Structural Studies on the Lipid Flippase MurJ

by

Alvin Chun Yin Kuk

Department of Biochemistry Duke University

Date:______Approved:

______Seok-Yong Lee, Supervisor

______Richard G. Brennan

______Pei Zhou

______Harold P. Erickson

______Margarethe Kuehn

Dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in the Department of Biochemistry in the Graduate School of Duke University

2018

ABSTRACT

Structural Studies on the Lipid Flippase MurJ

by

Alvin Chun Yin Kuk

Department of Biochemistry Duke University

Date:______Approved:

______Seok-Yong Lee, Supervisor

______Richard G. Brennan

______Pei Zhou

______Harold P. Erickson

______Margarethe Kuehn

An abstract of a dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in the Department of Biochemistry in the Graduate School of Duke University

2018

Copyright by Alvin Chun Yin Kuk 2018

Abstract

The of many important polysaccharides (including , lipopolysaccharide, and N-linked glycans) necessitates membrane transport of oligosaccharide precursors from their cytoplasmic site of synthesis to their site of assembly outside the cytoplasm. To address this problem, cells utilize transporters such as those of the multidrug/oligosaccharidyl-lipid/polysaccharide (MOP) superfamily to flip lipid-linked oligosaccharides across the cytoplasmic membrane. The MOP superfamily member MurJ has been shown to be the flippase that transports the lipid- linked peptidoglycan precursor lipid II, but the lack of structural information has limited our mechanistic understanding of the MurJ transport cycle. We determined the first structure of MurJ (MurJTA from Thermosipho africanus) to 2.0-Å resolution,

which assumed an inward-facing conformation unlike all other outward-facing

structures of MOP transporters. Our structural and mutagenesis studies provide insight

into a putative model of lipid II binding and an alternating-access mechanism of

transport. We followed up with crystal structures of MurJTA captured in inward-closed,

inward-open, inward-occluded and outward-facing conformations at 3.2, 3.0, 2.6 and

1.8-Å resolution, respectively. The structures reveal a lateral gating mechanism and

asymmetric rearrangements that drive transition to the Na+-bound outward-facing state,

providing a structural basis for lipid-linked oligosaccharide transport and a framework

for development of inhibitors.

iv

Dedication

To family, friends, colleagues, and mentors who have inspired and guided me on his journey

v

Contents

Abstract ...... iv

List of Tables ...... ix

List of Figures ...... x

List of Abbreviations ...... xii

Acknowledgements ...... xiv

1. Introduction ...... 1

1.1 Peptidoglycan Biosynthesis ...... 1

1.2 Role of MurJ in Lipid-Linked Oligosacccharide Transport ...... 5

1.3 Transport Mechanisms of Lipid Flippases ...... 8

1.4 Significance of Structural Studies on MurJ ...... 11

2. Crystal Structure of Thermosipho africanus MurJ (MurJTA) ...... 12

2.1 Overview ...... 12

2.2 Results ...... 13

2.2.1 Identification and Optimization of MurJTA for Crystallization ...... 13

2.2.2 Crystallization of MurJTA ...... 15

2.2.3 Structure Determination of MurJTA ...... 18

2.2.4 Overall Architecture of MurJTA ...... 25

2.2.5 The Hydrophobic Groove and Central Cavity ...... 27

2.2.6 Identification of Ions by Anomalous Scattering Experiments ...... 29

2.3 Discussion ...... 33

2.3.1 Model of Lipid II Binding ...... 33

vi

2.3.2 Importance of Alternating Access to MurJTA Function ...... 37

2.4 Materials and Methods ...... 40

2.4.1 Protein Expression and Purification ...... 40

2.4.2 Crystallization ...... 40

2.4.3 Data Collection and Structure Determination ...... 41

2.4.4 Docking of Lipid II and Structure Analysis ...... 42

2.4.5 Complementation Assay and Western Blot ...... 43

3. Structural Basis for the MurJ Transport Cycle ...... 45

3.1 Overview ...... 45

3.2 Results ...... 46

3.2.1 Conformational Landscape of MurJTA ...... 46

3.2.2 Inward-Facing Structures Reveal a Lateral-Gating Mechanism ...... 50

3.2.3 The Inward-Occluded Conformation of MurJTA ...... 52

3.2.4 The Outward-Facing Conformation of MurJTA ...... 56

3.2.5 Sodium Site in the Outward-Facing MurJTA Structure ...... 59

3.2.6 Probing the Conformation of MurJ in Detergent Micelles ...... 63

3.2.7 State-Dependent Changes in the Central Cavity ...... 67

3.3 Discussion ...... 69

3.3.1 Transport Mechanism of MurJ ...... 69

3.3.2 Relevance to Inhibitor Development Targeting MurJ ...... 71

3.4 Materials and Methods ...... 72

3.4.1 Protein Expression and Purification for the New Crystal Forms ...... 72

vii

3.4.2 Crystallization of the New Crystal Forms ...... 72

3.4.3 Data Collection and Structure Determination ...... 73

3.4.4 Docking of Lipid II and Simulation ...... 75

3.4.5 Protein Expression and Delipidation for In Vitro Assay ...... 77

3.4.6 In Vitro Cysteine Accessibility Assay ...... 78

4. Conclusions ...... 80

Appendix A: Alignment of MurJ Sequences ...... 83

References ...... 85

Biography ...... 99

viii

List of Tables

Table 1: Data collection and refinement statistics for the first MurJTA structure...... 24

Table 2: Data collection and refinement statistics for additional MurJTA structures ...... 49

Table 3: Sodium Fo − Fc omit peak heights and coordination distances ...... 61

ix

List of Figures

Figure 1: Membrane-associated steps of peptidoglycan biosynthesis...... 2

Figure 2: The MOP transporter superfamily ...... 6

Figure 3: Role of MurJ in peptidoglycan synthesis ...... 7

Figure 4: Ortholog screening strategy for MurJ ...... 14

Figure 5: Type I and II packing of membrane protein crystals ...... 16

Figure 6: Crystals of MurJTA grown in lipidic cubic phase ...... 17

Figure 7: Electron density calculation and the phase problem ...... 19

Figure 8: Merging data from multiple crystals enabled structure solution of MurJTA ...... 21

Figure 9: Experimental phasing of MurJTA from SeMet-substituted crystals ...... 23

Figure 10: Architecture of MurJTA ...... 26

Figure 11: Positive-inside electrostatics support the inward-facing topology ...... 27

Figure 12: The hydrophobic groove of MurJTA is linked to the central cavity ...... 28

Figure 13. Identification of chloride ions by bromide soak ...... 29

Figure 14: Deduction of a Zn2+ ion bound to the beginning of TM 7 ...... 30

Figure 15: Deduction of putative Ca2+ and Na+ ions ...... 32

Figure 16. Unmodeled electron density peaks in the portal and central cavity ...... 33

Figure 17: Model of lipid II binding to MurJTA ...... 35

Figure 18. MTSES-sensitive positions in the central cavity ...... 36

Figure 19: Alternating access is important for MurJ function...... 39

Figure 20: Four new crystal forms of MurJTA were obtained ...... 47

Figure 21: Structure determination of outward-facing MurJTA ...... 48

x

Figure 22: Crystal structures elucidate the conformation landscape of MurJTA ...... 48

Figure 23: Lateral gating mechanism in the inward-facing state ...... 51

Figure 24: The inward-occluded conformation of MurJTA ...... 54

Figure 25: Docking and MD simulation of lipid II to the inward-occluded structure ...... 55

Figure 26: The outward-facing conformation of MurJTA ...... 57

Figure 27: Structural features of the putative lipid- in the outward state...... 58

Figure 28: Sodium site in MurJTA ...... 60

Figure 29: Comparison of the sodium site in MurJTA and in PfMATE ...... 62

Figure 30: Cysteine accessibility profile of MurJTA by direct PEG-maleimide labeling .... 64

Figure 31: Accessibility profile of MurJTA in Na+ or in NMDG+ by inverse labeling ...... 66

Figure 32: State-dependent conformation changes in the central cavity ...... 68

Figure 33: Model of the MurJ transport mechanism ...... 70

Figure 34: Alignment of MurJ sequences...... 84

xi

List of Abbreviations

BME β-mercaptoethanol

Cymal-6 Cyclohexyl-hexyl maltoside, a non-ionic detergent

DDM Dodecyl maltoside, a non-ionic detergent

DM Decyl maltoside, a non-ionic detergent

DMNG Decyl maltoside neopentyl glycol, a non-ionic detergent

DTT Dithiothreitol

GlcNAc N-acetyl glucosamine

IPTG Isopropyl β-D-thiogalactopyranoside

LCP Lipidic cubic phase, a method for crystallizing membrane proteins

MATE Multidrug and toxic compound extrusion family

MBP Maltose-binding protein

MOP Multidrug/oligosaccharidyl-lipid/polysaccharide superfamily

MTSES Sodium 2-sulfonatoethyl methanethiosulfonate, a cysteine-reactive reagent

MurNAc N-acetyl muramic acid

MVF Mouse virulence family, contains MurJ

NBD 7-nitrobenzofurazan, a fluorophore

NMDG N-methyl-D-glucamine

PBP Penicillin-binding protein

PEG Polyethylene glycol

xii

PEG-Mal Methoxy-polyethylene glycol 5000 maleimide

POPE 1-Palmitoyl-2-oleoyl-sn-glycero-3-phosphoethanolamine, a phospholipid

PPX PreScission protease, specifically cleaves at LEVLFQ/GP (called a PPX site)

SAD Single-wavelength anomalous diffraction

SEDS Shape, elongation, division, and sporulation family, contains FtsW/RodA

SeMet Selenomethionine

TCEP Tris(2-carboxyethyl)phosphine

TM Transmembrane helix

UDP Uridine diphosphate

WT Wild-type

xiii

Acknowledgements

This work would not have been possible without the invaluable contribution from many people, to whom I give my heartfelt thanks. First and foremost, to my supervisor and mentor, Seok-Yong Lee, for his unwavering guidance and support—I owe much of my development as a scientist to him. To my labmates over the years, who have made the lab a joyful place even during difficult times: Ellene Mashalidis, Jiho Yoo,

Lejla Zubcevic, Marscha Hirschii, Ben Chung, Zachary Johnson, Yang Suo, Ying Yin, Aili

Hao, Nicholas Wright, and William Borschel. To my thesis committee for invaluable insight and feedback: Richard Brennan, Harold Erickson, Meta Kuehn, and Pei Zhou.

I would like to acknowledge Ellene Mashalidis and Aili Hao for performing complementation experiments, Aili Hao for lipid accumulation experiments, and

Ziqiang Guan for lipid mass spectrometry experiments. Their contributions have been essential for our current understanding of MurJ beyond a structural biology standpoint.

My thanks go to Natividad Ruiz (The Ohio State University) for sharing Escherichia coli strains NR1154 and NR1157. As well as to Will Thompson and Laura Dubois for protein mass spectrometry analysis, Nathan Nicely for coordinating synchrotron beamtime, and synchrotron staff at the Advanced Photon Source for assistance in data collection. Last but not least, to family and friends who have kept me going, both during the good times and the bad, on this wonderful journey.

xiv

1. Introduction

1.1 Peptidoglycan Biosynthesis

Peptidoglycan (PG) is a unique protective matrix enveloping the vast majority of bacterial cells, conferring rigidity and resistance to osmotic pressure as the main component of the bacterial wall. Our current understanding of peptidoglycan biosynthesis is built on the work of numerous individuals and research groups, as recently reviewed by [1, 2]. Peptidoglycan biosynthesis begins in the cytoplasm, where the water-soluble precursor UDP-MurNAc-pentapeptide is synthesized by the combined activities of the muramyl MurA-F [3]. The integral membrane

MraY transfers the phospho-MurNAc-pentapeptide moiety from UDP-MurNAc- pentapeptide to the lipid carrier undecaprenyl-phosphate, forming lipid I [4]. This step is notable as it translocates the soluble precursors to the inner leaflet of the cytoplasmic membrane (Figure 1). Subsequently, the peripheral membrane protein MurG adds a

GlcNAc moiety from UDP-GlcNAc, forming undecaprenyl-disphosphate-MurNAc- pentapeptide-GlcNAc, also known as lipid II [5]. Lipid II must be subsequently flipped from the cytoplasmic side to the periplasmic side of the membrane. Once flipped, the

GlcNAc-MurNAc-pentapeptide moiety of lipid II is incorporated into peptidoglycan by the glycosyltranferase activities of either shape, elongation, division and sporulation

(SEDs) proteins [6-8] or bifunctional (class A) penicillin-binding proteins (aPBPs) [9].

1

Glycan chains are then cross-linked between their stem peptides by transpeptidase activities of either bifunctional or monofunctional (class B) penicillin-binding proteins

(bPBPs) [10]. Finally, an undecaprenyl-diphosphate phosphatase such as BacA

regenerates the lipid carrier [11], which is flipped back into the cytoplasm by an unknown mechanism [12], possibly also mediated by BacA [13, 14]. Peptidoglycan biosynthesis has a proven track record as an excellent antibiotic target, with numerous drug scaffolds targeting this pathway (Figure 1).

Figure 1: Membrane-associated steps of peptidoglycan biosynthesis

While the cytoplasmic and periplasmic steps of peptidoglycan biosynthesis are

relatively well studied, little is known about the membrane transport steps—the size,

charge, and flexibility of lipid II would pose a substantial energy barrier to uncatalyzed

flipping. Indeed, spontaneous flipping of lipid II (albeit labeled with the fluorophore

2

NBD) in liposomes could not be detected over a 3 hour period [15]. In contrast, peptidoglycan polymerization in an actively growing E. coli cell requires an estimated

5,000 lipid II molecules to be flipped per second [12], out of a total pool of only 1,000-

2,000 molecules at any instant [16]. While this efficiency can be attributed to the increase in local substrate concentrations on the membrane—a major advantage of this lipid- linked system—it would not be achievable without dedicated transport machinery [17].

Two major candidates for the flippase were originally proposed: FtsW/RodA (part of the

SEDS family) [18, 19], and MurJ [20-22]. A third candidate (Amj), conditionally expressed under control of envelope-stress-response sigma factor σM, has been postulated to substitute the canonical flippase in Bacillus subtilis, but is only present in a

subset of [23].

FtsW/RodA are ~45 kDa integral membrane proteins containing 10

transmembrane helices [24, 25]. Lipid II flippase function was proposed based on data

from an in vitro flippase assay on FtsW [18, 19]. This claim was substantiated by indirect

evidence coming from close genetic linkage to other peptidoglycan synthesis genes [26,

27], essentiality and involvement in cell morphology [28, 29], as well as interaction with

the penicillin-binding protein FtsI [30-33]. However, FtsW-dependent flippase activity

could not be demonstrated in vivo [22], and the FtsW paralog RodA has recently been

characterized as a glycosyltransferase [6-8], which would also be consistent with its gene

location, importance to morphology, and interplay with PBPs.

3

MurJ is a ~55 kDa integral membrane protein containing 14 transmembrane helices [34] and is encoded by the gene mviN. MurJ has been estimated to be present at

~80 or ~670 copies per Escherichia coli cell when grown in minimal or rich medium, respectively [35]. Two groups proposed MurJ to be the lipid II flippase in 2008. Ruiz identified MurJ from a bioinformatics screen and showed that MurJ was essential, with

MurJ-depleted bacteria accumulating peptidoglycan precursors before lysing [20].

Inoue and colleagues isolated a temperature-sensitive mutant located immediately upstream of mviN (mutation suggested to reduce expression level), also finding

peptidoglycan precursor accumulation [21]. In 2014, Ruiz and colleagues showed MurJ- dependent flipping of radiolabeled lipid II in E. coli cells and spheroplasts [22]. They identified five positions that (when mutated to cysteine) were sensitive to the cysteine-

reactive reagent sodium 2-sulfonatoethyl methanethiosulfonate (MTSES), resulting in cell lysis or cell shape defects [22].

While an increasing body of evidence has supported MurJ as the lipid II flippase, controversy over the flippase identity has highlighted a considerable gap in knowledge of this critical step of peptidoglycan biosynthesis. Despite headway in genetics studies of the flippase, a dearth of structural information has limited our understanding of flippase identity and mechanism. The outward-facing cavity of MurJ has been probed extensively by the substituted cysteine accessibility method [34, 36], but the substrate- binding inward-facing conformation has not been elucidated. Determining crystal

4

structures of MurJ would pave the way for understanding MurJ function at the atomic level, filling in this gap of knowledge in peptidoglycan biosynthesis.

1.2 Role of MurJ in Lipid-Linked Oligosacccharide Transport

MurJ is also important from a membrane transport standpoint. Biosynthesis of many important polysaccharides (not just peptidoglycan, but also lipopolysaccharide,

N-linked glycans, etc.) necessitates membrane transport of oligosaccharide precursors

from their cytoplasmic site of synthesis to their site of assembly outside the cytoplasm.

To accomplish this task, cells utilize a to attach the precursor to a lipid

carrier, and a flippase to transport the lipid-linked precursor across the membrane. Such

lipid-linked transport systems are evolutionarily ancient and found across all domains

of life from bacteria to humans [37]. Transporters of the multidrug/oligosaccharidyl-

lipid/polysaccharide (MOP) superfamily constitute the flippase component of many

lipid-linked transport systems, and are distinct from other transporter superfamilies

such as the major facilitator superfamily. MOP transporters mediate the export of

numerous compounds of physiological or pharmacological importance (Figure 2),

including lipid-linked precursors for the biosynthesis of bacterial (MurJ, MVF family) [20-22], cell surface polysaccharides (Wzx, PST family) [38-41], eukaryotic N-

linked glycosylation (Rft1, OLF family, activity contested) [42-45], as well as expulsion of

drug molecules (the ubiquitous MATE family) [46].

5

Figure 2: The MOP transporter superfamily

The MOP superfamily member MurJ was previously shown to be the flippase for bacterial cell wall synthesis, transporting the lipid-linked peptidoglycan precursor lipid

II (Figure 3A) [20-22]. Lipid II is synthesized by the combined activities of MraY and

MurG, which respectively transfer phospho-MurNAc-pentapeptide and GlcNAc

moieties to the lipid carrier undecaprenyl-phosphate [47, 48], resulting in a large

molecule that is highly flexible and negatively charged (Figure 3B). Binding of lipid II to

the MurJ from Escherichia coli has been established by native mass spectrometry with an

in vitro dissociation constant of approximately 3 µM [49]. Upon flipping to the

6

Figure 3: Role of MurJ in peptidoglycan synthesis

(A) Lipid-linked transport system: (e.g. MraY, PDB 4J72 and MurG, 1F0K) attach the cargo to a carrier lipid, which is then flipped by a transporter. MurJ flips lipid-linked peptidoglycan precursor (lipid II) across the cytoplasmic membrane for bacterial cell wall synthesis. were rendered with QuteMol [50]. (B) Chemical structure of lipid II, highlighting the large size, charges, and flexibility.

7

periplasmic side of the membrane by MurJ, the GlcNAc-MurNAc-pentapeptide moiety is incorporated into peptidoglycan and the lipid carrier is recycled. While MurJ exemplifies the flippase component of an important lipid-linked oligosaccharide transport system, it is still not known how MurJ translocates the large, flexible, and negative-charged lipid II molecule across the membrane. A variety of mechanisms have

been put forth for other lipid flippases, as discussed next.

1.3 Transport Mechanisms of Lipid Flippases

Lipid transporters have been proposed to operate by a diverse variety of molecular mechanisms, which would be briefly reviewed here as they would be relevant to our discussion on MurJ. The ABC-superfamily flippase MsbA is among the most well-studied transporter systems and is characterized by “alternating access” between the inward-facing and outward-facing conformations. MsbA flips the core-lipid A anchor for lipopolysaccharide biosynthesis in Gram-negative bacteria [51, 52]. MsbA function is ATP-dependent and inhibited by vanadate [53], with conformational transitions during the ATP binding/hydrolysis cycle probed by spectroscopic experiments [54, 55]. Crystal structures of MsbA [56] and another ABC transporter,

Sav1866 [57], have provided insight into how they couple the ATP binding/hydrolysis cycle to an alternating-access mechanism that drives substrate transport. This is further substantiated by a recent cryo-EM structure of MsbA with electron density of the

8

putative substrate visualized deep in the inward-facing transport cavity, suggesting that the hydrophobic lipid A tails point up into the cavity apex rather than the headgroup

[58]. In contrast to MsbA, only the outward-facing state appeared to be important for the ABC-superfamily flippase PglK [59], which transports lipid-linked oligosaccharides

(LLOs) for a bacterial variant of N-linked glycosylation in Campylobacter jejuni [60].

Other lipid transporters operate by a so-called “credit card” mechanism, in which the transporter provides a hydrophilic groove for the passage of the lipid headgroup, while the hydrophobic lipid tail remains in the membrane. Such a mechanism would obviate the need for alternating-access transitions of the transporter, and is commonly observed for phospholipid scramblases. Examples of transporters proposed to operate by this model include opsin [61, 62], P4-ATPases [63], and scramblase members of the TMEM16 family [64, 65].

Finally, some lipid transporters utilize a “pore” mechanism distinct from the alternating-access and credit-card mechanisms, where an amphipathic translocation channel allows movement of lipids. This mechanism has been proposed for the outer- membrane asymmetry regulator MlaA [66] and the inner-membrane O-antigen polysaccharide Wzm-Wzt [67] transporter. In MlaA, this channel only spans part of the outer membrane span, but in Wzm-Wzt the channel is transmembrane with constrictions.

9

Although the alternating-access, credit card, and pore mechanisms of lipid

flipping are distinct, fundamentally they meet similar requirements that is required for

any lipid transport mechanism: (1) a translocation pathway for the hydrophilic

headgroup through the membrane, (2) topology and space considerations for the

hydrophobic tail moiety during headgroup translocation (be it bound to the transporter

or in the membrane), and (3) some selectivity mechanism to prevent unwanted flux of

solutes, solvent, or other lipids.

The mechanism of lipid II flipping by MurJ is not known: although MATE drug

exporters in the same MOP superfamily have been proposed to utilize an alternating-

access mechanism [68-72], no inward-facing structure has been determined for this

superfamily. While other MOP transporters have been characterized to be Na+ and/or

H+ coupled [40, 68, 73, 74], the energetic factor(s) that drive MurJ transport remain unknown. It has been shown that MurJ activity is dependent on membrane potential but not on the H+ gradient [75], but involvement of Na+ has not been ruled out. Taken

together, the mechanism of lipid II transport by MurJ and energy-coupling (if any is required) remains an important open question in the field.

10

1.4 Significance of Structural Studies on MurJ

The aforementioned genetics and in-vivo biochemical evidence point to an important role of MurJ in bridging the cytoplasmic and periplasmic steps of peptidoglycan biosynthesis, which remains a gap in our knowledge of the pathway.

Because MurJ is the flippase in an important prototype lipid-linked transport system, understanding MurJ structure and function would also provide insight into these transport systems and how they regulate the underlying biosynthetic pathways that govern cell physiology. Additional evidence from recent studies further underpin the importance of MurJ. Recruitment of MurJ to the septum drives the fast constriction step of cytokinesis in Staphylococcus aureus, suggesting that it is also a key player in bacterial cell division [76]. MurJ is the target of newly-discovered inhibitors such as humimycins

[77, 78] and phage M lysis protein [79], both of which block its activity to exert their bactericidal effects. To better understand the lipid II flippase mechanism, we first set out to determine crystal structures of MurJ. Our structural analyses, together with functional studies probing the in vitro conformational state and in vivo transport competency of MurJ mutants, would allow us to formulate a transport model for MurJ.

Taken together, this work would provide a framework for future biochemical studies and inhibitor development targeting this important membrane protein.

11

2. Crystal Structure of Thermosipho africanus MurJ (MurJTA)

2.1 Overview

To better understand the transport mechanism of MurJ, we set out to determine a

crystal structure of MurJ. We screened 18 MurJ orthologs from hyperthermophile

bacteria and identified MurJ from Thermosipho africanus (MurJTA) as a promising candidate. We crystallized MurJTA by the lipidic cubic phase (LCP) method and

determined experimental phases to 3.5-Å resolution by single-wavelength anomalous

diffraction (SAD) using selenomethionine(SeMet)-substituted protein. We determined

the final structure of native MurJ to 2.0-Å resolution.

The crystal structure of MurJTA revealed an inward-facing conformation, unlike the outward-facing conformation of all known crystal structures in the MOP superfamily. A hydrophobic groove is formed by two C-terminal transmembrane helices, which leads into a large central cavity that is mostly cationic. Our studies not only provide the first structural glimpse of MurJ but also suggest that alternating access is important for MurJ function, which may be applicable to other MOP superfamily transporters. Results for this chapter were published in the February 2017 issue of

Nature Structural and Molecular Biology [80]. I would like to acknowledge the

invaluable contribution from Ellene Mashalidis for the MurJ complementation assay.

12

2.2 Results

2.2.1 Identification and Optimization of MurJTA for Crystallization

Our first objective was to determine the molecular structure of a MurJ ortholog, which would provide the impetus for further mutagenesis and functional studies.

Because MurJ is a 55 kDa protein and is not known to oligomerize, we reasoned that X- ray crystallography would be a suitable technique for structure determination. Given that membrane proteins are often poorly-expressed, unstable in vitro, and refractory to crystallization, we adopted a screening approach among hyperthermophile orthologs to maximize our chances of finding a promising candidate for crystallographic studies. We first screened 10 hyperthermophile orthologs for expression in E. coli and stability in dodecyl maltoside (DDM) detergent micelles. We identified these orthologs by a

bioinformatics approach (Figure 4), including MurJ sequences from three bacterial phyla

(Phylum Aquificae, Phylum Deinococcus-Thermus, and Phylum Thermotogae).

Synthesized genes were cloned into a custom pET26 expression vector such that

the final expression construct was His10-MBP-PPX-MurJ-PPX-His10, containing maltose-

binding protein (MBP) for solubility enhancement, polyhistidine (His10) tags for affinity capture, and flanked by PreScission protease sites (PPXs). The majority of the orthologs did not express, or expressed very poorly, in E. coli C41 (DE3) at 1 L scale. However,

MurJ from Thermosipho africanus (MurJTA) expressed decently, albeit not spectacularly.

We were able to extract and purify MurJTA in DDM from these 1 L cultures. After

13

overnight removal of MBP by PPX cleavage, proteins were analyzed by gel filtration on a Superdex 200 column. A sharp monodisperse peak consistent with the predicted size of MurJ in a DDM micelle is indicative of a promising ortholog for crystallization.

MurJTA produced such a result from our initial screen (Figure 4B), but attempts to fine-

tune the screen around MurJTA did not identify additional promising orthologs.

Figure 4: Ortholog screening strategy for MurJ

(A) Ten MurJ sequences from hyperthermophile phyla were first selected and their genes were synthesized (denoted by circles containing “1”). MurJTA is denoted by an asterisk (*). (B) Orthologs were screened for expression in E. coli at (1 L culture scale) and stability in DDM micelles by gel filtration and SDS-PAGE. From results of the initial round of screening, 8 more orthologs were screened in a second round of screening (denoted by “2”), but additional promising orthologs were not identified.

14

Having identified MurJTA as a candidate that expressed decently and was stable in DDM micelles, we next optimized the protein preparation to make it more amenable to crystallization by two approaches: (1) detergent screening and (2) limited proteolysis analysis. Smaller micelles are generally more conducive to crystallization by vapor-

diffusion methods but are harsher on protein stability, and specific properties of the

detergent headgroup such as charge could matter in a protein-specific way [81, 82].

DDM-purified protein was diluted 10-fold into buffers containing various detergents,

including maltosides, glucosides, and zwitterionic detergents. Protein behavior was

subsequently evaluated by gel filtration. MurJTA was found to be stable in not just DDM,

but also the maltoside derivatives decyl maltoside (DM), decyl maltose neopentyl glycol

(DMNG), and Cymal-6. Limited proteolysis experiments were also performed to

identify protease-sensitive regions that could possibly indicate unstructured regions that

could disrupt crystal packing. Unlike some other proteins we had experience with, no

cleavage of MurJTA was observed even at the highest concentrations of trypsin or chymotrypsin used, and thus we did not pursue any termini or loop truncation.

2.2.2 Crystallization of MurJTA

Purified protein was concentrated and used for vapor diffusion crystallization

experiments, which yielded no initial hits in all the detergents tested. At the same time,

we performed crystal screening by the lipidic cubic phase (LCP) method [83, 84]. In

15

contrast to end-to-end (type II) packing of membrane protein micelles commonly seen in crystals grown by vapor diffusion (Figure 5), the LCP method provides a membrane-like

mesophase environment which promotes more optimal side-by-side (type I) packing

mediated by the host lipid mono-olein [82, 85]. Protein was mixed with molten mono-

olein in 2:3 w/w water/mono-olein ratio by the two-syringe method [84]. Boluses were

dispensed onto 96-well glass sandwich plates (which do not dehydrate over months) by

a robot and overlaid with precipitant solutions from a crystal screen made in-house.

Figure 5: Type I and II packing of membrane protein crystals

Illustration of type I and type II packing of a hypothetical membrane protein (in gray). Crystals obtained in LCP generally exhibit type I packing characterized by both top- down stacking and lateral lipid-mediated interactions. Crystals in detergent micelles often exhibit type II packing with fewer packing interactions. In this example, no protein-protein interactions are possible and thus is unable to produce type II crystals.

We obtained initial crystal hits by the lipidic cubic phase method. Crystals were small as expected for LCP (~20-50 µm in length) but were three-dimensional. These hits required calcium in the precipitant solution and were obtained in 30–45% (v/v) PEG 400,

200–400 mM CaCl2, pH 6.5–9.0. We were able to obtain crystals using DDM, DMNG, or

16

Cymal-6 to deliver protein to the mesophase. Upon cutting the glass coverslip, single crystals were harvested using a rigid micromount loop and flash-frozen in liquid

nitrogen without additional cryoprotectant. Crystals were screened using microfocused

X-ray beams at beamlines 24-ID-C and 24-ID-E (Advanced Photon Source). Because the

frozen mesophase obscures the crystals from view, each loop was exposed to the

microfocused beam in a grid scan to locate the crystals by diffraction. Single frames

were collected, from which a suitable data collection strategy was devised and executed.

Initial screens in DMNG produced chunky crystals that diffracted to 2.4-Å resolution,

for which complete datasets could be collected (Figure 6).

Figure 6: Crystals of MurJTA grown in lipidic cubic phase

One of the hits is shown, with crystals showing optical birefringence between crossed polarizers (left). Crystals were grown in a bolus of viscous mono-olein:water mesophase immersed in a drop of precipitant solution, sandwiched between two glass layers. Crystals were tiny (20-50 µm) as is usual for LCP crystals. These crystals diffracted to ~2.4-Å resolution (right), but spots were weak and those beyond 2.7-Å resolution (circle) were not visible to the eye. A lipid ring was observed at 4.5-Å resolution.

17

We attempted various approaches to improve crystal size and diffraction quality, including finer precipitant screens, additive screen, and addition of the putative substrate. Addition of lipid II to the mesophase improved the resolution from 2.4 Å to

2.0 Å and was performed for native crystals as follows: lipid II (BaCWAN, University of

Warwick) was added to mono-olein powder in a glass vial to a final concentration of 1.5 mM (in molten mono-olein). The mono-olein powder was dissolved in chloroform and mixed thoroughly to homogenize the lipids. This mixture was dried under argon and washed with pentane. Finally, organic solvent was removed by complete drying before the lipids were melted. Crystallization of MurJ using this lipid II-doped mono-olein and data collection were performed as described above.

2.2.3 Structure Determination of MurJTA

Electron density at every grid point xyz in the crystallographic unit cell is calculated by Fourier transform of all the structure factors for lattice planes hkl (Miller indices) reflecting the X-ray beam (Figure 7). While structure factor amplitudes, |Fhkl|,

are proportional to the square-root of reflection intensities (negative recorded intensity

values must be dealt with [86]), the phase component of each structure factor, hkl, is lost

α during the experiment, in what is known as the crystallographic phase problem [87].

Furthermore, phases dominate over structure factor amplitudes, and introduction of

inaccurate phase estimates readily leads to the problem of model bias [87].

18

Figure 7: Electron density calculation and the phase problem

To address the phase problem, we first attempted molecular replacement (MR).

For this approach, initial MR phases were obtained from a search model which should

have substantial similarity to what is inside the crystal, with the caveat of model bias

[88]. We utilized available crystal structures of MATE drug efflux pumps in the same

MOP superfamily as search models. MR was unsuccessful as the resulting electron density maps appeared completely biased by the search model and no new features were revealed. This was likely due to the low similarity between MurJ and these available structures (< 20% sequence identity). Efforts to improve the search model by removing loops and/or truncating sidechains did not improve the result.

We thus set out to obtain experimental phases, in which an experimental source of information such as anomalous scatterers is used for initial phase estimates. Since

MurJTA contains 9 methionines out of 475 residues in total, we first tried using single- wavelength anomalous diffraction (SAD) with selenomethionine(SeMet)-substituted protein [89] together with in-silico density modification techniques to break the phase ambiguity [90]. SeMet-substituted protein was expressed by autoinduction in the

19

methionine auxotroph E. coli B834 [91], and purified/crystallized as for native protein,

except with 2 mM Tris(2-carboxyethyl)-phosphine (TCEP) added to the final sample to protect from oxidation. While the SeMet-substituted crystals diffracted to 3.2-Å resolution when exposed to an X-ray beam with a wavelength (0.9791 Å) at the absorption peak near the Se K-edge, the anomalous signal was almost negligible.

Addition of up to 5 additional methionine residues (mutated from leucine or isoleucine) did not improve the anomalous signal. SeMet incorporation at some positions was detected by mass spectrometry (Proteomics and Metabolomics Shared Resource), although overall incorporation efficiency remained unknown due to poor peptide coverage. We realized that experimental phasing for LCP was a non-trivial undertaking—only 10 experimentally-phased structures from LCP crystals were available in 2014 [92]. We attribute our problem to weak diffraction from tiny LCP crystals and concomitant low anomalous signal, compounded by low symmetry (C2) and radiation damage.

The breakthrough was to merge datasets from multiple SeMet crystals, as pioneered by the Hendrickson group [93] and utilized for the phasing from LCP crystals of DgkA [92]. Because of the high isomorphism between crystals, we merged selected frames from 12 SeMet crystals to generate a dataset with over 50-fold average redundancy and signal extending to 3.2-Å resolution. This merged dataset exhibited stronger anomalous signal than any individual dataset (Figure 8A). Eight selenium sites

20

were located by SHELXD [94], and initial SAD phasing was performed to a high- resolution cutoff of 5.0 Å in Phenix.AutoSol [95, 96], producing a figure of merit of 0.5.

The solvent-flattened [90] map calculated with the merged data was of much better

quality than that from a single crystal (Figure 8B).

Figure 8: Merging data from multiple crystals enabled structure solution of MurJTA

(A) Merging data from multiple crystals increased the observable anomalous signal, represented in the y-axis by the parameter , which is the average anomalous difference divided by its estimated standard deviation. Random noise would have a of around 0.8. (B) This increase in anomalous signal translates to better initial phases and a map that allowed model building and phase extension. Both maps shown

21

are density-modified (solvent-flattened) Fo maps calculated up to 5.0-Å resolution and contoured to 1.0 σ, without any protein model phases. Selenium sites are shown as dots.

This map allowed us to utilize a MR-SAD and phase-extension strategy to

improve the phases (Figure 9). Because proteins in the MOP superfamily are known to

be α-helical, we used this a-priori information to dock ideal α-helical fragments into the

5.0-Å resolution map. Phases and map were recalculated, and the model was improved

and extended using the new map. This process was reiterated while gradually

increasing the resolution to 3.5 Å in Phenix.AutoSol [96] and the CCP4 program DM

[97]. Eventually, the backbone for all 14 transmembrane helices could be built. The

eight selenium positions, together with a-priori knowledge of the MOP transporter fold

from MATE crystal structures, allowed their unambiguous assignment. The helical

register was determined using SeMet positions when available, or other information

such as stretches of multiple bulky aromatic residues in the protein sequence.

Because the SeMet crystals were not very isomorphous with the native crystals,

we transferred the model and phases from the SeMet data to the high-resolution (2.0 Å)

native data by molecular replacement instead of phase extension. We refined the

structure against this native data to a final Rwork/Rfree of 18.9 / 21.4 (Table 1), good

geometry (98% Ramachandran favored, 0% Ramachandran outliers), and minimal

clashes (clashscore of 1). The final electron density map allowed us to unambiguously

22

build residues 4-470 of MurJTA out of a total of 475 residues, as well as place a shell of

ordered mono-olein and water molecules around the protein.

Figure 9: Experimental phasing of MurJTA from SeMet-substituted crystals

(A) Anomalous difference Fourier electron density peaks for selenium atoms are shown in magenta mesh contoured to 4 σ. (B) Initial density-modified SAD map calculated from the SeMet data up to 5.0 Å resolution without any model phases. (C) Phase-extended electron density map calculated from the SeMet data up to 3.5 Å resolution, shown here with the working model consisting of α-helical fragments. (D) The 2Fo − Fc electron density map calculated from native data up to 2.0 Å resolution with the final refined model, shown here as a close-up view of the central cavity. The map is contoured to 1.0 σ. Water molecules are shown as gray spheres.

23

Table 1: Data collection and refinement statistics for the first MurJTA structure

Native (PDB 5T77)a SeMetb Data collection Space group C2 C2 Cell dimensions a, b, c (Å) 94.16, 99.57, 74.66 93.86, 99.74, 75.67 α, β, γ (°) 90, 112.91, 90 90, 101.84, 90 Wavelength (Å) 0.9791 0.9791 Resolution (Å) 100–2.00 (2.07–2.00)c 70–3.15 (3.26–3.15)c Rpim (%) 6.5 (55.0) 4.2 (25.0) I/σ(I) 9.47 (1.06) 30.8 (4.10) CC1/2 0.99 (0.50) 1.00 (0.87) Completeness (%) 98.0 (94.0) 99.5 (95.0) Redundancy 4.2 (2.2) 50.4 (30.5)

Refinement Resolution (Å) 34.2–2.00 (2.07–2.00)c No. reflections 41992 (4018) Rwork / Rfree (%) 18.9 / 21.4 No. atoms 4404 Protein 3743 Mono-olein/PEG/ 453/48/5/1/3/2 Cl−/Zn2+/Ca2+/Na+ Water 149 B factors Protein 38.2 Mono-olein/PEG/ 68.5 Cl−/Zn2+/Ca2+/Na+ Water 43.2 R.m.s. deviations Bond lengths (Å) 0.003 Bond angles (°) 0.55 a Merged from 3 native crystals grown in LCP. b Merged from 12 SeMet crystals grown in LCP. c Values in parentheses are for highest-resolution shell.

24

2.2.4 Overall Architecture of MurJTA

Our structure of MurJTA is strikingly different from existing MATE structures.

The structure of MurJTA has the N-lobe (TMs 1–6) and C-lobe (TMs 7–12) arranged in an inward-facing N-shape conformation (Figure 10A) rather than the outward-facing V-

shape conformation observed in all existing MATE transporter structures [69-72, 98].

The inward-facing topology is supported by cytoplasmic orientation of the termini

consistent with MATE transporters, and by the positive-inside surface electrostatics

often seen for membrane proteins (Figure 11). Although strong pseudo-twofold

rotational symmetry is observed between the two lobes in the MATE structures [69, 70,

72], the symmetry between the N-lobe and C-lobe of MurJTA is significantly distorted

(Figure 10B). In particular, the conformations of TM 1 and TM 2 are very different from

those of the symmetry-related TM 7 and TM 8. TM 1 extends out from the transporter

core instead of interacting with the C-lobe as in other MATE structures, while TM 2 is

broken at the Gly/Ala–Glu–Gly–Ala motif that is conserved in Gram-negative bacteria

(Appendix A), allowing this region to protrude into the central cavity. In addition, both lobes of the core transport domain of MurJTA are substantially wider and shorter than those in MATE transporters.

25

Figure 10: Architecture of MurJTA

(A) Structure of MurJTA. Front view (left) and top view (right) are shown. The N-lobe is colored in blue, C-lobe in green, and TMs 13-14 in brown. The N-lobe and C-lobe are related by pseudo-twofold rotational symmetry with axis normal to the membrane. This symmetry is notably distorted at TMs 1, 2, 7, and 8. (B) Structural superposition between the N-lobe and C-lobe for MurJTA and PfMATE (PDB 3VVN).

26

Figure 11: Positive-inside electrostatics support the inward-facing topology

Electrostatic surface of MurJTA. Electrostatics were calculated by the Adaptive Poisson- Boltzmann Solver [99] and shown from −10 kT (red, anionic) to +10 kT (blue, cationic).

2.2.5 The Hydrophobic Groove and Central Cavity

C-terminal TMs 13 and 14 extend out from the transporter core and are oriented

beside TM 9. These two additional TMs, absent in MATE proteins, form a curved

groove (Figure 12A) that is ~20 Å long. This groove is lined mostly with hydrophobic

residues from TMs 1, 8, 9, and 14. The hydrophobic groove penetrates into the central

transport cavity via a membrane portal located between TMs 1 and 8 (Figure 12B). Most

of the residues lining the portal have small side chains, with the notable exception of

Arg18, which is essential for E. coli MurJ function [36]. Beyond the portal, the groove opens up into a large inward-facing central cavity ~20 Å wide and penetrating up to ~20

Å into the membrane (Figure 12C). This cavity can be subdivided into a proximal site

(directly connected to the hydrophobic groove) and a distal site (far from the groove).

27

The proximal site is highly positively charged (Figure 12D), mainly due to the presence of Arg 24, Arg 52 and Arg 255, counterparts of which were found to be important for the function of E. coli MurJ [36]. Taken together, the hydrophobic groove, portal, and cationic cavity are consistent with a lipid transport function.

Figure 12: The hydrophobic groove of MurJTA is linked to the central cavity

(A) TMs 13 and 14 (brown) form a hydrophobic groove (surface shown in gray). (B) The hydrophobic groove leads into the central cavity through a portal between TMs 1 and 8. (C) The central cavity is formed mainly by TMs 1, 2, 7, and 8. This large cavity can be subdivided into proximal and distal sites with respect to the groove. A chloride ion is found in the proximal site (cyan sphere). (D) The proximal site is strongly cationic, whereas the distal site is mildly anionic.

28

2.2.6 Identification of Ions by Anomalous Scattering Experiments

To gain insight into the energy-coupling mechanism of MurJ, we set out to

identify ions that were bound in our structure. We first located chloride ions by performing a bromide soak (Figure 13). The soaked crystal was exposed to an X-ray beam tuned to a wavelength (0.9197 Å) at the absorption peak near the Br K-edge, and an anomalous difference Fourier map was calculated which would show the positions of anomalous scatterers. One chloride ion was bound at the apex of the central cavity, in the proximal site near Arg24 and Arg255, which were critical for the function of both

MurJTA and E. coli MurJ [34, 36]. The chloride might be non-specifically bound due to the high concentration of chloride in the crystal drop (containing 400 mM CaCl2).

Figure 13. Identification of chloride ions by bromide soak

Chloride sites were identified by anomalous difference Fourier density peaks from Br- soaked crystals (λ = 0.92 Å, brown mesh, calculated to 3.5-Å resolution and contoured to 4.5 σ).

29

Because we observed strong electron density near two histidine residues on TM

7, we hypothesized that it could be from a divalent metal ion and performed anomalous

scattering experiments to deduce its identity (Figure 14). The energy required to eject an

inner shell electron increases with atomic number, resulting in characteristic absorption

edges for different elements [100]. At energies right above the edge (wavelengths below

the edge), anomalous differences are the highest due to phase shift of the scattered

photons from this absorption event, but rapidly drops once past the edge [101]. This

could be utilized to rule out lower atomic number elements. We exposed native

(unsoaked) crystals to X-rays at wavelengths of either 1.77 Å, 1.31 Å, or 0.98 Å, and

calculated anomalous difference Fourier maps to locate the anomalous scatterers at

Figure 14: Deduction of a Zn2+ ion bound to the beginning of TM 7

(A) Location of the putative Zn2+ ion. (B) Zn2+ was deduced from three factors. First, higher anomalous difference density was seen at λ = 0.98 Å (blue mesh) than at λ = 1.31 Å (green mesh) or λ = 1.77 Å (red mesh), ruling out many common metal ions including all other first-row d-block elements. All three maps were calculated to 3.5-Å resolution and contoured to 5.5 σ. (C) Second, the Zn2+ ion displayed tetrahedral coordination geometry to two histidines and two Cl− ions, ruling out anions. Third, the close coordination distance of 2.03 Å to the nitrogen of both histidines is suggestive of Zn2+.

30

each wavelength. There was a strong anomalous difference peak for the 0.98 Å data, but not for 1.31 Å or 1.77 Å, ruling out other common metal ions including those of first-row d-block elements before zinc (Zn), which has an absorption edge at 1.28 Å. This ion displayed restrained tetrahedral coordination geometry to two histidines and two Cl−

ions, ruling out anions due to electrostatic repulsion. Furthermore, the close

coordination distance of 2.03 Å to the nitrogen of both histidines also rules out many

other common metal ions. This was a surprising result because Zn2+ was never added to the sample during the protein purification and crystallization procedure. Because of its

strategic location at the start of the hinge helix TM 7, Zn2+ might influence gating between the inward-facing and outward-facing states.

Because other peaks were observed in the 1.77 Å anomalous difference Fourier

map, we used these peaks to deduce the identity of additional anomalous scatterers at

this wavelength such as Ca2+ (Figure 15). We deduced three Ca2+ ions from the presence of anomalous difference electron density peaks at 1.77 Å wavelength, as well as pentagonal bipyramidal coordination geometry. Two putative Na+ ions were also inferred from the octahedral coordination geometry with coordination distances of ~2.4

Å, as well as absence of anomalous difference peaks at 1.77 Å wavelength. However, these Ca2+ and Na+ ions are unlikely to function as the counter-transporting ions for lipid

II flipping, as they are located outside of the central cavity and are coordinated mostly

31

by water molecules. This does not rule out the involvement of Na+ or H+, as they are not only difficult to identify but also likely to be bound only in the outward-facing state.

Figure 15: Deduction of putative Ca2+ and Na+ ions

(A) Overlay of the anomalous difference Fourier maps from the data collected at 0.98 Å wavelength (blue) with that collected at 1.77 Å wavelength (red) reveals additional peaks in the latter. Both maps were calculated to 3.5 Å resolution and contoured to 3.8 σ. (B) Three Ca2+ ions were deduced from the anomalous difference peak at 1.77 Å wavelength (magenta), as well as the pentagonal bipyramidal coordination geometry. (C) Two Na+ ions were deduced from the octahedral coordination geometry and absence of anomalous difference density. Fo − Fc omit maps were contoured to 3.0 σ.

32

2.3 Discussion

2.3.1 Model of Lipid II Binding

Although we observed unmodeled electron density peaks at both the membrane portal and the central cavity (Figure 16), the quality of the density peaks was not sufficient for their unambiguous assignment. We therefore used a ligand docking approach to test whether the binding of lipid II to our MurJTA structure was feasible.

Figure 16. Unmodeled electron density peaks in the portal and central cavity

(A) The protein is sectioned on the dotted plane and viewed from the periplasm to visualize the portal. Non-protein 2Fo − Fc electron density peaks are shown as magenta mesh, contoured to 0.7 σ. Gray sticks denote mono-olein molecules. (B) The protein is viewed from the right to visualize the distal site of the central cavity. (C) Model of the pentapeptide (L-Ala-γ-D-Glu-L-Lys-D-Ala-D-Ala) built into the density in panel B, but which was not used for refinement.

33

In the docking result with the highest score, lipid II was docked into the groove and cavity with good geometry and feasible electrostatic interactions (Figure 17A). This

model places the undecaprenyl tail in the hydrophobic groove, the diphosphate-sugar

moieties in the cationic proximal site, and the pentapeptide in the distal site. Based on

this result, we could build the pentapeptide moiety into the unmodeled electron density

at the distal site (Figure 16C), but it was never used for refinement due to the

uncertainty of our assignment.

We next tested conserved residues in the portal and central cavity for their

importance to MurJ function by carrying out complementation assays as previously

described [20, 102] (Figure 17B). We used E. coli strain NR1154, in which endogenous

MurJ (murJEC) expression and thus cell survival was engineered to be arabinose- dependent [20]. Introduction of a plasmid containing murJTA under the control of an

IPTG-inducible promoter allowed growth of this strain in the absence of arabinose, provided that IPTG was present. Mutation of conserved residues Ser17, Arg18, Arg24,

Arg52, and Arg255 (Appendix A) at the proximal site led to loss of complementation, consistent with the idea that these residues are important for recognizing the diphosphate and/or sugar moieties of lipid II. Occlusion of the portal or changes in the shape of the groove also led to loss of complementation, suggesting disruption of lipid II binding (Figure 17C). All mutant MurJTA proteins were expressed except D235A and

N374A, which we removed from our analysis (Figure 17D).

34

Figure 17: Model of lipid II binding to MurJTA

(A) A model of lipid II (yellow sticks) docked into the structure of MurJTA. Residues that were mutated to alanine and assayed for complementation in B are shown as sticks. Residues essential for the function of MurJTA are colored red, nonessential residues are colored green, and residues whose replacement with Ala resulted in lack of expression are shown in cyan. (B) Functional complementation data for MurJTA in E. coli NR1154. Data shown are representative of three technical replicates. (C) Occluding the portal (with L259W) or perturbing the groove surface (with F256A) results in a protein that is unable to complement MurJ activity. (D) Expression of MurJTA mutants, as analyzed by Western blot of NR1154 membrane fractions probed by anti-FLAG antibody. Data shown are representative of three replicates. We observed no correlation between expression level and complementation outcome as long as proteins were expressed.

35

Ruiz and co-workers [22] found that although a single-cysteine mutant of E. coli

MurJ (MurJEC A29C) was functional, addition of the membrane-impermeant cysteine-

reactive agent 2-sulfonatoethyl methanethiosulfonate (MTSES) inhibited flippase activity

and led to cell lysis. They identified a total of five positions that showed a similar

MTSES-dependent phenotype in MurJEC [22, 34]. Because MurJTA can complement

MurJEC and essential residues are conserved between the two, we mapped these residues to the structure of MurJTA. Three of the residues (Phe49, Ser254, and Leu258) are located at the junction between the proximal and distal sites in the central cavity (Figure 18) but are not strongly conserved, which suggests that they do not directly participate in lipid

II binding. Consistent with our model, addition of a bulky adduct at these positions would likely obstruct lipid II binding.

Figure 18. MTSES-sensitive positions in the central cavity

Mapping of MTSES-sensitive positions from E. coli MurJ (shown as red sticks), for which cell lysis or cell shape defects were previously reported upon MTSES treatment of cysteine mutants, to the structure of MurJTA. The docked model of lipid II is shown as yellow sticks.

36

We caution against excessive interpretation of our docking and complementation results, as these experiments are not direct measurements of lipid II binding. It is not our intention to propose a definitive model of lipid II bound to MurJ, but just to provide insight into where the different moieties of lipid II could feasibly be located.

2.3.2 Importance of Alternating Access to MurJTA Function

Because our MurJTA structure is the first structure of a MOP transporter captured in an inward-facing conformation, we wondered about the possibility of an alternating- access transport mechanism. We observed from our inward-facing structure that TM 7 is bent by ~45° about Ser228, while it is straight in the outward-facing MATE structures.

As most MurJ and MATE sequences have either a serine or a threonine residue in this region (Appendix A), we hypothesized that bending of TM 7 about this residue might be relevant for the transition to the outward-facing state. We generated an outward-facing model of MurJTA by splitting the model at the middle of TM 7 (Ser228) and aligning the

N-terminal fragment (TMs 1–7a) and the C-terminal fragment (TMs 7b–12) as rigid bodies with the structure of PfMATE (PDB 3VVN), which we chose because it is the highest-resolution MATE structure published [70]. The resulting model aligned better at the C-lobe than at the N-lobe. Because of the substantial structural difference between

MurJ and MATE transporters, our outward-facing model generated based on the MATE structure should be interpreted with caution.

37

To test the validity of our outward-facing model, we mapped those five positions that are functionally sensitive to MTSES [22] to the inward-facing structure and

outward-facing model. All five positions are accessible from the periplasm in the

outward-facing model but not in the inward-facing structure, showing that our model is

consistent with the MTSES data (Figure 19A-B). Notably, mapping of two MTSES-

sensitive positions located at the periplasmic end of TM 1 (Ala29) or TM 8 (Ser248)

suggests that alternating access is important for lipid II flipping. In our inward-facing

MurJ structure, these positions are located at the interface between the N and C lobes,

and thus have important roles in occluding access to the periplasmic side. It is likely

that MTSES adduct formation at these sites disrupted the inward-facing state of MurJ,

thereby trapping MurJ in the outward-facing state, resulting in disruption of lipid II

flipping and cell lysis (Figure 19C-D).

These data suggest that both inward-facing and outward-facing states exist for

MurJ, and that alternating access would be important for lipid II flipping (Figure 19E).

In this transport model, the substrate is captured in the central cavity between the N-

lobe and the C-lobe. Rocker-switch motion of both lobes translocates lipid II across the

membrane, at which point lipid II is released. The hydrophobic groove formed by TMs

13 and 14 (brown) of MurJ facilitates capture of the undecaprenyl tail, which remains

associated with the groove during the transition to the outward-facing state.

38

Figure 19: Alternating access is important for MurJ function.

Previously reported MTSES-sensitive positions (red spheres) that resulted in loss of MurJ function after MTSES treatment [22] were mapped to the structure of MurJTA. (A) These residues are not accessible to the membrane-impermeable agent MTSES in the inward-facing structure. (B) In the outward-facing model, these residues are exposed to the periplasm, explaining their MTSES sensitivity. (C) Both Ala29 and Ser248 are located at the periplasmic gate in the inward-facing structure of MurJTA, suggesting that adducts at these positions would trap the protein in the outward-facing state (D) by sterically blocking the periplasmic gate from closing. (E) Proposed alternating-access model of lipid II flipping by MurJ. P, M, and G denote phosphate, MurNAc, and GlcNAc, respectively. Cyan circles denote the pentapeptide.

39

2.4 Materials and Methods

2.4.1 Protein Expression and Purification

The gene expressing full-length MurJ from T. africanus was synthesized (Bio

Basic Inc.) and cloned into a modified pET26 expression vector such that the final construct was His10-MBP-PPX-MurJ-PPX-His10. Protein was overexpressed in E. coli

C41(DE3) cells at 4 L scale for 3 hours at 37°C. Cell pellets were resuspended, disrupted

with a Microfluidizer, and membrane proteins were extracted by stirring with 30 mM

DDM for 1 hour at 4°C. After centrifugation to remove insoluble cell debris, the supernatant was affinity purified by TALON Co2+ resin. Eluted proteins were digested

overnight at 4 °C with 40 μg/mL PreScission protease to cleave MBP. PPX-treated

protein was purified by gel filtration on a Superdex 200 column in 20 mM Tris-HCl, pH

8.0, 150 mM NaCl, 2 mM DTT, and 0.3 mM DMNG. For preparation of

selenomethionine (SeMet)-labeled protein, the same plasmid was transformed into

B834(DE3) cells. Protein was overexpressed by autoinduction in presence of 125 mg/L

selenomethionine according to an established protocol [103] and purified as above.

2.4.2 Crystallization

We crystallized MurJTA in lipidic cubic phase (LCP) according to an established

protocol [84], with some modifications. Briefly, purified MurJTA was concentrated to

15 mg/mL, TCEP was added to 2 mM (for SeMet protein only), and protein was mixed

40

with molten mono-olein in a 2:3 (w/w) water:lipid ratio with a twin-syringe setup.

MurJ-reconstituted LCPs were dispensed on 96-well glass sandwich plates in 150-nL drops by a Gryphon LCP robot and overlaid with 1 μl of precipitant solution. Both native and SeMet MurJTA crystallized at room temperature in solutions containing 30–

45% (v/v) PEG 400, 200–400 mM CaCl2, pH 6.5–9.0. Crystals grew to full size in 4 weeks and were flash-frozen in liquid nitrogen without additional cryo-protectant. Addition of lipid II improved diffraction quality and was performed for native crystals as follows: L- lysine form of lipid II (University of Warwick) was added to mono-olein powder in a glass vial to a final concentration of 1.5 mM (in molten mono-olein). This mixture was dissolved and homogenized in chloroform, dried under argon, and washed with pentane. Organic solvent was completely removed before the lipids were melted. LCP trays were set up as described above. For the bromide soak experiment, a window was cut from the coverslip and precipitant solution was wicked away. Fresh precipitant solution (of the same composition except with CaCl2 replaced by CaBr2) was injected and the well was resealed with tape. Soaking lasted for 2 hours, after which crystals were harvested.

2.4.3 Data Collection and Structure Determination

The structure of MurJ was determined by single-wavelength anomalous dispersion (SAD). Data were collected at the NECAT 24-ID-C and 24-ID-E beamlines

41

(Advanced Photon Source) with a wavelength of 0.9791 Å and were processed using

HKL2000 [104]. Selected frames from 12 isomorphous SeMet crystals were merged together in HKL2000 to generate a data set with over 50-fold average redundancy and signal extending to 3.2-Å resolution. Heavy-atom sites were located by SHELXD [94], and phasing was performed by PHENIX AutoSol [96] initially to a high-resolution cutoff of 5.0 Å, producing a figure of merit of 0.50. We gradually extended phases to 3.5-Å

resolution using MR-SAD by placing idealized helical fragments into the density and density modification. For the high-resolution data, we merged selected frames from three native crystals crystallized in the presence of lipid II to generate a dataset with signal extending to 2.0 Å resolution. Phases were obtained from the SeMet model by

PHASER [105] and were improved by cycles of model-building in COOT [106] and

refinement in PHENIX.refine [107] to a final Rwork/Rfree of 18.9/21.4, good geometry (98%

Ramachandran favored, 0% outliers), and minimal clashes (clashscore of 1). For the bromide soak experiment, data were collected at beamline 24-ID-C with a wavelength of

0.9197 Å. Phases were obtained from the high-resolution model by PHASER [105], and

the anomalous difference Fourier map was calculated by PHENIX.maps [95].

2.4.4 Docking of Lipid II and Structure Analysis

Docking of lipid II to the structure of MurJTA was done with Autodock Vina

[108]. We chose the L-lysine form of lipid II for docking instead of the meso-

42

diaminopimelate (mDAP) form because the related bacterium Thermotoga maritima was shown to incorporate both L- and D-lysine into peptidoglycan instead of mDAP [109,

110]. Initial lipid II coordinates were generated from 2D geometry in PHENIX.eLBOW

[111] and stereochemistry was manually corrected in PHENIX.REEL [112], with reference to the NMR structure of lipid II in complex with nisin (PDB 1WCO)[113].

Effort was made to ensure that all 16 chiral centers had the correct R/S configuration, that the peptide and polyprenyl C=C bonds had the right cis/trans configuration and planarity, and that the sugars had the correct pucker and glycosidic linkage. In addition, the side chains of Arg18 and Glu57 of MurJTA were treated as flexible for docking, as they appeared to be flexible in our structure. The remaining 68 torsions were freely rotatable during docking. Docking was carried out over a search space of 30

× 30 × 36 Å covering the central cavity, portal, and hydrophobic groove. Surface electrostatics were calculated with APBS [99]. Molecular graphics were created in

PyMOL [114] and CueMol (www.cuemol.org). Sequence alignment was done by

PROMALS3D [115], and structure superposition was done in UCSF Chimera [116].

2.4.5 Complementation Assay and Western Blot

MurJTA was cloned into pEXT21 with a C-terminal FLAG tag and was subjected to site-directed mutagenesis using standard recombinant DNA methods. Expression of

MurJTA is controlled by the tac promoter in pEXT21 [102]. The pEXT21 plasmid bearing

43

wild-type MurJTA or a mutant was transformed into the conditional MurJ-depletion E. coli strain NR1154 [20], which was a gift from Natividad Ruiz (The Ohio State

University). The endogenous MurJ in NR1154 is placed under the control of the PBAD promoter; therefore, NR1154 growth depends on the presence of L-arabinose. D-fucose is an anti-inducer of endogenous E. coli MurJ expression in NR1154. Overnight cultures of transformants were diluted to an OD600 of 0.2 and grown to an OD600 of 0.6 in the presence of 80 μg/mL spectinomycin and 0.2% arabinose. IPTG was then added to the duplicate of each culture at a final concentration of 0.1 mM. Cultures were further incubated with shaking for 3 h at 37 °C before cells were harvested by centrifugation.

For the complementation assay, cells were washed three times with phosphate-buffered saline and normalized to an OD600 of 0.5. Each culture was then diluted to 0.02, subjected to ten-fold serial dilution, and spotted (5 μL) onto LB-agar plates supplemented with either 0.05% fucose or 0.05% fucose with 0.1 mM IPTG.

To check for protein expression, total membrane was isolated by ultracentrifugation and resuspended in Tris-buffered saline for Western blot. Samples were mixed with denaturing buffer (final 3 M urea and 2.5% SDS) and subjected to 3 min sonication before SDS-PAGE. Bands were transferred to a polyvinylidene difluoride membrane, which was probed with mouse anti-FLAG primary antibody

(1:1,200 dilution, Sigma F3165) and donkey anti-mouse secondary antibody (1:10,000

dilution, LI-COR 926-32212). Results were visualized on a LI-COR Odyssey CLx system.

44

3. Structural Basis for the MurJ Transport Cycle

3.1 Overview

We determined the first crystal structure of MurJTA, which exhibited an inward-

facing conformation unlike all existing outward-facing MOP transporter structures. This

structure and previously reported chemical genetics data [22] suggest that MurJ operates

by an alternating access mechanism to flip lipid II, but the conformation transitions and

energy-coupling mechanism that drive this transport remain unknown. To better

understand the MurJ transport mechanism, we carried out further structural studies to

capture MurJTA in different states, as well as functional studies probing the in vitro conformational state of MurJTA in detergent micelles.

We determined additional crystal structures of MurJTA captured in inward- closed, inward-open, inward-occluded and outward-facing conformations at 3.2, 3.0, 2.6 and 1.8-Å resolution, respectively. The structures reveal a lateral gating mechanism and state-dependent rearrangements in the cavity that could drive transition to the Na+-

bound outward-facing state. Our analyses provide a structural basis for lipid-linked

oligosaccharide transport and a framework for the development of inhibitors. Results

for this chapter will be published in a scientific journal (manuscript in preparation). I

acknowledge Aili Hao and Ziqiang Guan for their invaluable contributions to the in-vivo

complementation and functional assays which are in progress at the time of writing.

45

3.2 Results

3.2.1 Conformational Landscape of MurJTA

Because we observed a chloride ion in the central cavity of our previous crystal structure, we hypothesized that chloride might be restricting MurJTA to this crystal form and took precautions to remove chloride from cell lysis and protein purification buffers.

With this new protein purification procedure and in the presence of lipid II doped into the lipidic cubic phase, we obtained new crystal forms of MurJTA (Figure 20A). We

determined the crystal structures of MurJTA captured in three additional inward-facing conformations, which were resolved to 3.2, 3.0 and 2.6 Å, respectively. One crystal form that diffracted to 1.8-Å resolution required the N-lobe (TMs 1-7a) and C-lobe (TMs 7b-

14) of the published inward-facing structure (PDB ID: 5T77) as separate search models for successful molecular replacement phasing, which revealed MurJTA in an outward- facing conformation (Figure 21). We refined the structures to good geometry and minimal clashes (Table 2). The final electron density maps (Figure 20B) were of excellent

quality and allowed us to build almost all the residues unambiguously except at the

termini. Taken together, these new crystal forms represent four new conformations of

MurJTA: three inward-facing conformations distinct from the published structure, and

one outward-facing conformation (Figure 22). To gain insight into the transport

mechanism of MurJ, we will first compare the inward-facing conformations.

46

Figure 20: Four new crystal forms of MurJTA were obtained

(A) Crystals were obtained in lipidic cubic phase doped with lipid II. Crystal contacts are mostly formed by end-to-end stacking, as well as lateral interactions mediated by the host lipid mono-olein. TM 1 was not involved in crystal packing and thus its bending is unlikely to be an artifact. The asymmetric unit is colored, while crystallographic symmetry mates are shown in gray. (B) 2Fo − Fc composite omit electron density maps (contoured to 1 σ) calculated with the final models, omitting 5% of the model at a time. Maps were calculated to resolutions of (from left) 3.2 Å, 3.0 Å, 2.6 Å, and 1.8 Å.

47

Figure 21: Structure determination of outward-facing MurJTA

(A) LCP crystals that produced the outward-facing crystal form. (B) These crystals diffracted to 1.8-Å resolution. This screenshot was cropped at 2.0-Å resolution, beyond which spots were not visible to the eye. (C) Initial molecular replacement result using the N-lobe (TMs 1-7a) and C-lobe (TMs 7b-14) of our first published structure as search models. For this image, both 2Fo – Fc (blue, contoured to 1.5 σ) and Fo − Fc (green/red, contoured to +/− 3.0 σ) maps were blurred (B-factor +100) to reduce noise.

Figure 22: Crystal structures elucidate the conformation landscape of MurJTA

The first structure of MurJTA that was discussed in the previous chapter is shown on the left with four new crystal structures to the right. Resolutions were (from left) 2.0, 3.2, 3.0, 2.6, and 1.8 Å. 48

Table 2: Data collection and refinement statistics for additional MurJTA structures

Inward closeda Inward openb Inward occludedc Outwardd Data collection Space group P 2 2 21 C 2 P 2 21 21 C 2 Cell dimensions a, b, c (Å) 71.1, 101.8, 158.5 129.7, 105.5, 111.5 68.4, 80.3, 100.7 128.6, 57.4, 86.4 α, β, γ (°) 90, 90, 90 90, 125.3, 90 90, 90, 90 90, 100.7, 90

Wavelength (Å) 0.98 0.98 0.98 or 1.65c 0.98 Resolution (Å) 85.6–3.2 (3.3–3.2)e 74.7–3.0 (3.1–3.0)e 68.4–2.6 (2.7–2.6)e 84.9–1.8 (1.9–1.8)e I/σ(I) 9.48 (1.14) 11.41 (1.00) 13.29 (1.13) 22.71 (1.09)

CC1/2 1.00 (0.43) 1.00 (0.63) 0.95 (0.47) 0.97 (0.59)

Rpim 0.15 (0.94) 0.10 (0.83) 0.11 (>1.0) 0.076 (0.71) Redundancy 20.7 (21.3) 5.5 (5.1) 48.8 (38.3) 23.4 (23.5) Completeness (%) 99.4 (99.3) 99.0 (98.0) 99.8 (99.7) 99.9 (99.6) f Completeness (%) 89.6 (67.4) 82.7 (50.3) 99.0 (92.8) 88.5 (58.6) Ellipsoid completenessf (%) 100.0 (99.9) 99.2 (98.1) 100.0 (99.3) 100.0 (99.8)

Refinement Resolution (Å) 85.6–3.2 (3.3–3.2)e 74.7–3.0 (3.1–3.0)e 68.4–2.6 (2.7–2.6)e 84.9–1.8 (1.9–1.8)e No. reflections 17624 (1292) 20395 (1238) 17459 (1608) 51025 (3345)

Rwork / Rfree (%) 25.5 / 28.0 25.4 / 27.8 22.9 / 25.8 17.9 / 19.9 No. atoms Protein 7292 7241 3672 3775 Ions/monoolein/PEG 27 126 125 294 Water 0 0 0 156 B factors Protein 31.18 46.54 48.19 20.55 Ions/monoolein/PEG 32.77 45.10 58.48 48.24 Water N/A N/A N/A 28.35 R.m.s. deviations Bond lengths (Å) 0.005 0.005 0.003 0.003 Bond angles (°) 1.02 0.67 0.62 0.66 Ellipsoid completeness was calculated by STARANISO. All other statistics were calculated in PHENIX. a Merged from 3 crystals. b Merged from 2 crystals. c Merged from 7 crystals (2 collected at wavelength of 0.98 Å and 5 at 1.65 Å). d Merged from 6 crystals. e Values in parentheses are for highest-resolution shell. f After applying ellipsoidal truncation and anisotropy correction in STARANISO.

49

3.2.2 Inward-Facing Structures Reveal a Lateral-Gating Mechanism

Superposition of all the inward-facing structures show that they align very well at the periplasmic gate (Figure 23A), highlighting the importance of the periplasmic gate in stabilizing the inward-facing state. This gate is formed by hydrogen-bond interactions between the N-lobe and C-lobe (Figure 23B) concentrated at three locations:

between TM 1 and the TM 9-10 loop, between TM 7 and the TM 3-4 loop, and between the periplasmic ends of TM 2 and TM 8. The latter consists of three hydrogen bonds between Asp39 at TM 2 and G245(backbone nitrogen)/S248(backbone nitrogen, sidechain) at TM 8. The essential residue Asp39 functions as the hydrogen bond acceptor in all three interactions, suggesting that it plays a major role in stabilizing the periplasmic gate, consistent with the observations reported by others [34, 36, 117].

In contrast to the periplasmic gate, the structures diverge substantially at the cytoplasmic side (Figure 23C-D). Coordinated rearrangement of TM 1 (blue), TM 8

(green), and the loop between TMs 4 and 5 (TM 4-5 loop, brown) regulates the size of the membrane portal. Based on the conformation at these regions and the size of the membrane portal, we assigned two of the new structures as inward-closed and inward- open (Figure 23C). The Cα-Cα distance between Ser11 (TM 1) and Ser267 (TM 8) increased from 8.0 Å in the closed structure to 17.4 Å in the open structure. Our previously-published structure appeared to be an intermediate between these two conformations (Cα-Cα distance of 10.7 Å). TM 1 is bent outwards by almost 40° in the

50

Figure 23: Lateral gating mechanism in the inward-facing state

(A) Superposition of inward-facing structures (including non-crystallographic symmetry mates) highlights similarity at the periplasmic gate but divergence at the cytoplasmic gate. (B) The periplasmic gate is stabilized by hydrogen bonds between Gly245/Ser248 and the essential Asp39. (C) At the cytoplasmic side, coordinated rearrangement of TM 1 (blue) and TM 8 (green) controls portal dilation. The Cα-Cα distance between Ser11 (TM 1) and Ser267 (TM 8) increases from 8.0 Å in the inward-closed structure to 17.4 Å in the inward-open structure. (D) Rearrangement of the TM 4-5 loop (orange) and unwinding of the cytoplasmic ends of TMs 4/5 could induce bending of TM 1. The conserved Phe151 (Phe157 in E. coli MurJ) provides leverage to bend TM 1 (residues 1- 20) out into the membrane. The TM 4-5 loop is shown in sausage representation, with thickness proportional to conservation. Sidechains of Arg18 and Phe151 were not resolved in the inward-occluded structure. 51

open structure about the invariant Gly21, coordinated by rearrangement of the TM 4-5 loop and unwinding of the cytoplasmic ends of TM 4 and TM 5 (Figure 23D). The

conserved Phe151 (Phe157 in E. coli MurJ) at the TM 4-5 loop provides leverage to bend

TM 1 out into the membrane. Conserved residues Leu145, Asn146, and Pro158 appear

to participate in conformational change of the TM 4-5 loop. Because lipid II is believed

to access the central cavity of MurJ through the portal between TM 1 and TM 8 [80], the

different sizes of the portal could regulate entry of the large lipid II headgroup, thereby

providing a lateral-gating mechanism. Residues Arg18, Gly21 (TM 1) and Leu145,

Asn146, Phe151, Pro158 (TM 4-5 loop) are conserved in MurJ sequences but not MATE

drug efflux pumps in the same superfamily (Appendix A), suggesting the lateral gate to

be a specific adaptation for flippase function.

3.2.3 The Inward-Occluded Conformation of MurJTA

One of the inward-facing structures exhibited a partially-dilated membrane portal but was otherwise divergent from the rest of the inward-facing structures (Figure

24A). Specifically, the C-lobe is bent with the cytoplasmic section swung ~15° towards the N-lobe. Conserved residues Pro260, Pro300 and Gly340 serve as hinges on TMs 8, 9 and 10, respectively. In contrast, re-arrangements in the N-lobe are restricted to the middle section of TM 2, which also bends inwards towards the central cavity. These conformational changes appear to be aided by the flexible cytoplasmic half of TM 8,

52

which assumes a distinctly non-α-helical conformation, as well as the broken helix in

TM 2 formed by the G/A-E-G-A motif that is conserved in Gram negative MurJ

sequences (Figure 24B). Together these conformational changes position Glu57 of the

G/A-E-G-A motif within proximity (~ 4 Å) of Arg352 on TM 10 (or Lys368 one helical- turn up in E. coli MurJ), forming a thin gate between the N-lobe and C-lobe.

We observed unmodeled electron density peaks at the portal and in the central cavity above the thin gate (Figure 24C). Although we cannot unambiguously assign the electron density to lipid II, we speculate that this Glu57-Arg352 thin gate could help occlude the flexible lipid II molecule into the cavity prior to outward transition. To test whether this occluded cavity could accommodate the lipid II headgroup, we performed in-silico docking of lipid II and molecular dynamics simulation to emulate the rotational freedom of lipid II and MurJ sidechains in the docked complex (Figure 25). Our in-silico study suggests that the occluded cavity is large enough to accommodate the lipid II headgroup above the thin gate. Mutation of Glu57 to alanine led to loss of MurJTA function [80], consistent with the notion that the thin gate plays an important role in

transport, although experiments on Arg352 are ongoing at the time of writing. Based on the aforementioned conformational changes, presence of the thin gate, and unmodeled electron density, we assign this structure as an inward-occluded state of MurJ.

53

Figure 24: The inward-occluded conformation of MurJTA

(A) Rotation of the cytoplasmic half of C-lobe by ~15° from the inward-open structure (left) to the inward-occluded structure (right). Conserved residues Pro260, Pro300, and Gly340 serve as hinges. The middle segment of TM 2 also bends inwards by ~15°. Together these bring Glu57 (TM 2) and Arg352 (TM 10, Lys368 in E. coli MurJ) into proximity, which form a thin gate. TM 1 is hidden for clarity. (B) The conserved (G/A)- E-G-A motif (orange) allows S-shaped bending of TM 2 by breaking the helix. TM 8 assumes a more α-helical geometry (albeit still not ideal) in the inward-occluded structure than in the other inward structures. (C) Unmodeled 2Fo − Fc electron density at the portal and transport cavity of the inward-occluded structure, displayed here as purple mesh contoured to 0.8 σ.

54

Figure 25: Docking and MD simulation of lipid II to the inward-occluded structure

(A) Lipid II (yellow spheres) was docked to the inward-occluded structure by Autodock Vina. (B) The docking result was used as a starting point for molecular dynamics simulation by NAMD in a hydrated POPE membrane bilayer system. The equilibrated system is shown, with lipids visualized as spheres and waters denoted by red dots. (C) Cytoplasmic view of the docked complex (lipid II shown as yellow sticks, charged residues in the cavity shown as gray sticks), showing how the Glu57-Arg352 thin gate could help occlude the substrate. (D) Rotation of MurJ sidechain and lipid II torsions in a picosecond molecular dynamics simulation starting from the docked model (snapshots per 10 ps overlaid, total 100 ps). Surrounding lipids and waters are hidden for clarity.

55

3.2.4 The Outward-Facing Conformation of MurJTA

We next compared the outward-facing structure to the inward-occluded structure to elucidate the conformational changes that could drive outward transition.

The inward-to-outward state transition is associated with a rotation of the N- and C- lobes resulting in cytoplasmic gate closure and periplasmic gate opening. The N-lobe rotated by ~15° from the inward-occluded structure while the C-lobe rotated by only

~7.5°, possibly because the latter has already swung inwards in the occluded structure

(Figure 26A). The fulcrum of this rotation is located at TM 7, bridging the N- and C- lobes, which was bent by 90° in the inward-occluded structure but straightened in the outward-facing structure. This motion results in concomitant lowering of the TM 6-7 loop, which is amphipathic and partially embedded in the cytoplasmic leaflet of the membrane as it loops around the N-lobe. We postulate that tension from TM 7 straightening would lower the TM 6-7 loop in the membrane, providing the force required for N-lobe rotation and closure of the cytoplasmic gate. Aside from TM 7, TM 1 also straightens from its bent conformation in the inward-occluded structure, closing the lateral membrane portal (Figure 26B).

The outward-facing cavity was substantially shallower and narrower than the inward-facing cavity, supporting the idea that cavity shrinkage could be a mechanism to displace substrate into the periplasm, as was proposed for MATE transporters [68, 70]

(Figure 26C). We observed unmodeled electron density in the hydrophobic groove

56

Figure 26: The outward-facing conformation of MurJTA

(A) Straightening of TM 7 and concomitant lowering of the TM 6-7 loop was observed in the outward-facing structure relative to the inward-occluded structure. (B) Alignment of N- and C-lobes from the inward-occluded conformation (gray) to the outward-facing conformation (in color), showing rotation of not just TM 7 but also TM 1. (C) The outward-facing cavity is shallower and narrower than the inward-facing cavity. Electrostatic surface is shown from −10 kT (red, anionic) to +10 kT (blue, cationic). (D) The cytoplasmic gate is mainly stabilized by a hydrogen bond network between TMs 2 and 10, as well as salt bridges between the cytoplasmic loops.

leading into the outward-facing cavity, which might indicate the binding site of the undecaprenyl lipid tail of lipid II while the headgroup is on the periplasmic side (Figure

27). Because this density was weak and not completely connected, we cannot rule out the possibility of adventitious mono-olein molecules. Below the outward-facing central cavity is the cytoplasmic gate which is ~20 Å thick and formed by TMs 2, 4, 8 and 10

57

(Figure 26D). This gate is stabilized by hydrogen bonds between TM 2 and TM 10, again involving Arg352. While Arg352 forms the thin gate with Glu57 in the inward-occluded structure, it now interacts with Ser61, Ser62, and the backbone carbonyl of Gly58 in the

G/A-E-G-A helix break. This suggests importance of Arg352 to inward-to-outward conformational transition.

Figure 27: Structural features of the putative lipid-binding site in the outward state.

(A) Putative lipid-binding site. Yellow sticks denote a model of the undecaprenyl tail of lipid II, which was never used for structure refinement or map calculation. (B) The lipid is restrained by the hydrophobic groove and a short tunnel near the periplasmic side. (C, D) Unmodeled 2Fo − Fc electron density peaks (purple mesh, contoured to 0.7 σ) were observed in the hydrophobic groove (C), or in the tunnel (D) which is formed by hydrophobic residues Ile19, Leu22, Phe256, Leu259, and Phe316.

58

3.2.5 Sodium Site in the Outward-Facing MurJTA Structure

We noticed a strong Fo − Fc omit electron density peak (~19 σ) in the C-lobe

between TMs 7, 11, and 12 (Figure 28A). Coordination distances were within the

expected range for Na+ (~2.3 to 2.5 Å) in our 1.8 Å-resolution structure, and coordination

geometry is most consistent with Na+, as determined by the CheckMyMetal server [118].

Therefore, we assign this peak as Na+. Sodium is coordinated in trigonal bipyramidal

geometry, with an equatorial plane formed by Asp235, Asn374, Val390 (backbone

carbonyl) and axial positions occupied by Asp378 and Thr394. Asp235 is the only Na+- coordinating residue from TM 7, with the rest from either TM 11 or TM 12. Experiments to test the importance of these residues are in progress.

In light of this discovery, we re-analyzed all the other structures and found Na+

bound to the inward-occluded and inward-closed structures as well (Figure 28B). The

Na+ bound structures exhibited different degrees of TM 7 straightening, resulting in concomitant lowering of the N-terminal half of TM 7 and the TM 6-7 loop relative to the

Na+ site, suggesting that Na+-induced conformational change at Asp235 could be propagated down TM 7 (Figure 28C). Notably, the outward-facing structure displays the closest Na+ coordination distances for most of the Na+-coordinating residues (Table

3) and also the most straightening of TM 7. This suggests that Na+ coordination could

induce rearrangements in MurJ conformation, which would be investigated further by in-vitro cysteine accessibility experiments in the next section (3.2.6).

59

Figure 28: Sodium site in MurJTA

(A) Na+ is bound in the C-lobe (TMs 7, 11, 12) coordinated by Asp235, Asn374, Asp378, Val390 (backbone carbonyl), and Thr394. Fo − Fc omit density for Na+ is shown as magenta mesh contoured to 4.5 σ (peak height ~19 σ). 2Fo − Fc omit density is shown in gray mesh. Na+ is coordinated by trigonal bipyramidal geometry. (B) Na+ was also bound in the inward-occluded and inward-closed structures. (C) Na+-associated rearrangement could be propagated down TM 7.

60

Table 3: Sodium Fo − Fc omit peak heights and coordination distances

+ Structure Na Fo − Fc Na+ coordination distance (in Ångstrom) omit peak height Asp235 Asn374 Asp378 Val390 Thr394 OD1 OD1 OD1 O OG1 Inward closed 7 sigma 2.49 2.32 2.45 2.39 2.35 Inward occluded 9 sigma 2.44 2.43 2.40 2.40 2.41 Outward 19 sigma 2.45 2.26 2.42 2.33 2.33

Interestingly, recent structural re-analysis of PfMATE crystal structures showed

that one of them contains a Na+ site in the N-lobe [70, 119], consistent with spectroscopic

experiments on another MATE protein, NorM [74]. TM 7 in MurJTA is tucked into the

outward-facing cavity like TM 1 in the PfMATE structure [70] (Figure 29A). In contrast,

TM 1 is straight in the PfMATE structure without Na+ bound [70]. Taken together, Na+ coordination appears to be associated with rearrangement of TM 7 in MurJ or TM 1 in

PfMATE (Figure 29B). Relocation of the Na+ site from the N-lobe of MATE proteins to

the C-lobe of MurJ might be to accommodate their different substrates. Alternatively,

divergence in the location of the Na+ site could also reflect difference in energy-coupling mechanism between MurJ and MATE proteins.

61

Figure 29: Comparison of the sodium site in MurJTA and in PfMATE

(A) TM 7 in MurJTA is bent like TM 1 in the PfMATE structure which was recently reassigned to be also Na+-bound (PDB: 3VVO). In contrast, TM 1 is straight in the PfMATE structure without Na+ bound (3VVN). The Na+ site is located in the C-lobe of MurJ but in the N-lobe of PfMATE. (B) Superposition of TM 7 in MurJTA with TM 1 in PfMATE (bent conformation), showing decent alignment in the upper segment containing the Na+-coordinating residue Asp235(MurJTA) or Asp41(PfMATE), but deviation towards the lower segment. The all-atom r.m.s.d. between the two structures is ~1.47 Å over the entire TM but only ~0.55 Å in the upper segment.

62

3.2.6 Probing the Conformation of MurJ in Detergent Micelles

To probe the in-vitro conformational state of MurJTA, we designed single cysteine mutants at the cytoplasmic and periplasmic gates. Because MurJTA has no endogenous

cysteines, these mutants would manifest an upward shift in mass if the cysteine was

accessible to PEG-maleimide. We observed that single cysteine mutants at the

cytoplasmic gate were much more accessible to PEG-maleimide than those at the

periplasmic gate, suggesting that MurJTA adopts predominantly inward-facing conformations in detergent micelles (Figure 30). We also performed the experiments at

an elevated temperature (60°C) because MurJTA is from a hyperthermophilic bacterium.

We observed some changes in the accessibility profile at 60°C, including reduction of

accessibility at the residues comprising the cytoplasmic gate on TM 8 (Thr270, Ser274

and Ser277), indicating some extent of conformational sampling and not just a single

static inward-facing conformation of MurJTA in vitro.

63

Figure 30: Cysteine accessibility profile of MurJTA by direct PEG-maleimide labeling

(A) Single cysteine mutants of MurJTA (no endogenous cysteines) were expressed, purified, delipidated, and probed for accessibility in DDM micelles by a direct labeling strategy in potassium acetate (absence of sodium). Accessible cysteines would be labeled by PEG5000-maleimide (PEG-Mal), resulting in a mass shift up the gel. Experiments were performed at both 20°C and 60°C. (B) Mapping of these cysteine positions (for the 20°C data) to the inward-open and outward-facing crystal structures of MurJTA. Positions were considered inaccessible if only the lower band was observed, and accessible if the top band was observed and of higher intensity than the lower band. Remaining positions with both bands visible were considered partially accessible.

64

Having established the predominantly-inward conformation in DDM micelles, we next tested whether the presence of Na+ could shift the equilibrium towards an outward-facing conformation, as was suggested by our crystal structures. If Na+ were to

promote the outward-facing conformation, we would expect an increase in solvent

accessibility at the periplasmic gate. To test this hypothesis, we mutated single residues

at the periplasmic gate into cysteines (MurJTA has no endogenous cysteines) and probed

the accessibility of purified mutant proteins in buffer containing either Na+ or its substitute N-methyl-D-glucamine (NMDG+) (Figure 31). Cysteine mutants of Ala29,

Ser240 and Ile247 experienced a modest 1.5 to 2.5-fold increase in accessibility, but not other positions at or near the periplasmic gate (Met46, Ser113, Ser248). While the results are not conclusive evidence for a role of Na+ in outward transition, they are consistent with Na+-dependent rearrangements and provide support for Na+ as the bound ion.

Because the detergent-micelle environment is not an ideal mimic of a membrane bilayer [120], we will next repeat these cysteine accessibility assays in a membrane

environment, albeit it must remain amenable to testing of different ions. One approach

would be to reconstitute purified MurJTA into nanodiscs, which are water-soluble

nanoscale phospholipid bilayers stabilized by an encircling membrane scaffold protein

[121-123]. Alternatively, purified MurJTA could be reconstituted into model

phospholipid vesicles which would additionally allow ion gradients to be tested [124].

65

Figure 31: Accessibility profile of MurJTA in Na+ or in NMDG+ by inverse labeling

(A) Purified single cysteine mutants of MurJTA (no endogenous cysteines) were probed for accessibility in DDM micelles by the inverse labeling strategy in 150 mM sodium acetate versus in 150 mM N-methyl-D-glucamine acetate (NMDG+). Cysteines that were not accessible to 100 µM 2-sulfonatoethyl methanethiosulfonate (MTSES) would be labeled by PEG5000-maleimide (PEG-Mal), resulting in a mass shift up the gel. Relative accessibility was determined as detailed in Methods, with a larger number corresponding to higher accessibility. (B) Fold change in cysteine accessibility in Na+ versus in NMDG (n=5, error bars denote s.d.). Fold changes were compared to that for M46C, and P-values were calculated by paired two-tailed Student’s t-test (the abbreviation n.s. denotes not significant). (C) Mapping of these positions to the inward-open and outward-facing crystal structures of MurJTA.

66

3.2.7 State-Dependent Changes in the Central Cavity

We next analyzed differences in the central cavity of our MurJTA structures to gain insight into lipid substrate flipping. The inward-facing cavity was largest in the inward-open structure and deepest in the inward-occluded structure, while the outward-facing cavity resembled a narrow crevice (Figure 32A). We observed systematic rearrangement of Arg24, Asp25, and Arg255 (Arg270 in E. coli MurJ) in the cavity between the different crystal structures (Figure 32B-C). These residues are conserved (Arg24 is invariant), with both Arg24/Arg255 essential [22, 34] and Asp25 intolerant of certain substitutions [117]. This Arg24-Asp25-Arg255 triad is held in position by Asn162, Gln251, and a chloride ion in the inward-closed state. Arg255 is released from Gln251 in the inward-open state and swivels down, possibly to capture the diphosphate moiety of lipid II that enters through the portal. In the inward- occluded state, Arg24 and Arg255 are brought into unusual proximity (just 2.9 Å apart at the closest) on top of the unmodeled electron density. Since such close contact between two arginine side chains would experience unfavorable repulsion, we reasoned that they could be stabilized by electrostatic interactions with Asp25 and the diphosphate moiety of lipid II. The periplasmic gate opens in the outward conformation, disengaging

Arg255 and allowing substrate release. Because of their conserved and essential nature, as well as their coordinated rearrangement at the cavity apex in proximity to unmodeled density, we suggest Arg24-Asp25-Arg255 to be the putative substrate-binding triad.

67

Figure 32: State-dependent conformation changes in the central cavity

(A) Conformation changes in the MurJ transport cycle. The cavity surface is shown as dots. (B-C) Systematic rearrangement of the Arg24-Asp25-Arg255 triad in the cavity during the transport cycle. In the inward-occluded state, R24 and R255 are brought into unusual proximity (just 2.9 Å apart at the closest), with the unfavorable repulsion stabilized by electrostatic interactions with Asp25 and putatively the diphosphate moiety of lipid II (yellow sticks). The lipid-diphosphate model was never used for refinement or map calculation. 2Fo − Fc electron density (contoured to 1.0 σ) is shown as blue mesh with protein electron density masked. Numbers denote distances in Å.

68

3.3 Discussion

3.3.1 Transport Mechanism of MurJ

We determined the conformational landscape of the lipid flippase MurJ, and our analyses suggest that MurJ undergoes an ordered sequence through its transport cycle

(Figure 33). In the inward-open state, coordinated rearrangement of TM 1, TM 8 and the

TM4-5 loop regulates the size of the lateral membrane portal and thus the entry of lipid

II headgroup into the transport cavity. Upon lipid II binding, bending of the middle segment of TM 2 and closure of the C-lobe brings MurJ into an asymmetric inward- occluded conformation with a thin gate formed by Glu57 and Arg352. Subsequent rearrangement of TM 7, putatively due to Na+ binding, mediates outward transition,

upon which lipid II is displaced into the periplasmic side of the membrane. Factors such

as membrane potential [75] could serve to reset the protein to an inward-closed state.

Finally, release of Na+ into the cytoplasm and reopening of the lateral membrane portal would complete the transport cycle.

69

Figure 33: Model of the MurJ transport mechanism

In the inward-open state (top left), coordinated rearrangement of TM 1, TM 8, and the TM 4-5 loop opens the portal to allow lipid II entry. Binding of lipid II (involving the Arg24-Asp25-Arg255 triad at the cavity apex) transitions MurJ to the inward-occluded state, where a thin gate (Glu57-Arg352) is formed by C-lobe closure and TM 2 bending. Subsequent closure of the N-lobe is driven by rearrangement of TM 7 and concomitant lowering of the TM 6-7 loop (possibly in response to Na+ binding), completing transition to the outward-facing state, whereby lipid II is released. Factors such as membrane potential [75] could mediate return to the inward-closed state. After Na+ release into the cytoplasm, the portal is opened again for the next cycle. P, M, G denote phosphate, MurNAc, and GlcNAc respectively, while cyan circles denote pentapeptide.

70

3.3.2 Relevance to Inhibitor Development Targeting MurJ

Our outward-facing structure indicates importance of the cytoplasmic gate in occluding access to the outward-facing cavity from the cytoplasm. Consistent with this finding, resistant mutants to the humimycins, lipopeptides with antibacterial activity against some Gram-positive pathogens including methicillin-resistant Staphylococcus aureus [77], cluster mostly around the cytoplasmic side of MurJ below the putative lipid

II headgroup binding site [78]. While the mechanism of inhibition by humimycins is not currently known, it is conceivable that they could disrupt closure of the cytoplasmic gate and thus trap MurJ in an inward-facing conformation.

Our study also highlights the importance of TM 7 in the inward-to-outward

transition. Resistant mutants to the levivirus M lysis protein (LysM), which induces

bacterial cell lysis by inhibition of MurJ function, was previously mapped to the TM 2-

TM 7 region that is facing outside towards the membrane [79]. Cysteines positioned in the cavity displayed increased MTSES accessibility in cells upon LysM treatment, consistent with stabilization of the outward-facing state by LysM [75, 79]. As exemplified

by humimycins and LysM, our crystal structures of MurJTA provide a framework for understanding the inhibition mechanism of MurJ inhibitors and the development of these inhibitors into potential antibiotics.

71

3.4 Materials and Methods

3.4.1 Protein Expression and Purification for the New Crystal Forms

MurJTA was expressed as described in the previous chapter. Chloride was removed from all buffers used for cell lysis and protein purification, as detailed below.

Cell pellets from each liter of culture were resuspended in 10 mL lysis buffer (50 mM

HEPES-NaOH pH 8.0, 150 mM Na acetate). Cell lysis, cobalt affinity purification, and overnight PPX digestion were performed as described in the previous chapter, replacing chloride with acetate and Tris-HCl with HEPES-NaOH. PPX-treated MurJ was

exchanged to gel filtration buffer (20 mM HEPES-NaOH pH 8.0, 150 mM Na acetate, 2 mM DTT and 0.3 mM DMNG) and purified on a Superdex 200 column.

3.4.2 Crystallization of the New Crystal Forms

MurJTA was crystallized by LCP as described in the previous chapter with modifications. Purified, chloride-free MurJTA was concentrated to 20-25 mg/mL, before mixing with molten lipid II-doped monoolein which was prepared in the same way as described in the previous chapter. MurJ-reconstituted LCPs were dispensed on 96-well glass sandwich plates as 130 nL drops and overlaid with 1 µl of precipitant solution from crystallization screens made in-house and/or the commercial MemMeso HT-96

screen (Molecular Dimensions).

72

The salt composition of the precipitant solution was the key determinant of which crystal form was obtained, which was also influenced by pH to a lesser extent.

We did not notice any prep-to-prep variation in which crystal forms were obtained from each precipitant solution. Crystals of the inward-closed structure were obtained in 1 M

NaCl, 50 mM Na acetate-HCl pH 4.6, 40% PEG400. Crystals of the inward-open structure were obtained in 200 mM ammonium sulfate, 50 mM Tris-HCl pH 8.5, 20%

PEG400; or in 200 mM ammonium sulfate, 100 mM Na citrate tribasic-HCl pH 5.5, 40%

PEG400. Crystals of the inward-occluded structure were obtained in 100 mM Na citrate tribasic-HCl pH 5.0, 40% PEG200 (MemMeso A4). Crystals of the outward structure were obtained in 1 M NaCl, 50 mM HEPES-NaOH pH 7.5, 20% PEG400; or in 100 mM

MgCl2, 100 mM NaCl, 100 mM HEPES-NaOH pH 7.0, 30% PEG500 DME (MemMeso

E9); or in 100 mM MgCl2, 100 mM NaCl, 100 mM MES-NaOH pH 6.0, 30% PEG500

MME (MemMeso F9). Crystals grew to full size in 4 weeks and were flash-frozen in liquid nitrogen without additional cryo-protectant.

3.4.3 Data Collection and Structure Determination

We collected X-ray diffraction data at the NECAT 24-ID-C and 24-ID-E beamlines

with a wavelength of ~0.98 Å. For the inward-occluded structure, we attempted to

collect redundant data at a longer wavelength (1.65 Å) to locate anomalous difference

Fourier electron density peaks corresponding to the diphosphate moiety of lipid II, but

73

without success likely due to the inherent flexibility of lipid II (~68 rotatable torsions).

Nevertheless, the long-wavelength datasets were merged with the datasets collected at

0.98 Å for refinement as they were isomorphous and there was no noticeable radiation damage. Data were processed by XDS [125], and XDS-processed data from multiple isomorphous crystals were merged by POINTLESS and AIMLESS in BLEND [126-128].

Due to diffraction anisotropy in the high-resolution shells, merged data were subjected to anisotropy correction in STARANISO [129]. Anisotropy corrected data were used for molecular replacement using our previously published structure of MurJTA (PDB: 5T77) as the search model in PHASER [105]. For the outward structure, the search model was split into two fragments (residues 4-228 and residues 229-470) and a two-component

search yielded a strong MR solution with log-likelihood gain of 1588 and translation

function Z-score of 40. MR solutions were subjected to iterative rounds of model

building in COOT [106] and refinement in PHENIX.refine [107]. Structures were refined

to a final Rwork/Rfree of 25.5/28.0% (inward-closed structure, 3.2-Å resolution), 25.4/27.8%

(inward-open structure, 3.0-Å resolution), 22.9/25.8% (inward-occluded structure, 2.6-Å resolution), and 17.9/19.9% (outward structure, 1.8-Å resolution). The final models were refined to excellent geometry (>97% Ramachandran favored, 0% Ramachandran outliers,

<1% rotamer outliers) and minimal clashes (clashscore < 2.5). Crystals contained either a single MurJ molecule in the asymmetric unit (inward-occluded and outward structures),

74

or two MurJ molecules (inward-closed and inward-open structures). Structural alignments and molecular graphics were created in PyMOL [114].

3.4.4 Docking of Lipid II and Molecular Dynamics Simulation

The L-lysine form of lipid II was docked into the crystal structure of the inward-

occluded conformation as described in the previous chapter. Unresolved sidechains of

Arg18 and Lys53 were built manually with the least-clashing rotamer and subsequently

treated as flexible for docking. Docking was performed by Autodock Vina with a search

space of 24 x 24 x 38 Å encompassing the central cavity, portal, and groove [108].

Molecular dynamics simulations starting from the docked model were set up as

follows. The protein molecule from docking was processed by PDB2PQR to add all

hydrogens (at neutral pH) and all missing sidechains [130, 131]. The protein molecule

was embedded into a 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoethanolamine (POPE)

membrane of 140x120 Å with the membrane plugin in VMD, which also removes any

POPE molecules that overlap with the embedded protein [132]. This was performed

using the CHARMM all-hydrogen forcefield topology file for protein and lipids

(top_all27_prot_lipid.inp) [133-135].

The lipid II molecule in the docked conformation was processed by

PHENIX.ReadySet to add back all the hydrogens with protonation state set to neutral

pH, and the lipid II CHARMM forcefield topology file was generated by the SWISS-

75

PARAM server [136]. The embedded protein molecule, lipid II molecule, and POPE lipid bilayer were combined into a single file and solvated with TIP3P waters on both sides of the bilayer to an overall z-dimension of 100 Å using the Solvate plugin in VMD

[132]. Total charge of the system was neutralized with addition of ~150 mM NaCl by the

Autoionize plugin in VMD [132]. Waters that are in the hydrophobic membrane interior

(excluding those in the solvent-accessible cavity of MurJ) were manually deleted. This solvated and ionized system (~140,600 atoms) was then processed by the AutoPSF plugin in VMD to generate the psf and pdb files for MD simulation, using the aforementioned CHARMM forcefield topology files (53) [132].

Molecular dynamics simulations were performed by NAMD (58), with reference to the membrane proteins tutorial (www.ks.uiuc.edu/Training/Tutorials/science/ membrane/mem-tutorial.pdf). For all simulation runs, an integration time step of 1 fs was used, and all X-H bonds were set to rigid (both bond lengths and angles). A periodic cell of 140 x 120 x 100 Å was used with Particle Mesh Ewald electrostatics evaluation [137], wrapping both waters and POPE lipid molecules that diffuse out of the boundaries back into the opposite side of the cell. To mimic the disorder in an actual membrane, the lipid tails of POPE molecules were first melted while everything else

(including POPE headgroups) were fixed in position. Subsequently, gaps between the protein and POPE molecules were filled by a constrained equilibration run at 310 K and

1 atm, with energy minimization for 1000 steps and simulation for 100 ps with the fixed

76

atoms released but keeping harmonic constraints on the protein. Finally, a production run was performed at 310 K and 1 atm for 100 ps releasing all constraints, and coordinates were recorded at 1 ps intervals.

3.4.5 Protein Expression and Delipidation for In Vitro Assay

Because MurJTA has no endogenous cysteine residues, single cysteine mutants

could be engineered at either the cytoplasmic and/or periplasmic gates to probe the

protein conformation in vitro in the absence or presence of sodium. One liter of each

cysteine mutant was expressed as described for the structural studies. Proteins were

purified as described above but with changes. HEPES-NaOH was replaced with Tris- acetate, and sodium was replaced with either N-methyl-D-glucamine (NMDG) or

potassium. To protect cysteines from oxidation, 1 mM TCEP was used in all buffers.

As the presence of any cardiolipin that is carried over from E. coli lysate could be

inhibitory to MurJ function [49], we took measures to delipidate our MurJ samples. We

exploited the high isoelectric point of MurJTA (predicted to be 8.5-9.5) to delipidate

MurJTA by cation exchange. PreScission protease-treated proteins were diluted fivefold in buffer A (20 mM Tris-acetate pH 8.0, 1 mM DDM) and passed through 0.5 mL of pre- equilibrated sulfopropyl-sepharose fast-flow resin twice in a gravity column at room temperature. Buffer B contained 50 mM Tris-acetate pH 8.0, 300 mM NMDG, 1 mM

DDM). The resin was washed with 5 mL each of 16.7% buffer B, 33.3% buffer B, and

77

50% buffer B, before elution in 2 mL of 100% buffer B, divided into four 0.5 mL fractions, all at room temperature. Eluted protein was extremely pure (single band on SDS-PAGE) and mass spectrometry analysis indicated that all traces of cardiolipin had been successfully removed by the delipidation treatment. We did not perform delipidation for structural studies because the high purity came with drastically reduced recovery

(10-30 µg/L culture). Eluted proteins were concentrated to 3-5 µM before the assay.

3.4.6 In Vitro Cysteine Accessibility Assay

To probe the conformation of MurJTA by direct labeling in vitro, single-cysteine

mutants (no endogenous cysteines) were purified and delipidated in potassium acetate

buffer as detailed above. Mutants were diluted to ~1 µM in reaction buffer containing 20

mM HEPES-KOH pH 7.5, 1 mM DDM, and 150 mM potassium acetate (buffer was

added from 5x stock). Samples (9 µL) were equilibrated at either 20°C or 60°C for 10

minutes, then treated with 1 µL of methoxy-polyethylene glycol 5000 maleimide (PEG-

Mal, 10x stock in water) to a working concentration of 0.5 mM. An equivalent volume of

water was added to the negative control samples. Reactions proceeded at either 20°C or

60°C for 10 minutes and were quenched with 5 mM N-ethyl-maleimide (NEM, 10x stock

dissolved in ethanol) for 10 minutes at room temperature. Samples were mixed with 2x

denaturing buffer (60 mM Tris-HCl pH 6.8, 8 M urea, 4% w/v SDS, 20% v/v glycerol, and

0.01% w/v bromophenol blue), resolved by SDS-PAGE, and stained by Coomassie blue.

78

To probe accessibility changes in the presence of sodium by inverse labeling, single-cysteine mutants were purified and delipidated in NMDG buffer as detailed

above. Samples were diluted to ~1 µM in reaction buffer containing 20 mM Tris-acetate

pH 7.5, 1 mM DDM, and either 150 mM sodium acetate or 150 mM NMDG (buffers were

added from 5x stocks). Samples (9 µL) were equilibrated at 20°C for 10 minutes, then

treated with 1 µL of sodium 2-sulfonatoethyl methanethiosulfonate (MTSES, 10x stock in

water) to a working concentration of 100 µM. An equivalent volume of water was

added to the negative control samples. Because long MTSES incubation times were

found to reduce sensitivity by excessively blocking residues that should be inaccessible,

10 µL of 2x denaturing/labeling buffer was immediately added for a MTSES blocking

time of approximately 2 minutes. This 2x denaturing/labeling buffer contained 10 mM

PEG-Mal dissolved in the 2x denaturing buffer described above. Samples were

immediately homogenized by brief vortex and spun down. Denaturation and labeling

proceeded for 30 minutes under protection from light. Reactions were quenched by

addition of 1% (v/v) BME (from 10% stock). Samples were resolved by SDS-PAGE and

stained as described above. Gel band intensities were quantified in GelQuant.NET and

fold change was determined by dividing the relative accessibility in sodium by the

relative accessibility in NMDG. Relative accessibilities were calculated as follows:

79

4. Conclusions

I have determined the crystal structures of MurJTA in the inward-facing closed,

inward-facing open, inward-facing occluded, and outward-facing states, which together

elucidate the conformational transitions that occur during the MurJ transport cycle. The

inward-facing structures reveal a lateral gating mechanism for the entry of lipid

substrate into the central cavity. A “thin gate” was observed between the N-lobe and C-

lobe in the inward-occluded structure, which could be important for occluding the lipid

II molecule from the cytoplasm and/or for outward transition. The high-resolution

outward-facing structure enabled us to locate a sodium ion and assign the putative

sodium-binding site in MurJ proteins. Guided by the structures, our in vitro cysteine

accessibility studies show that MurJ assumes a predominantly inward-facing

conformation in DDM micelles, with some sodium-dependent conformational

rearrangements.

These MurJ structures highlight the importance of dynamic conformation

changes to the mechanism of lipid II flipping, which could be a design principle for

other flippases in general. In particular, TM 8 of MurJ undergoes dynamic

conformational changes through transient π or 310 helical elements throughout the transport cycle. These dynamic helical secondary structure changes were proposed to be important for gating of Transient Receptor Potential ion channels [138], but their

80

presence in our MurJ structures suggest that they are more universal for membrane transport proteins than were previously recognized.

We recognize that the importance of sodium binding (and its role in conformational transition) will require further clarification. At the time of writing, complementation and lipid II accumulation assays are in progress to examine the in-vivo functional competency of MurJTA sodium-site mutants. If sodium binding is important for MurJ function, we would expect mutants of sodium-coordinating residues to be non- functional and result in lipid II accumulation in cells depleted of endogenous MurJ.

Aside from sodium binding, factors responsible for the reported dependence of MurJ activity on membrane potential [75] also remain unknown and warrant further investigation.

The molecular determinants of substrate binding could be further investigated by mutating residues in the central cavity, portal, or hydrophobic groove. Substrate binding could be assayed by approaches such as native mass spectrometry as was previously done [49], or by other approaches such as isothermal titration calorimetry. It is notable that lipid II is sparingly soluble in water, so technical optimization (such as choice and concentration of detergents or organic solvents) will be required. We hypothesize that substrate binding (to the inward-facing state) could initiate outward transition. If this is true, we would expect residues at the periplasmic gate to become

81

more solvent accessible in the presence of increasing concentrations of lipid II. This can be probed by the in vitro cysteine accessibility assay established in this work.

In closing, recent studies from the perspectives of microbiology [78, 79], chemistry [49, 75], cell biology [76], and structural biology [80, 117] have all unequivocally supported the importance of MurJ to bacterial physiology. Our MurJ structures captured in multiple conformations, together with our structural analyses, provide a framework for future studies on this important bacterial protein.

82

Appendix A: Alignment of MurJ Sequences

83

Figure 34: Alignment of MurJ sequences

Positions shaded in red are identical in >90% of Gram-negative MurJ sequences in an alignment with 36 sequences, while those shaded gray are similar in >70% of the sequences. Asterisks (*) denote essential residues in MurJTA of which alanine mutants fail to complement E. coli MurJ. Carets (^) denote residues participating in Na+ coordination. Bullets (•) denote other residues discussed in this dissertation. MurJTA from Thermosipho africanus (UniProt ID: B7IE18) was aligned with 36 Gram-negative MurJ sequences from the UniRef50 library (no sequence is more than 50% identical to another). MATE transporters PfMATE and NorM-NG were also included in the alignment.

84

References

1. Dik, D.A., Fisher, J.F., and Mobashery, S. Cell-Wall Recycling of the Gram- Negative Bacteria and the Nexus to Antibiotic Resistance. Chem Rev, 2018. 118(12): p. 5952-5984.

2. Caveney, N.A., Li, F.K., and Strynadka, N.C. Enzyme structures of the bacterial peptidoglycan and wall teichoic acid biogenesis pathways. Curr Opin Struct Biol, 2018. 53: p. 45-58.

3. El Zoeiby, A., Sanschagrin, F., and Levesque, R.C. Structure and function of the Mur enzymes: development of novel inhibitors. Mol Microbiol, 2003. 47(1): p. 1- 12.

4. Ikeda, M., Wachi, M., Jung, H.K., Ishino, F., and Matsuhashi, M. The Escherichia coli mraY gene encoding UDP-N-acetylmuramoyl-pentapeptide: undecaprenyl- phosphate phospho-N-acetylmuramoyl-pentapeptide . J Bacteriol, 1991. 173(3): p. 1021-6.

5. Mengin-Lecreulx, D., Texier, L., Rousseau, M., and van Heijenoort, J. The murG gene of Escherichia coli codes for the UDP-N-acetylglucosamine: N- acetylmuramyl-(pentapeptide) pyrophosphoryl-undecaprenol N- acetylglucosamine transferase involved in the membrane steps of peptidoglycan synthesis. J Bacteriol, 1991. 173(15): p. 4625-36.

6. Meeske, A.J., Riley, E.P., Robins, W.P., Uehara, T., Mekalanos, J.J., Kahne, D., . . . Rudner, D.Z. SEDS proteins are a widespread family of bacterial cell wall polymerases. Nature, 2016.

7. Cho, H., Wivagg, C.N., Kapoor, M., Barry, Z., Rohs, P.D., Suh, H., . . . Bernhardt, T.G. Bacterial cell wall biogenesis is mediated by SEDS and PBP polymerase families functioning semi-autonomously. Nat Microbiol, 2016: p. 16172.

8. Emami, K., Guyet, A., Kawai, Y., Devi, J., Wu, L.J., Allenby, N., . . . Errington, J. RodA as the missing glycosyltransferase in Bacillus subtilis and antibiotic discovery for the peptidoglycan polymerase pathway. Nat Microbiol, 2017. 2: p. 16253.

85

9. Jackson, G.E. and Strominger, J.L. Synthesis of peptidoglycan by high molecular weight penicillin-binding proteins of Bacillus subtilis and Bacillus stearothermophilus. J Biol Chem, 1984. 259(3): p. 1483-90.

10. Pollock, J.J., Ghuysen, J.M., Linder, R., Salton, M.R., Perkins, H.R., Nieto, M., . . . Johnson, K. Transpeptidase activity of Streptomyces D-alanyl-D carboxypeptidases. Proc Natl Acad Sci U S A, 1972. 69(3): p. 662-6.

11. El Ghachi, M., Bouhss, A., Blanot, D., and Mengin-Lecreulx, D. The bacA gene of Escherichia coli encodes an undecaprenyl pyrophosphate phosphatase activity. J Biol Chem, 2004. 279(29): p. 30106-13.

12. Manat, G., Roure, S., Auger, R., Bouhss, A., Barreteau, H., Mengin-Lecreulx, D., and Touze, T. Deciphering the metabolism of undecaprenyl-phosphate: the bacterial cell-wall unit carrier at the membrane frontier. Microb Drug Resist, 2014. 20(3): p. 199-214.

13. El Ghachi, M., Howe, N., Huang, C.Y., Olieric, V., Warshamanage, R., Touze, T., . . . Caffrey, M. Crystal structure of undecaprenyl-pyrophosphate phosphatase and its role in peptidoglycan biosynthesis. Nat Commun, 2018. 9(1): p. 1078.

14. Workman, S.D., Worrall, L.J., and Strynadka, N.C.J. Crystal structure of an intramembranal phosphatase central to bacterial cell-wall peptidoglycan biosynthesis and lipid recycling. Nat Commun, 2018. 9(1): p. 1159.

15. Sanyal, S. and Menon, A.K. Flipping lipids: why an' what's the reason for? ACS Chem Biol, 2009. 4(11): p. 895-909.

16. van Heijenoort, Y., Gomez, M., Derrien, M., Ayala, J., and van Heijenoort, J. Membrane intermediates in the peptidoglycan metabolism of Escherichia coli: possible roles of PBP 1b and PBP 3. J Bacteriol, 1992. 174(11): p. 3549-57.

17. Ehlert, K. and Holtje, J.V. Role of precursor translocation in coordination of murein and phospholipid synthesis in Escherichia coli. J Bacteriol, 1996. 178(23): p. 6766-71.

18. Mohammadi, T., van Dam, V., Sijbrandi, R., Vernet, T., Zapun, A., Bouhss, A., . . . Breukink, E. Identification of FtsW as a transporter of lipid-linked cell wall precursors across the membrane. EMBO J, 2011. 30(8): p. 1425-32.

86

19. Mohammadi, T., Sijbrandi, R., Lutters, M., Verheul, J., Martin, N., den Blaauwen, T., . . . Breukink, E. Specificity of the transport of Lipid II by FtsW in Escherichia coli. J Biol Chem, 2014.

20. Ruiz, N. Bioinformatics identification of MurJ (MviN) as the peptidoglycan lipid II flippase in Escherichia coli. Proc Natl Acad Sci U S A, 2008. 105(40): p. 15553-7.

21. Inoue, A., Murata, Y., Takahashi, H., Tsuji, N., Fujisaki, S., and Kato, J. Involvement of an essential gene, mviN, in murein synthesis in Escherichia coli. J Bacteriol, 2008. 190(21): p. 7298-301.

22. Sham, L.-T., Butler, E.K., Lebar, M.D., Kahne, D., Bernhardt, T.G., and Ruiz, N. MurJ is the flippase of lipid-linked precursors for peptidoglycan biogenesis. Science, 2014. 345(6193): p. 220-222.

23. Meeske, A.J., Sham, L.T., Kimsey, H., Koo, B.M., Gross, C.A., Bernhardt, T.G., and Rudner, D.Z. MurJ and a novel lipid II flippase are required for cell wall biogenesis in Bacillus subtilis. Proc Natl Acad Sci U S A, 2015. 112(20): p. 6437-42.

24. Gerard, P., Vernet, T., and Zapun, A. Membrane Topology of the Streptococcus pneumoniae FtsW Division Protein. J Bacteriol, 2002. 184(7): p. 1925-1931.

25. Lara, B. and Ayala, J.A. Topological characterization of the essential Escherichia coli cell division protein FtsW. FEMS Microbiol Lett, 2002. 216(1): p. 23-32.

26. Tamaki, S., Matsuzawa, H., and Matsuhashi, M. Cluster of mrdA and mrdB genes responsible for the rod shape and mecillinam sensitivity of Escherichia coli. J Bacteriol, 1980. 141(1): p. 52-7.

27. Ishino, F., Jung, H.K., Ikeda, M., Doi, M., Wachi, M., and Matsuhashi, M. New mutations fts-36, lts-33, and ftsW clustered in the mra region of the Escherichia coli chromosome induce thermosensitive cell growth and division. J Bacteriol, 1989. 171(10): p. 5523-30.

28. Matsuzawa, H., Hayakawa, K., Sato, T., and Imahori, K. Characterization and genetic analysis of a mutant of Escherichia coli K-12 with rounded morphology. J Bacteriol, 1973. 115(1): p. 436-42.

87

29. Khattar, M.M., Begg, K.J., and Donachie, W.D. Identification of FtsW and characterization of a new ftsW division mutant of Escherichia coli. J Bacteriol, 1994. 176(23): p. 7140-7.

30. Datta, P., Dasgupta, A., Singh, A.K., Mukherjee, P., Kundu, M., and Basu, J. Interaction between FtsW and penicillin-binding protein 3 (PBP3) directs PBP3 to mid-cell, controls cell septation and mediates the formation of a trimeric complex involving FtsZ, FtsW and PBP3 in mycobacteria. Mol Microbiol, 2006. 62(6): p. 1655-73.

31. Fraipont, C., Alexeeva, S., Wolf, B., van der Ploeg, R., Schloesser, M., den Blaauwen, T., and Nguyen-Disteche, M. The integral membrane FtsW protein and peptidoglycan PBP3 form a subcomplex in Escherichia coli. Microbiology, 2011. 157(Pt 1): p. 251-9.

32. Mercer, K.L. and Weiss, D.S. The Escherichia coli cell division protein FtsW is required to recruit its cognate transpeptidase, FtsI (PBP3), to the division site. J Bacteriol, 2002. 184(4): p. 904-12.

33. Karimova, G., Dautin, N., and Ladant, D. Interaction network among Escherichia coli membrane proteins involved in cell division as revealed by bacterial two- hybrid analysis. J Bacteriol, 2005. 187(7): p. 2233-43.

34. Butler, E.K., Davis, R.M., Bari, V., Nicholson, P.A., and Ruiz, N. Structure- function analysis of MurJ reveals a solvent-exposed cavity containing residues essential for peptidoglycan biogenesis in Escherichia coli. J Bacteriol, 2013. 195(20): p. 4639-49.

35. Li, G.W., Burkhardt, D., Gross, C., and Weissman, J.S. Quantifying absolute protein synthesis rates reveals principles underlying allocation of cellular resources. Cell, 2014. 157(3): p. 624-35.

36. Butler, E.K., Tan, W.B., Joseph, H., and Ruiz, N. Charge requirements of lipid II flippase activity in Escherichia coli. J Bacteriol, 2014. 196(23): p. 4111-9.

37. Hvorup, R.N., Winnen, B., Chang, A.B., Jiang, Y., Zhou, X.F., and Saier, M.H., Jr. The multidrug/oligosaccharidyl-lipid/polysaccharide (MOP) exporter superfamily. Eur J Biochem, 2003. 270(5): p. 799-813.

88

38. Liu, D., Cole, R.A., and Reeves, P.R. An O-antigen processing function for Wzx (RfbX): a promising candidate for O-unit flippase. J Bacteriol, 1996. 178(7): p. 2102-7.

39. Islam, S.T., Fieldhouse, R.J., Anderson, E.M., Taylor, V.L., Keates, R.A., Ford, R.C., and Lam, J.S. A cationic lumen in the Wzx flippase mediates anionic O- antigen subunit translocation in Pseudomonas aeruginosa PAO1. Mol Microbiol, 2012. 84(6): p. 1165-76.

40. Islam, S.T., Eckford, P.D., Jones, M.L., Nugent, T., Bear, C.E., Vogel, C., and Lam, J.S. Proton-dependent gating and proton uptake by Wzx support O-antigen- subunit antiport across the bacterial inner membrane. MBio, 2013. 4(5): p. e00678- 13.

41. Liu, M.A., Morris, P., and Reeves, P.R. Wzx flippases exhibiting complex O-unit preferences require a new model for Wzx-substrate interactions. Microbiologyopen, 2018: p. e00655.

42. Helenius, J., Ng, D.T., Marolda, C.L., Walter, P., Valvano, M.A., and Aebi, M. Translocation of lipid-linked oligosaccharides across the ER membrane requires Rft1 protein. Nature, 2002. 415(6870): p. 447-50.

43. Frank, C.G., Sanyal, S., Rush, J.S., Waechter, C.J., and Menon, A.K. Does Rft1 flip an N-glycan lipid precursor? Nature, 2008. 454(7204): p. E3-4; discussion E4-5.

44. Sanyal, S., Frank, C.G., and Menon, A.K. Distinct flippases translocate glycerophospholipids and oligosaccharide diphosphate dolichols across the endoplasmic reticulum. Biochemistry, 2008. 47(30): p. 7937-46.

45. Rush, J.S., Gao, N., Lehrman, M.A., Matveev, S., and Waechter, C.J. Suppression of Rft1 expression does not impair the transbilayer movement of Man5GlcNAc2- P-P-dolichol in sealed microsomes from yeast. J Biol Chem, 2009. 284(30): p. 19835-42.

46. Kuroda, T. and Tsuchiya, T. Multidrug efflux transporters in the MATE family. Biochim Biophys Acta, 2009. 1794(5): p. 763-8.

47. Chung, B.C., Zhao, J., Gillespie, R.A., Kwon, D.Y., Guan, Z., Hong, J., . . . Lee, S.Y. Crystal structure of MraY, an essential membrane enzyme for bacterial cell wall synthesis. Science, 2013. 341(6149): p. 1012-6.

89

48. Ha, S., Walker, D., Shi, Y., and Walker, S. The 1.9 A crystal structure of Escherichia coli MurG, a membrane-associated glycosyltransferase involved in peptidoglycan biosynthesis. Protein Sci, 2000. 9(6): p. 1045-52.

49. Bolla, J.R., Sauer, J.B., Wu, D., Mehmood, S., Allison, T.M., and Robinson, C.V. Direct observation of the influence of cardiolipin and antibiotics on lipid II binding to MurJ. Nat Chem, 2018. 10(3): p. 363-371.

50. Tarini, M., Cignoni, P., and Montani, C. Ambient occlusion and edge cueing to enhance real time molecular . IEEE Trans Vis Comput Graph, 2006. 12(5): p. 1237-44.

51. Polissi, A. and Georgopoulos, C. Mutational analysis and properties of the msbA gene of Escherichia coli, coding for an essential ABC family transporter. Mol Microbiol, 1996. 20(6): p. 1221-33.

52. Zhou, Z., White, K.A., Polissi, A., Georgopoulos, C., and Raetz, C.R. Function of Escherichia coli MsbA, an essential ABC family transporter, in lipid A and phospholipid biosynthesis. J Biol Chem, 1998. 273(20): p. 12466-75.

53. Doerrler, W.T. and Raetz, C.R. ATPase activity of the MsbA lipid flippase of Escherichia coli. J Biol Chem, 2002. 277(39): p. 36697-705.

54. Dong, J., Yang, G., and McHaourab, H.S. Structural basis of energy transduction in the transport cycle of MsbA. Science, 2005. 308(5724): p. 1023-8.

55. Borbat, P.P., Surendhran, K., Bortolus, M., Zou, P., Freed, J.H., and McHaourab, H.S. Conformational motion of the ABC transporter MsbA induced by ATP hydrolysis. PLoS Biol, 2007. 5(10): p. e271.

56. Ward, A., Reyes, C.L., Yu, J., Roth, C.B., and Chang, G. Flexibility in the ABC transporter MsbA: Alternating access with a twist. Proc Natl Acad Sci U S A, 2007. 104(48): p. 19005-10.

57. Dawson, R.J. and Locher, K.P. Structure of a bacterial multidrug ABC transporter. Nature, 2006. 443(7108): p. 180-5.

58. Mi, W., Li, Y., Yoon, S.H., Ernst, R.K., Walz, T., and Liao, M. Structural basis of MsbA-mediated lipopolysaccharide transport. Nature, 2017. 549(7671): p. 233-237.

90

59. Perez, C., Gerber, S., Boilevin, J., Bucher, M., Darbre, T., Aebi, M., . . . Locher, K.P. Structure and mechanism of an active lipid-linked oligosaccharide flippase. Nature, 2015. 524(7566): p. 433-8.

60. Wacker, M., Linton, D., Hitchen, P.G., Nita-Lazar, M., Haslam, S.M., North, S.J., . . . Aebi, M. N-linked glycosylation in Campylobacter jejuni and its functional transfer into E. coli. Science, 2002. 298(5599): p. 1790-3.

61. Menon, I., Huber, T., Sanyal, S., Banerjee, S., Barre, P., Canis, S., . . . Menon, A.K. Opsin is a phospholipid flippase. Curr Biol, 2011. 21(2): p. 149-53.

62. Morra, G., Razavi, A.M., Pandey, K., Weinstein, H., Menon, A.K., and Khelashvili, G. Mechanisms of Lipid Scrambling by the G Protein-Coupled Receptor Opsin. Structure, 2018. 26(2): p. 356-367.e3.

63. Andersen, J.P., Vestergaard, A.L., Mikkelsen, S.A., Mogensen, L.S., Chalat, M., and Molday, R.S. P4-ATPases as Phospholipid Flippases-Structure, Function, and Enigmas. Front Physiol, 2016. 7: p. 275.

64. Brunner, J.D., Lim, N.K., Schenck, S., Duerst, A., and Dutzler, R. X-ray structure of a calcium-activated TMEM16 lipid scramblase. Nature, 2014. 516(7530): p. 207- 12.

65. Jiang, T., Yu, K., Hartzell, H.C., and Tajkhorshid, E. Lipids and ions traverse the membrane by the same physical pathway in the nhTMEM16 scramblase. Elife, 2017. 6.

66. Abellon-Ruiz, J., Kaptan, S.S., Basle, A., Claudi, B., Bumann, D., Kleinekathofer, U., and van den Berg, B. Structural basis for maintenance of bacterial outer membrane lipid asymmetry. Nat Microbiol, 2017. 2(12): p. 1616-1623.

67. Bi, Y., Mann, E., Whitfield, C., and Zimmer, J. Architecture of a channel-forming O-antigen polysaccharide ABC transporter. Nature, 2018. 553(7688): p. 361-365.

68. Miyauchi, H., Moriyama, S., Kusakizako, T., Kumazaki, K., Nakane, T., Yamashita, K., . . . Nureki, O. Structural basis for xenobiotic extrusion by eukaryotic MATE transporter. Nat Commun, 2017. 8(1): p. 1633.

91

69. He, X., Szewczyk, P., Karyakin, A., Evin, M., Hong, W.X., Zhang, Q., and Chang, G. Structure of a cation-bound multidrug and toxic compound extrusion transporter. Nature, 2010. 467(7318): p. 991-4.

70. Tanaka, Y., Hipolito, C.J., Maturana, A.D., Ito, K., Kuroda, T., Higuchi, T., . . . Nureki, O. Structural basis for the drug extrusion mechanism by a MATE multidrug transporter. Nature, 2013. 496(7444): p. 247-51.

71. Lu, M., Radchenko, M., Symersky, J., Nie, R., and Guo, Y. Structural insights into H+-coupled multidrug extrusion by a MATE transporter. Nat Struct Mol Biol, 2013. 20(11): p. 1310-7.

72. Lu, M., Symersky, J., Radchenko, M., Koide, A., Guo, Y., Nie, R., and Koide, S. Structures of a Na+-coupled, substrate-bound MATE multidrug transporter. Proc Natl Acad Sci U S A, 2013. 110(6): p. 2099-104.

73. Steed, P.R., Stein, R.A., Mishra, S., Goodman, M.C., and McHaourab, H.S. Na(+)- substrate coupling in the multidrug antiporter norm probed with a spin-labeled substrate. Biochemistry, 2013. 52(34): p. 5790-9.

74. Claxton, D.P., Jagessar, K.L., Steed, P.R., Stein, R.A., and McHaourab, H.S. Sodium and proton coupling in the conformational cycle of a MATE antiporter from Vibrio cholerae. Proc Natl Acad Sci U S A, 2018. 115(27): p. E6182-e6190.

75. Rubino, F.A., Kumar, S., Ruiz, N., Walker, S., and Kahne, D.E. Membrane Potential Is Required for MurJ Function. J Am Chem Soc, 2018. 140(13): p. 4481- 4484.

76. Monteiro, J.M., Pereira, A.R., Reichmann, N.T., Saraiva, B.M., Fernandes, P.B., Veiga, H., . . . Pinho, M.G. Peptidoglycan synthesis drives an FtsZ-treadmilling- independent step of cytokinesis. Nature, 2018. 554(7693): p. 528-532.

77. Chu, J., Vila-Farres, X., Inoyama, D., Ternei, M., Cohen, L.J., Gordon, E.A., . . . Brady, S.F. Discovery of MRSA active antibiotics using primary sequence from the human microbiome. Nat Chem Biol, 2016. 12(12): p. 1004-1006.

78. Chu, J., Vila-Farres, X., Inoyama, D., Gallardo-Macias, R., Jaskowski, M., Satish, S., . . . Brady, S.F. Human Microbiome Inspired Antibiotics with Improved beta- Lactam Synergy against MDR Staphylococcus aureus. ACS Infect Dis, 2018. 4(1): p. 33-38.

92

79. Chamakura, K.R., Sham, L.T., Davis, R.M., Min, L., Cho, H., Ruiz, N., . . . Young, R. A viral protein antibiotic inhibits lipid II flippase activity. Nat Microbiol, 2017. 2(11): p. 1480-1484.

80. Kuk, A.C.Y., Mashalidis, E.H., and Lee, S.Y. Crystal structure of the MOP flippase MurJ in an inward-facing conformation. Nat Struct Mol Biol, 2017. 24(2): p. 171-176.

81. Gutmann, D.A., Mizohata, E., Newstead, S., Ferrandon, S., Postis, V., Xia, X., . . . Byrne, B. A high-throughput method for membrane protein solubility screening: the ultracentrifugation dispersity sedimentation assay. Protein Sci, 2007. 16(7): p. 1422-8.

82. Michel, H. Crystallization of membrane proteins. Trends in Biochemical Sciences, 1983. 8(2): p. 56-59.

83. Landau, E.M. and Rosenbusch, J.P. Lipidic cubic phases: a novel concept for the crystallization of membrane proteins. Proc Natl Acad Sci U S A, 1996. 93(25): p. 14532-5.

84. Caffrey, M. and Cherezov, V. Crystallizing membrane proteins using lipidic mesophases. Nat Protoc, 2009. 4(5): p. 706-31.

85. Caffrey, M. A comprehensive review of the lipid cubic phase or in meso method for crystallizing membrane and soluble proteins and complexes. Acta Crystallogr F Struct Biol Commun, 2015. 71(Pt 1): p. 3-18.

86. French, S. and Wilson, K. On the treatment of negative intensity observations. Acta Crystallographica Section A, 1978. 34(4): p. 517-525.

87. Taylor, G.L. Introduction to phasing. Acta Crystallogr D Biol Crystallogr, 2010. 66(Pt 4): p. 325-38.

88. Evans, P. and McCoy, A. An introduction to molecular replacement. Acta Crystallogr D Biol Crystallogr, 2008. 64(Pt 1): p. 1-10.

89. Hendrickson, W.A., Horton, J.R., and LeMaster, D.M. Selenomethionyl proteins produced for analysis by multiwavelength anomalous diffraction (MAD): a

93

vehicle for direct determination of three-dimensional structure. Embo j, 1990. 9(5): p. 1665-72.

90. Wang, B.C. Resolution of phase ambiguity in macromolecular crystallography. Methods Enzymol, 1985. 115: p. 90-112.

91. Studier, F.W. Protein production by auto-induction in high density shaking cultures. Protein Expr Purif, 2005. 41(1): p. 207-34.

92. Li, D., Pye, V.E., and Caffrey, M. Experimental phasing for structure determination using membrane-protein crystals grown by the lipid cubic phase method. Acta Crystallogr D Biol Crystallogr, 2015. 71(Pt 1): p. 104-22.

93. Liu, Q., Zhang, Z., and Hendrickson, W.A. Multi-crystal anomalous diffraction for low-resolution macromolecular phasing. Acta Crystallogr D Biol Crystallogr, 2011. 67(Pt 1): p. 45-59.

94. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr A, 2008. 64(Pt 1): p. 112-22.

95. Adams, P.D., Afonine, P.V., Bunkoczi, G., Chen, V.B., Davis, I.W., Echols, N., . . . Zwart, P.H. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr D Biol Crystallogr, 2010. 66(Pt 2): p. 213-21.

96. Terwilliger, T.C., Adams, P.D., Read, R.J., McCoy, A.J., Moriarty, N.W., Grosse- Kunstleve, R.W., . . . Hung, L.W. Decision-making in structure solution using Bayesian estimates of map quality: the PHENIX AutoSol wizard. Acta Crystallogr D Biol Crystallogr, 2009. 65(Pt 6): p. 582-601.

97. Cowtan, K. 'dm': An automated procedure for phase improvement by density modification. Joint CCP4 and ESF-EACBM Newsletter on Protein Crystallography, 1994. 31: p. 34-38.

98. Mousa, J.J., Yang, Y., Tomkovich, S., Shima, A., Newsome, R.C., Tripathi, P., . . . Jobin, C. MATE transport of the E. coli-derived genotoxin colibactin. Nature Microbiology, 2016. 1: p. 15009.

99. Baker, N.A., Sept, D., Joseph, S., Holst, M.J., and McCammon, J.A. Electrostatics of nanosystems: application to microtubules and the ribosome. Proc Natl Acad Sci U S A, 2001. 98(18): p. 10037-41.

94

100. Moseley, H.G.J. LXXX. The high-frequency spectra of the elements. Part II. The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science, 1914. 27(160): p. 703-713.

101. Hendrickson, W.A. Anomalous diffraction in crystallographic phase evaluation. Q Rev Biophys, 2014. 47(1): p. 49-93.

102. Mohamed, Y.F. and Valvano, M.A. A Burkholderia cenocepacia MurJ (MviN) homolog is essential for cell wall peptidoglycan synthesis and bacterial viability. Glycobiology, 2014. 24(6): p. 564-76.

103. Fox, B.G. and Blommel, P.G. Autoinduction of protein expression. Curr Protoc Protein Sci, 2009. Chapter 5: p. Unit 5.23.

104. Otwinowski, Z. and Minor, W., [20] Processing of X-ray diffraction data collected in oscillation mode, in Methods in Enzymology. 1997, Academic Press. p. 307-326.

105. McCoy, A.J., Grosse-Kunstleve, R.W., Adams, P.D., Winn, M.D., Storoni, L.C., and Read, R.J. Phaser crystallographic software. Journal of Applied Crystallography, 2007. 40(4): p. 658-674.

106. Emsley, P., Lohkamp, B., Scott, W.G., and Cowtan, K. Features and development of Coot. Acta Crystallogr D Biol Crystallogr, 2010. 66(Pt 4): p. 486-501.

107. Afonine, P.V., Grosse-Kunstleve, R.W., Echols, N., Headd, J.J., Moriarty, N.W., Mustyakimov, M., . . . Adams, P.D. Towards automated crystallographic structure refinement with phenix.refine. Acta Crystallogr D Biol Crystallogr, 2012. 68(Pt 4): p. 352-67.

108. Trott, O. and Olson, A.J. AutoDock Vina: improving the speed and accuracy of docking with a new scoring function, efficient optimization, and multithreading. J Comput Chem, 2010. 31(2): p. 455-61.

109. Huber, R., Langworthy, T.A., König, H., Thomm, M., Woese, C.R., Sleytr, U.B., and Stetter, K.O. Thermotoga maritima sp. nov. represents a new genus of unique extremely thermophilic eubacteria growing up to 90°C. Archives of Microbiology, 1986. 144(4): p. 324-333.

95

110. Boniface, A., Bouhss, A., Mengin-Lecreulx, D., and Blanot, D. The MurE synthetase from Thermotoga maritima is endowed with an unusual D-lysine adding activity. J Biol Chem, 2006. 281(23): p. 15680-6.

111. Moriarty, N.W., Grosse-Kunstleve, R.W., and Adams, P.D. electronic Ligand Builder and Optimization Workbench (eLBOW): a tool for ligand coordinate and restraint generation. Acta Crystallogr D Biol Crystallogr, 2009. 65(Pt 10): p. 1074-80.

112. Moriarty, N.W., Draizen, E.J., and Adams, P.D. An editor for the generation and customization of geometry restraints. Acta Crystallogr D Struct Biol, 2017. 73(Pt 2): p. 123-130.

113. Hsu, S.T., Breukink, E., Tischenko, E., Lutters, M.A., de Kruijff, B., Kaptein, R., . . . van Nuland, N.A. The nisin-lipid II complex reveals a pyrophosphate cage that provides a blueprint for novel antibiotics. Nat Struct Mol Biol, 2004. 11(10): p. 963- 7.

114. Delano, W.L. The PyMOL Molecular Graphics System. 2002.

115. Pei, J., Kim, B.H., and Grishin, N.V. PROMALS3D: a tool for multiple protein sequence and structure alignments. Nucleic Acids Res, 2008. 36(7): p. 2295-300.

116. Pettersen, E.F., Goddard, T.D., Huang, C.C., Couch, G.S., Greenblatt, D.M., Meng, E.C., and Ferrin, T.E. UCSF Chimera--a visualization system for exploratory research and analysis. J Comput Chem, 2004. 25(13): p. 1605-12.

117. Zheng, S., Sham, L.T., Rubino, F.A., Brock, K.P., Robins, W.P., Mekalanos, J.J., . . . Kruse, A.C. Structure and mutagenic analysis of the lipid II flippase MurJ from Escherichia coli. Proc Natl Acad Sci U S A, 2018. 115(26): p. 6709-6714.

118. Zheng, H., Cooper, D.R., Porebski, P.J., Shabalin, I.G., Handing, K.B., and Minor, W. CheckMyMetal: a macromolecular metal-binding validation tool. Acta Crystallogr D Struct Biol, 2017. 73(Pt 3): p. 223-233.

119. Ficici, E., Zhou, W., Castellano, S., and Faraldo-Gomez, J.D. Broadly conserved Na(+)-binding site in the N-lobe of prokaryotic multidrug MATE transporters. Proc Natl Acad Sci U S A, 2018. 115(27): p. E6172-e6181.

96

120. Cross, T.A., Sharma, M., Yi, M., and Zhou, H.X. Influence of solubilizing environments on membrane protein structures. Trends Biochem Sci, 2011. 36(2): p. 117-25.

121. Ritchie, T.K., Grinkova, Y.V., Bayburt, T.H., Denisov, I.G., Zolnerciks, J.K., Atkins, W.M., and Sligar, S.G. Chapter 11 - Reconstitution of membrane proteins in phospholipid bilayer nanodiscs. Methods Enzymol, 2009. 464: p. 211-31.

122. Civjan, N.R., Bayburt, T.H., Schuler, M.A., and Sligar, S.G. Direct solubilization of heterologously expressed membrane proteins by incorporation into nanoscale lipid bilayers. Biotechniques, 2003. 35(3): p. 556-60, 562-3.

123. Denisov, I.G., Grinkova, Y.V., Lazarides, A.A., and Sligar, S.G. Directed self- assembly of monodisperse phospholipid bilayer Nanodiscs with controlled size. J Am Chem Soc, 2004. 126(11): p. 3477-87.

124. Johnson, Z.L. and Lee, S.Y. Liposome reconstitution and transport assay for recombinant transporters. Methods Enzymol, 2015. 556: p. 373-83.

125. Kabsch, W. XDS. Acta Crystallogr D Biol Crystallogr, 2010. 66(Pt 2): p. 125-32.

126. Foadi, J., Aller, P., Alguel, Y., Cameron, A., Axford, D., Owen, R.L., . . . Evans, G. Clustering procedures for the optimal selection of data sets from multiple crystals in macromolecular crystallography. Acta Crystallogr D Biol Crystallogr, 2013. 69(Pt 8): p. 1617-32.

127. Evans, P.R. and Murshudov, G.N. How good are my data and what is the resolution? Acta Crystallogr D Biol Crystallogr, 2013. 69(Pt 7): p. 1204-14.

128. Evans, P.R. An introduction to data reduction: space-group determination, scaling and intensity statistics. Acta Crystallogr D Biol Crystallogr, 2011. 67(Pt 4): p. 282-92.

129. Tickle, I.J., Flensburg, C., Keller, P., Paciorek, W., Sharff, A., and Vonrhein, C., Bricogne, G. STARANISO. Cambridge, United Kingdom: Global Phasing Ltd., 2017.

130. Dolinsky, T.J., Nielsen, J.E., McCammon, J.A., and Baker, N.A. PDB2PQR: an automated pipeline for the setup of Poisson-Boltzmann electrostatics calculations. Nucleic Acids Res, 2004. 32(Web Server issue): p. W665-7.

97

131. Dolinsky, T.J., Czodrowski, P., Li, H., Nielsen, J.E., Jensen, J.H., Klebe, G., and Baker, N.A. PDB2PQR: expanding and upgrading automated preparation of biomolecular structures for molecular simulations. Nucleic Acids Res, 2007. 35(Web Server issue): p. W522-5.

132. Humphrey, W., Dalke, A., and Schulten, K. VMD: visual molecular dynamics. J Mol Graph, 1996. 14(1): p. 33-8, 27-8.

133. MacKerell, A.D., Bashford, D., Bellott, M., Dunbrack, R.L., Evanseck, J.D., Field, M.J., . . . Karplus, M. All-atom empirical potential for molecular modeling and dynamics studies of proteins. J Phys Chem B, 1998. 102(18): p. 3586-616.

134. Schlenkrich, M., Brickmann, J., D. MacKerell, A., and Karplus, M., An Empirical Potential Energy Function for Phospholipids: Criteria for Parameter Optimization and Applications. 1996. 31-81.

135. Feller, S.E., Yin, D., Pastor, R.W., and MacKerell, A.D., Jr. Molecular dynamics simulation of unsaturated lipid bilayers at low hydration: parameterization and comparison with diffraction studies. Biophys J, 1997. 73(5): p. 2269-79.

136. Zoete, V., Cuendet, M.A., Grosdidier, A., and Michielin, O. SwissParam: a fast force field generation tool for small organic molecules. J Comput Chem, 2011. 32(11): p. 2359-68.

137. Darden, T., York, D., and Pedersen, L. Particle mesh Ewald: An N log(N) method for Ewald sums in large systems. The Journal of Chemical Physics, 1993. 98(12): p. ⋅ 10089-10092.

138. Zubcevic, L., Herzik, M.A., Jr., Chung, B.C., Liu, Z., Lander, G.C., and Lee, S.Y. Cryo-electron microscopy structure of the TRPV2 ion channel. Nat Struct Mol Biol, 2016. 23(2): p. 180-186.

98

Biography

Alvin Chun Yin Kuk attended the National University of Singapore initially majoring in architecture. After finding (building-scale) architecture not to be his cup of tea, he switched his interest to molecular architecture and graduated in 2013 with a

Bachelor of Science (first class honors). He began his doctoral studies at Duke

University in 2013 under the guidance of Seok-Yong Lee, working on the structural biology of membrane proteins.

Publications:

Kuk, A. C. Y., Hao, A., Guan, Z. & Lee, S.Y. Visualizing conformation transitions of the lipid flippase MurJ. (Tentative title, manuscript in preparation)

Yoo, J.*, Mashalidis, E.H.*, Kuk, A.C.Y.*, Yamamoto, K., Kaeser, B., Ichikawa, S., and Lee, S.Y. GlcNAc-1-P-transferase-tunicamycin complex structure reveals basis for inhibition of N-glycosylation. Nat Struct Mol Biol, 2018. 25(3): p. 217-224. *Equal contribution

Kuk, A.C.Y., Mashalidis, E.H., and Lee, S.Y. Crystal structure of the MOP flippase MurJ in an inward-facing conformation. Nat Struct Mol Biol, 2017. 24(2): p. 171-176.

99