naatm1: examining the mechanism of a heavy-metal abc exporter

134
e University of San Francisco USF Scholarship: a digital repository @ Gleeson Library | Geschke Center Master's eses eses, Dissertations, Capstones and Projects Spring 3-31-2019 NaAtm1: Examining the mechanism of a Heavy- metal ABC Exporter Dennis Hicks University of San Francisco, [email protected] Follow this and additional works at: hps://repository.usfca.edu/thes Part of the Biochemistry Commons is esis is brought to you for free and open access by the eses, Dissertations, Capstones and Projects at USF Scholarship: a digital repository @ Gleeson Library | Geschke Center. It has been accepted for inclusion in Master's eses by an authorized administrator of USF Scholarship: a digital repository @ Gleeson Library | Geschke Center. For more information, please contact [email protected]. Recommended Citation Hicks, Dennis, "NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter" (2019). Master's eses. 1204. hps://repository.usfca.edu/thes/1204

Upload: others

Post on 14-May-2022

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

The University of San FranciscoUSF Scholarship: a digital repository @ Gleeson Library |Geschke Center

Master's Theses Theses, Dissertations, Capstones and Projects

Spring 3-31-2019

NaAtm1: Examining the mechanism of a Heavy-metal ABC ExporterDennis HicksUniversity of San Francisco, [email protected]

Follow this and additional works at: https://repository.usfca.edu/thes

Part of the Biochemistry Commons

This Thesis is brought to you for free and open access by the Theses, Dissertations, Capstones and Projects at USF Scholarship: a digital repository @Gleeson Library | Geschke Center. It has been accepted for inclusion in Master's Theses by an authorized administrator of USF Scholarship: a digitalrepository @ Gleeson Library | Geschke Center. For more information, please contact [email protected].

Recommended CitationHicks, Dennis, "NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter" (2019). Master's Theses. 1204.https://repository.usfca.edu/thes/1204

Page 2: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

NaAtm1:

Examining the mechanism of a Heavy-metal ABC Exporter

A thesis presented to the faculty

of the Department of Chemistry

at the University of San Francisco

in partial fulfillment of the requirements for the degree of

Master of Science in Chemistry

Written by

Dennis Hicks

Bachelor of Science in Biochemistry

San Francisco State University

August 2019

Page 3: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

NaAtm1:

Examining the mechanism of a Heavy-metal ABC Exporter

Thesis written by Dennis Hicks

This thesis is written under the guidance of the Faculty Advisory Committee, and approved by all its members, has been accepted in partial fulfillment of the requirements for the degree of

Master of Science in Chemistry

at the University of San Francisco

Thesis Committee

Janet G. Yang, Ph.D. Research Advisor

Ryan M. West, Ph.D. Assistant Professor

Nicole M. Thometz, Ph.D. Assistant Professor

Marcelo F. Camperi, Ph.D. Dean, College of Arts and Sciences 

Page 4: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

Acknowledgements

They say it takes a village to raise a child and this can’t be more true for this thesis and my

time at USF. I would like to sincerely thank my advisor Dr. Janet Yang for her patience, support,

and guidance (along with all of the baby things for Rosie). Her immense knowledge and ability

to easily communicate complex ideas has made my education at USF very enjoyable. I would

like to give a special thanks to Jeff Oda, whose incredible knowledge of where stuff is in the

chemistry department has saved the Yang lab countless hours and an untold amount of money.

This thesis wouldn’t have included as much as it does without the help of all current and former

members of the Yang Lab and for that I am eternally grateful. Their help and company made life

in the Yang lab amazing.

I would also like to thank my wife Kylie, for her undying support and eternal patience. She

has no idea how grateful I am for her support and help during my time at USF, as well as taking

the nightshift with our daughter Rosie.

Page 5: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

iv  

Table of Contents

1 Introduction ..............................................................................................................................1

1.1 A Brief history of ABC transporters ........................................................................1

1.2 ABC Transporters in human health .........................................................................4

1.2.1 Multi-drug resistant (MDR) abilities of cancer and bacteria/fungi .........................4

1.2.2 Human ABC transporter related diseases ................................................................5

1.3 Challenges in organizing the ABC transporter family .............................................7

1.4 Basic structure and components of ABC transporters .............................................7

1.4.1 The transmembrane domains (TMDs) .....................................................................9

1.4.2 The nucleotide binding domains (NBDs) ................................................................9

1.4.2.1 Walker A motif (P-loop) ..................................................................................12

1.4.2.2 Walker B motif ................................................................................................13

1.4.2.3 A-loop ..............................................................................................................14

1.4.2.4 C Motif (ABC signature motif)........................................................................15

1.4.2.5 D-loop ..............................................................................................................16

1.4.2.6 H-loop ..............................................................................................................17

1.4.2.7 Q-loop ..............................................................................................................18

1.5 Accessory Domains ..........................................................................................................19

1.5.1 Extracytoplasmic domains .....................................................................................19

1.5.2 Membrane-embedded domains ..............................................................................20

1.5.3 Cytosolic regulatory domains ................................................................................20

1.5.4 Accessory catalytic domains ..................................................................................21

1.6 ABC Transporter classes........................................................................................21

1.7 ABC importer classes and proposed mechanisms .................................................22

1.7.1 Structure of Type I and II importers ......................................................................23

1.7.2 Mechanisms of Type I and II importers .................................................................25

1.7.3 Structure of Type III (ECF) importers ...................................................................29

1.7.3.1 Differences between ECF subgroups I and II ..................................................30

1.7.3.2 Type III substrates............................................................................................30

1.7.4 Mechanism of Type III importers ..........................................................................31

1.8 ABC exporter classes and proposed mechanism ...................................................32

Page 6: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

v  

1.8.1 ABCA ....................................................................................................................33

1.8.2 ABCB .....................................................................................................................33

1.8.3 ABCC .....................................................................................................................33

1.8.4 ABCD ....................................................................................................................34

1.8.5 ABCE .....................................................................................................................34

1.8.6 ABCF .....................................................................................................................34

1.8.7 ABCG ....................................................................................................................35

1.8.8 Proposed exporter mechanism ...............................................................................36

1.9 ABC non-porters ....................................................................................................37

1.10 Protein of interest: an ABC exporter found in mitochondria .................................37

1.10.1 Atm1 in Novosphingobium aromaticivorans .........................................................39

1.10.2 NaAtm1 substrate binding sites .............................................................................41

1.10.3 In vitro characterization of NaAtm1 activity .........................................................42

1.10.4 In vivo functionality of NaAtm1 ............................................................................43

1.11 Studies presented in this work ...............................................................................43

1.12 References ..............................................................................................................44

2 Protein Purification ...............................................................................................................54

2.1 Method theories .....................................................................................................54

2.1.1 Immobilized metal affinity chromatography (IMAC) ...........................................55

2.1.1.1 Poly-histidine tags ............................................................................................56

2.1.2 Size exclusion chromatography (SEC) ..................................................................57

2.1.3 Sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE) ...............59

2.2 Specialized instruments .........................................................................................62

2.2.1 Fast protein liquid chromatography (FPLC) instrument ........................................63

2.2.2 Spectrophotometer .................................................................................................64

2.3 Purification of NaAtm1 ..........................................................................................65

2.3.1 NaAtm1 purification overview ..............................................................................65

2.3.2 NaAtm1 protein expression ...................................................................................65

2.3.3 NaAtm1 protein purification ..................................................................................66

Page 7: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

vi  

2.3.4 NaAtm1 IMAC purification results .......................................................................67

2.3.5 NaAtm1 SEC purification results ..........................................................................69

2.3.6 Analysis of NaAtm1 ..............................................................................................70

2.4 Purification of MSP1D1 ........................................................................................72

2.4.1 MSP1D1 purification overview .............................................................................72

2.4.2 MSP1D1 protein expression ..................................................................................72

2.4.3 MSP1D1 protein purification .................................................................................73

2.4.4 MSP1D1 IMAC purification results ......................................................................74

2.4.5 MSP1D1 SEC purification results .........................................................................76

2.4.6 Analysis of MSP1D1 .............................................................................................77

2.5 Purification of TEV protease .................................................................................79

2.5.1 TEV protease purification overview ......................................................................79

2.5.2 TEV protease protein expression ...........................................................................79

2.5.3 TEV protease protein purification .........................................................................80

2.5.4 TEV protease IMAC purification results ...............................................................81

2.5.5 TEV protease SEC purification results ..................................................................83

2.5.6 Analysis of TEV protease ......................................................................................83

2.6 Conclusion .............................................................................................................85

2.7 References ..............................................................................................................86

3 Creating a realistic environment in vitro .............................................................................89

3.1 Introduction ............................................................................................................89

3.2 Types of environments ...........................................................................................90

3.2.1 Detergent micelles .................................................................................................91

3.2.2 Liposomes ..............................................................................................................92

3.2.3 Nanodiscs ...............................................................................................................93

3.3 Overview of nanodisc reconstitutions ....................................................................95

3.3.1 Preparation of nanodisc reagents ...........................................................................96

3.3.1.1 Membrane scaffold protein (MSP) ..................................................................96

3.3.1.2 1:3 PC:E lipids .................................................................................................97

Page 8: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

vii  

3.3.1.3 Bio-Beads .........................................................................................................99

3.3.2 Reconstitution methods ..........................................................................................99

3.3.2.1 Assembly and purification of small-scale reconstitutions ...............................99

3.4 Reconstitution Results .........................................................................................100

3.4.1 Identification of multiple species present in reconstitutions ................................100

3.4.1.1 Reconstitutions without NaAtm1: 0-3-60 ......................................................101

3.4.1.2 Reconstitutions without MSP1D1: 1-0-60 .....................................................102

3.4.1.3 Reconstitutions without additional lipids: 1-3-0 ............................................103

3.4.2 Conclusion for peak identification in small-scale reconstitutions .......................104

3.4.3 Reconstitution optimization .................................................................................104

3.4.3.1 NaAtm1:MSP1D1 optimization.....................................................................104

3.4.3.2 NaAtm1:lipid optimization ............................................................................105

3.4.3.3 Final NaAtm1 concentration optimization .....................................................107

3.4.4 Optimization conclusion ......................................................................................108

3.5 Determining the stoichiometry of nanodiscs using densitometry ........................108

3.6 Quantification of NaAtm1 nanodiscs via absorbance spectroscopy ....................110

3.7 Transition to an IMAC-based purification method ..............................................111

3.8 Optimization of His-tag cleavage by TEV protease ............................................112

3.9 Preparation of His-tag-cleaved MSP1D1 .............................................................113

3.10 Purification of NaAtm1 in MSP1D1 nanodisc using IMAC ...............................115

3.10.1 Assembly and purification of large-scale reconstitutions ....................................116

3.10.2 Analysis of large-scale reconstitutions ................................................................117

3.11 Conclusion ...........................................................................................................118

3.12 References ............................................................................................................119

4 Preliminary Kinetic Assays and Future Works ................................................................121

4.1 Introduction ..........................................................................................................121

4.2 Kinetic assays.......................................................................................................121

4.3 Future works ........................................................................................................123

4.3.1 Further kinetic assays ...........................................................................................123

Page 9: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

viii  

4.3.2 Detection of conformational changes ..................................................................124

4.4 Conclusion ...........................................................................................................125

4.5 References ............................................................................................................136

   

Page 10: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

  

Abstract

NaAtm1:

Examining the mechanism of a Heavy-metal ABC Exporter

Dennis Hicks

ATP-Binding Cassette (ABC) Transporters are a superfamily of integral membrane

proteins that have been found embedded in both the cellular membranes and internal organelles of

all species yet analyzed (1). This extreme conservation throughout the tree of life is due to their

vitally critical and broad task of transporting all manner of things into and out of the cell and

related organelles. This broadness of responsibility has led to an abundant diversity of this family,

which in turn, has allowed life to adapt to the many environments found on Earth. While the

importance and variability of this family is abundantly clear, many mechanistic and regulatory

questions remain. The diversity which has made this family so useful in the natural world has

resulted in a huge challenge in the scientific world: elucidating the many mechanisms this family

has developed to perform its many functions.

The scientific community has so far been able to identify overarching features of all ABC

transporters such as the Walker A and Walker B motifs (1); however, the mechanisms of individual

transporters remain undiscovered. In my research, I focus on a specific ABC transporter called

NaAtm1 (Novosphingobium aromaticivorans ABC transporter of mitochondria 1), which has been

shown to mediate transition metal export in its host species (2). While some transition metals (e.g.

iron and copper) are essential for certain cellular processes such as respiration and enzyme

catalysis, other metals (e.g. cadmium and mercury) can be very toxic. Due to this extreme

dichotomy, a cell’s ability to distinguish between friend and foe is of the utmost importance, and

therefore is a highly regulated process.

Page 11: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

  

My goal during this program was to purify NaAtm1 and create a realistic in-vitro

environment to analyze and better understand the full mechanism of this transporter, to determine

why it exports some metals while leaving others alone, and to understand the regulatory processes

that control it.

References

[1] Davidson, A. L.; Dassa, E.; Orelle, C.; Chen, J. Structure, function, and evolution of bacterial ATP-binding cassette systems. Microbiology and Molecular Biology Reviews 2008, 72(2), 317–364.

[2] Lee, J. Y.; Yang, J. G.; Zhitnitsky, D.; Lewinson, O.; Rees, D. C. Structural basis for heavy metal detoxification by an Atm1-type ABC exporter. Science 2014, 343(6175), 1133–11

Page 12: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

1  

CHAPTER 1

Introduction

In this work, the methodology to reconstitute a purified ABCB-type ATP Binding

Cassette transporter into a lipid bilayer mimetic environment is developed and optimized for use

in kinetic studies. This chapter focuses on the history and importance of ABC transporters,

transporter types and structure, as well as an overview of the protein of interest, NaAtm1.

1.1 A Brief History of ABC Transporters

The study of ABC transporters began in the late 1960’s with the study of membrane

transport systems for the uptake of nutrients in the gram-negative bacteria E. coli. At this time, all

cellular transport systems were known as permeases. Scientists divided this large group into two

smaller classes: 1) osmotic shock-sensitive transport systems and 2) osmotic shock-insensitive

transport systems.

The process of osmotic shock was first described by Harold Neu and Leon Heppel in 1965.

The process involves both a rapid change in the concentration of sucrose around a cell and the

partial destruction of the outer membrane resulting in the release of periplasmic proteins. Briefly,

gram-negative bacteria are placed in a solution that contains concentrated sucrose. The sucrose

draws water out of the cytoplasm shrinking the inner membrane and increasing the periplasmic

volume while the addition of membrane-lysing enzymes and chemicals create holes in the outer

membrane. The cells are then shifted to a solution containing no sucrose, which causes water to

rapidly enter the inner cell. This increase in cytoplasmic volume results in the rapid reduction of

periplasmic volume and pushes out many water-soluble periplasmic proteins through the lysed cell

wall [1, 2].

Page 13: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

2  

This technique of osmotic shock allowed scientists to develop the two-class theory of

shock-sensitive and insensitive permeases. Further work on these two classes led to the discovery

that the osmotic shock-sensitive nutrient permeases relied on a helper protein found solubilized

within the periplasm [3, 4]. These proteins shepherded nutrients to their respective transporters

found on the inner membrane and were later named substrate binding proteins (SBPs). It was the

resulting absence of these proteins from the expulsion of the periplasm during osmotic shock

which gave this class its name. Research also showed that this class of transporter directly derived

their energy from adenosine triphosphate (ATP) instead of from a concentration gradient, as is the

case with the osmotic shock-insensitive transport class [5]. This osmotic shock-sensitive class

would later become known as ATP-Binding Cassette (ABC) transporters.

By the early 1980’s, scientists had started cloning many genes that encoded for this class

of transporter; most notably the maltose transporter found in E. coli [6] and the histidine transporter

of S. Typhimurium [7]. At the same time, other researchers were looking into why some

mammalian cell lines –particularly certain cancer cell lines– had the unique ability to resist a wide

range of chemotherapy drugs. In 1976, researchers in Toronto discovered a large membrane

protein that was found in abundance in mammalian cells that were multi-drug resistant (MDR),

but not found in the corresponding wild-type cells (Figure 1-1). This protein was named

permeability glycoprotein (P-gp) [8] due to its ability to modulate drug permeability. Not long

after this discovery, P-gp was successfully cloned [9] and its amino acid sequence compared to

that of other transport proteins, both bacterial and mammalian.

The comparisons showed that sequences from certain sections of many transport proteins

were highly conserved. Two of the most highly conserved sequences are known as the Walker A

and Walker B motifs and were first described in ATP-requiring enzymes by Walker et al. in 1982

Page 14: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

3  

[10]. These two motifs were determined to be involved in the binding and catalysis of ATP,

respectively, and were the first hints at a very distant common ancestor, both among ABC

transporters and ATP-requiring enzymes in general. Further comparisons between mammalian and

bacterial transport proteins showed that these transporters were part of a large and diverse family

of proteins [11, 12, 13], and in the early 1990’s, the term ABC transporter was coined [14].

Figure 1-1. Timeline of ABC transporter research reproduced from Theodoulou et al. [15]. Notable discoveries include

the identification of P-gp in 1976 and the cause of CFTR mis-localization (F508Δ) in 1992.

With the help of cutting-edge technologies such as genetic sequencing, X-ray

crystallography, genomic libraries, and computer data analysis and modeling, the 1990’s and early

2000’s saw an explosion of new data and discoveries regarding this family (Figure 1-1). While

these new discoveries have led to a general understanding of the overall architecture of ABC

Page 15: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

4  

transporters, many questions remain unanswered. Most studied topics relating to ABC transporters

include substrate affinity, cycle mechanisms, intra-protein and protein-protein interactions, and

transporter kinetics. These topics are of immense interest, not only for their intellectual value, but

also their practical application for human health.

1.2 ABC Transporters in human health

1.2.1 Multi-drug resistant (MDR) abilities in cancer and bacteria/fungi

It was clear early on that at least one ABC

transporter, P-gp, played a major role in multi-drug

resistant cancers within mammals (Figure 1-2). P-gp,

also known as multidrug resistant protein 1 (MDR1),

has been shown to export a wide variety of xenobiotics

from human cells, including pharmaceutical drugs [18].

While cancer cells may be sensitive chemotherapeutics,

cells that overexpress P-gp can extrude these toxic

drugs, lowering the intracellular concentration and

ultimately rendering the drug ineffective. These cancer cells over-expressing P- gp are rendered

multidrug resistant.

While P-gp was first found in Chinese hamsters [8], homologs have been found in both

bacteria and fungi as well [17, 18]. Due in part to the over-prescription of antibiotics and the failure

of patients to use them correctly, many strains of bacteria and fungi have been able to develop their

own version of P-gp and other MDR transporters to remove a staggeringly wide range of

            Figure 1-2. Structure of P-gp as reported by

Aller et al. [16]. The red portion represents the component of P-gp surrounded by a lipid bilayer. Image from RCSB (3G5U). 

Page 16: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

5  

antibiotics, making the infections they cause virtually untreatable. Many scientists fear that we are

fast approaching the “post-antibiotic” era; a time at which all current antibiotics will no longer be

effective [20]. Along with antibiotic resistance, MDR presents a major hurdle in the development

of antivirals, antiparasitics, and antifungals. Fortunately, the study of ABC transporters offers a

possible solution to this challenging scenario. If scientists can elucidate the mechanisms of these

MDR transporters, they might also find a way to disable them.

1.2.2 Human ABC Transporter Related Diseases

While multi-drug resistant infections are of great importance to human health, so too is the

study of human ABC transporters. So far, about 50 human ABC transporters have been discovered

[21]. Of these, 14 have been shown to cause 13 different diseases (Table 1-1), the most notable

being cystic fibrosis.

Cystic fibrosis is caused by a point mutation in the ABCC7/CFTR transporter that is

involved in chloride ion transport in epithelial cells that are part of mucous membranes. These

membranes are abundant in the lungs, pancreas, and liver. When functioning normally, the

ABCC7/CFTR protein transports chloride ions to the cell surface. The chloride ions attract water

which results in a less viscous mucus that can be cleared from the lungs by cilia also located on

the cell surface. When the ABCC7/CFTR protein is mutated, chloride ions are unable to be

transported to the cell surface, resulting in less water accumulation and a thick mucus that is unable

to be moved by cilia and therefore accumulates within affected organs.

Symptoms of cystic fibrosis include mucus with an abnormal viscosity and sweat with a

high salt concentration [22]. These symptoms usually present in the lungs, pancreas and liver and

Page 17: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

6  

result in difficulty breathing, low amounts of digestive enzymes, and blockage of bile ducts,

respectively.

Table 1-1. List of current known diseases in which ABC transporters play a role [21]. Most notable diseases are

cystic fibrosis and Stargardt disease, caused by damaged ABCC7 and ABCA4 transporters respectively.

By understanding how these transporters work and what mutations in their structure lead

to their dysfunction, scientists may one day be able to repair them through gene editing. While this

technology is still in the very early stages of development, it has been used to treat cystic fibrosis

with some success [23].

Page 18: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

7  

1.3 Challenges in organizing the ABC transporter family

The ABC transporter superfamily is one of the largest and most diverse families of protein

in existence. They have been found embedded in both the cellular membranes and internal

organelles of all species yet analyzed [24]. This extreme conservation throughout the tree of life is

due to their vitally critical and broad task of transporting an extensive collection of molecules into

and out of the cell and related organelles. This broadness of responsibility has led to an abundant

diversity within this family, which in turn, has allowed life to adapt to the many environments

found on Earth. This same diversity has also posed an immense challenge to those seeking to

organize such a complex group of proteins for the purpose of making connections between

individual transporters and gaining significant insights regarding transporter form and function.

Before organization can begin, one needs first to understand the components that make up ABC

transporters.

1.4 Basic structure and components of ABC transporters

The typical ABC transporter consists of four core domains: two transmembrane domains

(TMD1 and TMD2) and two nucleotide binding domains (NBD1 and NBD2). These domains can

be connected in many different combinations (Figure 1-3). They can be found as four separate

subunits (tetramer), a heterodimer or homodimer consisting of one TMD and one NBD per subunit

(also known as a “half” transporter), a trimer consisting of a subunit containing two NBDs and

two TMD subunits, or a single polypeptide chain containing all four subunits [25]. While all ABC

transporters must include these four domains, some also include additional components, most

notably substrate binding proteins (SBPs) and regulatory domains.

Page 19: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

8  

Figure 1-3. Various domain combinations reproduced from the HDL Handbook (Third Edition) [26]. As a rule, tetramers

and trimers are only found in bacterial ABC transporters while eukaryotic ABC transporters are made up of either dimers or monomers.

In the following figures, the high-resolution x-ray crystallography structure of the NaAtm1

transporter is used as an example to highlight the basic domain architecture of ABC transporters

(Figure 1-4) . This transporter is the focus of this thesis, and its characteristics are explained in

greater detail at the end of this chapter.

Figure 1-4. TMDs and NBDs of NaAtm1 as reported by Lee et al. [33] NaAtm1 is a homodimer, with each monomer

(orange and green) consisting of one TMD and one NBD.

Page 20: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

9  

1.4.1 The transmembrane domains (TMDs)

Transmembrane domains form the translocation pathway for substrates to move through

the membrane and are typically comprised of between 5 and 10 amphipathic helices per subunit

(10-20 helices total) [24]. This pathway is only open to one side of the membrane at a time and

access to the pathways interior is alternated through the binding and hydrolysis of ATP in the

nucleotide binding domains (NBDs).

The NBDs drive the conformational change in the TMDs through a coupling helix – a short

alpha helix found on the cytosolic side of each TMD (Figure 1-5). This coupling helix fits into a

groove in the neighboring NBD and allows for the transfer of energy to induce the conformational

change in the TMD [27]. While these coupling helices differ depending on the class of ABC

transporter, in most cases each TMD is connected to one NBD resulting in two coupling helices

per transporter.

The TMDs of ABC transporters are extremely diverse but usually share a similar topology

within the same class [28]. This diversity is most likely due to the wide array of substrates

transported by ABC transporters.

1.4.2 The nucleotide binding domains (NBDs)

The nucleotide binding domains (NBDs) are also known as ATPase domains or ATP

Binding Cassettes and give this class of transporter its name [29]. They drive the conformational

change needed to translocate substrates across a membrane through the binding and hydrolysis of

ATP and are part of the family of P-loop NTPases that depend on magnesium ions for catalysis

[30].

Page 21: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

10  

There are always two nucleotide binding domains (NBD1 and NBD2) found within an

ABC transporter, with each domain consisting of two subdomains. These subdomains consist of a

RecA-like domain (found in other ATPases) and an alpha-helical domain (found only in ABC

transporters). Each subdomain in an NBD contributes to one aspect of the reaction (binding or

hydrolysis) and can only couple with the opposite subdomain on the corresponding NBD (Figure

1-5C). The functional nucleotide binding sites are located at the interface between these two

opposing subdomains.

Figure 1-5. A) Crystal structure of NaAtm1 as reported by Lee et al. [31]. B) A simplified diagram of NaAtm1showing

the four domains and coupling helices (unlabeled circles). C) A top-down view the two NBDs of NaAtm1 showing the RecA-like and α-helical subdomains.

Although there are always two NBDs and two ATP binding pockets, it is not always the

case that a transport cycle requires two ATP molecules. For example, while the OpuA transporter

hydrolyzes two molecules of ATP per substrate molecule [32], only one functioning active site is

enough to drive transport by TAP1. Interestingly, TAP1 contains a naturally occurring mutation

that results in one degenerate ATP binding site [33]. Additionally, when one site is artificially

mutated in the histidine transporter HisP2MQJ, the transporter can still function normally [34].

Page 22: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

11  

While the transmembrane domains of this family are diverse, the nucleotide binding

domains show very high conservation. There are seven highly conserved amino acid

sequences/motifs in NBDs that allow for the identification of ABC transporters, as described

below. Of these seven, only the C Motif (ABC signature motif) is located in the α-helical domain;

the other six are located in the RecA-like domain. In the case of the target protein NaAtm1, all

seven sequences appear within its structure (Figure 1-6) and are highlighted in the following

figures.

Figure 1-6. Crystal structure diagram top-down view of the NBDs of NaAtm1 [31]. Each NBD contains one RecA-like

domain and one α-helical domain. Each RecA-like domain contains a Walker A motif (yellow), Walker B motif (pink), A-Loop (black), D-Loop (red), H-Loop (magenta), and Q-loop (blue). Each α-helical domain contains the C motif (cyan). When ATP binds on the RecA-like domain (red circle) and a substrate binds in the substrate binding pocket (not shown), a conformational change takes place causing the α-helical domain on the opposite NBD (blue circle) to move toward the RecA-like domain, sandwiching the ATP between the two and allowing hydrolysis to take place.

Page 23: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

12  

1.4.2.1 Walker A Motif (P-loop)

The Walker A motif connects an α-helix and a ß-strand to form a pocket for ATP binding

within its NBD [35]. The consensus sequence for this motif varies slightly depending on the

source. The most common sequences are GXXXXGKS [36], GXXGXGKS [37],

(G/A)XXXXGK(T/S) [10], and GXXGXXK [38]. The X residues in these sequences signify highly

variable amino acids.

The most highly conserved residues in this motif are the N-terminal glycine and the C-

terminal lysine. The lysine has been shown to bind to the γ-phosphate of ATP and in some cases

is replaced by an arginine which can lower the activity [39]. A serine or threonine usually follows

the highly conserved lysine and provides a hydroxyl group to bind the divalent cation associated

with the bound nucleotide [40]. In NaAtm1, this sequence starts at residue 349 and ends at 401

(Figure 1-7). The sequence is GPSGAGKS and ends in a serine (S401) [31].

Figure 1-7. Cartoon diagram side view of the NaAtm1 Walker A motif (yellow) with the highly conserved residues

glycine (G394), lysine (K400), and serine (401) as reported by Lee et al. [31].

Page 24: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

13  

1.4.2.2 Walker B Motif

The Walker B motif coordinates the magnesium ion needed for catalysis. This ion interacts

with the ß- and ɣ- phosphates of ATP and helps to both stabilize and hydrolyze ATP. The general

consensus sequence for this motif is hhhhD(E) and consists of three components: four hydrophobic

residues (h) followed by an acidic negatively charged residue (usually aspartic acid) and then a

second acidic residue (usually glutamic acid). The second acidic residue (E) is commonly included

in the D-loop conserved sequence which immediately follows the Walker B motif. The aspartic

acid is believed to play a role in coordination of the magnesium ion while the acidic residue

immediately following it acts as a general base, polarizing an attacking water molecule involved

in the hydrolysis of the ɣ- phosphate in ATP [41]. In NaAtm1, this sequence starts at residue 518

and ends at 522 (Figure 1-8). The sequence is ILLFD(E), with the (E) overlapping into the D-

loop.

Figure 1-8. Cartoon diagram side view of the NaAtm1 Walker B motif (pink) with the highly conserved residue aspartic

acid (D522) as reported by Lee et al. [31].

Page 25: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

14  

1.4.2.3 A-loop

The A-loop is composed of a single aromatic residue that helps position the ATP through

stacking with the adenine ring of ATP [42]. This residue generally appears around 25 amino acids

upstream from the Walker A motif. In NaAtm1, the A-loop consists of a tyrosine 24 AA upstream

(Y370) (Figure 1-9).

Figure 1-9. Cartoon diagram side view of the NaAtm1 A-loop (black) with the aromatic residue tyrosine (D522) as

reported by Lee et al. [31].

Page 26: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

15  

1.4.2.4 C Motif (ABC signature motif)

This sequence is a signature of ABC transporters and is not found in other P-loop NTPases

and has a consensus sequence of LSGGQ. It is found in the α-helical subdomain and has been

shown to interact with the RecA-like domain on the opposing NBD to form a binding pocket. The

serine (S499) in this motif interacts through its oxygen with the ɣ- phosphate of ATP [43]. This

interaction may draw these two subunits together causing a conformational change in the NBDs

(going from an open position to a closed position) which results in a conformational change of the

TMDs. In NaAtm1 this sequence starts at residue 498 and ends at 502 (Figure 1-10). This motif

is slightly different than normally reported with a sequence of LSGGE.

Figure 1-10. Cartoon diagram side view of the NaAtm1 C Motif (cyan) with the active residue serine (S522) as reported

by Lee et al. [31].

Page 27: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

16  

1.4.2.5 D-loop

The D-loop is a sequence that immediately follows the Walker B motif and has a consensus

sequence of SALD or EATSALD (overlapping Walker B) depending on the source. In the

EATSALD sequence, the E corresponds to the second acidic residue in the Walker B motif,

glutamate. While this glutamate is involved in polarizing the catalytic water molecule in the active

site, the aspartic acid on the opposite end (D529) interacts with the P-loop (Walker A Motif) of

the opposing active site [44]. This interaction between the D-loop on one NBD and the P-loop of

the opposite NBD is believed to allow communication between the two active sites [45, 46]. In

NaAtm1 this sequence starts at residue 523 and ends at 529 (Figure 1-11). The sequence is

EATSALD.

Figure 1-11. Cartoon diagram side view of the NaAtm1 D-loop (red) with the active residues glutamic acid (E523) and

aspartic acid (D529) as reported by Lee et al. [31].

Page 28: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

17  

1.4.2.6 H-loop

The H-loop consists of a highly conserved histidine residue that forms a “hinge” or “lynchpin”

between an α-helix and a ß-sheet. Until recently, it was thought this residue acted as a non-reactive

“lynchpin” that only coordinated the aspartate at the end of the D-loop, the glutamate at the

beginning of the D-loop/end of the Walker B motif, the ɣ- phosphate on ATP, and the magnesium

ion through hydrogen bonding. New research suggests however that this residue also acts as a

proton transfer point by acting first as a general acid, and then as a general base during ATP

hydrolysis [47]. In NaAtm1 this residue is H554 (Figure 1-12).

Figure 1-12. Cartoon diagram side view of the NaAtm1 H-loop (magenta) with the highly conserved residue histidine

(H554) as reported by Lee et al. [31].

Page 29: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

18  

1.4.2.7 Q-loop

The Q-loop consists of a highly conserved glutamine found at the junction of the two NBD

subdomains and has the consensus sequence XXQXX where X represents any amino acid. The Q-

loop is the main interface between the NBD and TMD and its conformational changes allow the

conserved glutamine to form an active site when ATP is bound and to disrupt the active site when

ATP is hydrolyzed [29]. In NaAtm1 this sequence starts at residue 440 and ends at 444 (Figure 1-

13). The sequence is VPQDS.

Figure 1-13. Cartoon diagram side view of the NaAtm1 Q-loop (blue) with the highly conserved residue glutamic acid

(Q442) as reported by Lee et al. [31].

Page 30: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

19  

1.5 Accessory Domains

Accessory domains are transporter components that are not ubiquitous to all transporters -

indeed, the protein of interest in this thesis, NaAtm1, contains none of these- but when present,

play an important role in protein function. These domains can be found separate from, attached to,

or integrated in a transporter. These domains have been divided into four categories based on

location and function [48]. They are: extracytoplasmic domains, membrane-embedded domains,

cytosolic regulatory domains, and cytosolic catalytic domains.

1.5.1 Extracytoplasmic domains

These domains are found on the N-terminal (extracellular) end of transporters or in the

periplasm. There are three subgroupings of this domain: large extracytoplasmic domains (ECDs),

the N-terminal domains of ProW and ProU, and the substrate binding proteins (SBPs) found in

prokaryotes [48]. For purposes of simplicity, only substrate binding proteins (SBPs) will be

discussed in this paper.

ABC transporters that import substrates require Substrate Binding Proteins (SBPs) to act

as substrate shepherds. These proteins capture substrate and deliver it to the cytoplasmic side of

the TMDs. SBPs can be tethered to the ABC transporter, anchored to the membrane, or free

floating in the periplasm (only found in gram negative bacteria). These proteins have been shown

to be monomers with one binding site per SBP [49].

SBPs are generally composed of two lobes connected by hinge (although there are three-

lobed SBPs, of which the third lobes function is not known [50]) (Figure 1-14). This hinge allows

the two lobes to enclose a substrate similar to the way a Venus flytrap catches its prey [51]. The

importance of the SBPs will be discussed in the context of transport mechanisms in Section 1.9.

Page 31: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

20  

Figure 1-14. Crystal structure of the substrate binding protein (SBP) BtuF with vitamin B12 bound as reported by Borths

et al. [52]. The two lobes (blue and red) are attached to a hinge (green) that allow them to enclose the substrate vitamin B12.

1.5.2 Membrane-embedded domains

This set of accessory domains consists of the extra transmembrane helices seen in many

transporters that are not part of the “core” TMDs (i.e. helices that do not help form a translocation

pathway). While the function of many of these are unknown, some are proposed to act as anchors

for SBPs [53].

1.5.3 Cytosolic regulatory domains

These accessory domains are found on the cytosolic side of the membrane and are used to

regulate the activity of the transporter as well as transcriptional regulation. The mechanisms of

regulation vary by transporter and type of regulation. These regulations allow for the fine tuning

of nutrient uptake or product/waste expulsion within a cell. An example of this regulatory domain

can be found within the methionine importer of E. coli, MetNI. In this importer, methionine is both

a substrate and inhibitor, preventing the build-up of excess methionine within the cell.

Page 32: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

21  

1.5.4 Cytosolic catalytic domains

In some cases, the accessory cytosolic domain has a catalytic purpose. In the type I

secretion pathway, ABC transporters not only transport polypeptides targeted for export but have

the additional job of removing the leader peptides that denoted export as well. The signal sequence

in this case has a consensus sequence of LSXXELXXIXGG and is cleaved after the two conserved

glycine residues [54].

1.6 ABC transporter classes

The superfamily of ABC transporters is just that – a superfamily – and it can be a challenge

to divide such a large and diverse set of proteins. While all ABC transporters contain a core of four

domains (two NBDs and two TMDs), their cellular function allows for broad categorization. To

start, we can differentiate between importers, exporters, and non-porters. These groupings can be

further subdivided through structural homology, sequence similarity, and functional similarity.

1.7 ABC importer classes and proposed mechanisms

ABC importers are a group of ABC transporters that, as the name suggests, bring needed

materials into a cell and, with very few exceptions, are only found in bacteria and archaea [55].

Substrates include sugars, peptides, proteins, sterols, lipids, and ions. The majority of these

importers require a substrate binding protein (SBP) to function properly (part of the

extracytoplasmic accessory domains as described above). These SBPs led to the discovery of this

family (part of the osmotic-shock sensitive transport systems as described in the introduction). In

this class, the four core domains are only found as either a tetramer (as in the Nik, Mal, and Btu

systems), a fused TMD with two NBDs (Fhu), or two TMDs with a fused NBD (Rbs) (Figure 1-

Page 33: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

22  

15) [48]. Importers can also be subdivided based on their functionality into three subgroups: Type

I importers, Type II importers, and Type III (ECF) importers.

Figure 1-15. Cartoon drawing of the different core arrangements for the importer class of ABC transporter adapted

from Biemans-Oldehinkel et al. [48].

1.7.1 Structure of Type I and II importers

While these two types of importer are very similar in terms of composition and topology,

structural and mechanistic features divide them into two separate groups. Structurally, Type I

importers have a core of 5 transmembrane helices per TMD and commonly contain additional

TMD helices that wrap around its partner TMD (10-16 helices total, see Figure 1-16) [29]. The

two TMDs of the Type I importer face each other and the translocation pathway is found between

them. Type II importers have 10 helices per TMD (20 total), and do not contain any helices that

wrap around the partner TMD. These two TMDs are lined up, side-to-side facing opposite

directions.

Page 34: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

23  

Figure 1-16. Top-down view of the TMDs from a Type I and Type II importer. A) MalF and MalG TMD subunits of the

Type I transporter MalEFGK2 [56]. B) BtuC dimer of the Type II transporter BtuC2D2F [57].

In addition to the different TMD arrangements, Type I importers such as MalFGK2 contain

a substrate binding site within their TMDs (not to be confused with a Substrate Binding

Protein) while Type II importers do not. This is probably due to the types of substrates each

importer type transports as well as potential mechanistic differences between the two. Type I

importers generally only handle small molecules such as ions, small peptides, amino acids, and

mono-/oligosaccharides [58]. It is possible these small molecules have more room to move (and

escape) while in the translocation pathway, and the presence of a substrate binding site ensures

that the transport process is efficient by providing a checkpoint for regulation; if no substrate is

bound to the substrate binding site, ATP hydrolysis will not proceed. In contrast, Type II importers

Page 35: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

24  

handle large complex molecules such as vitamin B12 and heme, as well as inorganic ions [58]

which may have a lesser chance of escaping the translocation pathway.

Both the Type I and Type II importers contain an extracytoplasmic accessory domain: a

substrate binding protein (SBP) to shepherd substrates to their N-terminal openings. These

domains can be soluble within the periplasm (found in gram-negative bacteria), attached to the

cytoplasmic membrane either by a “lipid anchor” (gram-positive bacteria) or a transmembrane

peptide (archaea), or directly fused to one of the TMDs (resulting in two SBPs per importer) [48]

(Figure 1-17). In the glutamine-glutamate transporter system, GlnPQ, two SBPs are fused to a

single TMD resulting in four SBPs per importer [59].

Figure 1-17. Cartoon drawing of the different types of Substrate Binding Proteins (SBPs) reproduced from Biemans-

Oldehinkel et al. [48].

1.7.2 Possible mechanisms of Type I and II importers

While these two importer types contain similar core domains along with an SBP, an

increasing number of studies have demonstrated a diversity of transport mechanisms. The two

prevailing models for ABC importers are described below.

Page 36: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

25  

An impressive body of structural studies on the maltose transport system (MalFGK-E) and

vitamin B12 system (BtuCD-F) has provided evidence in support of an alternating access model for

transport [60 and 61]. In this model there are four general steps in the catalytic cycle, however

slight variations have been shown between different transporters (Figure 1-18).

Figure 1-18. The alternating access model for transport. This process starts at the center top and moves clockwise.

The cycle starts when a ligand-loaded substrate binding protein binds to the TMDs of the

transporter that is in a “resting state” (referred to as the “inward facing conformation”) with the

energy source ATP absent from the ATP active sites (Figure 1-18, top). The binding of the ligand-

loaded SBP brings the NBDs closer which promotes ATP binding. In the second step, the binding

Page 37: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

26  

of ATP triggers coordinated conformational changes in the TMDs, SBP and NBDs. In the TMDs,

the conformational change results in opening of the translocation pathway to the periplasm (the

“outward facing conformation”). In the SBP, the conformational change of the TMDs result in the

release of the ligand into the translocation pathway, where the ligand binds to the substrate binding

site. In the NBDs, the ATP-filled active sites come close enough for hydrolysis to occur. In the

third step ATP hydrolysis occurs; the hydrolyzed ATP (now ADP and Pi) cause the transporter to

return to the resting state (inward conformation) which results in the opening of the NBDs, release

of the ligand into the cytosol, and the release of the now empty SBP. The fourth step consists of

the release of ADP from the active sites to reset the cycle.

Page 38: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

27  

Figure 1-19. A second model for transport, based on the vitamin B12 importer, BtuCD-F [58].

A second “non-canonical” mechanism for transport is based on functional assays and, most

recently, mass spectrometry [58, 62, 63]. In this model, the cycle starts with the transporter in a

resting state (partially outward conformation) with no ATP bound and the substrate binding protein

firmly bound (Figure 1-19, top). The binding of one ATP molecule increases the affinity of a

second ATP binding and the presence of two ATP molecules partially destabilize the SBP by fully

opening the TMD (outward conformation). This destabilization allows for the ligand to fit in

between the SBP and the transporter translocation pathway. Although the TMDs do not have a

Page 39: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

28  

substrate binding site, the partial insertion of the ligand into the translocation pathway results in

the NBDs coming close enough for hydrolysis to occur. The hydrolysis results in the opening of

the NBDs (inward conformation), release of the substrate, and the re-binding of the SBP [58]. The

now-hydrolyzed ATP (ADP + Pi) are then released and the cycle is reset.

Interestingly, both of these models have been developed based on work from the same

transport systems, mainly MalFGK-E and BtuCD-F. Support for the alternating access model is

driven by the high-resolution crystal structures of several intermediate states throughout the

transport cycle. The non-canonical model, in contrast, is based on functional studies including

thermodynamic, kinetic, and mass spectrometry experiments. The techniques and experimental

conditions for these studies are fundamentally different, and their conflicting conclusions remain

to be resolved.

1.7.3 Structure of Type III (ECF) importers

If the Type I and II importers are considered fraternal twins, the Type III importer can be

considered the weird younger brother who may or may not be the milkman’s. This class of

transporter is also known as an energy coupling factor (ECF) type ABC transporter. While the

Type I and II importers have a pair of transmembrane domains that form a translocation pathway

for the substrate, Type III importer architecture is significantly different. Type III importers consist

of two distinct integral membrane proteins: (1) a substrate-binding component (or S-component)

with high substrate specificity and (2) a transmembrane component (or T-component) that

transfers the energy from the NBDs [64]. The three-component complex consisting of the T-

component and two NBDs is known as the ECF module. The Type III importer can itself be broken

Page 40: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

29  

into two subgroups based on chromosomal location of the genes that encode these two

components: ECF group I and ECF group II [65].

1.7.3.1 Differences between ECF subgroups I and II

Members of the ECF group I importers have genes that code for the S-component and ECF

module on the same operon which are thought to form a dedicated complex (i.e. a complex that

does not dissociate). In contrast, members of the ECF group II have genes for the ECF module

separate from multiple S-components that have different substrate specificities. These S-

components are thought to form interchangeable complexes with a single ECF module. This

modular system is an interesting departure from the Type I and Type II import systems and

potentially streamlines the number of protein components needed for uptake of different

substrates.

1.7.3.2 Type III substrates

The Type III importers have been shown to have high specificity for vitamins or vitamin

precursors including pantothenate, pyridoxine, thiamine and its precursors, methionine, cobalamin

precursors, queuosine, lipoate, niacin and folate [65]. In total, the S-components have been

classified into 21 protein families, which are associated with either ECF group I or group II T-

components (Figure 1-20). Some S-components form complexes with both dedicated and shared

ECF groups.

Page 41: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

30  

Figure 1-20. A list of known dedicated (green) and shared (pink and tan) Type III ECF transporters (bold) and their

substrates [65].

1.7.4 Type III transport mechanism

What is truly unique about this transporter is the movement of the S-component within the

lipid bilayer. Crystal structures have shown that while solitary S-components have substrate

binding sites close to the extracellular side of the membrane, the S-component can lie parallel to

the lipid bilayer with its binding site exposed to the cytosol when in complex with the ECF module.

This suggests that the re-orienting of the S-component plays a role in substrate transport (Figure

1-21) [65].

Page 42: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

31  

Figure 1-21. A proposed mechanism for the ECF-type transporter [61].

This mechanism varies slightly between Group I and Group II proteins only in that the S-

component is interchangeable in group II. The first step in the Group II mechanism is the joining

of a ligand-bound S-component and a free ECF module into a full complex.

1.8 ABC exporters

ABC exporters are found in all lifeforms yet studied. They are located on the cellular

membrane of both prokaryotic and eukaryotic species as well as in organellar membranes of

eukaryotic species including the ER, inner mitochondrial membrane, peroxisomal membrane, and

vacuolar membrane [48]. ABC exporters do not require a SBP to function as is the case with the

Type I and II importers. While many importers are found a tetramers or trimers, ABC exporters

are almost exclusively found as “half transporters” (a pair of polypeptides that each contain one

TMD and one NBD) or “full-length transporters” (a single polypeptide containing two TMDs and

two NBDs).

As previously stated, ABC importers are, with very few exceptions, found only in

prokaryotic life. This means that eukaryotes only possess ABC exporters and therefore the exporter

group of ABC transporters is usually referred to as the eukaryotic group of ABC transporters, even

though members of this group are also found in prokaryotes. This group has been broken into

Page 43: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

32  

seven families based on domain organization and primary structure similarity [66]. These families

are ABCA, ABCB, ABCC, ABCD, ABCE, ABCF, and ABCG. For simplicity, the descriptions

below focus on the human members of these families.

1.8.1 ABCA Transporters

The ABCA family was first described in 1994 by Luciani et al. [67] and consists of twelve

known members in humans. This group of transporters has been shown to be involved in lipid

transport and some members are hypothesized to be involved in pulmonary surfactant secretion

and macrophage lipid homeostasis.

1.8.2 ABCB Transporters

The ABCB family consists of eleven known members, seven of which are “half

transporters” and is the only family that contains both “half transporters” and “full transporters”.

This family contains the most famous of human ABC transporters – ABCB1. This transporter,

more commonly known as P-gp, acts as a localized general drug exporter and confers multi-drug

resistance (MDR). Some members of this family participate in the secretion of

phosphatidylcholine, cholesterol, and bile salts, while others are found in the mitochondria and are

involved in iron/sulfur synthesis and transport [66].

1.8.3 ABCC Transporters

This family contains twelve known human “full transporters” which are involved in drug

resistance, ion transport, nucleoside transport, and signal transduction. The most highly-studied

transporter in this family is the ABCC7 transporter, also known as CFTR. This modified

Page 44: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

33  

transporter acts as a chloride ion channel and results in cystic fibrosis when deformed or absent.

While there are many mutations that cause the CFTR protein to fail, the majority of those suffering

from cystic fibrosis have a mutation that causes misfolding in the protein that keeps it from

localizing to the cell surface.

1.8.4 ABCD Transporters

This family contains four genes that encode “half transporters”, all of which are expressed

only in the peroxisome. Recent studies suggest that these subunits can mix and match and might

be linked to fatty acid metabolism. It has been shown that ABCD1 plays a role in the X-linked

form of adrenoleukodystrophy (ALD) which is characterized by cells with an excess of unbranched

saturated fatty acids [66].

1.8.5 ABCE Transporters

There is currently only one human member of this family: ABCE1. It is considered a non-

porter due to the absence of a transmembrane domain. It is made up of two NBDs and an N-

terminal Fe-S binding domain. This protein inhibits a nuclease induced by interferon in

mammalian cells. It also has been shown to play a role in vertebrate translation initiation. ABCE1

is found in all characterized eukaryotes and archaea [68].

1.8.6 ABCF Transporters

There are three known human genes that encode 26 different known proteins for this

family. Like the ABCE transporters, these proteins lack any TMDs and are considered non-porters.

Many of the ABCF transporters are thought to play a role in inflammatory processes [69].

Page 45: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

34  

1.8.7 ABCG Transporters

This transporter class is unique in that it contains five genes that encode six reverse “half

transporters”. These subunits contain a NBD at the N-terminus and a TMD and the C-terminus—

the reverse of all other ABC genes. This family was first reported in Drosophila White (white-

eyed mutant fruit fly) and is known as the “White Half Transporters” [70]. In Drosophila White,

this first protein (white) found in white-eyed males can combine with two other proteins (brown

and scarlet) to form heterodimers which transport guanine and tryptophan respectively to the eye

cells and result in the eye color. While these proteins play a role in eye pigment in flies, in humans

they seem mainly involved in the transport of dietary sterols and cholesterol transport regulation

[71].

Page 46: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

35  

1.8.8 Proposed mechanism of exporters

The proposed general mechanism for exporters is similar to that of the canonical alternating

access model (Figure 1-22). In the case of exporters, the resting state is in an inward conformation.

The substrate binds at a substrate binding site in the TMDs, followed by the binding of two ATP,

or vice versa. The binding of substrate and ATP triggers a conformational change in the TMDs to

the outward conformation. The substrate is released to the exterior of the cell, and ATP is

hydrolyzed. The hydrolysis of the ATP resets the transporter back to the resting state [72].

Figure 1-22. Drawing of a proposed exporter mechanism [72].

Page 47: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

36  

1.9 ABC non-porters

Some proteins that contain both the Walker A and B motifs, as well as the ABC signature

motif are known to carry out other functions than transport. These types of “transporters” usually

do not contain a transmembrane domain such as the members of the eukaryotic groups ABCF and

ABCG. Many of these proteins such as Rad50 play a role in regulation and DNA repair [73]. Due

to the non-transporter-like nature of these proteins, they will not be discussed in depth.

1.10 Protein of interest: an ABC exporter found in mitochondria

This thesis focuses on a member of the ABCB exporter family found in mitochondrial

membranes. The first homolog of this exporter was discovered in the inner membrane of

mitochondria of Saccharomyces cerevisiae and was named Atm1p (ABC transporter of

mitochondria 1) [74]. Initial studies of Atm1p showed that the disruption of this transporter in vivo

led to very slow growth on rich media and no growth on minimal media. Further experiments with

Atm1p showed that the mitochondria lacked mature cytochromes (i.e. cytochromes that contained

a covalently bonded heme group) and that heme-containing proteins outside the mitochondria were

also affected in the same way. After ruling out other factors such as impaired heme attachment

reaction pathways and biosynthetic pathways, it was determined that Atm1p plays a critical role

in the maturation of heme-containing proteins within the cell [75].

Researchers then looked for what substrates might be transported by Atm1p and found that

in mitochondria that lacked functioning Atm1p, their “free” iron (iron not bound to heme or Fe-S

clusters) was 30 times greater than that of wild-type mitochondria. Of this, only a small fraction

was in a state that could be readily used for the synthesis of heme and Fe-S clusters. This suggested

that Atm1p mediates iron homeostasis, and that iron—or more likely an iron containing complex—

Page 48: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

37  

might be a natural substrate [75]. While the identity of Atm1p’s natural substrate remains unclear,

likely candidates include Fe-S complexes and heme (or some intermediate). The creation of these

cytosolic heme and iron sulfur cluster containing proteins are critical in many cellular processes

such as DNA synthesis and respiratory electron transfer chains. Due to this role in the maturation

of iron-containing proteins that are essential to cellular function, this is the only known eukaryotic

ABC transporter that is essential for cell growth [75].

Atm1p homologs have since been characterized in several other species, including ABCB6

and ABCB7 in humans and HMT1 in multiple species. In humans, the ABCB7 transporter plays

a role in transition metal homeostasis and iron-sulfur cluster maturation in the cytosol [76] while

the ABCB6 transporter may play a role in arsenic resistance [77]. When ABCB7 is damaged, it

results in X-linked sideroblastic anemia which is associated with spinocerebellar ataxia

accompanied by severe cerebellar hypoplasia [78]. In both C. elegans and D. melanogaster, HMT1

confers resistance to cadmium [79, 80].

In many of these homologs, the polypeptide glutathione (GSH, γ-Glu-Cys-Gly) has been

shown to play a role in transport of substrates (Figure 1-23) [81], although the mechanism of this

dependence is unclear. Glutathione is a key antioxidant and helps confer resistance to heavy metals

and xenobiotics [82, 83] so it is plausible that GSH acts as a co-substrate for Atm1p homologs,

with an affinity for a wide array of toxic materials as well as the ability to stimulate these

transporters.

Figure 1-23. Molecular structure of GSH.

Page 49: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

38  

1.10.1 Atm1 in Novosphingobium aromaticivorans

The focus of this work is on the Atm1p homolog

found in Novosphingobium aromaticivorans

(Novosphingobium aromaticivorans ABC transporter of

mitochondria 1). This species was first classified by

Takeuchi et al. in 2001 [84] and is an aerobic gram-negative

bacterium that can break down phenolic structures. It has

also been implicated in biliary cirrhosis in humans due to its

ability to express proteins similar to human proteins that

trigger T cell-mediated tissue destruction [86].

NaAtm1 was structurally and functionally characterized by Lee et al. in 2014. The high-

resolution crystal structure of NaAtm1 shows a homodimer “half transporter” in the inward

conformation (Figure 1-25). It has a total of 6 TMD helices per half and is roughly 120Å x 85Å in

size. Initial studies have shown that NaAtm1 shares ~45% sequence identity to the homologs

ABCB7, HMT1, and yeast Atm1 [31]. Five different crystal structures for this protein have been

reported (4MRN, 4MRP, 4MRR, 4MRV, 4MRS) with each crystal structure containing a different

bound ligand (Apo, Reduced glutathione (GSH), Selenomethionine, S-Mercury glutathione, and

Oxidized glutathione (GSSG) respectively) [31].

Figure 1-24. Microscopic image of N. Aromaticivorans [82].

 

Page 50: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

39  

Figure 1-25. Dimension of NaAtm1 along with locations of the primary (red) and secondary (black) substrate binding sites as reported by Lee et al. [31].

1.10.2 NaAtm1 substrate binding sites

Crystal structures have revealed two substrate binding sites (SBSs), with the main SBS

located about 5Å into the transmembrane region. The second lower affinity site is located about

5Å below the primary binding site. The primary binding site is formed by Asp316, Gly319, and

Met320 from TMD6/TMD12 and Asn296 and Gln272 from TMD5/TMD11, while the secondary

binding site is comprised of Arg206, 210 and 323.

Figure 1-26. A) Primary binding site bound with the substrate S-Hg(GSH)2 B) Secondary binding site bound with the

substrate GSSG, as reported by Lee et al. [31].

Page 51: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

40  

The relevance of these two sites is not yet fully understood; however, the relative

orientation could accommodate a glutathione complexed iron-sulfur cluster ([2Fe:2S](GS-)4)

(Figure 1-27) complex [31].

Figure 1-27. Structure of a glutathione complexed iron-sulfur cluster ([2Fe:2S](GS-)4). This may be a possible

substrate that can interact with both binding pockets found within NaAtm1.

1.10.3 In vitro characterization of NaAtm1 activity

In addition to structural studies, preliminary kinetic assays revealed that NaAtm1 has a

high affinity for the transition metals silver and mercury in complex with glutathione (Table 1-2).

Substrates that do not include glutathione did not stimulate activity in this transporter, consistent

with studies using homologs of NaAtm1 [78]. Interestingly, while it is accepted that GSH is a

critical component involved in transport, reduced glutathione (GSH) stimulated relatively low

transporter activity compared to oxidized glutathione (GSSG). This suggests that the natural

substrate of NaAtm1 contains more than one GSH, as shown in the binding of S-Hg(GSH)2 (Figure

1-26A).

Page 52: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

41  

Table 1-2. Kinetic constants for the ATPase activity of substrates in NaAtm1 as reported by Lee et al. [31]

To demonstrate that substrate was transported across a lipid bilayer, proteoliposomes

containing reconstituted NaAtm1 was assayed. In the presence of wild-type NaAtm1

proteoliposomes and ATP, increasing amounts of GSSG accumulated within the proteoliposomes

as a function of time. These results indicated that the increased ATPase activity of NaAtm1

correlated with transport of substrate across a lipid membrane.

1.10.4 In vivo functionality of NaAtm1

To test if NaAtm1 exports the substrates silver and mercury in vivo, metal sensitive strains

of E. coli were grown and subjected to metal salts (CuSO4, ZnSO4, CdCl2, AgNO3, and HgCl2).

No protection was observed against Cu2+, Zn2+, or Cd2+, however protection was conferred by

NaAtm1 for Ag2+ and Hg2+. These results suggest that Ag2+ and Hg2+ are capable of being exported

by NaAtm1 and therefore prolong the survival of an organism in the presence of these toxic

metals.

Page 53: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

42  

1.11 Studies presented in this work

This thesis describes the development and optimization of an NaAtm1 lipid bilayer system

designed for in vitro studies of transport. In Chapter 2, the general principles behind protein

purification are described, followed by procedures and results for three different proteins. Chapter

3 introduces the different lipid mimetic environments used in the study of ABC transporters. This

chapter details the experiments and findings for the reconstitution of NaAtm1 into a lipid lipid-

filled structure referred to as a nanodisc. In Chapter 4, data from preliminary kinetic studies are

presented, and future directions are discussed.

Page 54: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

43  

1.12 References

[3] Neu, H. C.; Heppel, L. A. The Release of Enzymes from Escherichia coli by Osmotic Shock and during the Formation of Spheroplasts. Journal of Biological Chemistry 1965, 240(1), 3685–3692.

[4] Bialecka-Fornal, M. Single-Cell Analysis of the Physiology of Mechanosensation in Bacteria, Ph. D. Dissertation, California Institute of Technology, Pasadena, CA, 2013.

[5] Boos, W. The Galactose Binding Protein and its Relationship to the β‐Methylgalactoside Permease from Escherichia coli. European Journal of Biochemistry 1969, 10(1), 66–73.

[6] Curtis, S. J. Mechanism of Energy Coupling For Transport of D-Ribose in Escherichia coli. Molecular and Cell Biology 1974, 120(1), 295–303.

[7] Berger, E.A.; Heppel, L.A. Different mechanisms of energy coupling in the shock-sensitive and shock-resistant amino acid permeases of Escherichia coli. Journal of Biological Chemistry 1974, 249(24), 7747–7755.

[8] Gilson E.; Higgins C. F.; Hofnung M.; Ames G. F.; Nikaido H. Extensive Homology between Membrane-associated Components of Histidine and Maltose Transport Systems of Salmonella typhimurium and Escherichia coli. Journal of Biological Chemistry 1982, 257(17), 9915–9918.

[9] Higgins, C. F.; Haag, P. D.; Nikaido, K.; Ardeshir, F.; Garcia, G.; Ames, G. F-L. Complete nucleotide sequence and identification of membrane components of the histidine transport operon of S. typhimurium. Nature 1982, 298(5876), 723–727.

[10] Juliano, R.; Ling, V. A surface glycoprotein modulating drug permeability in Chinese hamster ovary cell mutants. Biochimica et Biophysica Acta (BBA) – Biomembranes 1976, 455(1), 152–162.

[11] Riordan, J. R.; Deuchars, K.; Kartner, N.; Alon, N.; Trent, J.; Ling, V. Amplification of P-glycoprotein genes in multidrug-resistant mammalian cell lines. Nature 1985, 316(6031), 817–819.

Page 55: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

44  

[12] Walker, J. E.; Saraste, M.; Runswick, M. J.; Gay, N. J. Distantly related sequences in the α-and ß-subunits of ATP synthase, myosin, kinases and other ATP-requiring enzymes and a common nucleotide binding fold. The EMBO Journal 1982, 1(8), 945–951.

[13] Higgins, C. F.; Hiles, I. D.; Salmond, G. P. C.; Gill, D. R.; Downie, J. A.; Evans, I. J.; Holland, I. B.; Gray, L.; Buckel, S. D.; Bell, A. W.; Hermodson, M. A. A family of related ATP-binding subunits coupled to many distinct biological processes in bacteria. Nature 1986, 323(6087), 448–450.

[14] Gros, P.; Croop, J.; Housman, D. Mammalian multidrug resistance gene: Complete cDNA sequence indicates strong homology to bacterial transport proteins. Cell 1986, 47(3), 371–380.

[15] Gerlach, J. H.; Endicott, J. A.; Juranka, P. F.; Henderson, G.; Sarangi, F.; Deuchars, K. L.; Ling, V. Homology between P-glycoprotein and a bacterial haemolysin transport protein suggests a model for multidrug resistance. Nature 1986, 324(6096), 485–489.

[16] Chen, C.-J.; Chin, J. E.; Ueda, K.; Clark, D. P.; Pastan, I.; Gottesman, M. M.; Roninson, I. B. Internal duplication and homology with bacterial transport proteins in the mdr1 (P-glycoprotein) gene from multidrug-resistant human cells. Cell 1986, 47(3), 381–389.

[17] Theodoulou, F. L.; Kerr, I. D. ABC transporter research: going strong 40 years on. Biochem. Soc. Trans. 2015, 43(5), 1033–1040.

[18] Aller, S. G.; Yu, J.; Ward, A.; Weng, Y.; Chittaboina, S.; Zhuo, R.; Harrell, P. M.; Trinh, Y. T.; Zhang Q.; Urbatsch, I. L.; Chang, G. Structure of P-Glycoprotein Reveals a Molecular Basis for Poly-Specific Drug Binding. Science 2009, 323(5922), 1718–1722.

[19] Konings, W. N.; Poelarends, G. J. Bacterial multidrug resistance mediated by a homologue of the human multidrug transporter P-glycoprotein. IUBMB Life (International Union of Biochemistry and Molecular Biology: Life) 2002, 53(4-5), 213–218.

[20] Barabote, R. D.; Thekkiniath, J.; Strauss, R. E.; Vediyappan, G.; Fralick, J. A.; Francisco, M. J. S. XENOBIOTIC EFFLUX IN BACTERIA AND FUNGI: A GENOMICS UPDATE. Advances in Enzymology and Related Areas of Molecular Biology 2011, 237–306.

Page 56: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

45  

[21] Ueda, K.; Clark, D. P.; Chen, C.; Roninson, I. B.; Gottesman, M. M.; Pastan, I. The Human Multidrug Resistance (mdrl) Gene. Journal of Biological Chemistry 1987, 262(2), 505–508.

[22] Fair, R. J.; Tor, Y. Antibiotics and Bacterial Resistance in the 21st Century Perspectives in Medicinal Chemistry 2014, 6.

[23] Stefkova, J.; Poledne, R.; Hubacek, J. A. ATP-Binding Cassette (ABC) Transporters in Human Metabolism and Diseases Physiol 2004, 53(1), 235–243.

[24] Cystic fibrosis. https://www.mayoclinic.org/diseases-conditions/cystic-fibrosis/symptoms-causes/syc-20353700 (accessed Jul 14, 2019).

[25] Davies, J.; Burney; Alton, E. Gene therapy for the treatment of cystic fibrosis. The Application of Clinical Genetics 2012, 29.

[26] Davidson, A. L.; Dassa, E.; Orelle, C.; Chen, J. Structure, function, and evolution of bacterial ATP-binding cassette systems. Microbiology and Molecular Biology Reviews 2008, 72(2), 317–364.

[27] Holland, I.; Blight, M. A. ABC-ATPases, adaptable energy generators fuelling transmembrane movement of a variety of molecules in organisms from bacteria to humans. Journal of Molecular Biology 1999, 293(2), 381–399.

[28] Komoda, T. The HDL handbook: biological functions and clinical implications; Elsevier, Academic Press: London, 2017.

[29] Dawson, R. J. P.; Hollenstein, K.; Locher, K. P. Uptake or extrusion: crystal structures of full ABC transporters suggest a common mechanism. Molecular Microbiology 2007, 65(2), 250–257.

[30] Wilkens, S. Structure and mechanism of ABC transporters. F1000Prime Reports 2015, 7. [31] Beek, J. T.; Guskov, A.; Slotboom, D. J. Structural diversity of ABC transporters. The

Journal of General Physiology 2014, 143(4), 419–435.

Page 57: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

46  

[32] Vetter, I. R.; Wittinghofer, A. Nucleoside triphosphate-binding proteins: different scaffolds to achieve phosphoryl transfer. Quarterly Reviews of Biophysics 1999, 32(1), 1–56.

[33] Lee, J. Y.; Yang, J. G.; Zhitnitsky, D.; Lewinson, O.; Rees, D. C. Structural basis for heavy metal detoxification by an Atm1-type ABC exporter. Science 2014, 343(6175), 1133–1136.

[34] Patzlaff, J. S.; Heide, T. V. D.; Poolman, B. The ATP/substrate stoichiometry of the ATP-binding cassette (ABC) transporter OpuA. Journal of Biological Chemistry 2003, 278(32), 29546–29551.

[35] Procko, E.; Ferrin-Oconnell, I.; Ng, S.-L.; Gaudet, R. Distinct structural and functional properties of the ATPase sites in an asymmetric ABC transporter. Molecular Cell 2006, 24(1), 51–62.

[36] Nikaido, K.; Ames, G. F.-L. One intact ATP-binding subunit is sufficient to support ATP hydrolysis and translocation in an ABC transporter, the histidine permease. Journal of Biological Chemistry 1999, 274(38), 26727–26735.

[37] Deyrup, A. T.; Krishnan, S.; Cockburn, B. N.; Schwartz, N. B. Deletion and site-directed mutagenesis of the ATP-binding motif (P-loop) in the bifunctional murine ATP-sulfurylase/adenosine 5'-phosphosulfate kinase enzyme. Journal of Biological Chemistry 1998, 273(16), 9450–9456.

[38] Satishchandran, C.; Hickman, Y. N.; Markham, G. D. Characterization of the phosphorylated enzyme intermediate formed in the adenosine 5'-phosphosulfate kinase reaction. Biochemistry 1992, 31(47), 11684–11688.

[39] Thomas P. M.; Wohllk N.; Huang E.; Kuhnle U.; Rabl W.; Gagel R. F.; Cote G. J.

Inactivation of the first nucleotide-binding fold of the sulfonylurea receptor, and familial persistent hyperinsulinemic hypoglycemia of infancy. American Journal of Human Genetics 1996, 59(3), 510–518.

Page 58: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

47  

[40] Driscoll, W. J.; Komatsu, K.; Strott, C. A. Proposed active site domain in estrogen sulfotransferase as determined by mutational analysis. Proceedings of the National Academy of Sciences 1995, 92(26), 12328–12332.

[41] Delepelaire P. PrtD, the Integral Membrane ATP-binding Cassette Component of the Erwinia chrysanthemi Metalloprotease Secretion System, Exhibits a Secretion Signal-regulated ATPase Activity. Journal of Biological Chemistry 1994, 269(45), 27952–27957.

[42] Smith, C.; Rayment, I. Active site comparisons highlight structural similarities between myosin and other P-loop proteins. Biophysical Journal 1996, 70(4), 1590–1602.

[43] Geourjon, C.; Orelle, C.; Steinfels, E.; Blanchet, C.; Deléage, G.; Pietro, A. D.; Jault, J.-M. A common mechanism for ATP hydrolysis in ABC transporter and helicase superfamilies. Trends in Biochemical Sciences 2001, 26(9), 539–544.

[44] Ambudkar, S. V.; Kim, I.-W.; Xia, D.; Sauna, Z. E. The A-loop, a novel conserved aromatic acid subdomain upstream of the Walker A motif in ABC transporters, is critical for ATP binding. FEBS Letters 2005, 580(4), 1049–1055.

[45] Nikaido, H. How are the ABC transporters energized? Proceedings of the National Academy of Sciences 2002, 99(15), 9609–9610.

[46] Jones, P. M.; George, A. M. Role of the D-loops in allosteric control of ATP hydrolysis in an ABC transporter. The Journal of Physical Chemistry 2012, 116(11), 3004–3013.

[47] Hopfner, K. P.; Karcher, A.; Shin, D. S.; Craig, L.; Arthur, L. M.; Carney, J. P.; Tainer, J. A. Structural biology of Rad50 ATPase: ATP-driven conformational control in DNA double-strand break repair and the ABC-ATPase superfamily. Cell 2000, 101(7), 789– 800.

[48] Zaitseva, J.; Oswald, C.; Jumpertz, T.; Jenewein, S.; Wiedenmann, A.; Holland, I. B.; Schmitt, L. A structural analysis of asymmetry required for catalytic activity of an ABC-ATPase domain dimer. The EMBO Journal 2006, 25(14), 3432–3443.

[49] Zhou, Y.; Ojeda-May, P.; Pu, J. H-loop histidine catalyzes ATP hydrolysis in the E. coli ABC-transporter HlyB. Physical Chemistry Chemical Physics 2013, 15(38), 15811.

Page 59: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

48  

[50] Biemans-Oldehinkel, E.; Doeven, M. K.; Poolman, B. ABC transporter architecture and regulatory roles of accessory domains. FEBS Letters 2005, 580(4), 1023–1035.

[51] Quiocho, F. A.; Ledvina, P. S. Atomic structure and specificity of bacterial periplasmic receptors for active transport and chemotaxis: variation of common themes. Molecular Microbiology 1996, 20(1), 17–25.

[52] Tame; Murshudov, G.; Dodson, E.; Neil, T.; Dodson, G.; Higgins, C.; Wilkinson, A. The structural basis of sequence-independent peptide binding by OppA protein. Science 1994, 264(5165), 1578–1581.

[53] Mao, B.; Pear, M.R.; McCammon, J.A.; Quiocho, F.A. Hinge-bending in L-arabinose-binding protein. The "Venus's-flytrap" model. Journal of Biological Chemistry 1982, 257(3), 1131–1133.

[54] Borths, E. L.; Locher, K. P.; Lee, A. T.; Rees, D. C. The Structure of Escherichia Coli BtuF and Binding to Its Cognate ATP Binding Cassette Transporter. Proceedings of the National Academy of Sciences 2002, 99(26), 16642–16647.

[55] Boyd, D.; Manoil, C.; Beckwith, J. Determinants of membrane protein topology. Proceedings of the National Academy of Sciences 1987, 84(23), 8525–8529.

[56] Havarstein, L. S.; Holo, H.; Nes, I. F. The leader peptide of colicin V shares consensus sequences with leader peptides that are common among peptide bacteriocins produced by gram-positive bacteria. Microbiology 1994, 140(9), 2383–2389.

[57] El-Awady, R.; Saleh, E.; Hashim, A.; Soliman, N.; Dallah, A.; Elrasheed, A.; Elakraa, G. The Role of Eukaryotic and Prokaryotic ABC Transporter Family in Failure of Chemotherapy. Frontiers in Pharmacology 2017, 7.

[58] Licht, A.; Bommer, M.; Werther, T.; Neumann, K.; Hobe, C.; Schneider, E. Structural and Functional Characterization of a Maltose/Maltodextrin ABC Transporter Comprising a Single Solute Binding Domain (MalE) Fused to the Transmembrane Subunit MalF. Research in Microbiology 2019, 170(1), 1–12.

Page 60: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

49  

[59] Hvorup, R.; Goetz, B.; Niederer, M.; Hollenstein, K.; Perozo, E.; Locher, K. Asymmetry in the Structure of the ABC Transporter-Binding Protein Complex BtuCD-BtuF. Science 2007, 1387–1390.

[60] Fiorentino, F.; Bolla, J. R.; Mehmood, S.; Robinson, C. V. The Different Effects of Substrates and Nucleotides on the Complex Formation of ABC Transporters. Structure 2019, 27(4).

[61] Schuurman-Wolters, G. K.; Poolman, B. Substrate Specificity and Ionic Regulation of GlnPQ from Lactococcus lactis AN ATP-BINDING CASSETTE TRANSPORTER WITH FOUR EXTRACYTOPLASMIC SUBSTRATE-BINDING DOMAINS. Journal of Biological Chemistry 2005, 280(25), 23785–23790.

[62] Chen, J. Molecular Mechanism of the Escherichia Coli Maltose Transporter. Current Opinion in Structural Biology2013, 23(4), 492–498.

[63] Locher, K. Structure and Mechanism of ATP-Binding Cassette Transporters. Philosophical Transactions of the Royal Society B Biological Sciences 2008, 364(1514), 239–245.

[64] Bao, H.; Duong, F. Discovery of an Auto-Regulation Mechanism for the Maltose ABC Transporter MalFGK2. PLoS ONE 2012, 7(4).

[65] Lewinson, O. A Distinct Mechanism for the ABC Transporter BtuCD-BtuF Revealed by the Dynamics of Complex Formation. Nature Structural & Molecular Biology 2010, 17(3), 332–338.

[66] Slotboom, D. J. Structural and mechanistic insights into prokaryotic energy-coupling factor transporters. Nature Reviews Microbiology 2013, 12(2), 79–87.

[67] Rodionov, D.A.; Hebbeln, P.; Eudes, A.; ter Beek, J.; Rodionova, I.A.; Erkens, G.B.; Slotboom, D.J.; Gelfand, M.S.; Osterman, A. L.; Hanson, A. D.; Eitinger, T. A Novel Class of Modular Transporters for Vitamins in Prokaryotes. Journal of Bacteriology 2008, 191(1), 42–51.

[68] Dean, M. The human ATP-binding cassette (ABC) transporter superfamily. Genome Research 2001, 11(7), 1156–1166.

[69] Luciani, M. F.; Denizot, F.; Savary, S.; Mattei, M. G.; Chimini, G. Cloning of two novel ABC transporters mapping on human chromosome 9. Genomics 1994, 21(1), 150–159.

Page 61: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

50  

[70] Chen, Z.-Q.; Dong, J.; Ishimura, A.; Daar, I.; Hinnebusch, A. G.; Dean, M. The Essential Vertebrate ABCE1 Protein Interacts with Eukaryotic Initiation Factors. Journal of Biological Chemistry 2006, 281(11), 7452–7457.

[71] Vasiliou, V.; Vasiliou, K.; Nebert, D. W. Human ATP-binding cassette (ABC) transporter family. Human Genomics 2008, 3(3), 281.

[72] Morgan, T. H. SEX LIMITED INHERITANCE IN DROSOPHILA. Science 1910, 32(812), 120–122.

[73] Klucken, J.; Buchler, C.; Orso, E.; Kaminski, W. E.; Porsch-Ozcurumez, M.; Liebisch, G.; Kapinsky, M.; Diederich, W.; Drobnik, W.; Dean, M.; Allikmets, R.; Schmitz, G. ABCG1 (ABC8), the human homolog of the Drosophila white gene, is a regulator of macrophage cholesterol and phospholipid transport. Proceedings of the National Academy of Sciences 2000, 97(2), 817–822.

[74] Shintre, C. A.; Pike, A. C. W.; Li, Q.; Kim, J.; Barr, A. J.; Goubin, S.; Shrestha, L.; Yang, J.; Berridge, G.; Ross, J.; Stansfield, P. J.; Sansom, M. S. P.; Edwards, A. M.; Bountra, C.; Marsden, B. D.; Delft, F. V.; Bullock, A. N.; Gileadi, O.; Burgess-Brown, N. A.; Carpenter, E. P. Structures of ABCB10, a human ATP-binding cassette transporter in apo- and nucleotide-bound states. Proceedings of the National Academy of Sciences 2013, 110(24), 9710–9715.

[75] European Bioinformatics InstituteProtein Information ResourceSIB Swiss Institute of Bioinformatics. DNA repair protein RAD50. https://www.uniprot.org/uniprot/Q92878 (accessed May 18, 2019).

[76] Leighton, J.; Schatz, G. An ABC Transporter in the Mitochondrial Inner Membrane Is Required for Normal Growth of Yeast. The EMBO Journal 1995, 14(1), 188–195.

[77] Kispal, G.; Csere, P.; Guiard, B.; Lill, R. The ABC Transporter Atm1p Is Required for Mitochondrial Iron Homeostasis. FEBS Letters 1997, 418(3), 346–350.

[78] Lill, R.; Kispal, G. Abc Transporters In Mitochondria. ABC Proteins 2003, 515–531.

[79] Chavan, H.; Oruganti, M.; Krishnamurthy, P. The ATP-Binding Cassette Transporter ABCB6 Is Induced by Arsenic and Protects against Arsenic Cytotoxicity. Toxicological Sciences 2011, 120(2), 519–528.

Page 62: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

51  

[80] Pondarre, C.; Campagna, D. R.; Antiochos, B.; Sikorski, L.; Mulhern, H.; Fleming, M. D. Abcb7, The Gene Responsible for X-Linked Sideroblastic Anemia with Ataxia, Is Essential for Hematopoiesis. Blood 2007, 109(8), 3567–3569.

[81] Vatamaniuk, O. K.; Bucher, E. A.; Sundaram, M. V.; Rea, P. A. CeHMT-1, a Putative Phytochelatin Transporter, Is Required for Cadmium Tolerance In Caenorhabditis Elegans. Journal of Biological Chemistry 2005, 280(25), 23684–23690.

[82] Sooksa-Nguan, T.; Yakubov, B.; Kozlovskyy, V. I.; Barkume, C. M.; Howe, K. J.; Thannhauser, T. W.; Rutzke, M. A.; Hart, J. J.; Kochian, L. V.; Rea, P. A.; Vatamaniuk, O. K. DrosophilaABC Transporter, DmHMT-1, Confers Tolerance to Cadmium. Journal of Biological Chemistry 2008, 284(1), 354–362.

[83] Kuhnke, G.; Neumann, K.; Mühlenhoff, U.; Lill, R. Stimulation of the ATPase Activity of the Yeast Mitochondrial ABC Transporter Atm1p by Thiol Compounds. Molecular Membrane Biology 2006, 23(2), 173–184.

[84] Masip, L.; Veeravalli, K.; Georgiou, G. The Many Faces of Glutathione in Bacteria. Antioxidants & Redox Signaling 2006, 8(5-6), 753–762.

[85] Wang, L.; Ouyang, B.; Li, Y.; Feng, Y.; Jacquot, J.-P.; Rouhier, N.; Xia, B. Glutathione Regulates the Transfer of Iron-Sulfur Cluster from Monothiol and Dithiol Glutaredoxins to Apo Ferredoxin. Protein & Cell 2012, 3(9), 714–721.

[86] Takeuchi, M.; Hamana, K.; Hiraishi, A. Proposal of the Genus Sphingomonas Sensu Stricto and Three New Genera, Sphingobium, Novosphingobium and Sphingopyxis, on the Basis of Phylogenetic and Chemotaxonomic Analyses. International Journal of Systematic and Evolutionary Microbiology 2001, 51(4), 1405–1417.

[87] https://genome.jgi.doe.gov/portal/novar/novar.home.html (accessed Jul 14, 2019).

[88] Kaplan, M. M. Novosphingobium Aromaticivorans: A Potential Initiator of Primary Biliary Cirrhosis. The American Journal of Gastroenterology 2004, 99(11), 2147–2149.

Page 63: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

52  

CHAPTER 2

Protein Purification

When studying protein function from a biochemical perspective, an essential first step is

to isolate the protein of interest. The resulting protein must be reproducibly purified to a single

species with high yield. During my time at USF I have developed and improved multiple protein

purification methods for the purpose of purifying both single proteins (NaAtm1, MSP1D1, and

TEV protease), as well as protein complexes (NaAtm1 reconstituted into nanodiscs). This chapter

outlines the theories behind the primary techniques routinely employed during purification.

Additionally, key specialized instruments for protein purification are described. Lastly, the

methods used in the expression and purification of NaAtm1 and the accessory proteins MSP1D1

and TEV protease are detailed. The purpose of the accessory proteins MSP1D1 and TEV protease

will be described in Chapter 3.

2.1 Method theories

While the process of protein purification employs a myriad of materials and techniques,

this section will focus on two chromatography-based techniques used for preparative purposes and

one technique employed for analysis. The two chromatographic methods are immobilized metal

affinity chromatography (IMAC) and size exclusion chromatography (SEC). Both types of

chromatography are performed using an FPLC (Fast Protein Liquid Chromatography) instrument

which is described later in this chapter. These chromatography methods are paired with an analysis

method called sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE) which is

used to assess the purity of a protein sample.

Page 64: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

53  

Chromatography: “A process in which a chemical mixture carried by a liquid or gas is separated

into components as a result of differential distribution of the solutes as they

flow around or over a stationary liquid or solid phase” [1].

2.1.1 Immobilized metal affinity chromatography (IMAC)

IMAC is a method of chromatography that uses a metal (typically nickel or copper) that

has been immobilized to bind proteins of interest. IMAC was first developed by Porath et al. in

1975 as a novel way to separate proteins [2]. Originally, iminodiacetic acid (IDA) was used as a

metal immobilizer, however due to its weak binding of metals, a more effective immobilizer was

developed by Hochui et al. in 1987 [3] (Figure 2-1). This new immobilizer – Nitrilotriacetic acid

(NTA) – had the ability to bind to four of the six binding sites found in the coordination sphere of

a nickel ion rather than the three sites bound by IDA [4]. Commercial IMAC columns use

nitrilotriacetic acid-bound nickel, referred to as Ni-NTA columns.

Figure 2-1. Comparison of the NTA and IDA immobilizers interacting with nickel ions [4]

Page 65: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

54  

2.1.1.1 Poly-histidine tags

To selectively interact with the nickel ion described above, proteins of interest are

genetically modified to add a “His-tag”. His-tags were first described by Hochuli et al. in 1988 [5]

and are comprised of 6-10 histidine residues that have been placed at either the N- or C-terminus

of the protein of interest. For many proteins, protein folding and function are not significantly

affected by the addition of a His-tag to one of the termini. Multiple histidine residues are introduced

because the imidazole group on the side chain exhibits the highest affinity for metal ions of all the

amino acids [6] (Figure 2-2). Purification using a His-tagged protein and Ni-NTA can increase

target protein purity from less than 1% to up to more than 95% purity [4].

Figure 2-2. A nickel-loaded Nitrilotriacetic acid (Ni-NTA) (blue) bound to resin (blue sphere) binding to part of a poly-histidine tag (green) [7].

Once bound, the resin can be washed with the low concentrations of imidazole (10-50 mM)

to remove any impurities, followed by elution from the column at higher concentrations (300-500

mM). IMAC is routinely chosen as a primary purification method because it is robust, relatively

inexpensive, and can be used with a wide variety of buffer conditions.

Page 66: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

55  

2.1.2 Size exclusion chromatography (SEC)

Size exclusion chromatography is based on the use of a porous matrix to separate a

mixture of particles by their molecular weight. In this porous matrix, smaller components of a

mixture take longer to travel through the column than larger components. This is because the

smaller components are able to fit into a larger percentage of pores and are temporarily impeded.

This idea was first postulated by Synge and Tiselius [8] based on zeolite research by R. M. Barrer

[9]. The first columns were packed with potato or maize starch, which have a very low adsorption

of proteins [10]; however, they were not strong enough to withstand high linear velocities. Today,

the most commonly used packing materials (referred to as size exclusion media) are small porous

spheres made up of either cross-linked composite dextran, agarose, or polyacrylamide (Figure 2-

3) [11].

One way to understand this somewhat counter-intuitive technique is to imagine the size

exclusion column and its resin as a large hollow cylinder filled with whiffle balls. If a mixture of golf

balls and marbles were placed onto the top of this column, the marbles would be able to enter the

whiffle balls while the golf balls could not – they would have to go around the whiffle balls. In this

scenario the marbles have a longer path length through the column due to continually entering

and exiting the interior of the whiffle balls. This results in the golf balls (the larger component of

the mixture) to come out of the bottom of the cylinder first. Likewise, in size exclusion

chromatography, the largest protein components are eluted first.

One important feature of a SEC column is its void volume. The void volume of a column

is the volume at which an unretained molecule elutes. To extend the analogy above, the golf balls,

which do not get stuck in any whiffle balls, would elute at the void volume. Protein molecules

eluting at the void volume are larger than what the column can resolve and are most commonly

very large protein aggregates.

Page 67: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

56  

Figure 2-3. A) Sample of a size exclusion bead interacting with three different sized molecules. B) A size exclusion

column at different times (column 1 = T0, 2 = T1, 3 = T2). The colors correspond to the different components of the mixture. C) A sample chromatogram showing the elution order of largest to smallest [12].

While IMAC can yield a relatively pure sample enriched with the His-tagged protein, the

elution may also contain aggregates and other impurities. Size exclusion chromatography can

remove these impurities if the sizes are within the resolution ability of the resin.

2.1.3 Sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE)

Electrophoresis: “The movement of suspended particles through a medium (such as paper or

gel) under the action of an electromotive force applied to electrodes in contact

with the suspension” [13].

SDS-PAGE is a method of electrophoresis created by U.K. Laemmli in 1970 [14] that

denatures proteins using sodium dodecyl sulfate (SDS) and then separates them based on

Page 68: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

57  

molecular mass by passing them through a polyacrylamide gel [15]. Polyacrylamide is used

because it is chemically inert and produces different pore sizes at different concentrations.

Figure 2-4. Components and composition of a typical SDS-PAGE experiment [16].

Proteins are first prepared for SDS-PAGE by denaturation using SDS, dithiothreitol (DTT)

and heat. SDS denatures both the secondary and non-disulfide-linked tertiary structures, while

DTT is a powerful reducing agent that breaks disulfide bonds [17]. Proteins must be denatured in

order to ensure movement based on molecular mass and not on protein dimension. During this

process, all proteins are turned into “rods” with a diameter of roughly one amino acid so that only

their lengths determine how quickly they pass through the gel. In addition to denaturing the protein,

SDS gives proteins a uniform net negative charge which is essential for unidirectional movement

through the polyacrylamide gel. SDS binds amino acids with a general ratio of one SDS molecule

to two amino acids. This “mass-to-charge” ratio is roughly the same for all molecules and allows

electrical force to be applied to all molecules in a relative way.

Page 69: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

58  

The gel plate is comprised of polyacrylamide gel thinly sandwiched between glass or

plastic plates. The gel is made up of two parts called the “stacking gel” and “running gel” (Figure

2-4, right). The stacking gel is located at the upper portion of the gel and contains the sample wells.

It is comprised of a lower concentration of polyacrylamide to produce a large pore size mixed with

Tris-HCl buffer that has a lower pH (pH=6.8) than both the Tris-glycine running buffer (pH=8.3)

and the running gel (pH=8.8) (Figure 2-4, left) This difference in pH is essential in producing a

thin, uniform band of protein that is required for the running gel.

When a charge is applied to the system, negatively charged glycine ions in the running

buffer are forced into the stacking gel and encounter its lower pH environment. At pH=6.8, glycine

is a zwitterionic molecule (neutrally charged) which results in very slow mobility through the

stacking gel. The stacking gel also contains chloride ions (from the Tris-HCl) which are negatively

charged and run much faster through the gel than the neutral glycine. As the chloride ions move

towards the anode, an area of positive counter ions is created which results in a large voltage

gradient between the chloride ions and glycine. All of the proteins are within this gradient and are

“stacked” based on their electrophoretic mobility into thin layers at the boundary between the

stacking gel and running gel [18].

The running gel starts at the bottom of the stacking gel and continues to the bottom of the

gel plate. It contains a higher concentration of polyacrylamide to produce a smaller pore size and

has a higher pH than that of the stacking gel. Upon reaching the running gel, the stacked protein

sample encounters the dramatic change in pore size and pH. The increase in pH (to pH=8.8) results

in glycine returning to a negative charge and moving much faster, leaving the sample protein to

make its way through the gel alone. The smaller pore size causes the proteins in the sample to

separate based on their mass to charge ratio [19].

Page 70: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

59  

Figure 2-5. Mechanics of the stacking gel and loading gel. The stacking gel (A) is located above the running gel and has a pH of 6.8. At this pH, glycine molecules contain no net charge and move much more slowly than the chloride ions, creating a voltage gradient. Proteins (red lines) in this gel are pulled along behind the fast moving chloride ions and “stacked” at the boundary between the stacking gel and running gel. The running gel has a pH of 8.8 which causes the glycine to have a negative charge and results in both glycine and the chloride ions leaving the protein to move on its own. In this gel, the speed of the protein is inversely related to its size, with small proteins moving faster than large proteins [18].

Ultimately, SDS-PAGE separates the different proteins within a sample by molecular

weight, and in combination with various staining methods, allows for detection of these molecules

(Figure 2-6). This technique is commonly used to assess the results of each step in a purification

process. SDS-PAGE analysis can identify the enrichment or loss of a target protein, allowing for

optimization and troubleshooting of the method. Finally, the purity and yield of a final protein

sample can be visually assessed.

Page 71: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

60  

Figure 2-6. SDS-PAGE gel comprised of different samples taken during the purification of NaAtm1. Samples are more fully described in Section 2.3.6.

2.2 Specialized instruments

This section will focus on two specialized instruments used during the purification process:

a fast protein liquid chromatography (FPLC) instrument and a microvolume spectrophotometer.

Page 72: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

61  

2.2.1 Fast Protein Liquid Chromatography (FPLC) instrument

The FPLC instrument is essential

for purifying proteins. The AKTA Pure

(GE Healthcare) employed in this study

has several features including two

independent pumps, a column valve, an

injector port, and a fractionator for

collecting samples. It records system

pressure, pre-column pressure, delta

column pressure, absorbance at 280 nm,

and conductivity. The specific commercially available columns used in this work are as follows:

1. Ni-NTA HisTrap column:

HisTrap columns contain Ni-NTA resin (as described in the IMAC section above) and are

used to purify His-tagged proteins. The columns utilized here were either 1 mL or 5 mL in

volume, with a reported capacity of approximately 40 mg His-tagged per mL of resin.

2. HiPrep 26/10 Desalting:

This column is used for group separation of proteins from much smaller molecules. It is

most commonly used for buffer exchange after a particular chromatographic step. The

“HiPrep” name reflects the nature of its use; the maximum sample volume of 10 mL allows

for buffer exchange of larger scale protein preparations.

3. HiLoad 16/600 Superdex 200 pg

 

Figure 2‐7. ÄKTA Pure FPLC made by 

General Electric [20]. 

Page 73: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

62  

This size exclusion column is ideal for protein separation in the 10,000 - 600,000 Da range.

It has the ability to handle up to 5mL of sample per run and is commonly used as a

secondary purification method in large scale preparations, employed to remove

contaminants of significantly differing size from that of the protein of interest.

4. Superdex 200 Increase 10/300 GL:

This size exclusion column is ideal for protein separation in the 10,000 - 600,000 Da range

however its maximum sample volume is much smaller, only 0.5 mL and thus is used for

smaller scale preparations.

2.2.2 Spectrophotometer

A microvolume spectrophotometer was

used to determine the concentrations of all purified

proteins by measuring the proteins absorbance at

280 nm. An accurate concentration can be obtained

by using the proteins extinction coefficient (a

number derived using the total amount of amino

acids that absorb at 280 nm found within the target

protein). The nanophotometer is used because It can accurately measure absorption using 2uL of

protein, allowing for the conservation of the small amount of final purified protein that is produced

during one purification run (around 1 mL total).

 

Figure 2-8. ImplenTM N50 Nanophotometer [21]. 

Page 74: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

63  

2.3 Purification of NaAtm1

2.3.1 NaAtm1 purification overview

To isolate the ABC exporter NaAtm1 for

biochemical studies, an over-expression and

purification method was employed based on a

previously published procedure [23]. This

protocol was modified as detailed below to yield

higher purity and increased yield. Using this

optimized protocol, ~10 mg of purified NaAtm1

per 30 g of cells was routinely obtained.

2.3.2 NaAtm1 protein expression

The DNA plasmid containing full-length NaAtm1 was a gift of Douglas Rees (California

Institute of Technology). The gene was encoded in a pJL-H6 ligation-independent cloning vector,

which added a 6x histidine tag on the carboxy terminus of the protein. The plasmid was

transformed into BL21-Gold (DE3) E. coli (Aglient) using standard heat shock procedures.

Transformed E. coli were selected for on agar plates containing 100 µg/mL ampicillin. Single

colonies were then used to inoculate 5 mL of LB media containing 100 µg/mL ampicillin. Cultures

were incubated at 37 °C while shaking at 225 rpm for six hours. After six hours, each culture was

transferred to a Fernbach flask containing 900 mL of ZYM-5052 auto-induction media (1% (w/v)

tryptone, 0.5% (w/v) yeast extract, 0.4% (v/v) glycerol, 0.2% (w/v) lactose, and 0.05% (w/v)

dextrose), 100 mL of 10x phosphate buffer (0.17 M KH2PO4, 0.72 M K2HPO4), and 100 µg/mL

ampicillin. Four Fernbach flasks were incubated at 37 °C while being shaken at 180 rpm overnight.

The next morning the cell cultures were centrifuged at 18,370 xg for 20 minutes to pellet the cells.

Figure 2-9. Structure of NaAtm1 [22].

Page 75: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

64  

Approximately 30 g of cells were routinely obtained from 4 L of culture. Harvested cells were

either immediately used in the purification process or flash frozen using liquid nitrogen and stored

at -80 °C until needed.

2.3.3 NaAtm1 protein purification

NaAtm1 was purified using an ÄTKA Pure FPLC System (GE Healthcare) via

immobilized metal affinity chromatography followed by size exclusion chromatography. All

samples, lysates, and eluents were kept on ice or at 4 °C unless otherwise noted. 30 g of cells were

homogenized in 300 mL of NaAtm1 buffer (20 mM Tris-HCl, pH 7.5, 100 mM NaCl, and 0.05%

n-Dodecyl β-D-maltoside (DDM, Anatrace)). 10 mg Deoxyribonuclease I (Sigma), 10 mg

lysozyme (Sigma), 1.5 mL of 200 mM PMSF, and 3 g of DDM (1% w/v) were added to the cells

and was stirred for 20 minutes. The suspension was sonicated on ice using a flat tip sonicator with

20% bursts of 50 kHz for 45 seconds followed by 30 seconds of cooling for 7 cycles for a total of

5.25 minutes of sonication. The lysate was then centrifuged at 28,300 xg for 20 minutes. The

clarified supernatant was collected and 4 M imidazole (pH 8) was added to a final concentration

of 25 mM. The lysate was then injected onto a 5 mL Ni-NTA HisTrap HP column (GE Healthcare)

and washed with 75 mL of NaAtm1 buffer containing 35 mM imidazole. The column was eluted

with 20 mL of NaAtm1 buffer containing a final concentration of 350 mM imidazole. Peak

fractions were collected and loaded onto a HiPrep 26/10 Desalting column (GE Healthcare)

equilibrated in NaAtm1 buffer to remove imidazole. Peak fractions were collected, pooled

(typically 10-12 mL total) and frozen in liquid nitrogen and stored at -80 ˚C. The next day, the

eluate was thawed, filtered with a 0.2 µm filter, and concentrated to 5 mL using an Amicon Ultra-

15 centricon with a 100,000 kD molecular weight cutoff (Millipore) and injected onto a Superdex

200 pg 16/600 sizing column (GE Healthcare). Peak fractions (eluting at approximately 60 mL

Page 76: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

65  

elution volume) were collected, pooled, and concentrated as before until the final protein

concentration was approximately 15 mg/mL (extinction coefficient 152,640 M-1cm-1). Final

protein was flash frozen in liquid nitrogen and stored at -80 °C until needed.

2.3.4 NaAtm1 IMAC purification results

Immobilized metal affinity chromatography purification was carried out using both the

High Performance (HP) and Fast Flow (FF) versions of the 5 mL HisTrap (Ni-NTA) columns

produced by GE Healthcare. Results from these purifications showed that the HP HisTrap was

able to capture roughly twice the amount of NaAtm1 as the FF HisTrap (Figure 2-10). NaAtm1

purification consisted of four phases: loading, washing, eluting and desalting. During this

purification process, imidazole is used to prevent non-specific binding during loading, remove

contaminants during washing, and elute the target protein during elution. During the loading

phase, 4 M imidazole was added to roughly 300 mL of clarified supernatant containing NaAtm1,

bringing the total imidazole concentration to 25 mM. The supernatant was flowed over a HisTrap

column. This phase always produced a steady UV reading of around 2300 mAU, representing the

abundance of proteins in the cellular lysate. Once the loading phase was complete, a washing phase

using 35 mM imidazole and lasting 15 column volumes (CV) took place. This phase resulted in a

rapid decline in the UV reading from ~2300 mAU down to roughly 20 mAU. This decrease

indicates that the vast majority of non-specific-bound protein had been removed from the column

and suggests that only NaAtm1 is left bound to the column. Once washed, the protein was removed

from the column during the elution phase using 350 mM imidazole. This phase produced a sharp

peak of between 600 mAU and 2000 mAU followed by a steady UV reading of about 100 mAU

due to the high imidazole concentration (Figure 2-10).

Page 77: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

66  

Figure 2-10. Comparison of the elution peaks for WT NaAtm1 using 5 mL HisTrap FF (blue) and 5 mL HisTrap HP

(orange) Ni-NTA columns. The HisTrap HP consistently produced elution peaks more than twice as large as the HisTrap FF given the same amount of starting material.

The elution peak was immediately flowed over a HiPrep 26/10 Desalting column to

separate NaAtm1 from the imidazole used to elute the protein from the HisTrap column. Imidazole

needs to be removed immediately as it can interfere with enzyme activity and may destabilize the

protein. This process always produced a bell-curve shape starting around 10 mL and ending around

20 mL (Figure 2-11).

 

0

500

1000

1500

2000

2500

3000

145 165 185 205 225 245

Absorban

ce 280nm (mAU)

Volume (mL)

HisTrap FF vs. HP

HisTrap FF

HisTrap HP

Page 78: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

67  

Figure 2-11. Imidazole separation from NaAtm1 using a Size exclusion “Desalting” column. This

chromatogram shows only the peak of NaAtm1.  

2.3.5 NaAtm1 SEC purification results 

Following elution from the desalting column, a secondary purification process using size

exclusion chromatography was performed. This process uses a 120 mL HiLoad 16/600 Superdex

200 pg size exclusion column to remove both impurities and aggregates of NaAtm1.

Chromatograms for this process usually display two peaks at 50 mL and 64 mL (Figure 2-12). The

50 mL peak, representing the void volume of the column, contained aggregated NaAtm1 (data not

shown), while the 64 mL peak is consistent with the size of monodisperse NaAtm1 homodimers.

0

100

200

300

400

500

600

700

800

900

1000

5 10 15 20 25 30 35

Absorban

ce 280nm (mAU)

Volume (mL)

Page 79: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

68  

Figure 2-12. Size exclusion chromatogram of secondary purification of wild-type NaAtm1.

2.3.6 Analysis of NaAtm1

Samples were collected throughout the purification process and analyzed using SDS-

PAGE. As seen in samples of the lysate, pellet, and supernatant, NaAtm1 was successfully

overexpressed in cells (Figure 2-13, red arrow); however, as NaAtm1 is a membrane protein, the

majority of expressed protein remained in the pellet, despite the presence of high concentrations

of detergent (Figure 2-13, green arrow). The decrease of the NaAtm1 in the flow through indicated

that majority of NaAtm1 selectively bound to the Ni-NTA resin, and the wash sample contained

mostly non-specific proteins. While the eluate for Ni-NTA and desalt columns contained multiple

proteins, the NaAtm1 band was considerably enriched relative to all other bands. Finally, the

elution sample taken following SEC showed the final purified product with minimal

contamination.

 

0

50

100

150

200

250

300

350

0 20 40 60 80 100 120

Absorban

ce 280nm (mAU)

Volume (mL)

Page 80: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

69  

 

Figure 2-13. SDS-PAGE analysis of NaAtm1.  

Lane Name Description1 Ladder Molecular weight ladder (Bio-rad)2 Lysate Cellular lysate post-sonification.3 Pellet Insoluble cellular components, taken immediately after centrifugation

of cell lysate.4 Supernatant Soluble proteins, taken immediately after centrifugation of cell lysate.5 Flow Through Proteins that did not bind to HisTrap column 6 Wash Low imidazole wash of His-Trap column7 Fractions Pooled peak fractions from high imidazole elution 8 Desalt Pooled peak fractions from desalting column9 Final (1 µL) 1 µL of the pooled and concentrated fractions from size exclusion

column 10 Final (1 µg) 1 µg of the pooled and concentrated fractions from size exclusion

column

Page 81: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

70  

2.4 Purification of MSP1D1

MSP1D1 is an accessory protein

used in the assembly of nanodiscs, a

lipidic environment developed for the

study of integral membrane proteins

(Figure 2-14, orange and yellow). A full

description of its history, structure, and

implementation can be found in Chapter

3: Creating a Realistic Environment In

Vitro.

2.4.1 MSP1D1 purification overview

The Sligar Group at the University of Illinois Urbana Champaign originally cloned

MSP1D1 and developed methods for over-expression and purification. This protocol, however, is

extremely reagent and labor intensive (requiring seven different buffers), and thus an alternative

method was developed, as described here. This new MSP1D1 purification strategy was based on

NaAtm1, and showed similar yield and purity as the Sligar method. Using this more efficient and

streamlined protocol, ~50 mg of purified MSP1D1 per 15 g of cells was routinely obtained.

2.4.2 MSP1D1 protein expression

The DNA plasmid containing full-length MSP1D1 was a gift of Douglas Rees (California

Institute of Technology). The gene was encoded in a pJL-H6 ligation-independent cloning vector

which added a 6x histidine tag on the carboxy-terminus of the protein. The plasmid was

transformed into BL21-Gold (DE3) E. coli (Aglient) using standard heat shock procedures.

 Figure 2-14. Structure of an empty Nano-disc

containing two MSP1D1 proteins (yellow and orange) surrounding a lipid bilayer (gray) [24].

Page 82: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

71  

Transformed E. coli were selected for on agar plates containing 50 µg/mL kanamycin. Single

colonies were then used to inoculate 5 mL of LB media containing 50 µg/mL kanamycin. Cultures

were incubated at 37 °C while shaking at 225 rpm overnight. The next morning, each culture was

transferred to a fernbach flask containing 900 mL of Terrific Broth (TB) media (2.0% (w/v)

tryptone, 2.4% (w/v) yeast extract, and 0.4% (v/v) glycerol), 100 mL of 10x phosphate buffer

(0.017 M KH2PO4, 0.072 M K2HPO4), and 50 µg/mL kanamycin. Four Fernbach flasks were

incubated at 37 °C while being shaken at 180 rpm until the optical density of all flasks were ~0.5.

Once all flasks reached ~0.5 OD, the lactose analog IPTG was added to a final concentration of 1

mM to induce protein overexpression. The cultures were grown for an additional three hours

before being harvested via centrifugation at 18,370 xg for 20 minutes. Approximately 15 g of cells

were routinely obtained from 4 L of culture. Harvested cells were either immediately used in the

purification process or flash frozen using liquid nitrogen and stored at -80 °C until needed.

2.4.3 MSP1D1 protein purification

MSP1D1 was purified using an ÄTKA Pure FPLC System (GE Healthcare) via

immobilized metal affinity chromatography followed by size exclusion chromatography. All

samples, lysates, and eluents were kept on ice or at 4 °C unless otherwise noted. ~15 g of cells

were homogenized in 150 mL of MSP1D1 buffer (40 mM Tris-HCl, pH 7.5 and 300 mM NaCl).

10 mg Deoxyribonuclease I (Sigma), 10 mg lysozyme (Sigma), and 1.5 mL of 200 mM PMSF

were added to the cells and was stirred for 10 minutes. The suspension was sonicated on ice using

a flat tip sonicator with 20% bursts of 50 kHz for 45 seconds followed by 30 seconds of cooling

for 7 cycles for a total of 5.25 minutes of sonication. The lysate was then centrifuged at 28,300 xg

for 20 minutes. The clarified supernatant was collected and 4 M imidazole (pH 8) was added to a

final concentration of 20 mM. The lysate was then injected onto a 5 mL Ni-NTA HisTrap HP

Page 83: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

72  

column (GE Healthcare) and washed with 75 mL of MSP1D1 buffer containing 35 mM imidazole.

The column was eluted with 20 mL of MSP1D1 buffer containing a final concentration of 350 mM

imidazole. Peak fractions were collected and loaded onto a HiPrep 26/10 Desalting column (GE

Healthcare) equilibrated in MSP1D1 buffer to remove imidazole. Peak fractions were collected,

pooled (typically 10-12 mL total) and frozen in liquid nitrogen and stored at -80 ˚C. The next day,

the eluate was thawed, filtered with a 0.2 µm filter, and concentrated to 5 mL using an Amicon

Ultra-15 centricon with a 10,000 kD molecular weight cutoff (Millipore) and injected onto a

Superdex 200 pg 16/600 sizing column (GE Healthcare). Peak fractions (eluting at approximately

75 mL elution volume) were collected, pooled, and concentrated as before until the final protein

concentration was approximately 15 mg/mL (extinction coefficient 21,430 M-1cm-1). Final

protein was flash frozen in liquid nitrogen and stored at -80 °C until needed.

 

2.4.4 MSP1D1 IMAC purification results

Immobilized metal affinity chromatography purification was carried out using a High

Performance (HP) 5 mL HisTrap (Ni-NTA) columns produced by GE Healthcare. MSP1D1

purification consisted of four phases: loading, washing, eluting and desalting. The resulting

profiles are similar to those seen in the NaAtm1 purification procedure. During the loading phase,

roughly 150 mL of clarified supernatant containing overexpressed MSP1D1 was flowed over a

HisTrap column. This phase always produced a steady UV reading of around 2600 mAU,

representing the abundance of proteins in the cellular lysate. Once the loading phase was complete,

a washing phase lasting 15 column volumes (CV) took place. This phase resulted in a rapid decline

in the UV reading from ~2600 mAU down to roughly 20 mAU. This step was necessary to remove

any non-specific proteins bound to the column. Once washed, the protein was removed from the

column during the elution phase. This phase produced a peak of around 2000 mAU followed by a

Page 84: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

73  

UV reading of about 100 mAU due to the high imidazole concentration (Figure 2-15). The height

of the absorbance peak indicated a high concentration of protein was present in the eluate.

Figure 2-15. Primary purification of MSP1D1 using a 5 mL HisTrap HP column.

Once eluted, the peak fractions were immediately flowed over a HiPrep 26/10 Desalting

column to separate MSP1D1 from the high concentration of imidazole used in Ni-NTA elution.

This process always produced a bell-curve shape starting around 10 mL and ending around 25 mL

(Figure 2-16).

0

500

1000

1500

2000

2500

3000

0 50 100 150 200 250 300

Absorban

ce 280nm (mAU)

Volume (mL)

Page 85: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

74  

Figure 2-16. Imidazole separation from MSP1D1 using a size exclusion-type “desalting” column.  

2.4.5 MSP1D1 SEC purification results

Following elution from the desalting column, a secondary purification process using size

exclusion chromatography was performed. This process employed a 120 mL HiLoad 16/600

Superdex 200 pg size exclusion column to remove any remaining impurities and aggregates of

MSP1D1. Chromatograms for this process routinely displayed two peaks at 50 mL and 75 mL

(Figure 2-17). The protein eluting at the 50 mL void volume was comprised of aggregated

MSP1D1 (data not shown), while the peak at 75 ml consisted of purified MSP1D1. A faint shoulder

was normally seen after the main peak eluted. These fractions were not collected and not part of

the purified MSP1D1.

0

200

400

600

800

1000

1200

1400

1600

1800

0 5 10 15 20 25 30 35

Absorban

ce 280nm (mAU)

Volume (mL)

Page 86: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

75  

Figure 2-17. Size exclusion chromatogram of secondary purification of MSP1D1

2.4.6 Analysis of MSP1D1

Samples were collected throughout the purification process and analyzed using SDS-

PAGE. The purification of MSP1D1 can be followed using these samples (Figure 2-18, red arrow).

The absence of the MSP1D1 in the flow through indicated that MSP1D1 selectively bound to the

Ni-NTA resin, and the wash sample contained only non-specific proteins. While the eluate for Ni-

NTA and desalt columns contained multiple proteins, the MSP1D1 band was considerably

enriched relative to all other bands. Finally, the elution sample taken following SEC showed the

final purified product with minimal contamination.

0

100

200

300

400

500

600

700

800

900

1000

0 20 40 60 80 100 120 140 160

Absorban

ce 280nm (mAU)

Volume (mL)

Page 87: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

76  

Lane Name Description1 Ladder Molecular weight ladder (Bio-rad)2 Lysate Cellular lysate post-sonification.3 Pellet Insoluble cellular components, taken immediately after

centrifugation of cell lysate.4 Supernatant Soluble proteins, taken immediately after centrifugation of cell

lysate.5 Flow Through Proteins that did not bind to HisTrap column 6 Wash Low imidazole wash of His-Trap column7 Fractions Pooled peak fractions from high imidazole elution 8 Desalt Pooled peak fractions from desalting column9 Final (1 µL) 1 µL of the pooled and concentrated fractions from size exclusion

column 10 Final (1 µg) 1 µg of the pooled and concentrated fractions from size exclusion

column

 

Figure 2-18. SDS-PAGE of MSP1D1  

 

Page 88: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

77  

2.5 Purification of TEV protease

TEV Protease is a highly specific protease that

is used to remove engineered His-tags from proteins

produced in a heterologous expression system. This

protein and its application are described in detail in

Chapter 3: Creating a Realistic Environment in Vitro.

2.5.1 TEV protease purification overview

The TEV protease purification strategy follows a procedure developed by the Narlikar

Group at the University of California, San Francisco. Using a modified version of this protocol,

~10 mg of purified TEV protease per 15 g of cells was obtained.

2.5.2 TEV protease protein expression

The DNA plasmid containing full-length S219V mutant TEV protease was a gift from

Geeta Narlikar. The gene was encoded in a pMal-C2 ligation-independent cloning vector which

added a 6x histidine tag on the carboxy terminus of the protein. The plasmid was transformed into

BL21-Gold (DE3) E. coli (Aglient) using standard heat shock procedures. Transformed E. coli

were selected for on agar plates containing 100 µg/mL ampicillin. Single colonies were then used

to inoculate 5 mL of LB media containing 100 µg/mL ampicillin. Cultures were incubated at 37

°C while shaking at 225 rpm overnight. The next morning, each culture was transferred to a

Fernbach flask containing 1 L of 2x Luria Broth (LB) media (2% (w/v) tryptone, 1% (w/v) yeast

extract, and 2% (v/v) NaCl) and 100 µg/mL ampicillin. Four Fernbach flasks were incubated at 37

°C while being shaken at 180 rpm until the optical density of all flasks were ~0.5. Once all flasks

 

Figure 2-19. Cartoon model of TEV protease [25].

Page 89: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

78  

reached ~0.5 OD, the lactose analog IPTG was added to a final concentration of 0.4 mM to induce

protein production. The flasks were then incubated at 18 °C while shaking at 180 rpm overnight.

The next morning, the cell cultures were centrifuged at 18,730 xg for 20 minutes to pellet the cells.

Approximately 15 g of cells were obtained from 4 L of culture. Harvested cells were either

immediately used in the purification process or flash frozen using liquid nitrogen and stored at -

80 °C until needed.

2.5.3 TEV protease protein purification

TEV protease was purified using an ÄTKA Pure FPLC System (GE Healthcare) via

immobilized metal affinity chromatography followed by size exclusion chromatography. All

samples, lysates, and eluents were kept on ice or at 4 °C unless otherwise noted. ~15 g of cells

were homogenized in 150 mL of TEV protease lysis buffer (1x PBS, 300 mM KCl, 10% glycerol,

and 7.5 mM Imidazole pH 7.5). 10 mg Deoxyribonuclease I (Sigma) and 10 mg lysozyme (Sigma)

were added to the cells and was stirred for 10 minutes. The suspension was sonicated on ice using

a flat tip sonicator with 20% bursts of 50 kHz for 45 seconds followed by 30 seconds of cooling

for 7 cycles for a total of 5.25 minutes of sonication. The lysate was then centrifuged at 28,300 xg

for 20 minutes. The clarified supernatant was collected and 4 M imidazole (pH 8) was added to a

final concentration of 20 mM. The lysate was then injected onto a 5 mL Ni-NTA HisTrap HP

column (GE Healthcare) and washed with 75 mL of TEV protease lysis buffer containing 35 mM

imidazole. The column was eluted with 20 mL of TEV protease elution buffer (25 mM HEPES,

100 mM KCl, 10% glycerol, and 500 mM Imidazole pH 7.5) containing a final concentration of

350 mM imidazole. Peak fractions were collected and loaded onto a HiPrep 26/10 Desalting

column (GE Healthcare) equilibrated in TEV protease storage buffer (25 mM HEPES, 300 mM

Page 90: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

79  

KCl, 10% glycerol, and 2 mM DTT pH 7.5) to remove imidazole. Peak fractions were collected,

pooled (typically 10-12 mL total), frozen in liquid nitrogen and stored at -80 ˚C. The next day, the

eluate was thawed, filtered with a 0.2 µm filter, and concentrated to 5 mL using an Amicon Ultra-

15 centricon with a 10,000 kD molecular weight cutoff (Millipore) and injected onto a Superdex

200 pg 16/600 sizing column (GE Healthcare). Peak fractions (eluting at approximately 75 mL

elution volume) were collected, pooled, and concentrated as before until the final protein

concentration was approximately 15 mg/mL (extinction coefficient 36,130 M-1cm-1). Final

protein was flash frozen in liquid nitrogen and stored at -80 °C until needed.

 

2.5.4 TEV protease IMAC purification results

Immobilized metal affinity chromatography purification was carried out using a High

Performance (HP) 5 mL HisTrap (Ni-NTA) columns produced by GE Healthcare. TEV protease

purification consisted of four phases: loading, washing, eluting and desalting. During the loading

phase, roughly 150 mL of clarified supernatant containing TEV protease was flowed over a

HisTrap column. This phase produced a steady UV reading of around 2500 mAU. Once the loading

phase was complete, a washing phase lasting 15 column volumes (CV) took place. This phase

resulted in a rapid decline in the UV reading down to roughly 20 mAU. After washing, the TEV

protein was removed from the column during the elution phase. This phase produced a broad peak

of around 900 mAU.

Page 91: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

80  

Figure 2-20. Primary purification of TEV using a 5 mL HisTrap HP column.

Once eluted, the fractions containing the elution peak were immediately flowed over a

HiPrep 26/10Desalting column to remove imidazole from the TEV protease This process always

produced a bell-curve shape starting around 10 mL and ending around 25 mL (Figure 2-21).

 Figure 2-21. Imidazole separation from TEV protease using a Size exclusion “Desalting” column.

0

500

1000

1500

2000

2500

3000

0 50 100 150 200 250

Absorban

ce 280nm (mAU)

Volume (mL)

0

100

200

300

400

500

600

0 5 10 15 20 25 30 35

Absorban

ce 280nm (mAU)

Volume (mL)

Page 92: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

81  

2.5.5 TEV protease SEC purification results 

Following elution from the desalting column, a secondary purification process using size

exclusion chromatography was performed. This process uses a 120 mL HiLoad 16/600 Superdex

200 pg size exclusion column to remove both impurities and aggregates of TEV protease. The

chromatogram for this process displayed multiple peaks with the largest eluting at 90 mL (Figure

2-22). For this purification, only the largest peak was evaluated; no analysis was done on the

smaller peaks.

 Figure 2-22. Size exclusion chromatogram of secondary purification of TEV protease.

 

2.5.6 Analysis of TEV Protease

Samples were collected throughout the purification process and analyzed using SDS-

PAGE. As seen in samples of the lysate, pellet, and supernatant, TEV protease was successfully

overexpressed in cells (Figure 2-23, red arrow). The absence of the TEV protease in the flow

through indicated that TEV protease selectively bound to the Ni-NTA resin, and the wash sample

contained only non-specific proteins. While the eluate for Ni-NTA and desalt columns contained

multiple proteins, the TEV protease band was considerably enriched relative to all other bands.

‐20

30

80

130

180

230

280

330

0 20 40 60 80 100 120 140 160

Absorban

ce 280nm (mAU)

Volume (mL)

Page 93: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

82  

Finally, the elution sample taken following SEC showed the final purified product with minimal

contamination.

 

Lane Name Description1 Ladder Molecular weight ladder (Bio-rad)2 Lysate Cellular lysate post-sonification.3 Pellet Insoluble cellular components, taken immediately after centrifugation

of cell lysate.4 Supernatant Soluble proteins, taken immediately after centrifugation of cell lysate. 5 Flow Through Proteins that did not bind to HisTrap column 6 Wash Low imidazole wash of His-Trap column7 Fractions Pooled peak fractions from high imidazole elution 8 Desalt Pooled peak fractions from desalting column9 Final (1 µL) 1 µL of the pooled and concentrated fractions from size exclusion

column 10 Final (1 µg) 1 µg of the pooled and concentrated fractions from size exclusion

column

 

Figure 2-23. SDS-PAGE analysis of a TEV protease purification.

Page 94: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

83  

2.6 Conclusion

This chapter highlights some of the indispensable tools used in protein purification, which

is the essential first step in most biochemical studies. Immobilized metal affinity chromatography

(IMAC) and size exclusion chromatography (SEC), in combination with sodium dodecyl sulfate

polyacrylamide gel electrophoresis (SDS-PAGE), were used in the purification of three different

proteins overexpressed in E.coli: NaAtm1, MSP1D1, and TEV protease. Using experimental

conditions optimized for each protein, all three were successfully obtained with high purity and

yield. With these components in-hand, the next step was to utilize MSP1D1 and TEV protease to

create a realistic environment to study NaAtm1, as described in Chapter 3.

Page 95: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

84  

2.7 References

[1] Chromatography. https://www.merriam-webster.com/dictionary/chromatography (accessed May 18, 2019).

[2] Porath, J.; Carlsson, J.; Olsson, I.; Belfrage, G. Metal chelate affinity chromatography, a new approach to protein fractionation. Nature 1975, 258(5536), 598–599.

[3] Hochuli, E.; Döbeli, H.; Schacher, A. New metal chelate adsorbent selective for proteins and peptides containing neighbouring histidine residues. Journal of Chromatography 1987, 411, 177–184.

[4] Lecture 7: Affinity Chromatography -II. https://nptel.ac.in/courses/102103017/pdf/lecture%207.pdf (accessed May 18, 2019).

[5] Hochuli, E.; Bannwarth, W.; Döbeli, H.; Gentz, R.; Stüber, D. Genetic Approach to Facilitate Purification of Recombinant Proteins with a Novel Metal Chelate Adsorbent. Nature Biotechnology 1988, 6(11), 1321–1325.

[6] Bornhorst, J. A.; Falke, J. J. Purification of proteins using polyhistidine affinity tags. Methods in Enzymology Applications of Chimeric Genes and Hybrid Proteins Part A: Gene Expression and Protein Purification 2000, 245–254.

[7] Chemistry 41. http://www.dartmouth.edu/~chem41/lab.html#WeekTwo (accessed Jun 20, 2019).

[8] Synge, R. L. Fractionation of hydrolysis products of amylose by electrokinetic ultrafiltration in an agaragar jelly. Journal of Biochemistry 1950, 2, 41–42.

[9] Barrer, R. M.; Zeolites as adsorbents and molecular sieves. Annual Reports on the Progress of Chemistry 1944, 41, 31–46.

Page 96: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

85  

[10] Hong, P.; Koza, S.; Bouvier, E. S. P. Size-Exclusion Chromatography for the Analysis of Protein Biotherapeutics and their Aggregates. Journal of Liquid Chromatography & Related Technologies 2012, 35, 2923–2950.

[11] Guide to Gel Filtration or Size Exclusion Chromatography. Harvard Apparatus. https://www.harvardapparatus.com/media/harvard/pdf/Guide+for+Gel+Filtration.pdf (accessed May 18, 2019).

[12] Abidin, R. S. Murine polyomavirus VLPs as a platform for cytotoxic T cell epitope-based influenza vaccine candidates. thesis, 2016.

[13] Electrophoresis. https://www.merriam-webster.com/dictionary/electrophoresis (accessed May 18, 2019).

[14] Laemmli, U. K. Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature 1970, 227(5259), 680-685

[15] Caprette, D. R. https://www.ruf.rice.edu/~bioslabs/studies/sds-page/gellab2.html (accessed May 18, 2019).

[16] Methods for Working with Proteins. http://www.siumed.edu/~bbartholomew/course_material/protein_methods.htm (accessed Jun 20, 2019).

[17] DTT DTTRO. https://www.sigmaaldrich.com/catalog/product/roche/dttro?lang=en®ion=US (accessed May 18, 2019).

[18] The principle and Procedure of Polyacrylamide Gel Electrophoresis (SDS-PAGE) – How Biotech. https://howbiotech.com/the-principle-and-procedure-of-polyacrylamide-gel-electrophoresis-sds-page/ (accessed May 18, 2019).

Page 97: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

86  

[19] How SDS-PAGE Works. https://bitesizebio.com/580/how-sds-page-works/ (accessed May 18, 2019).

[20] Optimal configuration of ÄKTA™ pure 25 for small-scale SEC. https://cdn.gelifesciences.com/dmm3bwsv3/AssetStream.aspx?mediaformatid=10061&destinationid=10016&assetid=17984 (accessed Jun 20, 2019).

[21] NanoPhotometer-N60-N50-spectrophotometer. https://www.implen.de/product-page/nanophotometer-n60-n50-spectrophotometer/ (accessed Jun 20, 2019).

[22] Lee, J. Y.; Yang, J. G.; Zhitnitsky, D.; Lewinson, O.; Rees, D. C. Structural basis for heavy metal detoxification by an Atm1-type ABC exporter. Science 2014, 343(6175), 1133–1136.

[23] Ritchie, T.; Grinkova, Y.; Bayburt, T.; Denisov, I.; Zolnerciks, J.; Atkins, W.; Sligar, S. Reconstitution of Membrane Proteins in Phospholipid Bilayer Nanodiscs. Methods in Enzymology Liposomes, Part F 2009, 211–231.

[24] Phospholipid Bilayer Nanodiscs. http://depts.washington.edu/wmatkins/nanodiscs.html (accessed Jun 20, 2019).

[25] Phan, J.; Zdanov, A.; Evdokimov, A. G.; Tropea, J. E.; Peters, H. K.; Kapust, R. B.; Li, M.; Wlodawer, A.; Waugh, D. S. Structural Basis for the Substrate Specificity of Tobacco Etch Virus Protease. Journal of Biological Chemistry 2002, 277(52), 50564–50572.

Page 98: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

87  

CHAPTER 3

Creating a Realistic Environment in Vitro

3.1 Introduction

In order to accurately dissect the mechanism of any protein of interest, experiments must

be performed under conditions that mimic the native environment as closely as possible. In order

to understand how NaAtm1 drives transport in a membrane, we aimed to incorporate NaAtm1 into

a robust and reliable lipid bilayer system.

I have always thought of ABC transporters as the doorways of the cell. Indeed, they are

devices that allow selective passage through a lipid bilayer (wall), they have mechanisms (NBDs)

that require energy to physically change their 3D structure (door handles), and they have conserved

amino acid sequences that allow these conformational changes to take place (hinges).

Unfortunately, as with most doors, when they are removed from their wall they no longer act as

doors. If one wanted to study how these doors opened and closed, it would be advantageous to

make a temporary frame, wall, or room to put them in. There are three common environments used

to mimic a lipid bilayer: detergent micelles, liposomes, and nanodiscs.

Page 99: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

88  

Figure 3-1. Analogies to the different environments created for membrane protein experiments A) A bare membrane protein: without a supporting structure the protein will not function – like a door without a frame. B) A membrane protein surrounded by a detergent micelle: a minimum supporting structure that allows the protein to function – like a flimsy door frame around a door. C) A protein surrounded by a liposome: a full replica of a cell – like building an entire room for a door (more effort than necessary). D) A membrane protein surrounded by a nanodisc: acts as a wall that can give the door proper structure to function while keeping the structure simple and robust.

3.2 Types of Environments

To maintain its inclusion within a lipid bilayer, integral membrane proteins, including ABC

transporters, have many hydrophobic residues lining the exterior surface of the transmembrane

domains. If a transporter is removed from the lipid bilayer and placed into an aqueous environment,

these hydrophobic residues will rearrange to fold in on themselves or will drive the aggregation of

multiple protein molecules. To maintain the structure of transporters after extraction from the

native cell membrane, scientists have developed several lipid bilayer-like environments.

Page 100: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

89  

3.2.1 Detergent micelles

A detergent micelle is a self-assembling collection of surfactants –

amphipathic molecules that contain a hydrophilic polar head group and a

hydrophobic non-polar tail. When a critical concentration of these

molecules is reached, they assemble in an organized way that creates a

sphere of polar head groups with the non-polar tails on the interior of this sphere (Figure 3-2).

If present above the critical micelle concentration, detergents can cover the hydrophobic

portions of a transporter to prevent deformation and aggregation. While this alleviates some of the

issues involved with isolating a membrane protein in aqueous solution, it does have limitations.

The architecture of the surrounding micelle is fundamentally different from a bilayer, introducing

artificial contacts between the hydrophobic surfaces of the protein and its environment. These

interactions could alter native structure and dynamics of the transporter and therefore change its

activity.

Additionally, the micelle environment does not allow for the ability to directly monitor the

movement of substrate across a membrane. Given the lack of a discrete compartment in which

substrate can accumulate, experiments to measure transporter activity are limited to indirect

strategies, such as the stimulation of ATPase activity or detection of conformational change within

the protein.

Figure 3-2. Structure

of a Micelle [1].

Page 101: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

90  

3.2.2 Liposomes

Liposomes are artificially created vesicles – spheres composed of

a closed phospholipid bilayer (Figure 3-3). Phospholipids are similar to

detergents in that they have a polar head and a hydrophobic tail

component, however their chemical structure is very different. One major

difference is that detergents only have one hydrophobic tail while

phospholipids have two (Figure 3-4). The ultimate effect of the two lipid tails is the formation of

a lipid bilayer, and if there is enough bilayer, it is thermodynamically favored to close upon itself

when in aqueous solution. There are both advantages and disadvantages to using liposomes over

micelles. One major advantage is that liposomes have a true phospholipid bilayer that more closely

resembles the natural membrane environment. Another advantage is that due to the closed nature

of the environment (presence of the internal cavity), transporters that are placed in a liposome can

be evaluated for substrate transport.

Figure 3-4. Structures of the phospholipid phosphatidylcholine (PC) and the detergent sodium dodecyl sulfate (SDS).

Figure 3-3. Structure of a liposome [1].

Page 102: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

91  

Figure 3-5. Liposome containing a 1:1 ratio of inward and outward facing ABC transporters. In this scenario the ABC transporters are importers.

Unfortunately, creating and maintaining liposomes can be very challenging due to their

size. A major experimental challenge in using ABC transporters in a liposome is its orientation

within the membrane. While many scientists assume an equal distribution (50% outward, 50%

inward, Figure 3-5), the specific activity of the transporter is difficult to accurately measure.

Furthermore, the insertion of membrane proteins into a liposome is demanding in practice – large

amounts of protein must be available for reconstitution procedures which may require multiple

rounds of optimization given a particular transporter.

3.2.3 Nanodiscs

Nanodiscs are small sections of phospholipid bilayer held together by an alpha-helical

protein belt. This belt is a genetically modified form of human apolipoprotein A1 (Figure 32a).

Apolipoprotein A1 is a major protein component of HDL particles found in plasma and plays a

major role in lipid transport throughout the body [2]. It is comprised of a globular region followed

by multiple amphipathic α-helices (Figure 3-6B). Nanodiscs were first reported by Baybert et al.

Page 103: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

92  

in 1998 [3], with the modified apolipoprotein described as a “Membrane Scaffold Protein” (MSP).

All MSPs have been modified by replacing the globular region with a histidine tag and TEV

protease cleavage site, and each contain varying numbers of α-helices. By varying the number of

α-helices contained in the MSP, the final disc diameter can be modulated to accommodate different

sizes of transmembrane proteins. Membrane scaffold proteins are now widely available in

diameters ranging from 7 to 17 nm.

Figure 3-6. A) Structure of a nanodisc containing a transmembrane protein (green), surrounded by a phospholipid bilayer (red spheres with gray tails), held together by two modified Apolipoproteins (MSPs) (purple). B) Organization of the naturally occurring Apolipoprotein A1 and two genetically modified MSPs, MSP1D1 and MSP1E3D1 [4].

In contrast to liposomes, nanodiscs are relatively easy to generate in a self-assembling

process. As with the detergent micelle, this system is open, which makes it impossible to determine

if a substrate is being transported through the protein or if it is merely stimulating activity. Even

with this drawback, they are less labor intensive and more durable than liposomes while creating

a more realistic environment than detergent micelles. While detergent micelles stabilize isolated

membrane proteins in an aqueous solution and are relatively easy to create, there have been

discrepancies in enzyme activity based on the choice of mimetic environment in general. For

example, BtuCD transporters reconstituted into nanodiscs have been shown to have an almost 6-

fold decrease in uncoupled hydrolysis compared to transporters in detergent micelles [5]. It is

Page 104: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

93  

believed that these discrepancies come from the detergents causing the protein to adopt a structure

that is different from the one it takes in lipid. To address this issue and to create a more realistic

environment to test NaAtm1, we aimed to insert NaAtm1 into nanodiscs as described below.

3.3 Overview of nanodisc reconstitutions

NaAtm1 nanodiscs are created in a self-assembling reaction that requires purified

Membrane Scaffold Protein (MSP), micelles containing a mixture of lipid and detergent, and

purified NaAtm1 in detergent micelles. When these components are mixed using an optimized

molar ratio of NaAtm1-MSP1D1-Lipid, small polystyrene “Bio-Beads” are added. These beads

absorb the detergent from both the purified NaAtm1 as well as the lipid/detergent micelles while

leaving the lipid component untouched. This results in the accumulation of lipid around NaAtm1

and other lipids. The MSP1D1 then wraps around the lipid to form both full discs (containing

embedded transporter) and empty discs (Figure 3-7). Finally, the full disc species is isolated via

chromatography.

Figure 3-7. Overview of the nanodisc reconstitution process [6, 7]. This process consists of combining the target

membrane protein with MSP, lipid/detergent micelles, and Bio-Beads. The Bio-Beads absorb the detergent from the lipid/detergent micelles as well as the detergent micelles around the target membrane protein. This process results in the accumulation of lipid around the membrane protein and other lipids resulting in full and empty nanodisc.

Page 105: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

94  

3.3.1. Preparation of nanodisc reagents

3.3.1.1 Membrane scaffold protein (MSP)

The NaAtm1 reconstitutions described here employs MSP1D1,

which was chosen due to its availability and final diameter of ~9.7 nm.

At ~9.7 nm, the lipidic area accommodates only one NaAtm1. MSP1D1

contains an N-terminal His-tag (7x histidine residues), a TEV protease

cleavage site, and ten α-helices (Figure 3-9B). Two MSP1D1 proteins

wrap around the lipid bilayer form one disc that measures roughly 9.8 nm in diameter and 4.6 nm

(lipid bilayer only) in height. The purification of this protein can be found in Chapter 2.

Figure 3-9. A) The amino acid sequence of MSP1D1. B) Comparison of Apolipoprotein A1 and MSP1D1 [4].

3.3.1.2 1:3 PC:E lipids

The lipid mixture used in reconstitutions was based on previous studies using ABC

transporters [8]. A 1:3 ratio of chicken egg phosphatidylcholine (PC) to E. coli polar lipid extract

is a common starting point for reconstitutions. The composition of each of these components is

shown in Figure 3-10.

 Figure 3-8: Cartoon rendering of MSP1D1 [7].

Page 106: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

95  

Figure 3-10. A) Natural lipid mixture found in chicken egg PC. B) Composition of E. coli polar extract. C) Representative structure of phosphatidylcholine (PC) from chicken egg [9, 10].

To prepare this lipid mixture, commercially available chloroform stocks of egg PC and E.

coli polar lipid extract were mixed with a final mass of 100 mg of chicken egg PC to 300 mg of E.

coli polar extract (Avanti Polar Lipids). Lipids were transferred to a 500 mL balloon flask and

rotovaped to remove chloroform. Once dry, the flask was wrapped in tinfoil to prevent

photodegredation, and was placed in a desiccator overnight under vacuum to remove any

remaining chloroform.

The next morning, 15 mL of Reconstitution Buffer A (20 mM Tris HCl (pH 7.5) and 100

mM NaCl) was added to the flask. The mixture was then placed under argon to prevent oxidation

and shaken overnight at room temperature at 120 rpm. The following day, the resuspended lipids

were collected, and 5 mL of Reconstitution Buffer A was added to the flask to recover any

remaining lipid. The resuspended lipids were then brought to a final concentration of 20 mg/mL

in Reconstitution Buffer A. This lipid stock was then placed on ice and sonicated using a using a

Branson Sonifier 250 with an output of 4.5 and a duty cycle of 20% for six cycles of 15 seconds

Page 107: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

96  

sonification with a 45 second cool-down period in between. Once sonicated, the lipids were

divided into 100 µL aliquots, flash frozen in liquid nitrogen, and stored at -80 °C.

For use in reconstitutions, lipids were further processed on a smaller scale. Because

nanodiscs form upon removal of detergent, both from the hydrophobic region of the membrane

protein and from the lipid/detergent micelles by Bio-beads, the type of detergent used in this

process is very important. The detergent sodium cholate is used to create lipid/detergent micelles

for use in this process to due sodium cholates ability to be quickly and easily removed from the

mixture - allowing for plenty of lipid to be present for the membrane protein to surround itself with

once out of its detergent micelle. 4 mg of sodium cholate was added to 2 mg of lipid and brought

up to a 1 mL total volume using NaAtm1 Buffer A. This mixture was vortexed and then placed in

a waterbath sonicator for 30 minutes. The ultimate composition of the lipid mixture used in

reconstitutions was 2 mg/mL 1:3 (w/w) egg PC: E. coli polar lipid extract, 4 mg/mL sodium

cholate, and approximately 20 mM Tris, 100 mM NaCl, and 0% DDM.

3.3.1.3 Bio-Beads

Dry Bio-Beads (Amberlite® XAD-2) were purchased from Bio-Rad. ~15 mL of dry Bio-

Beads were first washed with 100% methanol in a 50 mL conical tube. After inverting multiple

times, the Bio-Beads were allowed to settle before removal of the methanol. The process was

repeated using an additional round of 100% methanol, 2 rounds of 95% ethanol, 2 rounds of

ddH2O, and finally 2 rounds Reconstitution Buffer A. Washed Bio-Beads were stored in a minimal

volume of Reconstitution Buffer A (enough to prevent beads from drying out) and stored at 4 °C

until needed.

Page 108: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

97  

3.3.2 Reconstitution methods

3.3.2.1 Assembly and purification of small-scale reconstitutions

As shown in published literature, each integral membrane protein requires an optimized

ratio of protein-of-interest, scaffolding protein, and lipid. To determine the proper conditions for

assembly of NaAtm1 nanodiscs, a series of small-scale reconstitutions were prepared and

analyzed.

In general, small-scale reconstitutions were approximately 300 µL in volume and prepared

in 1.5 mL microcentrifuge tubes. Reagents were added in the order of lipid, MSP1D1,

Reconstitution Buffer A (20 mM Tris-HCl and 150 mM NaCl pH 7.5), and NaAtm1. 50 µL of wet

Bio-Beads was then added, and the mixture placed on an Adams Nutator at 4 °C overnight.

The following day, the reconstitution was centrifuged for 1 minute at maximum speed in a

microcentrifuge to pellet the Bio-Beads. The supernatant was then applied to a 0.22 µm filter to

remove aggregates. The sample was injected over a Superdex 200 Increase 10/300 GL sizing

column at 1 mL/min. Elution fractions were collected for subsequent analysis.

3.4 Reconstitution results

3.4.1 Identification of multiple species present in reconstitutions

Initial reconstitutions were created using a 1-3-60 molar ratio of NaAtm1-MSP1D1-lipid

with a final transporter concentration of 10 µM, as recommended in published protocols [11]. This

reconstitution yielded three peaks (Figure 3-11) at ~12 mL, ~13.75 mL, and ~15.75 mL. A sample

of each peak was analyzed using SDS-PAGE (not shown) and it was determined that Peak 1

contained both NaAtm1 and MSP1D1 while Peaks 2 and 3 contained only MSP1D1. This result

suggested that the Peak 1 contained NaAtm1 successfully embedded in a nanodisc; however, more

testing was needed to verify the composition of this species. For this section, Peak 1 refers to the

Page 109: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

98  

peak that elutes at ~12 mL elution volume, Peak 2 refers to the peak that elutes at ~13.75 mL

elution volume, and Peak 3 refers to the peak that elutes at ~15.75 mL elution volume.

Figure 3-11. 1-3-60 NaAtm1-MSP1D1-Lipid reconstitution with a final NaAtm1 concentration of 10 µM and a final

volume of 300 uL.

3.4.1.1 Reconstitution without NaAtm1: 0-3-60

To aid in identifying the multiple peaks, different versions of the 1-3-60 reconstitutions

were prepared with one component removed (i.e. 0-3-60, 1-0-60, 1-3-0) at 5 µM final transporter

concentration.

A 0-3-60 NaAtm1-MSP1D1-lipid reconstitution yielded two peaks at the same elution

volume as Peaks 2 and 3 from the 1-3-60 reconstitution, with no peak at the same elution volume

as Peak 1. A sample of each peak was analyzed using SDS-PAGE (data not shown), and it was

determined that both peaks contained only MSP1D1, as expected. This result demonstrated that

NaAtm1 was a critical component of the species in Peak 1 seen in the 1-3-60

reconstitution. Furthermore, this suggested that the Peak 1 did not contain MSP1D1 aggregates,

excess nanodisc without transporter (“empty disc”), or MSP1D1 dimers or monomers.

Page 110: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

99  

Figure 3-12. Size exclusion chromatogram of a 0-3-60 5 µM 300 µL reconstitution. This chromatogram shows two

peaks that match Peaks 2 and 3 respectively in the 1-3-60 chromatogram (Figure 3-11).

3.4.1.2 Reconstitution without MSP1D1: 1-0-60

A reconstitution with a molar ratio of 1-0-60 of NaAtm1-MSP1D1-lipid was then created.

This reconstitution yielded a peak at the void volume, which is most likely a collection of non-

functional masses of NaAtm1 transporters. In the absence of the lipid bilayer enclosed by

MSP1D1, removal of detergent by Bio-Beads would drive aggregation of the hydrophobic regions

of NaAtm1. This result implies that MSP1D1 is a critical component of Peak 1 in the 1-3-60

reconstitution.

Figure 3-13. Size exclusion chromatogram of a 1-0-60 5 µM 300 µL reconstitution. This chromatogram shows a small

peak at the void volume for the column used. This suggests that in the absence of MSP1D1, NaAtm1 cannot remain soluble after removal of detergent

Page 111: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

100  

3.4.1.3 Reconstitution without additional lipids: 1-3-0

A reconstitution with a molar ratio of 1-3-0 was then created. This reconstitution yielded

two peaks that matched Peak 1 and Peak 3 in the 1-3-60 reconstitution. This interesting result

might be due to the presence of a small amount of lipid around both NaAtm1 and MSP1D1 that

was not removed during purification. The combination of the small lipid stores around each protein

could be enough to create an intact nanodisc. These observations suggest that lipid is a critical

factor in Peak 2 but not in Peak 3. This leads to the conclusion that Peak 2 forms only in the

presence of both MSP1D1 and lipid. In addition, the 0-3-60 reconstitution showed that NaAtm1 is

not a critical component of Peak 2 or 3. Together, these findings suggest that Peak 2 is empty

nanodisc and Peak 3 is excess MSP1D1.

Figure 3-14. Size exclusion chromatogram of a 1-3-0 5 µM 300 µL reconstitution. This chromatogram shows two peaks

that correspond to Peak 1 and 3 respectively in the 1-3-60 reconstitution (Figure 3-11).

Page 112: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

101  

3.4.2 Conclusion for peak identification in small-scale reconstitutions

In summary, a 1-3-60 ratio of nanodisc components resulted in three peaks resolved by size

exclusion chromatography: NaAtm1 successfully reconstituted nanodiscs (Peak 1), empty

nanodiscs (Peak 2), and excess MSP1D1 that had not formed nanodiscs (Peak 3) (Figure 3-15).

Figure 3-15. A comparison of size exclusion chromatograms of reconstitutions with and without NaAtm1.

3.4.3 Reconstitution optimization

Once the different species in the initial 1-3-60 reconstitution were identified, the reconstitution

method was optimized to maximize the yield of NaAtm1 nanodiscs. To accomplish this, three

different variables were altered: NaAtm1:MSP1D1 molar ratio, NaAtm1:Lipid molar ratio, and

final NaAtm1 concentration.

3.4.3.1 NaAtm1:MSP1D1 optimization

Five reconstitutions were created using different molar ratios of NaAtm1:MSP1D1 (1-3-

60, 1-4-60, 1-5-60, 1-6-80, and 1-7-80 NaAtm1-MSP1D1-Lipid). Each reconstitution was

analyzed using size exclusion chromatography and overlaid for comparison (Figure 3-16). These

data showed that the height of Peak 1 significantly increased when the ratio was changed from 1:3

Page 113: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

102  

to 1:4, but for higher ratios (1:5-1:7) there was no appreciable difference in Peak 1 height. It was

also observed that as the molar ratio of NaAtm1:MSP1D1 increased, both Peak 2 and Peak 3

increased as well. This suggests that at a molar ratio of 1:4, the reaction has gone to completion

and that NaAtm1 is now the limiting reagent. From this conclusion, the molar ratio 1:4

NaAtm1:MSP1D1 was chosen for future reconstitutions.

Figure 3-16. Comparison of reconstitutions created using different molar ratios of NaAtm1:MSP1D1.

3.4.3.2 NaAtm1:Lipid optimization

Four reconstitutions were created using different molar ratios of NaAtm1:Lipid (1-5-50, 1-

5-60, 1-5-80, and 1-5-120 NaAtm1-MSP1D1-Lipid). The NaAtm1:MSP1D1 molar ratio of 1:5

was chosen instead of 1:4 to increase the prominence of Peak 3 within this experiment. By using

a ratio that produces a more noticeable Peak 3, the resulting change in the peak is visible while

changing the NaAtm1:Lipid ratio. Analysis via size exclusion chromatography showed that the

height of Peak 1 increased from 1:50 to 1:60 (data not shown) but for higher ratios (1:80, 1:120)

Page 114: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

103  

there was no appreciable difference in Peak 1 height. It was also observed that as the molar ratio

of NaAtm1:lipid increased, the Peak 2 increased while Peak 3 decreased.

These data suggest that as the amount of lipid increases, the formation of nanodisc is driven

to completion, dictated by the amount of scaffolding protein. When NaAtm1 is the limiting reagent

for primary product formation (full nanodisc), the secondary product (empty nanodisc)

accumulates. This data is consistent with the initial conclusions (Section 3.4.1) that Peak 2 contains

empty nanodisc and Peak 3 contains excess MSP1D1. For the optimization of these

reconstitutions, the goal was to maximize the NaAtm1 nanodiscs identified in Peak 1, while

minimizing the empty nanodiscs found in of Peak 2. The results suggest that these parameters are

best met with a molar ratio 1:60 NaAtm1:Lipid.

Figure 3-17. Comparison of reconstitutions created using different molar ratios of NaAtm1:Lipid.

Page 115: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

104  

3.4.3.3 Final NaAtm1 concentration optimization

The last condition that was explored was the final NaAtm1 concentration of all of the

components at a fixed ratio, which was necessary for scaling up to larger amounts of nanodiscs.

Reconstitutions of 1-5-80 were prepared at 5 µM, 10 µM, 15 µM, and 20 µM NaAtm1. For

example, the 15 µM sample would consist of 15 µM NaAtm1, 75 µM MSP1D1, and 1200 µM

lipids. All reconstitutions were tested in a 300 µL and analyzed by size exclusion chromatography

(Figure 3-18).

Figure 3-18. Comparison of reconstitutions created using different final concentrations of NaAtm1.

As expected, the height of all peaks were directly correlated with final protein

concentration. With multiple repeats, however, it was observed that reconstitutions at high

concentrations were less reproducible than at lower concentrations. High concentration

reconstitutions often resulted in extremely low amounts of total recovered material (data not

shown). One explanation is that the high concentration of material would be more easily prone to

Page 116: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

105  

aggregation, which would be removed by routine sample filtration before the analysis on the size

exclusion column. Due to the difficulty in reproducibility at higher final transporter concentrations,

5 µM final NaAtm1 concentration was used for subsequent reconstitutions.

3.4.4 Optimization conclusion

The results of these experiments show that an optimized reconstitution for NaAtm1 in

MSP1D1 using 1:3 PC:E lipid consists of a molar ratio of 1-4-60 with a final NaAtm1

concentration of 5 µM.

3.5 Determining the stoichiometry of nanodiscs using densitometry

Once the reconstitution ratio was optimized, the resulting purified nanodiscs were analyzed

to determine the molar ratio of NaAtm1 to MSP1D1. This was necessary for two reasons: (1) to

ensure that each nanodisc contained only one NaAtm1, and (2) to create a method for

quantification of nanodiscs via absorbance spectroscopy.

To accomplish this, nanodiscs were analyzed by SDS-PAGE and stained using the

fluorescent protein gel stain SYPRO Ruby, which is highly quantitative and sensitive to low

protein concentrations (Figure 3-19). For these densitometry experiments, a standard curve was

generated for both NaAtm1 and MSP1D1 based on the molecular weight of each (135361 g/mol

and 24793 g/mol, respectively). The protein solutions used to prepare the standard curve were

quantified using the calculated extinction coefficient of each (152,640 M-1 cm-1 and 24130 M-1

cm-1, respectively). Serial dilutions of the purified NaAtm1 nanodiscs samples were prepared

identically, with SDS and DTT present to denature and separate protein complexes. As seen in

Figure 3-19, both MSP1D1 and NaAtm1 were present in the isolated nanodiscs.

Page 117: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

106  

Figure 3-19. SDS-PAGE analysis using SYPRO Ruby stain.

The intensities of the protein bands in the “NaAtm1 Standard” and “MSP1D1 Standard”

were quantified to create a standard curve for each protein (Figure 3-20). The standard curve for

NaAtm1 was then used to determine how much NaAtm1 was present in each lane of the

“Reconstitution” serial dilutions (Figure 3-19, “Reconstitution” top protein band) while the

MSP1D1 standard curve was used to determine how much MSP1D1 was present in each lane of

the “Reconstitution” serial dilutions (Figure 3-19, “Reconstitution” bottom protein band).

Figure 3-20. A) NaAtm1 standard curve with an R2 value of 0.9916. B) MSP1D1 standard curve with an R2 value of

0.9564

Page 118: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

107  

The amount of each component (NaAtm1 and MSP1D1) in each lane was then compared

to determine the NaAtm1:MSP1D1 ratio (w/w). The average NaAtm1:MSP1D1 ratio (w/w) over

multiple lanes was 2.75:1 (Table 3-1). This experimentally determined w/w ratio is remarkably

similar to the ideal calculated w/w ratio of 2.73:1, which corresponds to a 1:2 mol ratio of

NaAtm1:MSP1D1. This result strongly suggests that the nanodiscs produced during this

experiment contain a molar ratio of 1:2 NaAtm1:MSP1D1.

Label: NaAtm1 calculated using

Standards (µg): MSP1D1 calculated using Standards (µg):

Ratios:

Lane 19 1.645 0.595 2.766

Lane 20 1.618 0.592 2.733

Lane 21 1.581 0.577 2.740

Lane 22 1.290 0.470 2.744

Lane 23 1.137 0.418 2.718

Lane 24 0.924 0.308 2.997

Lane 25 0.711 0.253 2.810

Lane 26 0.503 0.201 2.499

Average: 2.751

Table 3-1. Representative comparison of component parts quantified by densitometry

3.6 Quantification of NaAtm1 nanodiscs via absorbance spectroscopy

Based on the results of the densitometry analysis above, a simple “correction factor” could

be calculated to allow for the quantification of NaAtm1 nanodiscs via absorbance measurement at

280 nm. Using samples containing the same amount of NaAtm1 transporter, the absorbance of

NaAtm1 in detergent micelles was compared to that NaAtm1 in nanodiscs. It was observed that

the nanodisc samples showed a consistent 1.33-fold increase in absorbance when compared to

detergent-solubilized samples. Thus, using the calculated extinction coefficient of NaAtm1 alone,

the concentration of the NaAtm1-MSP1D1-lipid complex can be determined. For example, an

Page 119: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

108  

absorbance value corresponding to 500 nM NaAtm1 in detergent would translate into 667 nM

NaAtm1 in nanodisc. This correction factor will be appropriate for quantification of samples for

functional assays.

3.7 Transition to an IMAC-based purification method

Unfortunately, the results seen in table 3-1 were not consistent across multiple tests. A

higher-than-expected result with a w/w ratio of ~1:3 was often seen. The most likely cause of this

in consistency was the presence of empty nanodisc (Peak 2) that was unable to be separated from

Peak 1 using size exclusion chromatography. To rectify this issue, size exclusion chromatography

was replaced by an IMAC-based purification method for the isolation of NaAtm1 nanodiscs

following reconstitution.

To isolate NaAtm1 nanodiscs from empty nanodiscs, the His-tag on MSP1D1 can be

removed via proteolytic cleavage. Using this strategy, NaAtm1 nanodiscs could bind to Ni-NTA

resin due to the His-tag on NaAtm1, but empty discs would remain unbound and removed from

the mixture (Figure 3-21).

Page 120: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

109  

Figure 3-21. IMAC purification with MSP1D1 containing His-tag produces two products, both with His-tag. Cleavage of

the MSP1D1 His-tag results in one His-tagged product and one tagless product.

To remove the His-tag on MSP1D1, we took advantage of a TEV proteolytic cleavage site

between the His-tag and MSP1D1 coding sequence. Due to the high cost of commercially

available TEV, and the need for large quantities, we decided to express and purify TEV in-house.

Once available, the TEV: MSP1 reaction ratio was optimized, and subsequent purification of His-

tag-cleaved was performed.

3.8 Optimization of His-tag cleavage by TEV protease

TEV protease was purified as described in Chapter 2. Once purified, an initial test of

protease activity was conducted using MSP1D1. Manufacturers recommend 1 µg of commercially

available TEV protease for every 15 µg of target protein, which translates into approximately a

1:17 TEV: MSP1D1 molar ratio. To assess if lower amounts of TEV could be used to obtain

Page 121: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

110  

efficient cleavage, four reactions were created with varying amounts of MSP1D1. In these

reactions, 1 µg of TEV protease was applied to 15 µg, 50 µg, 100 µg, and 200 µg of MSP1D1.

Samples of each reaction were taken immediately upon mixing and quickly flash frozen in liquid

nitrogen for analysis. The cleavage reactions were then allowed to proceed overnight at 4 °C. The

next day, samples were taken and analyzed via SDS-PAGE (Figure 3-22).

Figure 3-22. Cleavage reactions with varying ratios of TEV: MSP1D1. T0 = 0 hours, T1 = 24 hours.

The results showed that after 24 hours, both the 1:15 (w/w) and 1:50 (w/w) samples showed

complete cleavage of MSP1D1 (removal of upper band) while the 1:100 (w/w) showed almost

complete cleavage and the 1:200 (w/w) showed incomplete cleavage of MSP1D1 (presence of

upper band). From this test it was determined that using 1:50 (w/w) was the best ratio to use for

future reactions.

3.9 Preparation of His-tag-cleaved MSP1D1

After determining the proper ratio of TEV protease to MSP1D1, a method was created to

purify isolate cleaved MSP1D1 from uncleaved MSP1D1.

Following cleavage for 24 hours at 4 °C, reactions were purified by IMAC using Ni-NTA

resin. In this case however, the isolation process was somewhat unconventional. In most cases,

the target protein contains the His-tag and is retained on the column. A low concentration of

imidazole (15-25 mM) is present to prevent non-specific binding by contaminating proteins but

Page 122: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

111  

also results in the loss of some target protein as well. This small loss of His-tagged target protein

is preferable to having contaminants in the final product. The His-tagged target protein is then

eluted at higher concentration imidazole, typically 350-400 mM. This general procedure was

applied to the purification of NaAtm1, MSP1D1, and TEV protease. In contrast, for purification

of the cleaved MSP1D1, the contaminating proteins are retained on the column, and the target

protein is collected in the “flow through” (Figure 3-23).

Figure 3-23. Normal IMAC vs Reverse IMAC

To optimize the final purity of the “flow through” which contains the target protein, the

imidazole level needs to be low enough that all unwanted his-tagged contaminants are retained on

the column, even at the loss of a small amount of target protein due to non-specific binding. To

determine the optimum imidazole concentration, 10 mM imidazole, 15 mM imidazole, and no

imidazole were added to cleavage reactions. Once the “flow through” was complete, the remaining

bound contaminant proteins were eluted using 350 mM Imidazole. Both “flow-through” and

Page 123: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

112  

elution peaks were analyzed for the presence of His-tag cleaved MSP1D1. The results showed that

an imidazole concentration of 15 mM was ideal to prevent non-specific binding of cleaved

MSP1D1, and also retain the contaminant tagged MSP1D1 on the column. Ultimately, a large

amount of tagless MSP1D1 was recovered for subsequent reconstitutions.

Figure 3-24. Comparison of His-tag separations using 0 mM, 10 mM, and 15 mM imidazole. For each run a constant concentration of noted imidazole was using during the “flow through”. Once complete, 350 mM imidazole was used during an elution phase for analysis purposes.

3.10 Purification of NaAtm1 in MSP1D1 nanodisc using IMAC

Once purified, His-tag-cleaved MSP1D1 was used in the creation of large-scale

reconstitutions using the optimized molar ratio of 1-4-60 NaAtm1-MSP1D1-Lipid with a final

volume of 1 mL and a final NaAtm1 concentration of 5 µM.

Page 124: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

113  

3.10.1 Assembly and purification of large-scale reconstitutions

Large-scale reconstitutions were created using the same procedure as described in section

3.3.2.1 with minor changes. These changes included replacing His-tagged MSP1D1 with His-tag-

cleaved MSP1D1, increasing the final volume from 300 µL to 1 mL, and increasing the Eppendorf

tube from 1.5 mL to 5 mL. It was normal for three to five 1 mL reconstitutions to be created at

once.

Following overnight incubation, reconstitutions were centrifuged at 3,500 rpm for 2

minutes to pellet the Bio-Beads. The reconstitution mixtures were then pooled and filtered using a

0.22 µm filter to remove aggregates. Imidazole was added to the filtered mixture to create a final

concentration of 15 mM. The sample was then flowed over a 1 mL HisTrap HP (Ni-NTA) column,

using the column in the traditional way to capture His-tagged full nanodisc (Figure 3-25). The

target protein was eluted and fractions collected. Pooled fractions were then concentrated to under

1.5 mL using a 100kDa concentrator before being injected over a 5 mL HiTrap Desalting column

to remove imidazole. Given this new method of purification, no size exclusion chromatography

was necessary. IMAC fractions were collected, aliquoted, and flash frozen in liquid nitrogen for

use and further analysis.

Page 125: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

114  

Figure 3-25. IMAC chromatogram of nanodisc separation using a 1 mL HisTrap HP Ni-NTA column.

3.10.2 Analysis of large-scale reconstitutions

A 500 µL sample of the desalted protein was thawed and flowed over a Superdex 200

Increase 10/300 GL size exclusion column for confirmation of a single species. The results showed

a single peak around 12.1 mL (Figure 3-26). When compared to a small-scale reconstitution of 1-

4-60 NaAtm1-MSP1D1-Lipid 300 µL, Peak 1 in the small-scale reconstitution overlaps nicely

with the single peak from the IMAC purified nanodisc, and the contaminating Peaks 2 and 3 are

completely removed (Figure 3-27). This suggests that the IMAC purification process works well

for the creation of ultra-pure NaAtm1 in MSP1D1 nanodisc.

Page 126: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

115  

Figure 3-26. Full nanodisc from IMAC purification analyzed using SEC.

Figure 3-27. Comparison of IMAC separated nanodisc (blue) and small-scale 1-4-60 reconstitution.

3.11 Conclusion

In this Chapter I have described multiple structures that can be used to mimic a lipid bilayer

- the natural environment of NaAtm1 and ABC transporters in general. These structures include

detergent micelles, liposomes, and nanodiscs. Nanodiscs were chosen for studies with NaAtm1

due to their relative robustness compared to liposomes and their capacity to more closely mimic a

lipid bilayer than detergent micelles. Through iterative modifications of both reconstitution

component ratios and purification methods, I have created a method for the assembly of NaAtm1-

MSP1D1 complexed nanodiscs. Importantly, this system is scalable to the larger quantities needed

for future functional studies.

Page 127: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

116  

3.12 References

[26] Structure of the plasma membrane. https://www.khanacademy.org/science/high-school-biology/hs-cells/hs-the-cell-membrane/a/structure-of-the-plasma-membrane (accessed Aug 1, 2019).

[27] APOA1 apolipoprotein A1 [Homo sapiens (human)] - Gene - NCBI. https://www.ncbi.nlm.nih.gov/gene/335 (accessed Aug 1, 2019).

[28] Bayburt, T. H.; Carlson, J. W.; Sligar, S. G. Reconstitution and Imaging of a Membrane Protein in a Nanometer-Size Phospholipid Bilayer. Journal of Structural Biology 1998, 123(1), 37–44.

[29] Membrane Scaffold Protein 1E3D1 M7074. https://www.sigmaaldrich.com/catalog/product/sigma/m7074?lang=en®ion=US (accessed Aug 1, 2019).

[30] Kim, J. The Molecular Mechanism of the Escherichia Coli vitamin B12 Transporter BtuCD-F: Real-time Observation of the Transporter in Motion. thesis, 2012.

[31] Lee, J. Y.; Yang, J. G.; Zhitnitsky, D.; Lewinson, O.; Rees, D. C. Structural basis for heavy metal detoxification by an Atm1-type ABC exporter. Science 2014, 343(6175), 1133–1136.

[32] Mazhab-Jafari, M. T.; Marshall, C. B.; Smith, M. J.; Gasmi-Seabrook, G. M.; Stathopulos, P. B.; Inagaki, F.; Kay, L. E.; Neel, B. G.; Ikura, M.; Oncogenic and RASopathy-associated K-RAS mutations relieve membrane-dependent occlusion of the effector-binding site. PNAS 2015, 112(21), 6625–6630.

[33] Borths, E. L.; Poolman, B.; Hvorup, R. N.; Locher, K. P.; Rees, D. C. In Vitro Functional Characterization of BtuCD-F, TheEscherichia ColiABC Transporter for Vitamin B12Uptake†. Biochemistry 2005, 44(49), 16301–16309.

[34] E. coli Extract Polar. https://avantilipids.com/product/100600 (accessed Aug 1, 2019).

Page 128: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

117  

[35] Egg PC. https://avantilipids.com/product/840051 (accessed Aug 1, 2019).

[36] Bao, H.; Duong, F.; Chan, C. S. A Step-by-Step Method for the Reconstitution of an ABC Transporter into Nanodisc Lipid Particles. Journal of Visualized Experiments 2012, No. 66.

Page 129: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

118  

CHAPTER 4

Preliminary Kinetic Assays and Future Work

4.1 Introduction

The ultimate goal of our NaAtm1 studies is to understand the molecular mechanism by

which this protein drives transport of substrates across lipid membranes. The incorporation of

NaAtm1 into nanodiscs is an important foundational step in this pursuit. While the structural

studies in detergent have provided a wealth of information on NaAtm1, careful kinetic and

thermodynamic assays in lipidic environments will significantly further our understanding of its

mechanism. Questions include: what is the natural substrate for NaAtm1, how quickly does

NaAtm1 transport this substrate, and what is the affinity for this substrate? Using both DDM-

solubilized NaAtm1 and the newly-formed NaAtm1 nanodiscs, two preliminary experiments were

conducted.

4.2 Kinetic assays

First, the rate of ATP hydrolysis was measured as a function of ATP concentration. This

canonical experiment provides three key parameters when fit to a Michaelis-Menten model: (1)

the turnover number (kcat), which is the rate constant for ATP hydrolysis at saturating concentration

(2) the Michaelis constant (Km), which is the concentration of the substrate at which the turnover

rate is half the calculated maximum velocity, and (3) the Hill coefficient (n), which is a measure

of cooperativity between the two ATP sites. Using the system described in Chapter 2, the activity

of detergent-solubilized NaAtm1 was monitored in the absence of substrate. Best-fit values were

kcat = 14.42 min-1, Km = 36 µM, and n = 1.36. The latter value suggests that the binding of the first

ATP molecule to NaAtm1 increases the affinity for the second molecule, a feature that is common

Page 130: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

119  

amongst ABC transporters. This experiment was consistent with published parameters, which had

a slightly lower turnover rate of 8.8 ± 0.8 min-1 at room temperature, while our data was collected

at 37 °C.

A few months later, however, the result was startlingly different, with a turnover rate of

~54 min-1. (Figure 4-1). During the troubleshooting process, all reagents were made fresh,

including the preparation of NaAtm1, however the results remained consistently 4-fold higher. It

is still unclear what has caused this dramatic change in activity and what it means for further kinetic

assays.

Figure 4-1. Comparison of the kobs for an initial assay (circle) and subsequent experiments (triangle).

Next, the kinetic activity of NaAtm1 nanodiscs was directly compared to NaAtm1 in

detergent. Surprisingly, the results were identical. Most ABC transporters show drastic inhibition

of uncoupled hydrolysis upon insertion into a lipid bilayer, and thus further experimentation will

determine if NaAtm1 is an exception to this case.

Page 131: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

120  

Figure 4-2. A) Kinetic assay for NaAtm1 in detergent. B) Kinetic assay for NaAtm1 in nanodisc. The kcat, Km, and n for

both assays are startlingly similar.

4.3 Future work

While my time at USF is at an end, this project still has many more avenues of inquiry.

Now that methods of purification and nanodisc assembly have been created, there is a solid

foundation for further work. Types of future experiments include further kinetic assays combined

with mutational analyses, and EPR or fluorescence spectroscopy tests.

4.3.1 Further kinetic assays

Detailed kinetic studies will yield important insights into the mechanism of NaAtm1, such

as the coordination between the two NBDs, substrate preferences, and the efficiency of transport.

While NaAtm1 mutants have not been discussed in this thesis, mutant constructs of NaAtm1 have

been created to help answer these questions. These are K400A (Walker A mutant to remove ability

to bind ATP), E523Q (Walker B mutant to remove ability to hydrolyze ATP), and K142Q (an

equivalent of R216Q [1] that should prohibit binding of the substrate GSSG). The incorporation

kcat (min-1) Km (µM) n

NaAtm1 in detergent: 54.19 238 1.896 NaAtm1 in nanodisc: 54.59 267 1.743

Page 132: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

121  

of these mutants into kinetic assays, in addition to comparisons between detergent micelles and

nanodiscs, will allow us to further explore NaAtm1 function.

4.3.2 Detection of conformational changes

Currently, there is only one high-resolution crystal structure of NaAtm1, much like a single

snapshot of the enzyme in motion. To detect conformational changes that take place during the

transport cycle, a cysteine mutation was introduced which enables fluorescent labeling at the

reactive thiol side chain (NaAtm1 does not contain any naturally-occurring cysteines). Residue

S526 was selected as a labeling site because it is located in the NBDs, which is believed to move

substantially and is within the critical distance needed for use in fluorescence resonance energy

transfer (FRET) experiments. This technique will allow us to monitor changes in the relative

distance between the NBDs as a function of nucleotide-state and substrate binding. We could also

compare the flexibility of the protein in detergent micelles versus lipid bilayers. Furthermore, spin-

labeling of the cysteine may allow us to follow these motions using electron paramagnetic

resonance spectroscopy, another technique that has been commonly used to detect conformational

changes in proteins. Lastly, introduction of cysteine residues at other locations will allow us to

construct a more complete picture of NaAtm1 conformations during the transport cycle.

Page 133: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

122  

Figure 4-3. Side view (A) and bottom view (B) of the S526C mutation site (red) that will be used for FRET analysis [2].

4.4 Conclusion

This thesis has detailed ABC transporters, their diversity, and their importance to human

health. It focuses on the ABCB-type importer NaAtm1 - an exporter shown to export the heavy

metals silver and mercury in vivo [2]. The main goal of this work was to create a robust lipidic

environment in which to study NaAtm1. This mimetic environment came in the form of nanodiscs

- a self-assembling lipid bilayer disc that surrounds NaAtm1.

To accomplish this, the purification of NaAtm1 and the scaffolding protein MSP1D1 were

conducted and optimized. Through careful and systematic modification of the nanodisc

components, an ideal reconstitution ratio of NaAtm1:MSP1D1:lipids was determined, and the

proper stoichiometry of 2 MSP1D1 per 1 NaAtm1 was verified. Additionally, an improved method

for the enrichment of NaAtm1 nanodiscs was developed based on affinity chromatography. This

work lays the foundation for future studies on elucidating the mechanism of NaAtm1 in a lipidic

environment.

Page 134: NaAtm1: Examining the mechanism of a Heavy-metal ABC Exporter

 

123  

4.5 References

[37] Schaedler, T. A.; Thornton, J. D.; Kruse, I.; Schwarzlander, M.; Meyer, A. J.; Van Veen, H. W.; Balk, J.; A conserved mitochondrial ATP-binding cassette transporter exports glutathione polysulfide for cytosolic metal cofactor assembly. Journal of Biological Chemistry 2014, 289(34), 23264–23274.

[38] Lee, J. Y.; Yang, J. G.; Zhitnitsky, D.; Lewinson, O.; Rees, D. C. Structural basis for heavy metal detoxification by an Atm1-type ABC exporter. Science 2014, 343(6175), 1133–1136.