faster folding18,19.

These proteins have also been successfully implemented as FRET acceptors20.

Developing a second-generation CH49Y sensor with an alternate FRET pair could, in

conjunction with existing CH49Y, allow simultaneous heme quantitation in multiple parasite

compartments. In principle, parasites expressing these sensors could enable quantitative study of

heme fluxes between organelles both under normal physiological conditions but also under

stresses imposed by antimalarial drugs. Such a technique was recently demonstrated in HeLa

cells using Zn2+ biosensors based on cyan/yellow and red/green FRET pairs21 and could

potentially be accessible here. In any case, each additional donor-acceptor pair would need to be

111

evaluated empirically, as substituting donor-acceptor pairs can adversely affect FRET biosensor

performance22.

Finally, the heme-dependent quenching of ECFP detailed in Chapter 2 presents further

opportunities for FRET biosensor development. FRET quantitation by fluorescence lifetime is

advantageous because lifetime is independent from fluorescence intensity, making measurements

robust to changes in excitation intensity, inner filter effects, and detector sensitivity23. Given that

heme has been shown to be an effective quencher of fluorescence in other protein fusions24,25,

fusing the heme binding domain from CH49Y or CHY to a suitable fluorophore could generate a

robust lifetime sensor for heme. An optimal fluorophore would have high brightness and a

monoexponential decay curve, as has been described for the cyan fluorescent protein mTFP126.

This type of sensor could potentially be used in conjunction with recently-developed high

throughput lifetime measurements27,28 to screen for heme-perturbing antimalarials.

4.4 Heme sensing in other biological systems

Finally, the genetically-encodable heme biosensor described in Chapter 2 should be broadly

applicable across biological systems. For example, the well-studied helminth Caenorhabditis

elegans has generated considerable interest as a model system for heme trafficking and

homeostasis, and has been used to elucidate the function of multiple eukaryotic heme

transporters29,30. Applying this sensor to C. elegans should facilitate detailed, quantitative studies

of heme transport within cells and tissues. Additionally, acquisition of heme iron from the host

is an important virulence determinant in pathogenic organisms31. Using the heme biosensor to

quantify labile heme in these organisms could further understanding of heme homeostasis and

also contribute to screening efforts devoted to discovering new antibiotics.

112

4.5 Synthetic Receptor Development

In our proof-of-concept studies, we were able to restore significant amounts of sialic

acid-containing receptors to the surface of neuraminidase-treated red blood cells. However, only

enzymatic reattachment by an α2,3(O)-sialyltransferase was able to rescue the neuraminidase-

induced invasion defect. One key question that remains to be addressed is why the sialic acid

attached with other methods could not function as a receptor for parasite invasion. Two potential

explanations include (1) that the synthetic receptors did not bind to parasite ligands with

sufficient affinity, and (2) that binding to the synthetic receptors created a junction that could not

facilitate downstream processes necessary for invasion to proceed.

The first hypothesis could be tested by performing in vitro binding assays to recombinantly-

expressed parasite ligands. The domain of EBA-175 that mediates binding to sialic acid (Region

II or RII) has been recombinantly expressed and shown to bind to erythrocytes, where binding is

sensitive to neuraminidase treatment1. A similar assay could be employed here – recombinant

EBA-175 RII could be fluorescently labeled and incubated with treated red blood cells. Binding

could be quantified by flow cytometry, and compared between untreated red blood cells, red

blood cells treated with neuraminidase, and neuraminidase-treated cells where sialic acid had

been attached via the methods described in Chapter 3.

Invasion arrest following junction formation could be observed using microscopy. Giemsa

staining during the invasion process could identify merozoites that had attached to the red blood

cell surface but had not invaded successfully. Electron microscopy could be used to examine the

junction between the merozoite and the red blood cell in more detail, yielding structural

information that could identify precisely where the invasion process is blocked.

113

If the tested synthetic receptor structures did not bind EBA-175 with sufficient affinity to

promote invasion, multiple strategies could be used to identify suitable receptor structures with

higher binding affinity. Sialic acid analogues have been shown to display potent inhibition of

EBA-175 binding to glycophorin A by ELISA2. In principle, these analogues could be

synthesized with aminooxy handles and attached to the red blood cell surface using the methods

developed in Chapter 3. Additional glycan structures with high affinity for EBA-175 could be

identified by screening glycan arrays3 where accumulation of fluorescently-labeled protein could

identify glycans with high binding affinity. Azido-functionalized glycans used in array synthesis4

could then be immobilized on the red blood cell surface using the oxime ligation method

described in Chapter 3, combined with commercially-available bifunctional linkers that contain

aminooxy and alkyne chemical handles.

A defect in synthetic receptor-mediated invasion occurring downstream from initial

merozoite binding would suggest several additional hypotheses. In the case of the immobilized

glycophorins, the size or steric properties of the immobilized receptor complex may impede the

formation of other ligand-receptor interactions necessary for invasion to proceed. In addition, the

hydrophobic transmembrane domains of the extracted glycophorins may promote unfavorable

protein-protein interactions. To address these possibilities, sialylated glycophorin peptides could

be prepared by protease-treating intact red blood cells or extracted glycophorin protein, and then

purifying using anion exchange or lectin affinity chromatography. These purified peptides could

then be immobilized on the red cell surface using the biotin-NeutrAvidin system described in

Chapter 3.

More generally, if binding to the synthetic sialic acid receptors does not promote other

ligand-receptor interactions necessary for productive invasion, combinations of receptors could

114

be investigated. Along with the sialic acid receptors described in Chapter 3, binding peptides

could be immobilized that facilitate ligand-receptor interactions that are known to be

indispensible to invasion. If these binding peptides were presented similarly to the synthetic

sialic acid receptors, perhaps this would increase their simultaneous accessibility to parasite

ligands and facilitate invasion.

4.6 Towards synthetic receptor use for in vitro culture of P. vivax

Finally, the immobilization strategies detailed in Chapter 3 could be applied to a more P.

vivax-like model system. In the closely-related parasite P. knowlesi, like in P. vivax, interactions

between the Duffy antigen and the parasite Duffy-binding protein (DBP) are required for

efficient invasion into human red blood cells. The crystal structure of PkDBP has been solved,

and the region of the Duffy antigen necessary for binding has been identified. This PkDBP-

binding peptide could be synthesized with the appropriate biotin- or aminooxy- handle and

immobilized on the surface of Duffy-negative human red blood cells to facilitate P. knowlesi

invasion. Given that P. vivax also expresses a homologous Duffy-binding protein, these

experiments could be directly translated to P. vivax.

Taken together, the outcomes of further synthetic receptor development in P. falciparum and

P. knowlesi could provide valuable insights toward facilitating P. vivax invasion into mature red

blood cells. If improving ligand-receptor affinity proves advantageous in P. falciparum, this

would suggest that efforts should be directed toward identifying high-affinity binders for the P.

vivax reticulocyte-binding proteins. This could be facilitated by affinity maturation of peptide

libraries using established yeast- bacterial- or bacteriophage-display strategies and recombinant

PvRBP domains. Additionally, pulldown experiments with recombinant PvRBP and peptides

115

released from reticulocytes via proteolysis may be useful in identifying potential PvRBP

receptors.

4.7 Conclusions

This thesis describes the development of novel chemical biology tools to address the growing

issue of drug resistance in the parasites responsible for the vast majority of malaria disease

burden. Using a genetically-encoded heme biosensor, we were able to directly link the

antimalarial activity of chloroquine to the perturbation of cytosolic labile heme for the first time.

We also were able to quantify a cytosolic labile heme pool that remains remarkably stable

throughout blood-stage parasite development, despite the large flux of hemoglobin-heme that

occurs during hemoglobin degradation in the trophozoite stage. Given the historical success of

chloroquine and the apparent lack of resistance to its proposed mechanism of action, perturbation

of cytosolic labile heme is likely to represent a promising phenotype to identify in future screens

of antimalarial compounds. Further insights into parasite heme metabolism enabled by this

biosensor may present additional opportunities for chemotherapeutic intervention.

Finally, by developing generalizable methods to immobilize synthetic receptors on the red

blood cell surface, we demonstrated that supplying the necessary receptor could promote parasite

invasion into cell types that were otherwise inaccessible. More broadly, this suggests that using

synthetic receptors can be useful to elucidate structural and biochemical requirements for red cell

receptors to productively engage parasite ligands. Future designs, when presented in appropriate

biochemical contexts, could promote the invasion of P. vivax into mature red blood cells and

potentially facilitate practical in vitro culture. Together, these tools present new opportunities to

gain fundamental insights into parasite biology and mechanisms of antimalarial drug action,

facilitating discovery of new generations of antimalarial drugs.

116

4.8 References

1. Pandey, K. C. et al. Bacterially expressed and refolded receptor binding domain of Plasmodium falciparum EBA-175 elicits invasion inhibitory antibodies. Mol Biochem Parasitol 123, 23–33 (2002).

2. Bharara, R., Singh, S., Pattnaik, P., Chitnis, C. E. & Sharma, A. Structural analogs of sialic acid interfere with the binding of erythrocyte binding antigen-175 to glycophorin A, an interaction crucial for erythrocyte invasion by Plasmodium falciparum. Mol Biochem Parasitol 138, 123–129 (2004).

3. Blixt, O. et al. Printed covalent glycan array for ligand profiling of diverse glycan binding proteins. Proc Natl Acad Sci USA 101, 17033–17038 (2004).

4. Blixt, O. & Razi, N. Chemoenzymatic synthesis of glycan libraries. Meth Enzymol 415, 137–153 (2006).

5. Sato, S., Rangachari, K. & Wilson, R. J. M. Targeting GFP to the malarial mitochondrion. Mol Biochem Parasitol 130, 155–158 (2003).

6. Waller, R. F., Reed, M. B., Cowman, A. F. & McFadden, G. I. Protein trafficking to the plastid of Plasmodium falciparum is via the secretory pathway. EMBO J 19, 1794–1802 (2000).

7. Nagaraj, V. A. et al. Malaria parasite-synthesized heme is essential in the mosquito and liver stages and complements host heme in the blood stages of infection. PLoS Pathog 9, e1003522 (2013).

8. Kuhn, Y., Rohrbach, P. & Lanzer, M. Quantitative pH measurements in Plasmodium falciparum-infected erythrocytes using pHluorin. Cell Microbiol 9, 1004–1013 (2007).

9. Sigala, P. A., Crowley, J. R., Hsieh, S., Henderson, J. P. & Goldberg, D. E. Direct Tests of Enzymatic Heme Degradation by the Malaria Parasite Plasmodium falciparum. J Biol Chem 287, 37793–37807 (2012).

10. Goldfless, S. J., Wagner, J. C. & Niles, J. C. Versatile control of Plasmodium falciparum gene expression with an inducible protein-RNA interaction. Nat Commun 5, 5329 (2014).

11. Armstrong, C. M. & Goldberg, D. E. An FKBP destabilization domain modulates protein levels in Plasmodium falciparum. Nature Methods 4, 1007–1009 (2007).

12. Pasini, E. M., van den Ierssel, D., Vial, H. J. & Kocken, C. H. M. A novel live-dead staining methodology to study malaria parasite viability. Malar J 12, 190 (2013).

13. Spangenberg, T. et al. The open access malaria box: a drug discovery catalyst for neglected diseases. PLoS ONE 8, e62906 (2013).

14. Banning, C. et al. A flow cytometry-based FRET assay to identify and analyse protein-protein interactions in living cells. PLoS ONE 5, e9344 (2010).

15. Rizzo, M. A., Springer, G. H., Granada, B. & Piston, D. W. An improved cyan fluorescent protein variant useful for FRET. Nature Biotechnology 22, 445 (2004).

16. Goedhart, J. et al. Structure-guided evolution of cyan fluorescent proteins towards a quantum yield of 93%. Nat Commun 3, 751 (2012).

17. Woehler, A. Simultaneous Quantitative Live Cell Imaging of Multiple FRET-Based Biosensors. PLoS ONE 8, e61096 (2013).

18. Griesbeck, O., Baird, G. S., Campbell, R. E., Zacharias, D. A. & Tsien, R. Y. Reducing the environmental sensitivity of yellow fluorescent protein. Mechanism and applications. J Biol Chem 276, 29188–29194 (2001).

19. Nagai, T. et al. A variant of yellow fluorescent protein with fast and efficient maturation

117

for cell-biological applications. Nature Biotechnology 20, 87–90 (2002). 20. van der Krogt, G. N. M., Ogink, J., Ponsioen, B. & Jalink, K. A comparison of donor-

acceptor pairs for genetically encoded FRET sensors: application to the Epac cAMP sensor as an example. PLoS ONE 3, e1916 (2008).

21. Miranda, J. et al. New Alternately Colored FRET Sensors for Simultaneous Monitoring of Zn2+ in Multiple Cellular Locations. PLoS ONE 7, (2012).

22. Palmer, A. E., Qin, Y., Park, J. G. & McCombs, J. E. Design and application of genetically encoded biosensors. Trends in Biotechnology 29, 144–152 (2011).

23. PIETRASZEWSKA-BOGIEL, A. & GADELLA, T. W. J. FRET microscopy: from principle to routine technology in cell biology. J Microsc 241, 111–118 (2011).

24. Takeda, S., Kamiya, N. & Nagamune, T. A novel protein-based heme sensor consisting of green fluorescent protein and apocytochrome b(562). Anal Biochem 317, 116–119 (2003).

25. Koga, S. et al. Development of a heme sensor using fluorescently labeled heme oxygenase-1. Anal Biochem 433, 2–9 (2013).

26. Padilla-Parra, S. et al. Quantitative Comparison of Different Fluorescent Protein Couples for Fast FRET-FLIM Acquisition. Biophysical Journal 97, 2368–2376 (2009).

27. Paterson, M. J. et al. A fluorescence lifetime-based assay for serine and threonine kinases that is suitable for high-throughput screening. Anal Biochem 402, 54–64 (2010).

28. Alibhai, D. et al. Automated fluorescence lifetime imaging plate reader and its application to Förster resonant energy transfer readout of Gag protein aggregation. J Biophotonics 6, 398–408 (2013).

29. Rajagopal, A. et al. Haem homeostasis is regulated by the conserved and concerted functions of HRG-1 proteins. Nature 453, 1127–1131 (2008).

30. Chen, C., Samuel, T. K., Sinclair, J., Dailey, H. A. & Hamza, I. An intercellular heme-trafficking protein delivers maternal heme to the embryo during development in C. elegans. Cell 145, 720–731 (2011).

31. Pishchany, G. & Skaar, E. P. Taste for blood: hemoglobin as a nutrient source for pathogens. PLoS Pathog 8, e1002535 (2012).