185
MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR CHLORIDE CHANNEL _______________________________________ A Dissertation presented to the Faculty of the Graduate School at the University of Missouri-Columbia _______________________________________________________ In Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy _____________________________________________________ by KANG-YANG JIH Dr. Tzyh-Chang Hwang, Dissertation Supervisor JULY 2012

MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF

THE CFTR CHLORIDE CHANNEL

_______________________________________

A Dissertation

presented to

the Faculty of the Graduate School

at the University of Missouri-Columbia

_______________________________________________________

In Partial Fulfillment

of the Requirements for the Degree

Doctor of Philosophy

_____________________________________________________

by

KANG-YANG JIH

Dr. Tzyh-Chang Hwang, Dissertation Supervisor

JULY 2012

Page 2: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

The undersigned, appointed by the dean of the Graduate School, have examined the

dissertation entitled

MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR CHLORIDE

CHANNEL

presented by Kang-Yang Jih,

a candidate for the degree of doctor of philosophy,and hereby certify that, in their

opinion, it is worthy of acceptance.

Professor Tzyh-Chang Hwang

Professor Kevin Gillis

Professor Mike Hill

Professor Luis Polo-Parada

Professor Xiaoqin Zou

Page 3: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

ii

ACKNOLEDGEMENTS

I gratefully acknowledge people who had assisted and instructed me during the

past three years, including my committee members, colleagues in Dr. Hwang’s lab,

my family and friends. The works I about to describe in this thesis cannot have been

done without their kind supports.

I would like to express my deepest gratitude to my dissertation advisor, Dr.

Tzyh-Chang Hwang. Dr. Hwang is an enthusiastic mentor who effortlessly guided me

with his critical thoughts and erudite knowledge, shaping me into a critical-thinking,

independent scientist. He also inspired me to appreciate the true beauty of science,

the excitement in exploring the unknowns.

I would like to extend my gratitude to my dissertation committee members, Dr.

Kevin Gillis, Dr. Mike Hill, Dr. Luis Polo-Parada and Dr. Xiaoqin Zou. They attended

my committee meetings and provided valuable inputs to my thesis project as well as

offer insightful feedbacks to my manuscripts.

Also, I want to express my gratitude to Dr. Yoshiro Sohma in Keio University,

Japan. Dr. Sohma provides assistance in computer simulation and modeling parts of

this project.

Many thanks to the past and current member of Hwang lab, Dr. Yonghong Bai,

Cindy Chu, Xiaolong Gao, Shenghui Hu, Dr. Zoia Kopeikin, Dr. Min Li, Dr. Yumi

Nakamura, Dr. Ming-Feng Tsai and Jingyao Zhang and Dr. Silvia Bompadre, who is

now an Assistant Professor in Department of Physics. Especially I would like to

Page 4: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

iii

thank Dr. Mingfeng Tsai, who was a senior PhD student when I joined the lab, he

patiently taught me the experimental skills and provided critical insights for my first

two projects. Special thanks to Cindy Chu and Dr. Min Li who made all the DNA

construct and did all the molecular biology experiments for my projects. It is a

wonderful experience working with you all.

Finally, I would like to thank my parents and my partner, Chinghui Hsu. They

had always been supportive and encouraging during my Ph.D. study. Having a loved

one far away from home must be very difficult for them. After completing my PhD

works, I will return to Taiwan and try to make up for the lost time.

Page 5: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

iv

TABLE OF CONTENTS

ACKNOLEDGEMENTS………………………………………………………………………………………ii

LIST OF FIGURES……………………………………………………………………………………………vii

LIST OF TABLES……………………………………………………………………………………………….x

ABSTRACT………………………………………………………………………………………………………xi

CHAPTERS

1. INTRODUCTION

1-1. Overview and clinical significance of CFTR……………………………….......1

1-2. Structure and functions of CFTR…………………………………………………..3

1-3. Coupling between ATP hydrolysis and gating………………………………..5

1-4. Fishing kinetic states with non-hydrolytic ATP analogs ………………...9

1-5. CFTR pharmacology…………………………………………………………………...11

1-6. References…………………………………………………………………………………13

2. IDENTIFICATION OF A NOVEL POST-HYDROLYTIC STATE IN CFTR

GATING

2-1. Abstract……………………………………………………………………………………….18

2-2. Introduction………………………………………………………………………………...19

2-3 Material and Methods……………………………………………………………….….24

2-4 Results………………………………………………………………………………………...27

Page 6: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

v

2-5 Discussion…………………………………………………………………………………...43

2-6 Supplementary information…………………………………………………………52

2-7 References…………………………………………………………………………………..60

3. NON-INTEGRAL STOICHIOMETRY IN CFTR GATING, SEEING IS

BELIEVING

3-1. Abstract……………………………………………………………………………………….64

3-2. Introduction………………………………………………………………………………...65

3-3 Material and Methods…………………………………………………………………..69

3-4 Results………………………………………………………………………………………...72

3-5 Discussion…………………………………………………………………………………...90

3-6 Supplementary information………………………………………………………..105

3-7 References………………………………………………………………………………....112

4. THE MOST COMMON CYSTIC FIBROSIS ASSOCIATED MUTATION

DESTABILIZES THE DIMERIC STATE OF THE NUCLEOTIDE-BINDING

DOMAINS OF CFTR

4-1. Abstract…………………………………………………………..………………………...117

4-2. Introduction……………………………………………………..……………………….119

4-3 Material and Methods…………………………………………..……………............122

4-4 Results…………………………………………………..…………………………………..125

4-5 Discussion…………………………………………..……………………………………..141

4-6 Supplementary information………..………………………………………………148

4-7 References…………………………………………………………………………………151

5. FUTURE DIRECTIONS

5-1 Overview…………………………………………………………………………………...156

Page 7: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

vi

5-2 Unraveling the mechanism of CFTR potentiator, Vx-770………….…..158

5-3 Potential impacts of the energetic coupling model in CF

gene therapy…………………………………………………………………………….165

5-4 Roles of positive charge in the transmembrane domains…………….168

5-5 References………………………………………………………………………………..170

VITA…………………………………………………………………………………………………………....173

Page 8: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

vii

LIST OF FIGURES

Figure Page

1-1. The prevailing gating model for CFTR before 2012…………………………………...7

2-1. A hypothetic model of CFTR gated by ATP………………………………………………22

2-2. PPi captures a short-lived, post-hydrolytic state……………………………………...29

2-3. Identification of a post-hydrolytic state by AMP-PNP……………………………....33

2-4. ADP competes with PPi or AMP-PNP for state X………………………………………35

2-5. Differential modulation of the C2 state and state X………………………………….38

2-6. Single-channel ligand exchange for W401F-CFTR…………………………………..40

2-7. The mean open time of W401F-CFTR is [ATP] dependent………………………44

2-8. A revised CFTR gating scheme showing hypothetical conformational

transitions that takes place during an opening burst………………………………51

2-S1. Single-channel ligand exchange for WT-CFTR………………………………………...55

2-S2. PPi locked open WT-CFTR 5 s after ATP washout…………………………………...57

2-S3. Open time histograms for WT- and W401F-CFTR…………………………………..58

2-S4. Kinetic model and parameters for computer simulation in Figure 2-2C…...59

3-1. Cysless/R352C-CFTR reveals two different open states with distinct

conductance level…………………………………………………………….…………………….73

3-2. Hydrolysis triggers the O1 → O2 transition…………………………………………….78

3-3. Non-strict coupling between ATP hydrolysis and gating cycle…………………82

3-4. Modulation of the re-entry frequency by stabilizing the O2 state……….…….86

3-5. A new gating model for CFTR gating……………………………………………….………91

Page 9: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

viii

3-6. R352C shorten the locked-open time of hydrolytic deficient CFTR

mutant…………………………………………………………………………………………………..97

3-S1 The interburst dwell time histogram of R352C-CFTR……………………………108

3-S2 Dwell time histogram for O1, O2 state and opening burst in

Cysless/R352C- and R352C-CFTR………………………………………………………..109

3-S3 Gating kinetics for W401F/R352C-CFTR……………………………………………...110

4-1. Difference in locked-open time induced by PPi in WT-CFTR and

ΔF508-CFTR………………………………………………………………………………………………….127

4-2. Comparison of the locked-open time of E1371S- and

∆F508/E1371S-CFTR……………………………………………………………………………………..128

4-3. Increase of the locked-open time for ΔF508-CFTR channels by

gain-of-function mutations or the high affinity ATP analog P-

ATP………………………………………………………………………………………………….….130

4-4. ATP/P-ATP ligand exchange for WT-CFTR and ΔF508-CFTR………………………..132

4-5. Effects of “solubilizing mutations”, F494N/Q637R, on WT- and

ΔF508-CFTR channels……………………………………………………….……………………137

4-6. Effects of deletion of the regulatory insertion (ΔRI) on WT- and

ΔF508-CFTR channels……………………………………………………….…………………...138

4-7. Summary of effects of solubilizing mutations or ΔRI on locked-open

time and the time constant of the slow phase current rise upon

ATP/P-ATP ligand exchange…………………………………………………………………………..139

4-8. Effect of P-dATP on F494N/Q637R/ ΔF508-CFTR and ΔRI/ΔF508-CFTR……..….140

4-S1. Properties of ΔF508-CFTR expressed in CFPAC-1 cells………………………………...148

4-S2. Western blot analysis for WT, ∆F508, ∆RI/∆F508 and ∆F508/Sol

Page 10: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

ix

(∆F508/F494N/Q637R) ……………………………………………………………………….…149

5-1. Vx-770 increases ATP-independent activity in WT-CFTR………………………161

5-2. [ATP]-dependent mean open time of WT-CFTR in the presence

of Vx-770……………………………………………………………………………………………..162

5-3. Representative traces of R352C-CFTR in the presence of Vx-770…………..163

5-4. MTSET modification dramatically increases the current of

Cysless/I344C/∆NBD2-CFTR……………………………………………………….………167

Page 11: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

x

LIST OF TABLES

Table Page

3-1. Summary of opening events by different gating patterns in three

CFTR mutants ………………………………………………………………………………………………….76

5-1. Summary of different types of opening events for R352C-CFTR in the

presence of 200 nM Vx-770 …………………………………………………………………………….166

Page 12: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

xi

MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR

CHLORIDE CHANNEL

Kang-Yang Jih

Dr. Tzyh-Chang Hwang, Dissertation Advisor

Abstract

Cystic fibrosis transmembrane conductance regulator (CFTR) is the only ATP

binding cassette (ABC) protein that functions as an ion channel. The clinical

importance of CFTR lies in the fact that its malfunction causes the lethal genetic

disease, cystic fibrosis (CF). Like other ABC proteins, CFTR contains four canonical

domains— two transmembrane domains (TMDs) that form the ion permeation

pathway and two nucleotide binding domains (NBDs) that utilize the free energy

from ATP-hydrolysis to drive the gating cycle. The prevailing model in the field

dictates that ATP hydrolysis and the gating cycle are strictly coupled. In this study,

we have identified a post-hydrolytic state by using non-hydrolytic ATP analogs as

baits. As this state may accommodate an ATP molecule to initiate another hydrolysis

reaction without the necessity of gate closure, a non-integral stoichiometry between

ATP hydrolysis and gating cycle is posited. This hypothesis was reaffirmed by

studying a mutant CFTR that allows us to visualize ATP hydrolysis. Based on our

new findings, we proposed an energetic coupling model for CFTR gating.

Importantly, understanding the gating mechanism of CFTR may help us to decipher

the pathogenesis of CF at a molecular level. We discovered that the most common

CF-associated mutant, ∆F508-CFTR, destabilizes the NBD dimer, which likely links

to its low open probability. Furthermore, the energetic coupling model could explain

the action of the now clinically applied CF potentiator, Vx-770 (Kalydeco).

Page 13: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

1

CHAPTER 1

INTRODUCTION

1-1 Overview and clinical significance of CFTR

Cystic fibrosis (CF) is one of the most common lethal genetic disorder among

Caucasians. Although CF was first described in the 1930s (Andersen, 1938), the

cause of this disease remained unclear until the discovery of the CFTR gene in the

late 1980s (Riordan et al., 1989). To date, more than 1600 mutations in CFTR have

been identified in CF patients (www.genet.sickkids.on.ca/cftr/app). Cystic fibrosis is

a systemic disease that affects multiple organs and systems, including the

respiratory track, pancreas, gastrointestinal track and reproductive system.

(Knowles et al., 1983; Rosenstein and Cutting, 1998; Rowe et al., 2005). For patients

with CF, the defective function of CFTR results in reduced chloride permeability

across the apical membrane of the epithelial cells. In sweat glands, malfunction of

CFTR impairs chloride reabsorption (Gadsby et al., 2009), resulting in an elevated

sweat [chloride], which is considered the gold standard for CF diagnosis. Impaired

CFTR also causes thickening of the mucus in respiratory tracts (Engelhardt et al.,

1992), an ideal condition for bacteria growth. Subsequent repeated inflammation,

infection and pulmonary exacerbation eventually lead to respiratory failure, the

leading cause of death in CF patient (Rosenstein and Cutting, 1998; Rowe et al.,

2005). Since the discovery of the CFTR gene, tremendous efforts have been

Page 14: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

2

dovetailed to investigate the normal function of CFTR as well as how different

mutations cause CFTR dysfunction at a cellular and molecular level with the hope

that knowledge gained from these studies can provide clues for designing new drugs

to treat patients with CF.

In another aspect of the broad spectrum of research, CFTR attracts much

attention because of its unique identity in the phylogenetic classification. CFTR is a

member of the ABC Protein Superfamily (Riordan et al., 1989), an ancient protein

family that encompasses mostly active transporters (Dean and Annilo, 2005). CFTR

is the only ABC protein that functions as an ion channel (Bear et al., 1992). As recent

structure/function studies have yielded results that blur the boundary between

transporters and ion channels (Chen and Hwang, 2008; Gadsby et al., 2009; Feng et

al., 2010), insights from studying CFTR function can be extended to illuminate the

difference and similarities between channels and transporters at large. Despite

numerous high-resolution crystallographic structures providing snapshots of the

overall structures of ABC transporters in different conformational states (Dawson

and Locher, 2006; Hollenstein et al., 2007; Pinkett et al., 2007; Aller et al., 2009;

Oldham and Chen, 2011), very little is known about how these states shuttle

between each other. Contrary to transporters, ion channels can be scrutinized by

electrophysiological techniques that provide exceptional temporal resolution at the

single molecule level. Thus, biophysical studies of CFTR channels nicely fill the void

of our relatively poor understanding in the molecular mechanisms underlying

transporter kinetics.

Page 15: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

3

1-2 Structure and functions of CFTR

CFTR, a bona-fide member of ABC Protein Superfamily, contains the four

canonical domains found in most other ABC proteins— two transmembrane

domains (TMDs) that form the ion permeation pathway and two nucleotide binding

domains (NBDs) that harvest free energy from ATP hydrolysis to drive the gating

cycle (see schemes in Fig. 1). In addition, CFTR holds a unique regulatory domain (R

domain) that is not found in any other ABC proteins. The R domain contains

multiple serine and threonine residues that can be phosphorylated by protein

kinase A (PKA). The R domain is known to inhibit channel activity in the

unphophorylated state and release it inhibition after PKA phosphorylation (Rich et

al., 1991; Hwang et al., 1994; Mense et al., 2006; Wang et al., 2010).

Although there is no doubt that the two TMDs form the ion permeation pathway,

the pore architecture of CFTR remains poorly understood. The conformation of

TMDs and its motion during gating are of particular interest as this line of

information may offer insights into how a transporter can evolve to an ion channel.

For ABC transporters, a drastic conformational change in their TMDs has been

noticed by comparing the crystal structures solved under different conditions. It has

been postulated that gating of CFTR involves a similar scale of conformational

changes (Chen and Hwang, 2008; Jordan et al., 2008; Gadsby, 2009). By using the

substituted cysteine accessibility method (SCAM), recent studies identified the pore-

lining residues in the transmembrame helixes (TM) 1, 6, 12 (Bai et al., 2010; El Hiani

and Linsdell, 2010; Bai et al., 2011; Wang et al., 2011), which serve as the stepping

Page 16: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

4

stone for more detailed understanding of the pore architecture. Furthermore, by

demonstrating significant state-dependent accessibility to the thiol reagents for

some residues in TM6 and 12, Bai and his colleagues provided solid evidence for a

large-scale conformational change in TM 6 and 12 during gating (Bai et al., 2010,

2011) that is consistent with what has been proposed for ABC transporters. Another

mystery in the CFTR pore is the nature of the gate, or gates. For transporters, they

require at least two gates to function efficiently (Gadsby, 2009) whereas one gate is

sufficient for ion channels. The prevailing hypothesis posits that the cytoplasmic

gate was degenerated during protein evolution so that only one external gate exists

to regulate ionic flow in CFTR (Chen and Hwang, 2008; Jordan et al., 2008; Gadsby,

2009). This degraded transporter hypothesis for CFTR remained speculative until a

recent study showed that residues deep inside the pore can be readily accessed by

thiol reagents applied from the intracellular side in both the open and closed states

(Bai et al., 2011), a critical piece of evidence that supports the nonexistence of a

cytoplasmic gate.

The NBD is considered the engine of ABC proteins. It utilizes the free energy of

ATP hydrolysis to drive the transport/gating cycle. For CFTR, it is generally

accepted that the formation of a NBD dimer upon ATP binding is required for gate

opening (Vergani et al., 2005) and ATP hydrolysis in the ATP binding pocket (ABP) 2

triggers gate closing (Vergani et al., 2003; Bompadre et al., 2005). Notably, the NBD

dimer in CFTR as well as in other members of the ABCC subfamily reveals an

asymmetric catalytic ability: only one of the two ATP binding sites is capable of

hydrolyzing ATP. As a result, after NBD dimer formation, only the ATP molecule in

Page 17: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

5

the catalytic site, or ABP2, will be hydrolyzed. For CFTR, the asymmetric catalysis of

ATP hydrolysis results in a partial separation of the NBD dimer in which the head of

NBD1 and the tail of NBD2 remain attached (ABP1) with one ATP sandwiched in

between while the opposite side (ABP2, formed by the head of NBD2 and the tail of

NBD1) of the NBD dimer is detached and the ABP2 is vacated (Tsai et al., 2010).

That partial separation of the NBD dimer leads to the gate closure suggests that a

complete separation of the NBD dimer is not required for closing the gate (Tsai et al.,

2010). Surprisingly, such closed state with a partial NBD dimer is extremely stable

with a dwell time of ~ 30 s before the ATP in ABP1 eventually dissociates and the

NBDs are completely separated (Tsai et al., 2010). The aforementioned gating

mechanism of CFTR is summarized in Figure 1-1. To distinguish the two closed

states, the one with a partial NBD dimer was named the C2 state and the one with

two completely separated NBDs was named the C1 state (Tsai et al., 2009; Tsai et al.,

2010).

1-3 Coupling between ATP hydrolysis and gating

Numerous previous reports unequivocally concluded that ATP hydrolysis

assumes a critical role in facilitating gate closure as the open time is drastically

prolonged in hydrolysis deficient mutant channels (Carson et al., 1995; Zeltwanger

et al., 1999; Ikuma and Welsh, 2000; Vergani et al., 2003; Bompadre et al., 2005).

But it is not clear how the free energy from ATP hydrolysis contributes to such

open-closed transition. Unlike a transport cycle, in which the free energy input is

Page 18: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

6

strictly required for driving at least one irreversible reaction in order to pump the

cargo against the concentration gradient, gating of almost all ion channels is

considered an equilibrium process. As a result, it was proposed that for CFTR the

free energy harvested from ATP hydrolysis simply shifts the thermodynamic

equilibrium to favor the closed state but the transition between the open and closed

state is still in equilibrium. This was known as the reversible model (Aleksandrov et

al., 2000; Aleksandrov et al., 2002; Aleksandrov et al., 2009). On the contrary, the

irreversible model proposed that the ATP hydrolysis evokes an irreversible

structural change that inevitably leads to the closed state (Csanady et al., 2006;

Csanady, 2009; Csanady et al., 2010). This decade-long debate was tentatively

settled recently. By analyzing the open dwell time histogram, Csanady and his

colleagues (Csanady et al., 2010) showed a bi-modal distribution of the open time

histogram with a paucity of short opening events, supporting that closing of the gate

involves at least two steps, presumably ATP hydrolysis and subsequent gate closure.

This latest report effectively disputes the reversible model wherein the closing

process is simply the backward reaction of the one-step opening process.

Fitting their open dwell time distribution with a simple four-state model,

Csanady et al. (2010) also concluded that the ATP hydrolysis cycle and the gating

cycle are strictly coupled. This strict coupling hypothesis was also supported by

many previous studies. First, it has been shown that hydrolysis-deficient CFTR

mutants and non-hydrolysable ATP analogs significantly delay channel closing

(Gunderson and Kopito, 1994; Hwang et al., 1994; Carson et al., 1995; Vergani et al.,

Page 19: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

7

Figure 1-1

Figure 1-1. The prevailing gating model for CFTR before 2012. Before applying

ATP, the channel resides in a closed state that has its two NBDs separated (C1 state).

Upon ATP application, two ATPs occupy the two ATP binding sites (ABP1, ABP2) in

the NBDs and induces channel opening (O state). Once opened, the one ATP in the

ABP2 will be hydrolyzed and released, but the other ATP in the ABP1 remains

bound. ATP hydrolysis also triggers a partial separation of the NBD dimer and leads

to gate closure (C2 state). The C2 state is extremely stable with a constant of 30 s

before returning to the C1 state. Therefore, in the continuous presence of ATP, the

channel rarely visits the C1 state.

NBD1

NBD2

TMDs

ATP ADP + Pi

C2 C2'

OO'

hydrolysis

ABP2

ABP1

C1

Page 20: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

8

2003; Bompadre et al., 2005). Second, asymmetrical gating transitions suggested

that the gating cycle is driven by ATP hydrolysis (Gunderson and Kopito, 1995).

Third, by comparing WT and hydrolysis-deficient mutant CFTR, Csanady et al. (2010)

were able to conclude that for WT-CFTR, most closing events are the consequence of

ATP hydrolysis. Nonetheless, CFTR can function fairly well under an equilibrium

condition when ATP hydrolysis is abolished. Hydrolytic deficient mutants such as

K1250R-, D1370N-CFTR reamin ATP responsive and their Po is even higher than

that of WT-CFTR. So how exactly does ATP hydrolysis affect gating? Does it triggers

an irreversible conformational change by re-directing the channel to another

pathway that is fundamentally different from that in equilibrium gating, like that

proposed by previously Csanady et al. (Csanady et al., 2010)? Or does it simply offer

a fast track that accelerates the closing process but the gating motion involved in

such pathway is similar to the equilibrium pathway in general? It is hard to tackle

this question because ATP hydrolysis is not directly observed in

electrophysiological recordings. While microscopic kinetic studies offer a simple

and elegant approach to approach this issue, the limited resolution in detecting

transient states and rapid transitions between states also make this approach

vulnerable to potential technical faults. In Chapter 2, I will show my work in

identifying a post-hydrolytic state in CFTR’s gating cycle that shakes the foundation

of the prevailing strict coupling idea. Later in Chapter 3, by directly visualizing ATP

hydrolysis within each opening burst, I will show evidence that led us to propose a

new energetic coupling model to better illustrate CFTR gating. The essence of such

energetic coupling model is that the CFTR gating cycle is basically an equilibrium

Page 21: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

9

process that obeys microscopic reversibility, the role of ATP hydrolysis simply

provides a shortcut that allows the channel to bypass an otherwise slow transition

in order to accelerate gate closing.

1-4 Fishing kinetic states with non-hydrolytic ATP analogs

It had been known for more than a decade that non-hydrolytic ATP analogs, such as

pyrophosphate (PPi) and Adenylyl-imidodiphosphate (AMP-PNP), when applied

together with ATP, lead to opening events that can last for tens of seconds as if the

channel is locked in the open state (Gunderson and Kopito, 1994; Hwang et al., 1994;

Carson et al., 1995). Such locked-opening events are similar to that seen in the

hydrolysis deficient CFTR (Vergani et al., 2003; Bompadre et al., 2005). However,

when applied alone, PPi or AMP-PNP elicits very little current (Gunderson and

Kopito, 1994; Hwang et al., 1994; Carson et al., 1995).

Interestingly, my colleague (Tsai et al., 2009) found that if PPi is applied alone

soon after ATP washout, it is still capable of lock-opening the CFTR channel with

considerably high efficacy. However, such a robust effect of PPi vanished if the same

concentration of PPi was applied minutes after ATP washout. These results suggest

that ATP can somehow prime the CFTR channel for better responsiveness to

subsequent PPi application, as if the channels can “remember” that it has been

treated by ATP. In terms of channel kinetics, priming by the ATP implies there are at

least two different closed states that are distinguishable by different responsiveness

to PPi: a primed state that still “remember” being treated by ATP thus responding to

Page 22: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

10

PPi robustly and a non-primed state that already “forget” being activated by ATP so

that it respond to PPi poorly. The configuration of the two closed states was further

determined by ligand exchange experiments that showed rapid ligand substitution

in ABP2 whereas the ligand in ABP1 remained trapped for ~ 50 s. (Tsai et al., 2010).

It was concluded that the ATP-primed closed state that emerges immediately after

ATP washout bears a partial NBD dimer with one ATP bound in ABP1 (the C2 state

in Figure 1-1). Following the C2 state, the poorly PPi responsive closed state bears

two completely separated NBDs with no ATP bound (the C1 state in Figure 1-1).

This study demonstrated a nice example of fishing out “invisible” functional states

by applying non-hydrolytic ATP analogs such as PPi as baits at different time point

throughout the gating cycle. By using the similar strategy, I identified a post-

hydrolytic state. And surprisingly, identification of such inconspicuous state had

shaken the widely accepted strict coupling gating model. This series of work will be

presented in Chapter 2 in detail.

Besides serving as baits to catch invisible states, non-hydrolytic ATP analogs

also help decipher functional perturbations caused by disease related mutations.

For example, it is generally accepted that a full NBD dimer is required for opening

the channel; regardless of whether the channel is opened by ATP or other ATP-like

analogs. Thus these non-hydrolytic ATP analogs allow us to probe, despite indirectly,

the stability of NBD dimer by measuring the current relaxation time constant of the

PPi or AMP-PNP locked-open channel and to quantitatively examine effects of

mutations on the NBD dimer. In Chapter 4, I will elaborate how the dimer stability is

affected by the most common CF-associated mutation, ∆F508.

Page 23: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

11

1-5 CFTR pharmacology

Understanding the fundamental mechanism of CFTR gating is more than just

satisfying intellectual curiosity; it could serve practical purposes in guiding us to

develop new strategies to improve the function of many common CF related

mutations, including the most common ∆F508 and the third most common G551D

mutations, both assuming a much reduced open probability (Dalemans et al., 1991;

Haws et al., 1996; Hwang et al., 1997; Bompadre et al., 2007; Ostedgaard et al.,

2007). As a result, searching for CFTR potentiators has become a rational and

realistic approach in the pursuit of controling CF. Many small molecules had been

identified as CFTR potentiators in the past decade (Hwang et al., 1997; Ai et al., 2004;

Berger et al., 2005; Zhou et al., 2005; Van Goor et al., 2006; Yu et al., 2011); however,

few of them have come even close to clinical application.

Recently, there was a significant breakthrough in CFTR pharmacology. A small

molecule CFTR potentiator, Vx-770 (Kalydeco), was identified by high-throughput

drug screening assays (Van Goor et al., 2009). It exhibits extremely high affinity (Kd

= ~ 50 nM) for CFTR and considerably high efficacy (> 10 fold current increase in

G551D-CFTR), making it a suitable candidate for clinical use. Consistent with in vitro

studies, subsequent clinical trials showed promising outcomes (Accurso et al., 2010;

Ramsey et al., 2011) and FDA has recently approved its clinical use in treating CF

patients carrying the G551D mutation. Although Vx-770 is being prescribed to CF

patient every day, very little is known about its mechanism of action.

Previous studies suggested that Vx-770 potentiates CFTR activity by extending

the duration of opening burst in both WT and many disease-related mutant

Page 24: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

12

channels, such as G551D (Van Goor et al., 2009; Yu et al., 2012). According to the

traditional strict coupling model (Figure 1-1), there are only two points of

intervention for prolonging the open time, namely slowing down ATP hydrolysis (O

→ O’) or slowing down the post-hydrolysis NBD dimer separation (O’ → C2), but

none of these can offer a suitable explanation for the effect of Vx-770 on G551D

since ATP-induced NBD dimerization probably is not present in this mutant channel

(Bompadre et al., 2008). On the other hand, the energetic coupling model proposed

in Chapter 3 can provide a rational and simple explanation for how Vx-770 functions

in both WT- and G551D-CFTR (see details in Chapter 3).

The importance of elucidating how Vx-770 works lies in the fact that currently

only patients carrying the G551D mutation, which accounts for less than 5% of total

CF cases, are benefiting from this new drug. The mechanism by which Vx-770 exerts

its effect could provide a pivotal knowledge for improving the existing Vx-770

compound or perhaps for designing new drugs. Furthermore, according to the

clinical trials (Accurso et al., 2010; Ramsey et al., 2011), Vx-770 does not actually

cure CF, it simply attenuate its symptom. Therefore it remains an urgent task to

identify drugs with higher potency or drugs that complement Vx-770 in order to

reach the ultimate goal of curing CF. In Chapter 5, I will discuss our preliminary

findings about the mechanism of Vx-770 and propose related projects that worth

investing.

Page 25: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

13

1-6 References

Accurso, F.J., S.M. Rowe, J.P. Clancy, M.P. Boyle, J.M. Dunitz, P.R. Durie, S.D. Sagel, D.B.

Hornick, M.W. Konstan, S.H. Donaldson, R.B. Moss, J.M. Pilewski, R.C.

Rubenstein, A.Z. Uluer, M.L. Aitken, S.D. Freedman, L.M. Rose, N. Mayer-

Hamblett, Q. Dong, J. Zha, A.J. Stone, E.R. Olson, C.L. Ordonez, P.W. Campbell,

M.A. Ashlock, and B.W. Ramsey. 2010. Effect of VX-770 in persons with cystic

fibrosis and the G551D-CFTR mutation. N Engl J Med. 363:1991-2003.

Ai, T., S.G. Bompadre, Y. Sohma, X. Wang, M. Li, and T.C. Hwang. 2004. Direct effects

of 9-anthracene compounds on cystic fibrosis transmembrane conductance

regulator gating. Pflugers Arch. 449:88-95.

Aleksandrov, A.A., L. Aleksandrov, and J.R. Riordan. 2002. Nucleoside triphosphate

pentose ring impact on CFTR gating and hydrolysis. FEBS Lett. 518:183-188.

Aleksandrov, A.A., X. Chang, L. Aleksandrov, and J.R. Riordan. 2000. The non-

hydrolytic pathway of cystic fibrosis transmembrane conductance regulator

ion channel gating. J Physiol. 528 Pt 2:259-265.

Aleksandrov, A.A., L. Cui, and J.R. Riordan. 2009. Relationship between nucleotide

binding and ion channel gating in cystic fibrosis transmembrane conductance

regulator. J Physiol. 587:2875-2886.

Aller, S.G., J. Yu, A. Ward, Y. Weng, S. Chittaboina, R. Zhuo, P.M. Harrell, Y.T. Trinh, Q.

Zhang, I.L. Urbatsch, and G. Chang. 2009. Structure of P-glycoprotein reveals

a molecular basis for poly-specific drug binding. Science. 323:1718-1722.

Andersen, D.H. 1938. CYSTIC FIBROSIS OF THE PANCREAS AND ITS RELATION TO

CELIAC DISEASE. Am J Dis Child. 56:334-339.

Bai, Y., M. Li, and T.C. Hwang. 2010. Dual roles of the sixth transmembrane segment

of the CFTR chloride channel in gating and permeation. J Gen Physiol.

136:293-309.

Bai, Y., M. Li, and T.C. Hwang. 2011. Structural basis for the channel function of a

degraded ABC transporter, CFTR (ABCC7). J Gen Physiol. 138:495-507.

Bear, C.E., C.H. Li, N. Kartner, R.J. Bridges, T.J. Jensen, M. Ramjeesingh, and J.R.

Riordan. 1992. Purification and functional reconstitution of the cystic fibrosis

transmembrane conductance regulator (CFTR). Cell. 68:809-818.

Berger, A.L., C.O. Randak, L.S. Ostedgaard, P.H. Karp, D.W. Vermeer, and M.J. Welsh.

2005. Curcumin stimulates cystic fibrosis transmembrane conductance

regulator Cl- channel activity. J Biol Chem. 280:5221-5226.

Page 26: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

14

Bompadre, S.G., J.H. Cho, X. Wang, X. Zou, Y. Sohma, M. Li, and T.C. Hwang. 2005.

CFTR gating II: Effects of nucleotide binding on the stability of open states. J

Gen Physiol. 125:377-394.

Bompadre, S.G., M. Li, and T.C. Hwang. 2008. Mechanism of G551D-CFTR (cystic

fibrosis transmembrane conductance regulator) potentiation by a high

affinity ATP analog. J Biol Chem. 283:5364-5369.

Bompadre, S.G., Y. Sohma, M. Li, and T.C. Hwang. 2007. G551D and G1349D, two CF-

associated mutations in the signature sequences of CFTR, exhibit distinct

gating defects. J Gen Physiol. 129:285-298.

Carson, M.R., S.M. Travis, and M.J. Welsh. 1995. The two nucleotide-binding domains

of cystic fibrosis transmembrane conductance regulator (CFTR) have distinct

functions in controlling channel activity. J Biol Chem. 270:1711-1717.

Chen, T.Y., and T.C. Hwang. 2008. CLC-0 and CFTR: chloride channels evolved from

transporters. Physiol Rev. 88:351-387.

Csanady, L. 2009. Application of rate-equilibrium free energy relationship analysis

to nonequilibrium ion channel gating mechanisms. J Gen Physiol. 134:129-

136.

Csanady, L., A.C. Nairn, and D.C. Gadsby. 2006. Thermodynamics of CFTR channel

gating: a spreading conformational change initiates an irreversible gating

cycle. J Gen Physiol. 128:523-533.

Csanady, L., P. Vergani, and D.C. Gadsby. 2010. Strict coupling between CFTR's

catalytic cycle and gating of its Cl- ion pore revealed by distributions of open

channel burst durations. Proc Natl Acad Sci U S A. 107:1241-1246.

Dalemans, W., P. Barbry, G. Champigny, S. Jallat, K. Dott, D. Dreyer, R.G. Crystal, A.

Pavirani, J.P. Lecocq, and M. Lazdunski. 1991. Altered chloride ion channel

kinetics associated with the delta F508 cystic fibrosis mutation. Nature.

354:526-528.

Dawson, R.J., and K.P. Locher. 2006. Structure of a bacterial multidrug ABC

transporter. Nature. 443:180-185.

Dean, M., and T. Annilo. 2005. Evolution of the ATP-binding cassette (ABC)

transporter superfamily in vertebrates. Annual review of genomics and

human genetics. 6:123-142.

El Hiani, Y., and P. Linsdell. 2010. Changes in accessibility of cytoplasmic substances

to the pore associated with activation of the cystic fibrosis transmembrane

conductance regulator chloride channel. J Biol Chem. 285:32126-32140.

Page 27: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

15

Engelhardt, J.F., J.R. Yankaskas, S.A. Ernst, Y. Yang, C.R. Marino, R.C. Boucher, J.A.

Cohn, and J.M. Wilson. 1992. Submucosal glands are the predominant site of

CFTR expression in the human bronchus. Nature genetics. 2:240-248.

Feng, L., E.B. Campbell, Y. Hsiung, and R. MacKinnon. 2010. Structure of a eukaryotic

CLC transporter defines an intermediate state in the transport cycle. Science.

330:635-641.

Gadsby, D.C. 2009. Ion channels versus ion pumps: the principal difference, in

principle. Nature reviews. Molecular cell biology. 10:344-352.

Gadsby, D.C., A. Takeuchi, P. Artigas, and N. Reyes. 2009. Review. Peering into an

ATPase ion pump with single-channel recordings. Philosophical transactions

of the Royal Society of London. Series B, Biological sciences. 364:229-238.

Gunderson, K.L., and R.R. Kopito. 1994. Effects of pyrophosphate and nucleotide

analogs suggest a role for ATP hydrolysis in cystic fibrosis transmembrane

regulator channel gating. J Biol Chem. 269:19349-19353.

Gunderson, K.L., and R.R. Kopito. 1995. Conformational states of CFTR associated

with channel gating: the role ATP binding and hydrolysis. Cell. 82:231-239.

Haws, C.M., I.B. Nepomuceno, M.E. Krouse, H. Wakelee, T. Law, Y. Xia, H. Nguyen, and

J.J. Wine. 1996. Delta F508-CFTR channels: kinetics, activation by forskolin,

and potentiation by xanthines. Am J Physiol. 270:C1544-1555.

Hollenstein, K., D.C. Frei, and K.P. Locher. 2007. Structure of an ABC transporter in

complex with its binding protein. Nature. 446:213-216.

Hwang, T.C., G. Nagel, A.C. Nairn, and D.C. Gadsby. 1994. Regulation of the gating of

cystic fibrosis transmembrane conductance regulator C1 channels by

phosphorylation and ATP hydrolysis. Proc Natl Acad Sci U S A. 91:4698-4702.

Hwang, T.C., F. Wang, I.C. Yang, and W.W. Reenstra. 1997. Genistein potentiates

wild-type and delta F508-CFTR channel activity. Am J Physiol. 273:C988-998.

Ikuma, M., and M.J. Welsh. 2000. Regulation of CFTR Cl- channel gating by ATP

binding and hydrolysis. Proc Natl Acad Sci U S A. 97:8675-8680.

Jordan, I.K., K.C. Kota, G. Cui, C.H. Thompson, and N.A. McCarty. 2008. Evolutionary

and functional divergence between the cystic fibrosis transmembrane

conductance regulator and related ATP-binding cassette transporters. Proc

Natl Acad Sci U S A. 105:18865-18870.

Knowles, M.R., M.J. Stutts, A. Spock, N. Fischer, J.T. Gatzy, and R.C. Boucher. 1983.

Abnormal ion permeation through cystic fibrosis respiratory epithelium.

Science. 221:1067-1070.

Page 28: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

16

Mense, M., P. Vergani, D.M. White, G. Altberg, A.C. Nairn, and D.C. Gadsby. 2006. In

vivo phosphorylation of CFTR promotes formation of a nucleotide-binding

domain heterodimer. EMBO J. 25:4728-4739.

Oldham, M.L., and J. Chen. 2011. Crystal structure of the maltose transporter in a

pretranslocation intermediate state. Science. 332:1202-1205.

Ostedgaard, L.S., C.S. Rogers, Q. Dong, C.O. Randak, D.W. Vermeer, T. Rokhlina, P.H.

Karp, and M.J. Welsh. 2007. Processing and function of CFTR-DeltaF508 are

species-dependent. Proc Natl Acad Sci U S A. 104:15370-15375.

Pinkett, H.W., A.T. Lee, P. Lum, K.P. Locher, and D.C. Rees. 2007. An inward-facing

conformation of a putative metal-chelate-type ABC transporter. Science.

315:373-377.

Ramsey, B.W., J. Davies, N.G. McElvaney, E. Tullis, S.C. Bell, P. Drevinek, M. Griese, E.F.

McKone, C.E. Wainwright, M.W. Konstan, R. Moss, F. Ratjen, I. Sermet-

Gaudelus, S.M. Rowe, Q. Dong, S. Rodriguez, K. Yen, C. Ordonez, and J.S. Elborn.

2011. A CFTR potentiator in patients with cystic fibrosis and the G551D

mutation. N Engl J Med. 365:1663-1672.

Rich, D.P., R.J. Gregory, M.P. Anderson, P. Manavalan, A.E. Smith, and M.J. Welsh.

1991. Effect of deleting the R domain on CFTR-generated chloride channels.

Science. 253:205-207.

Riordan, J.R., J.M. Rommens, B. Kerem, N. Alon, R. Rozmahel, Z. Grzelczak, J. Zielenski,

S. Lok, N. Plavsic, J.L. Chou, and et al. 1989. Identification of the cystic fibrosis

gene: cloning and characterization of complementary DNA. Science.

245:1066-1073.

Rosenstein, B.J., and G.R. Cutting. 1998. The diagnosis of cystic fibrosis: a consensus

statement. Cystic Fibrosis Foundation Consensus Panel. The Journal of

pediatrics. 132:589-595.

Rowe, S.M., S. Miller, and E.J. Sorscher. 2005. Cystic fibrosis. N Engl J Med. 352:1992-

2001.

Tsai, M.F., M. Li, and T.C. Hwang. 2010. Stable ATP binding mediated by a partial

NBD dimer of the CFTR chloride channel. J Gen Physiol. 135:399-414.

Tsai, M.F., H. Shimizu, Y. Sohma, M. Li, and T.C. Hwang. 2009. State-dependent

modulation of CFTR gating by pyrophosphate. J Gen Physiol. 133:405-419.

Van Goor, F., S. Hadida, P.D. Grootenhuis, B. Burton, D. Cao, T. Neuberger, A. Turnbull,

A. Singh, J. Joubran, A. Hazlewood, J. Zhou, J. McCartney, V. Arumugam, C.

Decker, J. Yang, C. Young, E.R. Olson, J.J. Wine, R.A. Frizzell, M. Ashlock, and P.

Page 29: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

17

Negulescu. 2009. Rescue of CF airway epithelial cell function in vitro by a

CFTR potentiator, VX-770. Proc Natl Acad Sci U S A. 106:18825-18830.

Van Goor, F., K.S. Straley, D. Cao, J. Gonzalez, S. Hadida, A. Hazlewood, J. Joubran, T.

Knapp, L.R. Makings, M. Miller, T. Neuberger, E. Olson, V. Panchenko, J. Rader,

A. Singh, J.H. Stack, R. Tung, P.D. Grootenhuis, and P. Negulescu. 2006. Rescue

of DeltaF508-CFTR trafficking and gating in human cystic fibrosis airway

primary cultures by small molecules. American journal of physiology. Lung

cellular and molecular physiology. 290:L1117-1130.

Vergani, P., S.W. Lockless, A.C. Nairn, and D.C. Gadsby. 2005. CFTR channel opening

by ATP-driven tight dimerization of its nucleotide-binding domains. Nature.

433:876-880.

Vergani, P., A.C. Nairn, and D.C. Gadsby. 2003. On the mechanism of MgATP-

dependent gating of CFTR Cl- channels. J Gen Physiol. 121:17-36.

Wang, W., Y. El Hiani, and P. Linsdell. 2011. Alignment of transmembrane regions in

the cystic fibrosis transmembrane conductance regulator chloride channel

pore. J Gen Physiol. 138:165-178.

Wang, W., J. Wu, K. Bernard, G. Li, G. Wang, M.O. Bevensee, and K.L. Kirk. 2010. ATP-

independent CFTR channel gating and allosteric modulation by

phosphorylation. Proc Natl Acad Sci U S A. 107:3888-3893.

Yu, H., B. Burton, C.J. Huang, J. Worley, D. Cao, J.P. Johnson, Jr., A. Urrutia, J. Joubran, S.

Seepersaud, K. Sussky, B.J. Hoffman, and F. Van Goor. 2012. Ivacaftor

potentiation of multiple CFTR channels with gating mutations. Journal of

cystic fibrosis : official journal of the European Cystic Fibrosis Society.

Yu, Y.C., H. Miki, Y. Nakamura, A. Hanyuda, Y. Matsuzaki, Y. Abe, M. Yasui, K. Tanaka,

T.C. Hwang, S.G. Bompadre, and Y. Sohma. 2011. Curcumin and genistein

additively potentiate G551D-CFTR. Journal of cystic fibrosis : official journal of

the European Cystic Fibrosis Society. 10:243-252.

Zeltwanger, S., F. Wang, G.T. Wang, K.D. Gillis, and T.C. Hwang. 1999. Gating of cystic

fibrosis transmembrane conductance regulator chloride channels by

adenosine triphosphate hydrolysis. Quantitative analysis of a cyclic gating

scheme. J Gen Physiol. 113:541-554.

Zhou, Z., X. Wang, M. Li, Y. Sohma, X. Zou, and T.C. Hwang. 2005. High affinity

ATP/ADP analogues as new tools for studying CFTR gating. J Physiol.

569:447-457.

Page 30: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

18

CHAPTER 2

IDENTIFICATION OF A NOVEL POST-

HYDROLYTIC STATE IN CFTR GATING

This chapter has been modified from my manuscript published in J. Gen Physiol.

139:359-370, by Kang-Yang Jih, Yoshiro Sohma, Min Li, and Tzyh-Chang Hwang.

According to their web site, http://jgp.rupress.org/misc/terms.shtml, I retain the

copyright for this work and am allowed to alter and build upon this work.

2-1. Abstract

ABC transporters, ubiquitous proteins found in all kingdoms of life, catalyze

substrates translocation across biological membranes using the free energy of ATP

hydrolysis. Cystic fibrosis transmembrane conductance regulator (CFTR) is a unique

member of this superfamily in that it functions as an ATP-gated chloride channel.

Despite different in function, recent studies suggest that the CFTR chloride channel

and the exporter members of the ABC Protein Family may share an evolutionary

origin. While ABC exporters harness the free energy of ATP hydrolysis to fuel a

transport cycle, for CFTR, ATP-induced dimerization of its nucleotide binding

domains (NBDs) and subsequent hydrolysis-triggered dimer separation are

proposed to be coupled respectively to the opening and closing of the gate in its

transmembrane domains (TMDs). In this study, by using non-hydrolyzable ATP

Page 31: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

19

analogs, such as pyrophosphate (PPi) or adenylyl-imidodiphosphate (AMP-PNP) as

baits, we captured a short-lived state (state X), which distinguishes itself from the

previously identified long-lived C2 closed state by its fast response to these

nonhydrolyzable ligands. As state X is caught during the decay phase of channel

closing upon washout of the ligand ATP but before the channel sojourns to the C2

closed state, it likely emerges after the bound ATP in the catalysis-competent site

has been hydrolyzed and the hydrolytic products have been released. Thus this

newly identified post-hydrolytic state may share a similar conformation of NBDs as

the C2 closed state (i.e., a partially separated NBDs and a vacated ATP binding

pocket). The significance of this novel state in understanding the structural basis of

CFTR gating is discussed.

2-2. Introduction

Cystic Fibrosis Transmembrane conductance Regulator (CFTR), the culprit

behind the fatal genetic disease cystic fibrosis (CF) (Riordan et al., 1989), is a

member of the ABC (ATP binding cassette) transporter superfamily whose members

exist throughout the biological universe. This family of integral membrane proteins

is characterized by an evolutionarily conserved topology consisting of two

transmembrane domains (TMDs) that form the cargo translocation pathway, and

two cytosolic nucleotide binding domains (NBDs) serving as engines to drive the

transport process. Recent structural and functional studies of ABC proteins have led

to a hypothesis that the formation and separation of an NBD dimer are coupled to

Page 32: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

20

the conformational changes in TMDs to complete a transport cycle (Vergani et al.,

2005; Dawson and Locher, 2006; Hollenstein et al., 2007; Ward et al., 2007; Khare et

al., 2009).

While nearly all ABC proteins assume the function of active transport, CFTR, a

bona fide member of this superfamily, is unique in that it is an ATP-gated chloride

channel (Bear et al., 1992). Nonetheless, numerous studies have indicated that CFTR

shares similar architecture and mechanism of action with other ABC transporters.

For example, the crystal structures of CFTR’s two NBDs are virtually

indistinguishable from those of other ABC transporters (Lewis et al., 2004 and #

3GD7, PDB). It has also been shown for CFTR that ATP binding induces the

formation of a canonical NBD dimer seen in ABC transporters (Vergani et al., 2005;

Mense et al., 2006). More recent studies (Bai et al., 2010, 2011) of CFTR’s gating

conformational changes in its TMDs provided evidence supporting the notion that

CFTR evolves from a primordial ABC exporter by simply removing its cytoplasmic

gate (i.e., degraded transporter hypothesis). Therefore, mechanistic studies of how

CFTR’s NBDs control gating transitions bear a broad implication as the insights

gained may help decipher the complex transport mechanism of all ABC proteins

(Csanady, 2010).

For all ABC proteins, ATP is the source of the free energy that drives the

transport cycle. Recent crystallographic studies provide snapshots of these proteins

in different conformations — an “inward-facing” (Ward et al., 2007; Aller et al., 2009)

configuration with separated NBDs and an “outward-facing” (Dawson and Locher,

Page 33: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

21

2006; Ward et al., 2007) configuration with dimerized NBDs. However, it remains

unclear about how they shuttle between the two states. One major impediment to

directly scrutinizing such conformational changes is the lack of techniques with high

temporal resolutions. On the other hand, CFTR, as an ion channel, fills the void. It is

generally accepted that the ATP-dependent NBD dimerization and separation

control the gate of CFTR. Since ATP interacts exclusively with the NBDs, the gating

signal upon ATP binding and subsequent NBD dimerization must be allosterically

transmitted to the TMDs to initiate the gating cycle. The current model for CFTR

gating (Figure 2-1) dictates a strict coupling between the gating cycle and ATP

hydrolysis cycle: opening of the channel into a burst and subsequent termination of

the burst are synchronized to the dimerization and separation of the NBD dimer.

This model derived mostly from single-channel kinetic studies—albeit with its

simplicity and elegance—is susceptible to potential technical faults. As the

transitions between different conformational states could be rapid and transient in

nature, classical kinetic analysis may not be sensitive enough to capture all relevant

states and thus fail to detect slight deviation from the proposed scheme.

In the current study, by using the rapid ligand-exchange protocols developed in

our laboratory (Tsai et al., 2009; Tsai et al., 2010b), we were able to capture a

transient state by locking the post-hydrolytic channel into a prolonged bursting

state (i.e., locked-open state) with non-hydrolyzable ATP analogs, such as

pyrophosphate (PPi) or adenylyl-imidodiphosphate (AMP-PNP) (Gunderson and

Kopito, 1994; Hwang et al., 1994). In patches yielding macroscopic CFTR currents,

when switching the ligand directly from ATP to PPi or AMP-PNP, we observed an

Page 34: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

22

Figure 2-1

Figure 2-1. A hypothetic model of CFTR gated by ATP. (A) A scheme illustrating

ATP-dependent gating mechanism of CFTR. This scheme is synthesized based on

several latest publications on CFTR gating (Csanady et al., 2010; Tsai et al., 2010b;

Szollosi et al., 2011). For channel opening, two NBDs dimerize upon ATP binding to

Page 35: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

23

the catalysis-competent site (ABP2) in the C2 state, which harbors a partially

dimerized NBD with an ATP molecule bound to the catalysis-incompetent site

(ABP1). For channel closure, first the ATP in ABP2 is hydrolyzed to convert the pre-

hydrolytic open state (O) to a post-hydrolytic open state (O’). Then, it is the

separation of the NBD dimer and the dissociation of hydrolytic products that

coincide with gate closure. In the continuous presence of millimolar ATP, CFTR

rarely shuttles back to the C1 state, which bears completely separated NBDs. (B) A

cartoon depicting how the proposed CFTR gating scheme in (A) correlates an

experimentally observed opening and closing of a single CFTR channel to the

molecular events in its NBDs and TMDs. The red box encompasses the open channel

conformations.

Page 36: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

24

unusual bi-phasic response to this experimental maneuver, suggesting the existence

of two distinct functional states with different responsiveness to these non-

hydrolyzable ligands: one (C2 closed state characterized previously in (Tsai et al.,

2009)) and the other, state X, with a faster response to PPi or AMP-PNP. These two

states can be further differentiated by using mutations or the high-affinity ATP

analog, N6-phenyethyl-ATP (P-ATP). In patches containing one single CFTR channel,

switching the ligand from ATP to a very brief (1-s) application of PPi in the open but

not the closed state resulted in a direct transition into a locked-open state without

transiting to a long interburst closure. These results suggest that state X, like the

previously characterized C2 closed state, bears a partially separated NBD dimer and

a vacated ATP binding site 2 (ABP2). Although more studies are needed to

determine whether this newly identified state is an open state or a short-lived

closed state, our results could provide novel insight into the structural basis of CFTR

gating and potentially impact our understanding of the functional mechanism for

other ABC proteins.

2-3. Material and Methods

Cell culture and transient expression system

PolyFect transfection reagent (QIAGEN) was used to cotransfect CFTR cDNA and

pEGFP-C3 (Clontech, Palo Alto, CA), encoding the green fluorescence protein into

Chinese hamster ovary (CHO) cells. CHO cells were grown at 37°C in Dulbecco’s

Modified Eagle’s Medium supplemented with 10% fetal bovine serum. One day

Page 37: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

25

before transfection, cells were trypsinized and cultured in 35-mm tissue culture

dishes. After transfection, cells were cultured at 27°C for at least 2 days before

electrophysiological experiments were performed.

Mutagenesis

All mutations were constructed by QuickChange XL kit (Stratagene, La Jolla, CA)

according to manufacturer’s protocols and then sequenced to confirm the mutation

(DNA core, University of Missouri) on cDNA.

Electrophysiological recordings

Glass chips carrying the transfected cells were transferred to a chamber located on

the stage of an inverted microscope (IX51, Olympus). Membrane patches were

excised into an inside-out mode after the seal resistance was > 40 GΩ. After excision,

the pipette was perfused by 25 IU PKA and 2.75 mM ATP until the CFTR current

reached a steady state, all other solutions containing ATP applied thereafter

contained 10 IU PKA to maintain the phosphorylation level. An EPC10 amplifier

(HEKA, Lambrecht/Pfalz, Germany) was used to record electrophysiological data at

room temperature at a -60 mV holding potential. The data were filtered on-line at

100 Hz with an eight-pole Bessel filter (LPF-8, Warner Instruments), and digitized

to a computer at a sampling rate of 500 Hz. The inward current was inverted for

clear data presentation. The resistance of pipettes for patch-clamp experiments was

2 to 4 MΩ in the bath solution. The pipettes were prepared from borosilicate

Page 38: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

26

capillary glass using a Flaming/Brown-type micropipette puller (P97, Sutter

Instrument Co.) and then polished with a homemade microforge. All inside-out

patch experiments were performed with a fast solution exchange perfusion system

(SF-77B, Warner Instruments). The dead time of solution change is ~30 ms (Tsai et

al., 2009).

Chemicals and composition solutions

The pipette solution contained (in mM): 140 methyl-D-glucamine chloride (NMDG-

Cl), 2 MgCl2, 5 CaCl2, and 10 HEPES (pH 7.4 with NMDG). Cells were perfused with a

bath solution containing (in mM): 145 NaCl, 5 KCl, 2 MgCl2, 1 CaCl2, 5 glucose, 5

HEPES and 20 sucrose (pH 7.4 with NaOH). For inside-out configuration, the

perfusion solution contained (in mM): 150 NMDG-Cl, 2 MgCl2, 10 EGTA and 8 Tris

(pH 7.4 with NMDG).

MgATP, PPi and PKA were purchased from Sigma-Aldrich. N6-(phenylethyl)-ATP (P-

ATP) was purchased from Biolog Life Science Institute. Adenylyl-imidodiphosphate

(AMP-PNP) was purchased from Roche Applied Science. PPi and MgATP were stored

in 200 mM and 250 mM stock solution respectively at -20°C. P-ATP was stored in 10

mM stock at -70°C. AMP-PNP was store at -70°C and prepared before used. All

chemicals were diluted to the concentration indicated in each figure using perfusion

Page 39: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

27

solution and the pH was adjusted to 7.4 with NMDG. For solutions containing AMP-

PNP or PPi, an equal concentration of MgCl2 was added to the solution.

Data analysis and statistics

Igor Pro program (Wavemetrics, Lake Oswego, OR) was used to calculate the

steady-state mean current amplitude and the current relaxation time constant.

Current relaxation was fitted with a single exponential function using a Levenberg-

Marquardt-based algorithm within the Igor Pro program. Channel kinetics was

analyzed with a program developed by Csanady (Csanady, 2000) on traces that

contain 3 or less channels. Kinetic modeling and computation simulations were

described in Kopeikin et al. (Kopeikin et al., 2010).

Results are shown as mean ± SEM. Student’s t-test was performed for statistical

analysis using Excel (Microsoft). P < 0.05 was considered statistically significant.

2-4. Results

Identification of a new post-hydrolytic state

Once phosphorylated by PKA, CFTR channels are opened by ATP; upon ATP

removal, CFTR will enter a relatively stable (dwell time = ~ 25 s) closed state (so-

called C2 state) that possesses a partial NBD dimer wherein the head of NBD2 and

the tail of NBD1 (i.e., ATP binding pocket (ABP) 2) are separated but the opposite

Page 40: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

28

side of the NBD dimer (ABP1) remain attached (Tsai et al., 2009; Tsai et al., 2010b;

Szollosi et al., 2011). Channels in the C2 state is characterized by its capability to be

locked-open by non-hydrolyzable ATP analogs such as PPi or AMP-PNP (Tsai et al.,

2009; Tsai et al., 2010b) (Figure 2-2 A and B); however, the lock-opening rate is

very slow (time constant of current rising, τ = 4.76 ± 0.42 s, n = 8, Figure 2-2 B,

inset). Since the closing rate of the CFTR channel estimated from the current

relaxation upon ATP removal (τrel = 380 ± 40 ms, n = 8, Figure 2-2 B) is much faster,

it is expected that if one switches the ligand directly from ATP to PPi, very few

channel can be locked open during the current decay phase. Intuitively speaking,

this is because the very short time (~1 s) of current decay upon removing ATP is

insufficient for a significant number of the closed channels to respond to PPi. Thus,

nearly all channels are expected to close first upon ligand switches before they start

to be opened by PPi. Indeed, when we performed computer simulations based on

the model of scenario 1 (Figure w-2 A) wherein the C2 state is the only state in the

gating cycle that can be locked-open by PPi, the simulated trace shows a decay of the

macroscopic current to near the baseline (to the C2 state) before a slowly rising

current is seen (Figure 2-2 C). In contrast, when we carried out such an experiment,

the result turned out aberrant from this prediction. As shown in Figure 2-2 D, upon

ligand switching, the currents plummeted immediately but stopped in the middle

rather than decay to the baseline, followed by a slow rising phase with a time

constant (τ= 4.48 ± 0.35 s, n = 11) indistinguishable from that of channels locked-

open by PPi from the C2 state (Figure 2-2 D, inset). Upon removal of PPi, all the

currents decrease mono-exponentially with a time constant of 27.65 ± 2.88 s (n = 7),

Page 41: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

29

Figure 2-2

Figure 2-2. PPi captures a short-lived, post-hydrolytic state. (A) A cartoon

depicting the mechanism based in Figure 2-1 A by which PPi locks open CFTR. O:

open state with a dimerized NBDs, C2: closed state with a partially separated NBDs,

LO: locked-open state with PPi occupying the second ATP binding pocket, ABP2

(B,D,F,G) Macroscopic current of WT-CFTR channels was activated by ATP to a stead

state before carrying out different ligand-switch protocols: (B) washout for 10 s

before applying PPi, (D) direct switch from ATP to PPi, (F) direct switched from ATP

to a 1-s PPi pulse (G) washout for 3 s before applying a 1-s PPi pulse. (C) Computer

simulation of macroscopic currents based on scenario 1. (E) A cartoon depicting a

revised CFTR gating model, wherein state X can respond to PPi rapidly. Insets in (B)

and (D) show the current rising phase upon PPi application; τrepresents the

relaxation time constant obtained by fitting the current rise with a single

Page 42: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

30

exponential function (mean ± SEM were specified in the main text). Bars above each

trace mark the perfused ligand denoted on the very left (applied to every figures).

Page 43: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

31

which is almost identical to the lifetime of locked-open channels from the C2 closed

state (24.15 ± 2.67 s, n = 6). This result suggests that there are at least two post-

hydrolytic states that can be distinguished by their responsiveness to PPi; one,

named state X, can be more readily locked open within our solution-exchange time

and thus prevent the current from dropping to the baseline, and the other, the

previously identified C2 state, responds to PPi much more sluggishly (scenario 2,

Figure 2-2 E). Since the lifetime of the locked-open state is constant no matter from

which state, C2 or X, the channels become locked open, we conclude that a single

locked-open conformation is attained, i.e., a channel with a dimerized NBDs wherein

ABP1 is occupied by ATP while ABP2 is taken by PPi (Tsai et al., 2009).

This newly identified state can be further probed by applying PPi for a brief

duration that is deemed too short for channels in the C2 state to respond. Figure 2-2

F shows such an experiment where ATP was switched to 10 mM PPi for just 1

second. This maneuver indeed effectively prevented a complete current decay seen

upon removal of ATP and at the same time eliminated the slow rising phase seen in

Figure 2-2 D. In contrast, the application of the same 1-s pulse of PPi 3 s after ATP

washout resulted in only minuscule current (Figure 2-2 G), indicating that 1-s pulse

application of PPi is indeed too short for the C2 state to respond to a measurable

extent, and that this newly identified state X is short-lived. Similar results were

obtained with AMP-PNP instead of PPi as a non-hydrolyzable ligand (Figure 2-3).

The results in Figure 2-2 and 2-3 not only demonstrate a novel state with a

distinct responsiveness to PPi or AMP-PNP, the fact that this state vanishes within

Page 44: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

32

seconds after ATP washout (Figure 2-2 G and 2-3 C) also suggests that it bears an

energetically unstable conformation. The time when state X emerges is also

intriguing in that it is not present before opening of the channel with ATP, and it

ceases to exist just 3 s after closing of the channel following ATP hydrolysis. If we

accept the idea that the open channel at least initially harbors an NBD dimer with

both ABPs occupied (Vergani et al., 2005), this unique timing of state X surfacing

leads to a conclusion that state X represents a post-hydrolytic state appearing prior

to the C2 closed state as depicted in Figure 2-2 E (see Supplemental Discussion for

details). More important, the fact that PPi or AMP-PNP can occupy ABP2 while the

channel resides in state X indicates that the NBD dimer must have separated to the

extent that can accommodate a large ligand like AMP-PNP.

Competition between ADP and PPi (or AMP-PNP) for ABP2 of state X

The proposition that PPi or AMP-PNP enters ABP2 in state X to lock open CFTR

suggests that the hydrolytic products, ADP and Pi, have been released. Thus, one

expects that when applying ADP altogether with PPi or AMP-PNP, the effectiveness

of these non-hydrolyzable ligands should be decreased as ADP may compete with

PPi or AMP-PNP for the binding site. Using the protocol shown in Figure 2-2 F, we

compared the fraction of locked-open channels by three different concentrations (1,

2 and 10 mM) of the non-hydrolyzable ligands in the presence or absence of 1 mM

ADP. Although at lower concentrations of PPi the measurements were not as

accurate due to the smaller signal observed (i.e., the fraction of the locked-open

Page 45: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

33

Figure 2-3

Figure 2-3. Identification of a post-hydrolytic state by AMP-PNP. Macroscopic

current of WT-CFTR channels were activated by ATP to a steady state before

carrying out similar ligand exchange protocols as shown in Figure 2-2: (A) washout

for 10 s and then applied AMP-PNP, (B) directly switched to a 1-s AMP-PNP pulse (C)

washout for 3 s and then applied a 1-s AMP-PNP pulse.

Page 46: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

34

channels), we indeed found significant inhibitory effects of ADP in all three

experimental conditions (Figure 2-4). These results again support the notion that in

state X, the NBDs most likely assume a partial dimeric conformation, where the head

of NBD2 and tail of NBD1 is disengaged to expose the ATP binding site in ABP2

similar to the configuration of the previously identified C2 state (Tsai et al., 2010b).

Differential modulation of state X and the C2 state

Although both state X and the C2 state may harbor conformationally similar

NBDs (i.e., a partial NBD dimer) as revealed by their responsiveness to AMP-PNP or

PPi, they must differ in the overall conformation since the rate of their response to

these non-hydrolyzable analogs is not the same. Here we provide evidence that

these two states can be differentially modulated by different experimental

maneuvers. First we used N6-phenylethyl-ATP (P-ATP), a high affinity hydrolyzable

ATP analogue (Zhou et al., 2005), as the initial ligand and carried out the protocols

shown in Figure 2-2 B and 2 F to isolate the fast and slow PPi responsive states

respectively. Interestingly, the proportion of the fast locked-open current relative to

the current prior to washout remained unchanged when directly switch the ligand

from P-ATP to a 1-s PPi pulse (46 ± 3 %, n = 8 for P-ATP versus 49 ± 2 %, n = 13 for

ATP), whereas the slow locked-open current relative to the ATP- activated current

was increased when applied PPi after 10-s washout (83 ± 11 %, n = 7 for P-ATP

versus 32 ± 2 %, n = 6 for ATP) (Figure 2-5 A). Thus, opening CFTR channels with P-

ATP improves the responsiveness of the C2 state (comparing to Figure 2-2 B), but

Page 47: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

35

Figure 2-4

Figure 2-4. ADP competes with PPi or AMP-PNP for state X. (A and B)

Macroscopic current of WT-CFTR was activated by ATP to a stead state before

switching the ligand to a 1-s pulse of PPi (A) or AMP-PNP (B) in the presence (right

panels) or absence (left panel) of 1 mM ADP. (C and D) Summery of the proportion

Page 48: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

36

of locked-open currents (I LO) relative to ATP activated currents (I ATP) in two

different PPi (C) or AMP-PNP (D) concentrations (mean ± SEM). * indicates P < 0.05.

Page 49: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

37

not state X, to PPi. On the other hand, mutating the residue Trp401 (W401), which

forms a ring-ring stacking interactions with ATP in ABP1 (Lewis et al., 2004; Zhou et

al., 2006), to phenylalanine significantly increased the propensity of state X in

response to PPi (i.e. a higher percentage of channels locked-open from state X (WT:

49 ± 2 %, n = 13, W401F: 59 ± 2 %, n = 13, P < 0.05)). But, this same mutation

dramatically reduced the lock-open current from the C2 state (0.08 ± 0.005 %, n =

6,Fig 5 B). Therefore, in opposite to P-ATP, the W401F mutation enhances the

responsiveness of state X to PPi but may destabilize the C2 state. These differential

modulations of the properties of state X and the C2 state reinforce our

interpretation that these two states cannot be conformationally identical.

State X is an open state or a very brief closed state

To further investigate the nature of state X, we performed single-channel ligand

exchange experiments, which allow close monitoring of the channel gating upon

ligand exchange. After witnessing the channel entering an open state, we switched

the perfusate

from ATP to PPi or AMP-PNP for just 1 s before washout. In this way, we can be

certain that the channel was exposed to PPi or AMP-PNP in the open state. Such

protocol could be technically difficult to perform with the wild-type (WT) channel

because its short mean open time (~300 ms) provides a limited operational window.

However, the conserved W401F mutation (see below), which significantly prolongs

the open time (Tsai et al., 2010a), makes such experiments more feasible.

Page 50: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

38

Figure 2-5

Figure 2-5. Differential modulation of the C2 state and state X. (A) WT-CFTR

channels were first activated by ATP and then the ligand was switched to P-ATP

until the current reached a steady state. Then the ligand was switched directly to a

1-s PPi pulse (upper panel) to assess the responsiveness of state X to PPi. P-ATP was

washed out for 10 s before applying PPi (lower panel) to assess the responsiveness

of the C2 state to PPi. (B) Similar protocol was conducted with W401F-CFTR

channels. The percentage of locked open current relative to ATP- (IATP) or P-ATP-

induced (IP-ATP) current was marked.

Page 51: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

39

In the presence of ATP, the W401F-CFTR channel exhibits similar behavior as WT-

CFTR: the channel opens into bursts that last for hundreds of milliseconds to

seconds, and each opening burst is separated by a long interburst closing also in the

range of hundreds of milliseconds (Figure 2-6). Upon ligand switches from ATP to 1-

s pulse of PPi or AMP-PNP (Figure 2-6 A and C), the channel can sojourn into a long-

lasting locked-open burst without entering into a long interburst closure (see

Discussion for details). For control, we carried out the same ligand exchange

protocol when the channel was closed following washout of ATP and did not see any

locked-open event (Figure 2-6 B and D). These results suggest that state X is either

an open state or a short-lived closed state that is indistinguishable from the flickers

within each opening burst. We measured the duration of each closed events within

that 1-s time window upon which PPi or AMP-PNP was applied. The maximal

duration among 68 events is 86 ms with an average value of 25.1 ± 2.4 ms,

interburst closed time (≥ ~400 ms). It is noteworthy that the success rate of this line

of experiments is lower for WT channels (see Supplemental Discussion for more

details), but similar observations were made for WT-CFTR (Figure 2-S1), indicating

that such open-X-locked-open transitions are not just idiosyncratic for W401F

mutant channels.

The lifetime of opening bursts forW401F-CFTR is [ATP] dependent

So far our results have laid out a picture that state X, like the C2 closed state, is

a state with a vacant ABP2 that is readily accessible to sizable ligands such as AMP-

Page 52: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

40

Figure 2-6

Figure 2-6. Single-channel ligand exchange for W401F-CFTR. (A and C) A single

W401F-CFTR was activated by ATP and the ligand was switched to a 1-s PPi (A) or

Page 53: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

41

AMP-PNP (C) pulse in the open state, the channel was locked-open after the short

application of PPi or AMP-PNP. The current trace was expanded in the red box to

better discern events during ligand switch. (B and D) The same 1-s PPi (B) or AMP-

PNP (D) pulse was applied when the channel resided in the close state; no locked-

open event was seen. Channels can be reactivated by ATP after long washout.

Page 54: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

42

PNP or PPi. But, unlike the C2 state, which can be very stable, state X is likely short-

lived (Figure 2-1 and which is similar to the mean lifetime of flickering closures

characterized previously for hydrolysis-deficient mutants (Bompadre et al., 2005)

but is indeed much shorter than the Figure 2-6). Since both the shape and size

between AMP-PNP and ATP are very similar, we reasoned that ATP is also capable

of binding to state X, re-dimerizes the NBD and brings the channel to the initial, pre-

hydrolytic open state (X → X’ → O in Figure 2-8). This hypothesis would predict that

the mean open burst time should increase by increasing [ATP] as more reentry

events (X → X’ → O in Figure 2-8) are likely to occur at higher [ATP]. This prediction

is self-explanatory if state X is an open state. Even if state X is a short-lived closed

state, this phenomenon should be observed as the analytic method used for

microscopic kinetic analysis of CFTR discards short-lived closings in order to extract

“ATP-dependent” events (see (Csanady and Gadsby, 1999; Vergani et al., 2005;

Csanady et al., 2006; Csanady et al., 2010; Kopeikin et al., 2010; Tsai et al., 2010a;

Cai et al., 2011; Erkens et al., 2011; ter Beek et al., 2011) and Discussion for details).

However, this [ATP]-dependent increase of the open burst time was not observed

for WT-CFTR (Winter et al., 1994; Zeltwanger et al., 1999; Vergani et al., 2003; Wang

et al., 2010) probably because state X in WT channels is very unstable or/and it

responds extremely poorly to ATP (See Supplemental Discussion for detail). For

reason unclear at this moment, the W401F mutation could somehow improve the

responsiveness of state X to AMP-PNP (Figure 2-3) and therefore presumably to

ATP as well. We therefore quantified the open burst time of W401F-CFTR at

different [ATP]. Consistent with our idea, there is indeed a discernible [ATP]-

Page 55: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

43

dependent increase of the mean open burst time when the [ATP] is above 1 mM

(Figure 2-7). Furthermore, the notion that a prolonged open burst time seen with

the W401F mutation is caused by reentry of the channel from state X predicts a

disparity between the mean open time estimated from microscopic analysis and the

relaxation time constant obtained from macroscopic current decay upon ATP

removal, which prohibits reentry to occur. Figure 2-7 B and C shows that when

fitting the current relaxation of W401F macroscopic current after washout of 2.75

mM ATP, the time constant is almost identical to the mean open time of W401F

channels in the presence of micromolar ATP.

2-5. Discussion

In the current manuscript, using nonhydrolyzable nucleotide analogs as “baits”,

we were able to capture a short-lived, post-hydrolytic state (state X, Figure 2-8).

This newly identified state distinguishes itself by its fast response to non-

hydrolyzable ligands, PPi or AMP-PNP, the telltale sign for the presence of an

exposed nucleotide binding site. In light of the evolutionary relationship between

CFTR and ABC transporters, our results has potential to not only bring new insights

to the coupling mechanism between ATP hydrolysis and gating transitions of CFTR,

but also bear potential structural/functional implications on the operational

mechanism of the widespread ABC transporters.

Page 56: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

44

Figure 2-7

Figure2-7. The mean open time of W401F-CFTR is [ATP] dependent. (A)

Microscopic current traces of W401F-CFTR in the presence of 10 mM (top), 1 mM

(middle) or 100 μM ATP. (B) Mean open time of W401F-CFTR at different [ATP]

(filled circles). Stars indicate P < 0.05 when compared to 10 mM [ATP] (c)

Macroscopic current of W401F-CFTR was activated by ATP to a steady state. The

current relaxation upon ATP washout was fitted with single exponential function

(red curve), τ = 377 ± 14 ms (n = 13).

Page 57: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

45

Of note, the ligand exchange method and subsequent interpretations of our data are

based on one well established theory: for WT-CFTR, channel opening is associated

with NBD dimerization that sandwiches two ATP molecules in the two ABPs

(Vergani et al., 2005), and two simple physical rules: First, for a ligand to leave or

enter an encaged space (i.e., NBD dimer interface) there must be a cleft equal or

larger than the size of the ligand.

Second, for ligand A to be substituted by ligand B, ligand A must leave first.

Thus, ligand exchanges between ATP and AMP-PNP during state X yield an

inevitable conclusion that state X harbors a conformation that features a vacated

ATP binding site. Since state X is caught during the current decay phase (Figure 2-1),

which is embedded with ATP hydrolysis and release of the hydrolytic products, but

before the channel transits to the previously characterized C2 closed state (Tsai et

al., 2009; Tsai et al., 2010b), we conclude that state X, similar to the C2 closed state,

bears a partially separated NBDs with an exposed ABP2. However, the partial NBD

dimeric conformation in state X may not be identical to the one in the C2 state. As

shown in Figure 2-5, the W401F mutation and the high-affinity ATP analog P-ATP,

both acting on ABP1, affect state X and the C2 state very differently. One simple

scenario is that separation of the NBD dimer is incremental so that proceeding of the

channel from state X to the C2 state is accompanied by a stepwise increase of the

separation of NBDs. This idea, albeit awaiting further experimental substantiation,

could account for the observation that state X exhibits a much faster response to PPi

or AMP-PNP than the C2 state (Figure 2-2) as if these ligands can re-dimerize the

NBDs of state X with ease.

Page 58: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

46

Perhaps the most important question pertaining to the newly identified state is

the functional status of its gate. As described above, state X is either an open state or

a brief closed state. In the following section, we will discuss in detail the

structure/function implications as well as the technical consequences out of these

two equally possible scenarios.

But first, it is necessary to delineate the generally adopted definition of “open”

and “closed” events in CFTR gating. It has been long held that CFTR channels open in

bursts; there are actually two distinct groups of closed events in CFTR’s gating: the

[ATP]-dependent one (i.e., interburst closure) with a time constant of hundreds of

milliseconds to seconds (Zeltwanger et al., 1999; Vergani et al., 2003), and the other

one buried in an opening burst with a time constant in tens of milliseconds that was

known as flickers (marked by stars in Figure 2-8). These flickering closings have

been long thought to be ATP independent as they can be easily discerned, in a

complete absence of ATP, within an opening burst of hydrolysis-deficient mutants

such as E1371S or K1250A (Powe et al., 2002; Bompadre et al., 2005; Vergani et al.,

2005; Kopeikin et al., 2010). Although the nature and the mechanism of such flickers

have not yet been rigorously studied, it was proposed that at least some of them are

due to pore blockage by anionic molecules in the solution (Zhou et al., 2001). These

flickering closings have been excluded by various means in the literature upon data

analysis to extract gating parameters that are associated with ATP binding and

hydrolysis (Csanady and Gadsby, 1999; Bompadre et al., 2005; Vergani et al., 2005;

Csanady et al., 2006; Csanady et al., 2010; Kopeikin et al., 2010; Tsai et al., 2010a;

Wang et al., 2010; Cai et al., 2011; Erkens et al., 2011; ter Beek et al., 2011).

Page 59: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

47

Therefore, the strict coupling model depicted in Figure 2-1 means that each opening

burst, which may contain several flickering closings, corresponds to hydrolysis of

one ATP molecule.

Scenario 1: state X is a transient closed state

Because we cannot distinguish between the presumed closed state X and other

ATP-independent flickers in the single channel traces, an accurate estimation of the

lifetime of state X is unattainable. However, as discussed in Results, should state X be

a closed state, its dwell time is unlikely much longer than that of flickering closings

(~ 25 ms in Figure 2-6 and in Bompadre et al., 2005b). Furthermore, the prolonged

open burst time in the presence of higher [ATP] (> 1 mM, P < 0.05 when comparing

1 mM and 10 mM) for W401F-CFTR indicates that the lifetime of state X as a closed

state must be indistinguishable from that of the flickers so that closed events

corresponding to state X are excluded by the analysis. In other words, the longer

“observable” openings (or the open bursts), which have been considered as

hydrolysis-coupled events (Csanady and Gadsby, 1999; Bompadre et al., 2005;

Vergani et al., 2005; Csanady et al., 2006; Csanady et al., 2010; Kopeikin et al., 2010;

Tsai et al., 2010a; Wang et al., 2010; Cai et al., 2011; Erkens et al., 2011; ter Beek et

al., 2011), in fact could contain several “true” opening events separated by the

closed state X. As explained in detail below, contrary to the model depicted in Figure

2-1, it is this “true” opening event that is coupled to hydrolysis of one ATP under the

premise that state X is a closed state.

Page 60: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

48

As described above, in order for state X to be accommodating to PPi or AMP-

PNP, it has to bear a partially separated NBD dimer with a vacated ABP2, to which

ATP should also be able to bind. Furthermore, PPi or AMP-PNP needs to act on this

presumed, barely visible closed state at an extremely fast rate; otherwise this closed

event will become more discernible in the single-channel recording. As the chemical

structure of AMP-PNP is nearly identical to that of ATP (Yount, 1975), and this ATP

analog can be found comfortably nestled in the ATP binding sites of ABC

transporters (Dawson and Locher, 2007), one comes to an ineluctable implication

that ATP can also readily re-open the channel from state X. Thus, contrary to the

long-held idea that ATP opens CFTR solely through the long interburst closure, one

has to conclude that some of the short-lived closures that are buried in an opening

burst can respond to ATP to enter the open state. Thus, ATP hydrolysis is not

coupled to the “observable” opening burst, but to the “true” opening that can

happen more than one time in a burst.

Scenario 2: state X is an open state.

If we interpret state X as an open state by discarding short-lived closures as

previously adopted (Csanady and Gadsby, 1999; Bompadre et al., 2005; Vergani et

al., 2005; Csanady et al., 2006; Csanady et al., 2010; Kopeikin et al., 2010; Tsai et al.,

2010a; Wang et al., 2010; Cai et al., 2011; Erkens et al., 2011; ter Beek et al., 2011),

one immediate consequence is a violation of one-to-one stoichiometry between

CFTR’s gating cycle and the ATP hydrolysis cycle since when the channel resides in

state X, the vacated ATP binding site is now accessible to a new ATP molecule, which

Page 61: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

49

subsequently re-dimerizes the NBDs and return the channel to the original open

state (X (open) →X’ (open) → O1 in Figure 2-8).

At first glance, this conclusion based on the assumption that state X is an open state

seems in disagreement with those reported in Csanady et al. (Csanady et al., 2010),

but, in fact, they are not in conflict with but complementary to what they had

discovered. The bimodal distribution in their open time histograms convincingly

demonstrated that closure of CFTR involves at least two steps, presumably ATP

hydrolysis and NBD dimer separation. Csanady and his colleagues (Csanady et al.,

2010) concluded that in WT-CFTR, the majority of channel closings are the

consequence of ATP hydrolysis rather than through a non-hydrolytic pathway. If

state X is an open state, our data lead us to propose that for those channels

undergoing hydrolysis-dependent closing, the hydrolysis per se does not guarantee

a gate closure and neither does the release of the hydrolytic products. It suggests a

delay of the gating signal transmission from NBDs to TMDs; the NBD dimer

separates prior to the gate closure, resulting in a post-hydrolytic open state (state X)

with a partially separated NBDs. Indeed, this addition of a post-hydrolytic open state

(state X) and a re-entry pathway to the pre-hydrolytic open state (O1 state) (X → X’

→ O1, Figure 2-8) still predicts an open time histogram with a bimodal distribution

similar to that reported in Csanady et al. (Csanady et al., 2010). By comparing WT-

and W401F-CFTR, the increasing reentry events (X → X’ → O1) results in differences

only in the mean open time and the tail post-hydrolytic open state with a full NBD

dimer, X: the newly identified state with partial separated NBD dimer. X’: one ATP

bound to the vacant ABP2 in state X. Stars mark the ATP-independent, short-lived,

Page 62: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

50

flickery closed events that are discounted in data analysis and interpretation (see

Discussion for details). The red box encompasses the reentry pathway that may

occur within each opening burst.

of the fitted curve (Figure 2-S2), which is supplanted with an increased number of

long opening bursts (> 1000 ms).

It is noteworthy that the idea of a post-hydrolytic open state was proposed 17

years ago. Gunderson and Kopito observed two open states with different single-

channel conductance in WT-CFTR (Gunderson and Kopito, 1995). Because the

appearance of the two open states follows a preferred order (i.e small conductance

→ large conductance), they proposed that by providing an input of the free energy,

ATP hydrolysis per se triggered such transition. Could the O2 state seen by

Gunderson and Kopito correspond to state X here? Further studies are needed to

provide the answer.

In addition, if separation of NBDs could occur before closure of the gate in

TMDs takes places, it would suggest that these two domains in CFTR may assume

certain degree of autonomy—movement of one domain does not have to be coupled

to the motion of the other. Then, instead of physical coupling of each TMD-NBD

complex as a rigid body (Ivetac et al., 2007; Khare et al., 2009), one will entertain the

possibility of an energetic coupling between NBDs and TMDs. Indeed, based on

thermodynamic analysis of CFTR gating, Csanady and his colleagues proposed that

the gate in TMDs opens after the two NBDs have dimerized (Csanady et al., 2006).

Also consistent with this idea, lately, we (Bai et al., 2010) showed that chemical

Page 63: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

51

Figure 2-8

Figure 2-8. A revised CFTR gating scheme showing hypothetical

conformational transitions that takes place during an opening burst. C1: closed

state with two separated NBDs, C2: partial NBD dimeric closed state, O1: pre-

hydrolytic open state, O1’: post-hydrolytic open state with a full NBD dimer, X: the

newly identified state with partial separated NBD dimer. X’: one ATP bound to the

vacant ABP2 in state X. Stars mark the ATP-independent, short-lived, flickery closed

events that are discounted in data analysis and interpretation (see Discussion for

details). The red box encompasses the reentry pathway that may occur within each

opening burst.

Page 64: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

52

modifications of an introduced cysteine in the TMDs can render the CFTR channel

almost completely ATP independent. As this dramatic effect was also observed in a

construct devoid of the entire NBD2 (unpublished observation), at least in this

extreme case, opening and closing of the gate in TMDs can take place without the

necessity of NBD dimerization.

In conclusion, by using non-hydrolytic ATP analogs as tools, we uncovered a

novel post-hydrolytic state that has a vacant ABP2 readily accessible to ATP.

Regardless of the status of the gate for this state, our results argue that each

previously defined opening burst in CFTR gating may entail hydrolysis of more than

one ATP. An imminent goal is to answer a crucial question: is state X an open or a

closed state? Answering this question could lead to a better understanding of the

coupling mechanism between TMDs and NBDs not only for CFTR, but also for ABC

transporters at large.

2-6. Supplementary information

State X is a post-hydrolytic state

In Figure 2-2, we proposed that state X is a short-lived state that emerges after

the pre-hydrolytic open state (the O state in Figure 2-2 E) and that it rapidly

dissipates into the C2 closed state. This proposition was made based on the

following rationales. First, Csanady and collegues (Csanady et al., 2010) showed

convincingly that more than 95% of the WT-CFTR channels opened by NBD

Page 65: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

53

dimerization close following hydrolysis of the ATP bound in the second ATP binding

pocket, or ABP2 (i.e., hydrolytic closing). In the pre-hydrolytic open state, two ATP

molecules are sandwiched by the NBD dimer (Vergani et al., 2005), therefore both

binding sites are inaccessible to a new ligand such as PPi or AMP-PNP until ATP

hydrolysis takes place and the hydrolytic products dissociate from ABP2. Thus, for

PPi or AMP-PNP to lock open the channel by binding to ABP2 (Tsai et al., 2009; Tsai

et al., 2010b), ATP hydrolysis and release of the hydrolytic products need to occur

first. Second, Tsai et al. (Tsai et al., 2009) reported that once the channel closes after

dissociation of the hydrolytic products from ABP2, it stays in a stable closed state

(C2 closed state) with a lifetime of ~ 25 s. This C2 closed state responds to PPi very

slowly and therefore cannot account for the fast phase of current response to PPi

upon ligand switches as shown in Figure 2-2 D. Third, our experiments showed that

state X, which bears the capability of a rapid response to PPi, vanishes within 3 s

after ATP washout and does not resurface until the channel is first opened by ATP

again (Figure 2-2 G). Although other more complicated schemes may explain our

data, the simplest scenario is to place state X in between the post-hydrolytic open

state and the C2 state.

Technical issues for single-channel ligand exchange

The single-channel ligand exchange experiments shown in Figure 2-6 and 2-S1

are technically challenging in two aspects: First, since most of the WT-CFTR

channels open for less than half a second, the time window for solution changes is

Page 66: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

54

very short, therefore in many experiments the channel simply closed upon ligand

switches from ATP to PPi or AMP-PNP and did not open again. Second, to avoid

significant baseline shifts due to mismatch of the ionic strength in solutions with

ATP or PPi/AMP-PNP, concentrations of these non-hydrolyzable ligands were kept

at 1 mM. Thus, although in macroscopic ligand exchange experiments (Figure 2-2 F),

1-s pulse application of 10 mM PPi lock-opens ~40% of the open WT channels, the

capture rate is much lower in single-channel experiments. The W401F mutation

offers two advantages for single-channel ligand exchange experiments. Compared to

WT-CFTR, W401F-CFTR channels respond more robustly to PPi. In addition, the

mean open time of W401F-CFTR is longer than that of WT-CFTR, thus providing a

longer time window for ligand exchange. Despite the technical difficulties described

above, such open-X-locked-open transitions were observed with WT-CFTR (Figure

2-S1), but the percentage of exchanges that result in a locked-open event is lower

compare to that of W401F-CFTR (7% in WT-CFTR vs 13% in W401F-CFTR).

Detection of [ATP] dependent open time

As elaborated in Discussion, no matter whether the state X is an open state or a

closed state, the “observable” open time (i.e., the open burst time) should become

longer at higher [ATP] as the frequency of the reentry event (X → X’ → O in Figure 2-

8) increased. However, Figure 2-7 also exemplifies the technical difficulties in

experimentally demonstrating this effect due to a wide patch-to-patch variation of

Page 67: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

55

Figure 2-S1

Figure 2-S1. Single-channel ligand exchange for WT-CFTR. (A) A single WT-

CFTR was activated by ATP and the ligand was switched to a 1-s PPi pulse in the

open state, the channel was locked-open after the short application of PPi. The

current trace was expanded in the red box to better discern the events during ligand

switch. (B) Same 1-s PPi pulse was applied when the channel was in the closed state;

no locked-open event was seen. Channels can be reactivated by ATP after long

washout, indicating that a functional channel remained in the membrane patch.

1 mM ATP1 mM PPi

1 mM ATP1 mM PPi

5 s

0.4

pA

2 s

0.4

pA

500 ms

1 mM ATP1 mM PPi

0.4

pA

AAAA

BBBB

Page 68: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

56

the open time. Thus, even with a large number of experiments, our results as well as

the results from other studies still showed an SEM around 50 -100 ms (Vergani et al.,

2003; Gadsby et al., 2006; Zhou et al., 2006; Csanady et al., 2010; Wang et al., 2010;

Erkens et al., 2011; ter Beek et al., 2011). In addition, state X responds to PPi more

sluggishly in WT- than in W401F-CFTR, suggesting that fewer reentry events take

place in WT-CFTR than in W401F-CFTR even with millimolar [ATP]. As a result, such

trend of [ATP] dependent open time was rarely seen in WT-CFTR. (Winter et al.,

1994; Zeltwanger et al., 1999; Vergani et al., 2003; Wang et al., 2010).

Short of experimental data demonstrating [ATP]-dependent open times, how

can we be sure that re-entry does occur with WT-CFTR? In other words, can we

extrapolate the observations for the W401F mutation to WT channels? The answer

to these questions is likely positive for the following reasons. First, W401F is a

conserved mutation. In previous studies, the gating characteristics are very similar

between W401F- and WT-CFTR, except that W401F-CFTR shows a longer mean

open time and PPi locked-open time (Tsai et al., 2010a; Tsai et al., 2010b). Therefore,

it is unlikely that this Trp to Phe mutation could result in a gating mechanism that is

entirely different from that of WT-CFTR. Indeed, in the current studies, all the

different experimental maneuvers only reveal quantitative but not qualitative

differences between WT- and W401F-CFTR. Second, as shown in Figure 2-3, AMP-

PNP was effective in locking open WT-CFTR, suggesting that the structurally similar

ATP may capture the state X as well. Therefore, it is likely that such reentry pathway

indeed exists in WT-CFTR.

Page 69: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

57

Figure 2-S2

Figure 2-S2. PPi locked open WT-CFTR 5 s after ATP washout. Macroscopic

current of WT-CFTR channels was activated by ATP to a stead state. 10 mM PPi was

applied 5 s after ATP washes out. The ratio of PPi locked-open current to ATP-

induced current is 38 ± 4 %, n = 7.

2.75 mM ATP

10 mM PPi

38 % of I ATP

Page 70: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

58

Figure 2-S3

Figure 2-S3. Open time histograms for WT- and W401F-CFTR. (A) Open time

histograms of WT-CFTR in the presence of 2.75 mM ATP. (B) Open time histograms

of W401F-CFTR in the presence of 10 mM ATP. Black lines: open time histograms

fitted with double exponential function.

1.51.00.5

1.51.00.5

Open time (s)

0

2

4

6

8

0

2

4

6

8

10%

of

tota

l e

ve

nts

% o

f to

tal e

ve

nts

AAAA

BBBB

n = 547

n = 447

Figure S3

Page 71: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

59

Figure 2-S4

Figure 2-S4. Kinetic model and parameters for computer simulation in Figure

2-2C. (A, B) The kinetic scheme (A) and parameters (B) used for computer

simulation. Parameters K1F/R and K2F/R were chosen based on the opening rate of

WT-CFTR (Zeltwanger et al., 1999; Vergani et al., 2005; Wang et al., 2010), K3F and

K4F were adapted from (Csanady et al., 2010), K4R were chosen based on the

spontaneous opening rate (Bompadre et al., 2005), K5R/F and K6R/F were chosen

based on the lock-opening rate and locked-open time for WT-CFTR ((Tsai et al.,

2009) and Figure 2-1).

C2 C2'

O1

AAAA

K1F

K1R

K2FK4FK4R K2R

K3F

C2P'

LO

K6F K6R

K5R

K5F

K7RK7F

C1

Rate WT Units

K1F

K1R

K2F

K2R

K3F

K4R

K5F

K5R

K6F

K6R

[s ]

[s x M ]

[s x M ]

[s ]

[s ]

[s ]

[s ]

[s ]

[s ]

-1

-1

-1

-1

-1

-1

-1

-1 -1

-1 -1

2

4.5

50

1 x 10

4 x 103

5

BBBB

ATP ADP + Pi PPi

1 x 105

0.4

0.1

10

K4F

4 x 10-3

3 x 10-2

[s ]-1

0.04

1 x 106

K7F

K7R

[s x M ]-1 -1

[s ]-1

[s ]-1

Page 72: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

60

2-7. References

Aller, S.G., J. Yu, A. Ward, Y. Weng, S. Chittaboina, R. Zhuo, P.M. Harrell, Y.T. Trinh, Q.

Zhang, I.L. Urbatsch, and G. Chang. 2009. Structure of P-glycoprotein reveals

a molecular basis for poly-specific drug binding. Science. 323:1718-1722.

Bai, Y., M. Li, and T.C. Hwang. 2010. Dual roles of the sixth transmembrane segment

of the CFTR chloride channel in gating and permeation. J Gen Physiol.

136:293-309.

Bai, Y., M. Li, and T.C. Hwang. 2011. Structural basis for the channel function of a

degraded ABC transporter, CFTR (ABCC7). J Gen Physiol. 138:495-507.

Bear, C.E., C.H. Li, N. Kartner, R.J. Bridges, T.J. Jensen, M. Ramjeesingh, and J.R.

Riordan. 1992. Purification and functional reconstitution of the cystic fibrosis

transmembrane conductance regulator (CFTR). Cell. 68:809-818.

Bompadre, S.G., J.H. Cho, X. Wang, X. Zou, Y. Sohma, M. Li, and T.C. Hwang. 2005.

CFTR gating II: Effects of nucleotide binding on the stability of open states. J

Gen Physiol. 125:377-394.

Cai, Z., Y. Sohma, S.G. Bompadre, D.N. Sheppard, and T.C. Hwang. 2011. Application

of high-resolution single-channel recording to functional studies of cystic

fibrosis mutants. Methods Mol Biol. 741:419-441.

Csanady, L. 2000. Rapid kinetic analysis of multichannel records by a simultaneous

fit to all dwell-time histograms. Biophys J. 78:785-799.

Csanady, L. 2010. Degenerate ABC composite site is stably glued together by trapped

ATP. J Gen Physiol. 135:395-398.

Csanady, L., and D.C. Gadsby. 1999. CFTR channel gating: incremental progress in

irreversible steps. J Gen Physiol. 114:49-53.

Csanady, L., A.C. Nairn, and D.C. Gadsby. 2006. Thermodynamics of CFTR channel

gating: a spreading conformational change initiates an irreversible gating

cycle. J Gen Physiol. 128:523-533.

Csanady, L., P. Vergani, and D.C. Gadsby. 2010. Strict coupling between CFTR's

catalytic cycle and gating of its Cl- ion pore revealed by distributions of open

channel burst durations. Proc Natl Acad Sci U S A. 107:1241-1246.

Dawson, R.J., and K.P. Locher. 2006. Structure of a bacterial multidrug ABC

transporter. Nature. 443:180-185.

Page 73: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

61

Dawson, R.J., and K.P. Locher. 2007. Structure of the multidrug ABC transporter

Sav1866 from Staphylococcus aureus in complex with AMP-PNP. FEBS Lett.

581:935-938.

Erkens, G.B., R.P. Berntsson, F. Fulyani, M. Majsnerowska, A. Vujicic-Zagar, J. Ter

Beek, B. Poolman, and D.J. Slotboom. 2011. The structural basis of modularity

in ECF-type ABC transporters. Nat Struct Mol Biol. 18:755-760.

Gadsby, D.C., P. Vergani, and L. Csanady. 2006. The ABC protein turned chloride

channel whose failure causes cystic fibrosis. Nature. 440:477-483.

Gunderson, K.L., and R.R. Kopito. 1994. Effects of pyrophosphate and nucleotide

analogs suggest a role for ATP hydrolysis in cystic fibrosis transmembrane

regulator channel gating. J Biol Chem. 269:19349-19353.

Gunderson, K.L., and R.R. Kopito. 1995. Conformational states of CFTR associated

with channel gating: the role ATP binding and hydrolysis. Cell. 82:231-239.

Hollenstein, K., D.C. Frei, and K.P. Locher. 2007. Structure of an ABC transporter in

complex with its binding protein. Nature. 446:213-216.

Hwang, T.C., G. Nagel, A.C. Nairn, and D.C. Gadsby. 1994. Regulation of the gating of

cystic fibrosis transmembrane conductance regulator C1 channels by

phosphorylation and ATP hydrolysis. Proc Natl Acad Sci U S A. 91:4698-4702.

Khare, D., M.L. Oldham, C. Orelle, A.L. Davidson, and J. Chen. 2009. Alternating access

in maltose transporter mediated by rigid-body rotations. Mol Cell. 33:528-

536.

Kopeikin, Z., Y. Sohma, M. Li, and T.C. Hwang. 2010. On the mechanism of CFTR

inhibition by a thiazolidinone derivative. J Gen Physiol. 136:659-671.

Lewis, H.A., S.G. Buchanan, S.K. Burley, K. Conners, M. Dickey, M. Dorwart, R. Fowler,

X. Gao, W.B. Guggino, W.A. Hendrickson, J.F. Hunt, M.C. Kearins, D. Lorimer,

P.C. Maloney, K.W. Post, K.R. Rajashankar, M.E. Rutter, J.M. Sauder, S. Shriver,

P.H. Thibodeau, P.J. Thomas, M. Zhang, X. Zhao, and S. Emtage. 2004.

Structure of nucleotide-binding domain 1 of the cystic fibrosis

transmembrane conductance regulator. EMBO J. 23:282-293.

Mense, M., P. Vergani, D.M. White, G. Altberg, A.C. Nairn, and D.C. Gadsby. 2006. In

vivo phosphorylation of CFTR promotes formation of a nucleotide-binding

domain heterodimer. EMBO J. 25:4728-4739.

Powe, A.C., Jr., L. Al-Nakkash, M. Li, and T.C. Hwang. 2002. Mutation of Walker-A

lysine 464 in cystic fibrosis transmembrane conductance regulator reveals

functional interaction between its nucleotide-binding domains. J Physiol.

539:333-346.

Page 74: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

62

Riordan, J.R., J.M. Rommens, B. Kerem, N. Alon, R. Rozmahel, Z. Grzelczak, J. Zielenski,

S. Lok, N. Plavsic, J.L. Chou, and et al. 1989. Identification of the cystic fibrosis

gene: cloning and characterization of complementary DNA. Science.

245:1066-1073.

Szollosi, A., D.R. Muallem, L. Csanady, and P. Vergani. 2011. Mutant cycles at CFTR's

non-canonical ATP-binding site support little interface separation during

gating. J Gen Physiol. 137:549-562.

ter Beek, J., R.H. Duurkens, G.B. Erkens, and D.J. Slotboom. 2011. Quaternary

structure and functional unit of energy coupling factor (ECF)-type

transporters. J Biol Chem. 286:5471-5475.

Tsai, M.F., K.Y. Jih, H. Shimizu, M. Li, and T.C. Hwang. 2010a. Optimization of the

degenerated interfacial ATP binding site improves the function of disease-

related mutant cystic fibrosis transmembrane conductance regulator (CFTR)

channels. J Biol Chem. 285:37663-37671.

Tsai, M.F., M. Li, and T.C. Hwang. 2010b. Stable ATP binding mediated by a partial

NBD dimer of the CFTR chloride channel. J Gen Physiol. 135:399-414.

Tsai, M.F., H. Shimizu, Y. Sohma, M. Li, and T.C. Hwang. 2009. State-dependent

modulation of CFTR gating by pyrophosphate. J Gen Physiol. 133:405-419.

Vergani, P., S.W. Lockless, A.C. Nairn, and D.C. Gadsby. 2005. CFTR channel opening

by ATP-driven tight dimerization of its nucleotide-binding domains. Nature.

433:876-880.

Vergani, P., A.C. Nairn, and D.C. Gadsby. 2003. On the mechanism of MgATP-

dependent gating of CFTR Cl- channels. J Gen Physiol. 121:17-36.

Wang, W., J. Wu, K. Bernard, G. Li, G. Wang, M.O. Bevensee, and K.L. Kirk. 2010. ATP-

independent CFTR channel gating and allosteric modulation by

phosphorylation. Proc Natl Acad Sci U S A. 107:3888-3893.

Ward, A., C.L. Reyes, J. Yu, C.B. Roth, and G. Chang. 2007. Flexibility in the ABC

transporter MsbA: Alternating access with a twist. Proc Natl Acad Sci U S A.

104:19005-19010.

Winter, M.C., D.N. Sheppard, M.R. Carson, and M.J. Welsh. 1994. Effect of ATP

concentration on CFTR Cl- channels: a kinetic analysis of channel regulation.

Biophys J. 66:1398-1403.

Yount, R.G. 1975. ATP analogs. Adv Enzymol Relat Areas Mol Biol. 43:1-56.

Zeltwanger, S., F. Wang, G.T. Wang, K.D. Gillis, and T.C. Hwang. 1999. Gating of cystic

fibrosis transmembrane conductance regulator chloride channels by

Page 75: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

63

adenosine triphosphate hydrolysis. Quantitative analysis of a cyclic gating

scheme. J Gen Physiol. 113:541-554.

Zhou, Z., S. Hu, and T.C. Hwang. 2001. Voltage-dependent flickery block of an open

cystic fibrosis transmembrane conductance regulator (CFTR) channel pore. J

Physiol. 532:435-448.

Zhou, Z., X. Wang, M. Li, Y. Sohma, X. Zou, and T.C. Hwang. 2005. High affinity

ATP/ADP analogues as new tools for studying CFTR gating. J Physiol.

569:447-457.

Zhou, Z., X. Wang, H.Y. Liu, X. Zou, M. Li, and T.C. Hwang. 2006. The two ATP binding

sites of cystic fibrosis transmembrane conductance regulator (CFTR) play

distinct roles in gating kinetics and energetics. J Gen Physiol. 128:413-422.

Page 76: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

64

CHAPTER 3

NON-INTEGRAL STOICHIOMETRY IN CFTR

GATING, SEEING IS BELIEVING

This chapter has been modified from my manuscript which had been submitted

to J. Gen Physiol., by Kang-Yang Jih, Yoshiro Sohma and Tzyh-Chang Hwang.

3-1. Abstract

Cystic fibrosis transmembrane conductance regulator (CFTR) is a unique

member of the ABC protein superfamily. Unlike most other ABC proteins that

function as active transporters, CFTR is an ATP-gated chloride channel. Opening of

CFTR’s gate is associated with ATP-induced dimerization of its two nucleotide

binding domains (NBD1 and NBD2) whereas gate closure is facilitated by ATP

hydrolysis-triggered partial separation of the NBDs. This generally held theme of

CFTR gating—a strict coupling between the ATP hydrolysis cycle and the gating

cycle—is put to the test by our recent finding of a short-lived, post-hydrolytic state

that can bind ATP and re-enter the ATP-induced original open state (Jih et al., 2012).

We accidentally found a mutant CFTR channel that exhibits two distinct open

conductance states, the smaller O1 state and the larger O2 state. In the presence of

ATP, the transition between the two states follows a preferred O1 → O2 order, a

telltale sign of a violation of microscopic reversibility, hence demanding an external

Page 77: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

65

energy input likely from ATP hydrolysis as such preferred gating transition was

abolished in a hydrolysis deficient mutant. Interestingly, we also observed

considerable amount of opening events that contains more than one O1 → O2

transition, indicating that more than one ATP molecule may be hydrolyzed within an

opening burst. We thus conclude a non-integral stoichiometry between the gating

cycle and ATP consumption. Our results lead to a new gating model conforming to

the classical allosteric mechanism: both NBDs and TMDs hold a certain degree of

autonomy whereas the conformational change in one domain will facilitate the

conformational change in the other domain.

3-2. Introduction

Cystic Fibrosis Transmembrane Conductance Regulator (CFTR) is a unique

member of the ABC (ATP binding cassette) Protein Superfamily because while most

members in this family are active transporters, CFTR is a bona fide ion channel.

Despite this functional disparity, the overall architecture of CFTR seems well

conserved. Canonical domains that a typical ABC transporter is made of, namely two

transmembrane domains (TMDs), each conjoined to a nucleotide binding domain,

are clearly preserved in CFTR. Recent published structures of CFTR’s two NBDs

reveal almost identical structural motifs seen in all ABC proteins (Lewis et al., 2004;

Dawson and Locher, 2006; Hollenstein et al., 2007; Aller et al., 2009 and

PDB#3GD7). Furthermore, functional studies in the past decade have provided

evidence that the operational mechanism of CFTR shares strikingly similar features

to at least the exporter members of the ABC Protein Family. For example, gating of

Page 78: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

66

CFTR is controlled by ATP binding and hydrolysis that induce respectively

association and dissociation of the two NBDs (Vergani et al., 2005) —a central

theme that operates in all ABC proteins. Upon ATP binding, CFTR’s NBD2 undergoes

an induced-fit conformational change (Szollosi et al., 2010) that has also been

proposed for ABC transporters based on comparison between the ATP-bound and

ATP-free crystal structures of individual NBD (Gaudet and Wiley, 2001; Karpowich

et al., 2001). Lately, Bai et al. (Bai et al., 2010, 2011) demonstrated that CFTR

employs evolutionarily conserved transmembrane segments for the construction of

its ion permeation pathway and that the conformational changes of CFTR’s TMDs

during gating transitions resemble what has been proposed for ABC transporters.

Thus unraveling the gating mechanism of CFTR could bear broader ramifications.

One of the most intriguing questions about CFTR is the role of ATP hydrolysis—

the hallmark for an active transport process—in driving gating conformational

changes. Two decades of research have gained enormous understanding about the

gating mechanism of CFTR. It is now well accepted that after phosphorylation by

PKA in the serine-rich regulatory domain (Gadsby et al., 1994; Bompadre et al.,

2005a; Csanady et al., 2005; Mense et al., 2006), the gate of CFTR is opened by NBD

dimerization induced by binding of ATP to its two NBDs (Vergani et al., 2003;

Vergani et al., 2005; Zhou et al., 2006). After ATP is hydrolyzed in the hydrolysis-

competent site (ATP bind pocket 2, ABP2), the channel sojourns into a closed state

(C2 state) with its NBDs in a partial dimeric configuration (Tsai et al., 2009; Tsai et

al., 2010b; Szollosi et al., 2011), in which the head of NBD1 and the tail of NBD2 (i.e.,

the catalysis-inert ABP1) remain associated with one ATP sandwiched in between

Page 79: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

67

while the catalysis-active ABP2 is separated and vacated. This C2 state has a dwell

time of ~30 s before entering the C1 state in which the NBD dimer completely

separates and the remaining ATP dissociates (Tsai et al., 2010b). As the role of each

ATP binding site and how ATP triggers the conformational change of NBDs become

well understood, one of the lingering puzzles in CFTR gating is how the gating

signals originated in the NBDs are transmitted to the TMDs to control the gate, or in

other words, how NBDs and TMDs communicate during gating?

Supported by crystal structures and computer simulations (Ivetac et al., 2007;

Khare et al., 2009), the popular rigid body hypothesis dictates a concurrent

movement of each NBD-TMD complex during the transport cycle of ABC

transporters. One functional implication of the rigid body movement hypothesis for

CFTR is a strict coupling between NBDs and TMDs with a one-to-one stoichiometry

between the ATP hydrolysis cycle and the gating cycle. This strict coupling

hypothesis has been proposed for more than a decade in the CFTR field and is

supported by many studies. For example, the drastic effect of non-hydrolyzable ATP

analogs or mutations (e.g., E1371S or K1250A) that abolish ATP hydrolysis on the

open time supports the notion that ATP hydrolysis is coupled to channel closure

(Gunderson and Kopito, 1994; Hwang et al., 1994; Carson et al., 1995; Vergani et al.,

2003; Bompadre et al., 2005b). An asymmetrical transition between gating states

suggests that the gating cycle of CFTR is driven by ATP hydrolysis (Gunderson and

Kopito, 1995). Recently, microscopic kinetic analysis that shows a paucity of short-

lived opening events in the open time histogram further reinforces the strict

coupling hypothesis (Csanady et al., 2010). However, in our latest report (Jih et al.,

Page 80: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

68

2012), by identifying a transient post-hydrolytic state that is capable of voyaging

into the pre-hydrolytic open state in response to ATP, we raised the possibility that

the movement of NBDs and the opening/closing of the gate may not be

synchronized. As described in Jih et al. (2012), the critical piece of experimental

evidence that will contest the strict coupling hypothesis is to affirm the existence of

a post-hydrolytic open state with a vacated ABP2 so that ATP can initiate an open-

to-open transition.

A recent fortuitous discovery provides us an opportunity to do just that. When

mutating the positively charged arginine at position 352 (located in the sixth

transmembrane segment, TM6) to cysteine, the mutant channel (R352C-CFTR)

features two distinct open states with unequal conductance. Carefully scrutinizing

single-channel recordings revealed a striking phenomenon: the appearance of these

two open states follows a preferred order similar to that reported for WT-CFTR by

Gunderson and Kopito 17 years ago (Gunderson and Kopito, 1995). In the presence

of ATP, most often the channel leaves the closed state to enter a small conductance

open state (named O1 state hereafter); however, the channel often returns to the

closed state from a large conductance open state (named O2 state hereafter). After

sorting all the opening events, we observed a dominant population that features a C

→ O1 → O2 → C transition (Table 3-1) whereas very rare C → O2 → O1 → C

transitions were seen in our recordings. Such time asymmetry indicates that these

three states are not at thermodynamic equilibrium (Richard and Miller, 1990;

Gunderson and Kopito, 1995). Instead, an input of free energy, likely coming from

ATP hydrolysis, maintains these transitions at a steady state. Consistent with this

Page 81: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

69

idea, ATP only induces C → O1 → C transitions in E1371S/R352C-CFTR, a

hydrolysis deficient mutant. The R352C mutant channel becomes a precious tool as

it allows us to distinguish the pre- (O1) and post-hydrolytic (O2) open states and

thus to “visualize” ATP hydrolysis taking place within each opening burst.

Interestingly, we observed opening bursts that include more than one O1→O2

transition, implicating that hydrolysis of more than one ATP molecule can take place

within an opening burst. Furthermore, the frequency of opening bursts that undergo

multiple rounds of ATP hydrolysis can be modulated by changing [ATP], maneuvers

that greatly enhance ATP-independent opening, or introducing mutation that affects

ATP binding to ABP1. Based on these results, we propose a new model for CFTR

gating that features an energetic coupling between the NBDs and the TMDs (details

elaborated in Discussion).

3-3. Material and Methods

Cell culture and transient expression system

Chinese hamster ovary (CHO) cells were cultured at 37 °C in Dulbecco’s Modified

Eagle’s Medium supplemented with 10% fetal bovine serum. The cells were

trypsinized and cultured in 35-mm tissue culture dishes 1 day before the

transfection. The cDNA of CFTR and pEGFP-C3 (Clontech, Palo Alto, CA), which

encodes green fluorescence protein, were cotransfected into CHO cells using

PolyFect transfection reagent (QIAGEN). The cells for electrophysiological

experiments were grown at 27 °C for two days after transfection.

Page 82: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

70

Mutagenesis

QuickChange XL kit (Stratagene, La Jolla, CA) were used to construct all the

mutations used in this study according to manufacturer’s protocols. The DNA

constructs were sequenced to confirm the mutation (DNA core, University of

Missouri) made on cDNA.

Electrophysiological recordings

Glass chips carrying the transfected cells were transferred to a chamber located on

the stage of an inverted microscope (IX51, Olympus). All the electrophysiological

data were recorded at room temperature with an EPC10 amplifier (HEKA,

Lambrecht/Pfalz, Germany). Borosilicate capillary glasses were used to produce

pipettes by using a Flaming/Brown-type micropipette puller (P97, Sutter

Instrument Co.). The pipettes were polished with a homemade microforge before

experiments. In the bath solution, the resistance of pipettes for patch-clamp

experiments was 2 to 4 MΩ. When the seal resistance was > 40 GΩ, membrane

patches were excised into an inside-out mode. After excision, 25 IU PKA and 2.75

mM ATP was perfused to the pipette until the CFTR current reached a steady state.

10 IU PKA was added to all other ATP-containing solutions applied thereafter to

maintain the phosphorylation level. The membrane potential was held at -60 mV

during recording unless specified. The data was filtered with an eight-pole Bessel

filter (LPF-8, Warner Instruments) with a 100 Hz cutoff frequency and digitized to a

computer at a sampling rate of 500 Hz. For clear data present, the inward current

Page 83: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

71

was inverted for clear data presentation. All inside-out patch experiments were

performed with a fast solution exchange perfusion system (SF-77B, Warner

Instruments). The dead time of solution change is ~30 ms (Tsai et al., 2009).

Chemicals and composition solutions

The 150 mM Cl- pipette solution contained (in mM): 140 methyl-D-glucamine

chloride (NMDG-Cl), 2 MgCl2, 5 CaCl2, and 10 HEPES (pH 7.4 with NMDG). The 375

mM Cl- pipette solution contained (in mM): 360 methyl-D-glucamine chloride

(NMDG-Cl), 2 MgCl2, 5 CaCl2, and 10 HEPES (pH 7.4 with NMDG). Cells were

perfused with a bath solution containing (in mM): 145 NaCl, 5 KCl, 2 MgCl2, 1 CaCl2,

5 glucose, 5 HEPES and 20 sucrose (pH 7.4 with NaOH). For inside-out configuration,

the 150 mM Cl- perfusion solution contained (in mM): 150 NMDG-Cl, 2 MgCl2, 10

EGTA and 8 Tris (pH 7.4 with NMDG), the 375 mM Cl- perfusion solution contained

(in mM): 375 NMDG-Cl, 2 MgCl2, 10 EGTA and 8 Tris (pH 7.4 with NMDG).

MgATP, Dithiothreitol (DTT) and PKA were purchased from Sigma-Aldrich. MgATP

and DTT were stored in 250 mM and 100 mM stock solution respectively at -20 °C.

2-trimethylaminoethyl (MTSET) was purchased from Toronto Research Chemicals

Inc. MESET was stored in 100 mM stock solution at -70°C and was diluted to

working concentration immediately before use. All chemicals were diluted to the

concentration indicated in each figure using perfusion solution and the pH was

adjusted to 7.4 with NMDG.

Page 84: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

72

Data analysis and statistics

Igor Pro program (Wavemetrics, Lake Oswego, OR) was used to calculate the

steady-state mean current amplitude and plot the histograms. Current relaxation

was fitted with a single exponential function using a Levenberg-Marquardt-based

algorithm within the Igor Pro program. A program developed by Csanady (Csanady,

2000) was used to measure channel kinetics on traces that contain 3 or less

channels. Kinetic modeling and computation simulations were described in

Kopeikin et al. (Kopeikin et al., 2010).

3-4. Results

Unique pattern of single-channel gating transitions in Cysless/R352C-CFTR

During our previous studies in scanning the pore lining residues in the 6th

transmembrane segment of TMDs (TM6), a unique feature of Cysless/R352C-CFTR

caught our attentions. Single channel recordings of this CFTR mutant revealed two

different open states with distinct single-channel amplitude (Figure 3-1A). However,

removing the positive charge at this position also reduced the single-channel

amplitude by half, and thus the signal to noise ratio was too compromised to extract

useful information about the transitions between these two conductance states from

our recordings. Since this positive charge was proposed to concentrate local

chloride ions in the internal vestibule of the pore (Bai et al., 2010), we reasoned that

increasing [Cl-] should overcome this technical hurdle. To our delight, elevating the

Page 85: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

73

Figure 3-1

Figure. 3-1. Cysless/R352C-CFTR reveals two different open states with

distinct conductance level. (A) Five representative traces and amplitude

604020

0

604020

0

80604020

0

80604020

0

604020

0

160120

8040

0

120

80

40

0

160120

8040

0

120

80

40

0

A

B

150 mM Cl

375 mM Cl

1 s

0.2

pA

C

O2O1

C

O2

O1

0.3

pA

1 s

1 2

5

3

4

C

O2

O1

C

O2

O1

C

O2

O1

C O1 O2

C

O2O1

C

O2O1

C

O2O1

C

O2O1

0.20 pA

0 0.2 0.4 pA

Co

un

tsC

ou

nts

Page 86: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

74

histograms from a patch that contained only one Cysless/R352C-CFTR channel. The

channel was fully phosphorylated (apply to all the following figures) and the traces

were recorded in the presence of 2.75 mM ATP. (B) Four representative traces and

amplitude histograms for Cysless/R352C-CFTR recorded in a condition similar to

that in (A), except that both pipette and perfusion solution contain 375 mM Cl- (see

Material and Methods for detail). High Cl- solution significantly augmented the

single-channel amplitude (0.17 ± 0.01 pA, n = 8 in 150 mM solutions versus 0.38 ±

0.01 pA, n = 11 in 375 mM, holding potential = - 60 mV for the larger conductance

state). The events in the red boxes show a preferred order of gating transitions (i.e.

C → O1 → O2 → C) and are expanded in Figure 3-3 for clearer view. Events following

the C → O1 → C transition are shown in the green boxes. One event showing the C →

O2 → C transition is highlighted in the blue box.

Page 87: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

75

chloride concentration in both the pipette and the perfusion solutions to 375 mM

nearly doubled the single-channel amplitude at – 60 mV holding potential and

significantly improved the signal to noise ratio (Figure 3-1B). Under such condition,

two open states with distinct single-channel amplitude (O1 and O2 states, referred

as the small and large conductance respectively) can be clearly defined in the

amplitude histogram (Figure 3-1B, next to each trace).

Like WT channels, this CFTR mutant opens into bursts with a lifetime of

hundreds of milliseconds to seconds and each opening burst is interrupted by very

brief closures of tens of milliseconds. Close inspections of the single channel traces

reveal an interesting pattern of gating transitions: in the opening bursts that harbor

both O1 and O2 states (e.g., events in red boxes in Figure 3-1B), when the channel

leaves the closed state, it often opens into the smaller O1 state first; whereas when

the channel returns to the closed state, it preferentially departs from the larger O2

state. Thus, the transition between O1 and O2 seems to be unidirectional as much

more C → O1 → O2 →C events were seen than C → O2 → O1 →C events (Table 3-1).

This asymmetry in the O1 ↔ O2 transition indicates a violation of detailed balance

resulted from a free energy input that drives the gating transition in a preferred

order as first proposed for CLC-0 (Richard and Miller, 1990). Similar gating

transition preference for WT-CFTR was reported by Gunderson and Kopito

(Gunderson and Kopito, 1995) under a special experimental condition (see

Discussion for details). For CFTR, the best candidate for the free energy source is

ATP hydrolysis as there is preponderant evidence that ATP hydrolysis fuels CFTR’s

gating cycle (Vergani et al., 2003; Bompadre et al., 2005b; Csanady et al., 2010). In

Page 88: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

76

Table 3-1

Table 3-1. Summary of opening events by different gating patterns in three

CFTR mutants. Five different gating patterns are illustrated on the top of the table.

The number of events and percentage for each pattern are displayed. The

corresponding CFTR mutants and ATP concentrations are shown on the left. #, note

that the (O1-O2)n category includes events that contain at least one O2 → O1

transition (such as C → O1 → O2 → O1 → O2 → C, C → O2 → O1 → O2 → C, C → O1 →

O2 → O1 → C, etc.).

O1 - O2 O1 O2 - O1O2 (O1 - O2)n

R352C

Cysless/R352C

R352C/W401F

Total

2.75 mM ATP

µ100 M ATP

2.75 mM ATP

100 M ATPµ

2.75 mM ATP

100 M ATPµ

834 (55%) 301 (20%) 173 (11%) 39 (3%) 169 (11%)

1246 (59%) 406 (19%) 281 (13%) 45 (2%) 121 (6%)

1516 (100%)

2099 (100%)

720 (45%) 290 (18%) 175 (11%) 42 (3%) 375 (23%) 1602 (100%)

733 (44%) 326 (19%) 122 (7%) 28 (2%) 474 (28%) 1683 (100%)

663 (56%) 216 (18%) 137 (12%) 32 (3%) 128 (11%) 1176 (100%)

1189 (54%) 367 (17%) 337 (15%) 60 (3%) 232 (11%) 2185 (100%)

#

Page 89: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

77

addition, we also observed opening bursts displaying only one conductance level, i.e.,

C → O1 → C (events in green boxes in Figure 3-1B) and C → O2 → C (events in blue

box in Figure 3-1B).

Notably, our results shown in Figure 3-1 are somewhat different from those

reported by Cui et al. (Cui et al., 2008). They showed that in addition to O1 and O2

(named s1 and s2 respectively in Cui et al., 2008), R352C-CFTR occasionally transits

to a full conductance level that is not different from that of WT-CFTR. In 4 patches

with ~ 25 minutes of overall recording time, we did not observe full conductance for

this mutant. The exact reason for this discrepancy remains unclear. It is possible

that the full conductance state, if exists, becomes extremely unstable in our

experimental system so that those events were filtered out. Alternatively, the

recordings shown in Cui et al. (2008) may actually contain more than one channel as

the mutation incidentally decreases the conductance by 50%.

ATP hydrolysis drives the O1 → O2 transition

Although our initial observations were made with Cysless/R352C-CFTR, this

unique pattern of gating transitions was also seen when we introduced the R352C

mutation into the WT background (Figure 3-2A and Table 3-1). One notable

difference between R352C- and Cysless/R352C-CFTR channels is that a higher

percentage of C → O1 → O2 → C transitions were present under the WT

background (Table 3-1). The significance of this finding will be elucidated below.

Page 90: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

78

Figure 3-2

Figure 3-2. Hydrolysis triggers the O1 → O2 transition. (A) Four representative

traces and amplitude histograms show the gating pattern of R352C-CFTR channel in

the presence of 2.75 mM ATP. The events in blue boxes are expanded in Figure 3-3

for clearer view. (B and C) Representative traces and amplitude histograms for

R352C/E1371S-CFTR in the presence (B) or absence (C) of 2.75 mM ATP. The traces

120

80

40

0

1500

1000

500

0

120100

80604020

0

160120

8040

0

150100

500

200150100

500

200150100

500

A R352C

C

O1O2

C

O1O2

C

O1O2

C

O1O2

C

O1O2

C

O1O2

B R352C/E1371S +ATP

C R352C/E1371S w/o ATP

C O1 O2

C O1 O2

Co

un

tsC

ou

nts

C O1 O2

1 s

5 s

1 s

0.4

pA

0.4

pA

0.4

pA

Co

un

ts

0 0.3 pA

0 0.4 pA

0 0.3 pA

1 2

3

4

5

C

O1O2

Page 91: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

79

were recorded in 375 mM pipette and perfusion solution (applied to Figures 3 and 4

as well). Similar observations were made in xx patches.

Page 92: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

80

To further test our hypothesis that the dominant O1 → O2 transition versus

O2 → O1 transition is the result of ATP hydrolysis, we engineered the E1371S

mutation into R352C-CFTR to abolish ATP hydrolysis (Vergani et al., 2003;

Bompadre et al., 2005b) and recorded ATP-dependent opening events. In Figure 3-

2B, each of the long-lasting ATP-induced opening bursts exhibits the O1

conductance level without transiting to O2 prior to channel closure (Figure 3-2B).

This lack of transitions to the O2 state for ATP-induced opening bursts is not

because the mutation, E1371S, eliminates the O2 state since in the absence of ATP,

the channel opens predominantly into bursts of a larger conductance level

corresponding to that of the O2 state (Figure 3-2C). Of note, in contrast to long-lived

ATP-dependent openings (Figure 3-2B), these spontaneous ATP-independent

openings were short-lived with opening durations in hundreds of milliseconds

(Figure 3-2C). The results presented so far bear three important implications: first,

the O1 → O2 transition signiaies an irreversible step associated with ATP hydrolysis.

Second, the C → O1 transition likely realects openings that results from ATP-induced

NBD dimerization (Vergani et al., 2005) not only because this transition dominates

when ATP is present, but also because this transition is usually followed by the

“hydrolysis step”, O1 → O2. Third, both C ↔ O1 and C ↔ O2 transitions are

reversible; based on the arguments made above, we propose that the O1 → C would

reflect non-hydrolytic closing and C ↔ O2 represent ATP-independent spontaneous

opening and closing of the gate. A simplified scheme (scheme 1) summarizes the

idea that the gating transition pattern observed in R352C-CFTR represents a cyclic

Page 93: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

81

steady state in which ATP hydrolysis drives a “clockwise” movement around the

state diagram (c.f., Richard and Miller, 1990; Gunderson and Kopito, 1995).

Non-strict coupling between ATP hydrolysis and gating cycle

If the biochemical interpretation of scheme 1 is valid, we conclude that each

C → O1 → O2 → C event (i.e., one gating cycle) in the single-channel recording

reflects hydrolysis of one ATP molecule. Table 3-1 shows that for both

Cysless/R352C and R352C mutant channels, C → O1 → O2 → C is the prevailing

transition; thus most gating events indeed follow the long-held one-to-one

stoichiometry between the gating cycle and ATP hydrolysis cycle. Interestingly

however, we also found that a significant number of opening bursts contains more

than one O1 → O2 transition (Figure 3-3, Table 3-1), implicating that hydrolysis of

more than one ATP molecule does take place within an opening burst. The existence

of O2 → O1 transition within an opening burst suggests that the post-hydrolytic

open state (O2) can re-enter the pre-hydrolytic O1 state before closing. Structurally,

it implies that the O2 state has already had its ABP2 vacated so that it is accessible

to ATP for initiating another hydrolysis cycle. This scenario predicts that the

C

O1O2

Scheme 1

hydrolysis

Page 94: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

82

Figure 3-3.

Figure 3-3. Non-strict coupling between ATP hydrolysis and gating cycle.

Opening events of Cysless/R352C- (A) or R352C-CFTR (B) containing one (Events #

1, 3, 4 in (A) and # 1-4 in (B)) or more (Events # 2, 5 in (A) and # 5 in (B)) O1 → O2

transitions. Events were extracted from traces in Figure 3-1B and 3-2A. Chosen

events were specified in boxes and numbered in Figure 3-1B and 3-2A.

1 2

3 4 5

400 ms

0.4

pA

A Cysless/R352C

100 ms

0.4

pA

1 2 3

4 5

B R352C

C

O1O2

C

O1O2

C

O1O2

C

O1O2

Page 95: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

83

frequency of O2 → O1 transition should be [ATP] dependent. Consistent with this

idea, fewer “re-entry”events (i.e., O1 → O2 → O1 → O2 transitions) were found at

lower [ATP] (Table 3-1). Based on this finding, a new pathway is amended to

scheme 1 to illustrate the re-entry events (O2 → O2 ATP → O1) in the presence of

ATP (scheme 2). These two newly added kinetic steps should be reversible as we

found no reason to presume an infusion of free energy in any of them.

Before we accept this new scheme that implicates a non-strict coupling

between the gating cycle and ATP hydrolysis cycle, we consider an alternative

scenario that may uphold the more parsimonious theory of one-to-one

stoichiometry. Since the limited bandwidth of our recordings renders events < 3 ms

(t10-90 = 0.3 / filter frequency) “invisible”, we cannot eliminate the possibility of the

existence of a very brief closure right after each O1 → O2 transition in those

opening bursts presumed to embody multiple rounds of ATP hydrolysis. However,

these non-discernible closures, if they exist, are not the same as the original closed

state (C in scheme 1) for the following reasons, (i) the closed state of R352C-CFTR

marked with stars in Figure 3-1 assumes a very long lifetime (~1 s , Figure 3-S1, vs.

300 – 400 ms for WT-CFTR); thus the probability of having this state with a lifetime

C

O1O2 Scheme 2hydrolysis

O2 ATP

Page 96: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

84

of < 3 ms buried in an opening burst is extremely small (< 0.003), (ii) for this idea to

be valid, one has to propose that ATP can open this presumed brief closed channel

at a rate that is 2 – 3 order faster than it does to the original closed state, and (iii) as

shown in the next section, multiple rounds of O1 → O2 transition can take place

within an opening burst in conditions when the O2 state is stabilized—a theme

predicted by scheme 2.

Modulation of the re-entry frequency by stabilizing the O2 state

An important implication of scheme 2 is that the lifetime of the O2 state plays

a crucial role in determining the frequency of the re-entry events (i.e., opening

bursts with multiple O1 → O2 transitions). At a given [ATP], the longer the channel

stays in the O2 state, the more likely an ATP would act to shove the channel back to

the O1 state. This idea is first supported simply by comparing the Cysless/R352C-

and R352C-CFTR results shown in Table 3-1. The Cysless/R352C mutant channel

has an O2 state that is about four times as long as that in the R352C-CFTR (Figure 3-

S2). Correspondingly, the percentage of opening bursts encompassing more than

one O1 → O2 transition is higher in Cysless/R352C (Table 3-1). Since the post-

hydrolytic O2 state also represents spontaneous opening of the closed channel in

the absence of ATP (C → O2 in scheme 2), to further support scheme 2, all we need

here is a channel that exhibits robust ATP-independent gating.

Page 97: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

85

Another serendipitous discovery in our lab grants us this opportunity. In Bai

et al. (2010), we showed that after Cysless/I344C-CFTR is modified by MTSET, the

Po of this CFTR mutant in the presence of ATP becomes virtually 1 (Bai et al., 2010).

More intriguingly, even minutes after ATP washout, the Po remains > 0.5 (compared

to < 0.01 for WT-CFTR (Bompadre et al., 2005b)) with visibly long opening bursts.

Based on scheme 2, we predict that, when removing the positive charge at position

352 in Cysless/I344C-CFTR, these long ATP-independent openings should show up

as the larger O2 state, and that, in the presence of ATP, an opening bursts could

contain numerous O1 → O2 transitions. Figure 3-4 confirms these predictions.

We introduced the R352Q mutation into the Cysless/I344C-CFTR channel.

Before MESET modification, Cysless/I344C/R352Q mutant channels behaved

similarly as Cysless /R352C-CFTR in the presence of ATP (Figure 3-4A). Two

distinct conductance levels representing the O1 and O2 states respectively can be

readily observed in the amplitude histogram. The preferred order of gating

transition, C → O1 → O2 → C, as well as opening bursts with the re-entry events

was also preserved (expanded traces in Figure 3-4A).

After MTSET modification of Cysless/I344C/R352Q, we indeed observed

robust ATP-independent openings (Figure 3-4B) with an open lifetime of 1.03 ± 0.30

s (n = 10), which is ~ 5 fold longer than the mean lifetime of the O2 state for

Cysless/R352C-CFTR. In addition, the single-channel amplitude is increased by

MTSET modification, an interesting yet unsurprising result since residue 344,

located two helical turns extracellular to position 352, also lines the chloride

Page 98: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

86

Figure 3-4

Figure 3-4. Modulation of the re-entry frequency by stabilizing the O2 state.

Representative traces and amplitude histograms for Cysless/I344C/R352Q-CFTR

under the following condition: (A) in the presence of 2.75 mM ATP; (B-C) after

MTSET modification, in the absence (B) or presence (C) of 2.75 mM ATP. Trace in (D)

highlights the O2 → O1 transition when 2.75 mM ATP was applied to the patch after

Page 99: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

87

witnessing an ATP-independent channel opening. (E) The amplitude of O1 and O2

states of Cysless/I344C/R352Q-CFTR before (the left bar) or after (the right bar)

modified by MTSET. The O1/O2 ratio is 0.6 ± 0.02, n = 8 before MTSET modification

versus 0.68 ± 0.02, n = 11 after MTSET modification. P < 0.05.

Page 100: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

88

permeation pathway (Bai et al., 2010). Once the patch was treated with ATP, the

single-channel amplitude of the initial opening step (O1) is smaller than that of ATP-

independent openings (O2). But multiple back-and-forth transitions between two

levels of conductance (O1 ↔ O2) are in display for several seconds before the

opening burst is terminated into a long closure (Figure 3-4C). Of note, just like

unmodified channel, there is a preference of channel closing from the large

conductance state (O2). Thus, once the O2 state is stabilized, the channel is trapped

in the lower half of scheme 2 with repetitive counterclockwise cycles driven by ATP

binding and hydrolysis.

To further verify the existence of the O2 → O2 ATP → O1 transition depicted

in scheme 2, we applied ATP after witnessing the ATP-independent opening

occurred (Figure 3-4D). Within a second upon the application of ATP, the channel

transits from the O2 state directly to the O1 state without a discernible transition to

a closed state. Subsequently in the continuous presence of ATP, multiple rounds of

O1 → O2 → O1 transition similar to that described in Figure 3-4C are presented.

(Similar results were obtained in 6 patches.) These results reaffirm the idea that in

the absence of ATP, the channel opens spontaneously into the post-hydrolytic O2

state, which can re-enter the O1 state upon ATP binding. By comparing the single-

channel amplitude before and after MTSET modification, we also noted that adding

a positively charged adduct at position 344 increases the current amplitude of the

O1 state slightly more than it does on the O2 state (Figure 3-4E). The significance of

this finding will be discussed.

Page 101: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

89

W401F mutation promotes O1 → O2 transition

In our latest paper (Jih et al., 2012), using non-hydrolyzable ATP analogs (PPi

or AMP-PNP) as baits, we identified a post-hydrolytic state (state X), which, like the

O2 state in the current report, can re-enter the pre-hydrolytic open state (O1) upon

ATP binding to the vacant ABP2. Moreover, we also raised the possibility that state X

could be a transient open state (Jih et al., 2012). Interestingly, a conserved mutation

in NBD1 (W401F) can enhance the responsiveness of state X to ATP as well as non-

hydrolyzable ATP analogs for reasons yet to be elucidated. Nevertheless, the striking

functional similarities between state X in Jih et al. (2012) and the O2 state in the

current study prompt us to test the effect of the W401F mutation on R352C-CFTR.

Figure 3-S3 shows a representative single channel trace of W401F/R352C-CFTR.

Compared to R352C-CFTR (Figure 3-2), this double mutant also exhibits a preferred

order of the gating transition. Quantitative analysis of gating events indeed

demonstrates a higher percentage of gating events with re-entry transitions in

W401F/R352C-CFTR (Table 3-1).

Kinetics of the O1 and O2 states

We next quantified the kinetics of these two open states by constructing the

dwell time histograms for individual O1 and O2 states as well as for overall open

state (O1 + O2). The results reveal that the R352C mutation significantly shortens

the total open time (~ 150 ms, Figure 3-S2 vs. ~400 ms for WT-CFTR (Vergani et al.,

2003; Bompadre et al., 2005a)). In addition, the lifetimes of both O1 and O2 states

Page 102: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

90

are significantly prolonged by mutating all endogenous cysteine residues (i.e.,

Cysless). Prolonged O1 and O2 dwell times were also found in W401F/R352C-CFTR,

but to a less extent. The effect of Cysless and W401F mutations in prolonging the

open time of R352C mutant channels is consistent with the observations made for

the same mutants in the WT background (Bai et al., 2010; Tsai et al., 2010a).

3-5. Discussion

An accidental discovery of the R352C mutation grants us the opportunity to

actually “see”—in real time—ATP hydrolysis taking place during CFTR gating as the

ordered transition between two distinct levels of open channel conductance (O1 and

O2) indicates an input of the free energy from ATP hydrolysis. Based on statistical

analyses of this unique gating feature, we conclude that although most of the ATP-

induced opening bursts are, as proposed previously (Csanady et al., 2010), strictly

coupled to one round of ATP hydrolysis, a considerable amount of opening bursts do

entail hydrolysis of more than one ATP molecule. This casual violation of one-to-one

stoichiometry is not predicted by previous gating mechanisms and a new gating

model (Figure 3-5) is therefore proposed in an attempt to explain current as well as

many previously published results.

Before we discuss the new gating model and its structure/function

implications, we remind our readers that a similar ordered open-to-open transition,

in fact, was first reported in WT-CFTR more than a decade ago (Gunderson and

Kopito, 1995). In their heavily filtered (10 Hz) single channel recordings, Gunderson

Page 103: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

91

Figure 3-5

Figure 3-5. A new gating model expanded from scheme 2 for CFTR. ATP

hydrolysis provides a shortcut from O1 to O2 state. The C, O1, O2, O2 ATP states are

C

O1O2

ADP + Pi

ATP

O1'

C ATP C AD

O2 ATP

Page 104: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

92

the same as described in scheme 2. The O1’ state is added to represent ATP

hydrolysis. The C ATP state is added to represent ATP binding prior to NBD

dimerization. The reason for adding the C AD state is described in Discussion.

Page 105: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

93

and Kopito (1995) observed two levels of open state conductance that appeared

mostly in a C → O1 → O2 → C order in the presence of ATP. Following that study,

Ishihara and Welsh (Ishihara and Welsh, 1997) showed that the difference in

conductance between the O1 and O2 states was due to a different degree of blockage

by the anionic buffer, MOPS, used in the bathing solution. After we made our

discovery with R352C-CFTR, we recorded WT-CFTR under conditions described in

these early reports, but did not observe two levels of open channel current. Further

studies are needed to uncover the causes of such discrepancy. Nonetheless, the

mutant R352C does offer the advantage of observing transitions between the O1 and

O2 states with a much better temporal resolution necessary for a more thorough

microscopic kinetic analysis.

A new gating model

As described in Results, our data led to a simple 4-state gating scheme

(scheme 2). Here, by amending three additional states, two implicitly embedded in

the transitions and the other more speculative one supported by previous studies as

well as data presented below, we propose a new gating model, which is nearly as

concise as scheme 2 but is more comprehensible and can potentially bring some

fresh insights into the structural mechanism of CFTR gating (Figure 3-5).

Upon construction of this model, we incorporated two generally accepted

views on CFTR gating: gate opening is triggered by ATP-induced NBD dimerization;

ATP hydrolysis in ABP2 leads to a partial separation of the NBD dimer and closes

Page 106: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

94

the gate (Vergani et al., 2005; Csanady et al., 2010; Tsai et al., 2010b). The basic

framework of this model is composed of three experimentally detectable states,

namely the C, O1 and O2 states, as well as the transitions between them as described

above. We reasoned that any ligand-induced conformational changes likely consist

of at least two steps: ligand binding and subsequent conformational change.

Therefore, for each ATP-dependent transition, C → O1 and O2 → O1 in scheme 2, it is

justifiable to insert at least one additional state in between to represent the ATP-

bound states, C ATP and O2 ATP. Since ATP hydrolysis is associated with the O1 →

O2 transition, we added a transition state depicting ATP hydrolysis. (Conceptually,

there should be more states representing kinetic steps for separation of the NBD

dimer and dissociation of the hydrolytic products.) As for the speculative C ATP

dimer state (simplified as C AD in Figure 3-5) at the upper right of the model, there

is really no direct evidence establishing its existence because this putative state is a

silent state (it doesn’t give out current) and its ATP binding sites are inaccessible to

probes such as nucleotide analogs (Tsai et al., 2009; Tsai et al., 2010b; Jih et al.,

2012). Thus, its existence can only be inferred. Indeed, through thermodynamic

analyses of CFTR gating, Csanady et al. (Csanady et al., 2006) proposed that NBD

dimerization occurs prior to gate opening. Next, as our data suggest that gate

closure (O2 → C) and the separation of NBD dimer (O1 → O2) do not occur

concurrently, it seems unlikely that opening of the gate and formation of the NBD

dimer have to be synchronized during the opening process. Furthermore, adding

this closed state into the model pleasingly explains some microscopic and

macroscopic kinetic data presented below.

Page 107: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

95

The idea that opening of CFTR by ATP involves three distinct steps, ATP

binding (C → CATP), NBD dimerization (CATP → CAD) and gate opening (CAD →

O1), predicts that mutations that disrupt either of these processes can decrease the

apparent opening rate. Indeed, mutations of the amino acid residue that directly

interacts with ATP have been reported to decrease the opening rate in an ATP-

dependent manner (Zhou et al., 2006); further, mutations located at the NBD dimer

interface have also been shown to lower the opening rate (Vergani et al., 2005).

Here, we show that mutating R352 also leads to a reduced opening rate (1 s-1 for

R352C-CFTR versus 2.5 s-1 for WT-CFTR (Vergani et al., 2003; Bompadre et al.,

2005a)). Since the R352 residue, located in TMDs, is unlikely involved directly in

ATP binding or in NBD dimerization, we reckoned mutations at this position could

decrease the apparent opening rate by affecting the CAD → O1 transition, a step

expected to take place in CFTR’s TMDs. Kinetically, the presence of the CAD state

also dictates that the gate can close, probably very briefly, even if the NBD dimer has

not separated yet (i.e., O1 → CAD). These events reflecting channel shuttling

between O1 and C AD, if exist, should be embedded in an opening burst. However,

“seeing” these events will be challenging as they are likely to be masked by the great

number of intraburst, voltage-dependent flickery closures, some of which are

probably caused by a pore blocking mechanism (Zhou et al., 2001) and are normally

eliminated by various means during kinetic analysis to extract ATP-dependent

events (Csanady, 2000; Bompadre et al., 2005b). Consistent with this idea, however,

previous studies (Bompadre et al., 2005b) did identify two distinguishable

populations of intraburst closed events embedded in a long opening burst of a

Page 108: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

96

hydrolysis-deficient mutant CFTR. It would be interesting to find out in the future

whether these events may signify events other than simple channel blockade.

The three-step opening process described above also means that closing of

the hydrolysis-deficient CFTR mutant channels upon removal of ATP will go through

three reverse steps, O1 → CAD, CAD → CATP, and CATP → C. It follows that

mutations that decrease the rate of NBD dimerization (CATP → CAD), e.g., T1246N

in Vergani et al. (2005), or those that reduce the rate for CAD → O1 (presumably the

R352C mutation), will accelerate the decay rate of macroscopic currents in

hydrolysis-deficient mutants upon ATP washout. Indeed, Vergani et al. (2005)

showed that the open burst time of the T1246N/K1250R-CFTR is much shorter than

that of K1250R-CFTR. Here, in Figure 3-6, we show that both R352C and T1246N

mutations significantly shorten the locked-open time of the hydrolytic deficient

E1371S-CFTR (Figure 3-6). The locked-open time is further shortened when

combining R352C and T1246N mutations together (Figure 3-6 D, E), suggesting that

the two mutations affects two different kinetic steps as described above.

Functional implications of the new gating model

One appealing feature of the proposed model is its symmetry when the

hydrolysis pathway is taken away. The whole scheme is then divided into three

reversible processes: ligand binding/unbinding (first horizontal transitions), NBD

association/dissociation (second horizontal transitions) and gate opening/closing

(three vertical transitions). One thermodynamic outcome from this equilibrium

Page 109: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

97

Figure 3-6

Figure 3-6. R352C shorten the locked-open time of hydrolytic deficient CFTR

mutant. Macroscopic current traces of E1371S-CFTR (A), R352C/E1371S-CFTR (B),

T1246N/E1371S-CFTR (C) and R352C/T1246N/E1371S-CFTR (D). In each pane, the

current was activated by PKA + ATP to a steady state and then PKA and ATP were

removed to allow the current to decay. The current relaxation was fitted with a

single exponential function resulting in the relaxation time constant for each mutant:

65.6 ± 10.1 s, n = 8 for E1371S-CFTR, 4.9 ± 0.8 s, n = 12 for R352C/E1371S-CFTR, 7.8

± 1.6 s, n = 7 for T1246N/E1371S-CFTR, and 2.27 ± 0.27 s, n = 6 for

R352C/T1246N/E1371S-CFTR. (E) Bar chart summarizing the results in (A) to (D).

80

60

40

20

0A B C D

A E1371S B R352C/E1371S

2.75 mM

ATP

C T1246N/E1371S D R342C/T1246N/E1371S

2.75 mMATP

E

(s)

τ

*

(7)

(12)(7)

(6)

**

#

#

100 s

5 p

A

20

pA

20 s

10

pA

2 p

A

20 s

100 s

Page 110: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

98

* indicates P < 0.05 compared to E1371S. # indicates P < 0.05 between two

designated data.

Page 111: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

99

scheme is the conclusion that the states harboring an NBD dimer (O1 and CAD) are

most energetically stable in the presence of ATP. (See Supplemental Discussion for a

more comprehensive elaboration.) Thus, ATP hydrolysis simply provides a shortcut

for the channel to rapidly escape the absorbing states to carry on the gating cycle.

By assigning NBD dimerization and gate opening as two distinct steps, this

model also explicitly states that they are energetically coupled: NBD dimerization

favors gate opening and vice versa. This interpretation then distinguishes CFTR

from other conventional ligand-gated channels, which couple the ligand binding

energy to the gating conformational change. Instead, CFTR operates as an NBD

dimerization-gated channel, although the fundamental principle is the same (Monod

et al., 1965). CFTR appears like a ligand-gated channel simply because ATP binding

accelerates the dimerization process. It is hence not surprising that mutations that

facilitate the formation of the NBD dimer also promote gate opening even in the

absence of ATP (Szollosi et al., 2010). Furthermore, when two NBDs are cross-linked

to form a head-to-tail dimer, macroscopic fluctuations reflecting opening and

closing of the gate can be discerned in ATP-independent currents (Mense et al.,

2006).

One interesting functional consequence out of the model is a novel

mechanism for prolonging the opening burst. Contrary to the strict coupling

hypothesis, which allows only two ways to lengthen the opening burst of CFTR: by

delaying either ATP hydrolysis or NBD dimer separation, our new model offers an

alternative mechanism: by promoting the re-entry pathway (O2 → O2ATP → O1 in

Page 112: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

100

Figure 3-5). As shown in Figure 3-4C, the opening burst can be dramatically

extended when the frequency of re-entry is increased by a stabilized O2 state.

Similarly, we previously reported an [ATP]-dependent increase of the opening burst

duration for W401F-CFTR (Jih et al., 2012), presumably because the mutation also

increases the re-entry frequency. Indeed, a higher percentage of re-entry events was

observed with the W401F/R352C mutant in the presence of ATP compared to

R352C-CFTR (Table 3-1, Figure 3-S3). Interestingly however, the lifetime of the O2

state is not greatly increased by the W401F mutation (Figure 3-S3C). This seemingly

puzzling observation actually makes sense as the model predicts that mutations in

NBDs could also boost the frequency of re-entry by increasing the rate of O2 → O2

ATP without the necessity of stabilizing the O2 state. Although more studies are

needed to unravel the functional significance of the newly discovered re-entry

mechanism in modulating CFTR gating, the mere existence of this pathway raises

the possibility that pharmacological reagents known as CFTR potentiators may

work through this novel mechanism. In particular, it is interesting to note that a

CFTR potentiator, Vx-770, lately approved to be used clinically for the treatment of

cystic fibrosis patients (Accurso et al., 2010; Ramsey et al., 2011), significantly

increases the Po of both WT-CFTR and disease-associated mutant channels, such as

G551D-CFTR, by prolonging the opening burst duration (Schnitzer et al., 2000; Yu et

al., 2012). In this regard, the R352C mutant may turn up to be an invaluable tool for

us to address the mechanism of action for this clinically important drug.

Page 113: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

101

R352’s role in channel conductance

One may ask: why removing the positive charge at position 352 results in a

state-dependent single channel conductance? Although a definitive answer to this

question awaits more studies, some speculations can be made based on what we

know about the structural elements of CFTR’s pore and the gating motion in TMDs.

For whatever mechanism accounting for this unique behavior, one needs to take

into considerations several basic properties of R352C-CFTR. First, even for the

larger O2 state, the conductance of R352C-CFTR is ~ 50 % of the WT (Figure 3-1),

suggesting that the positive charge at this position may play a crucial role in

attracting Cl- ions as proposed previously (Cui et al., 2008; Bai et al., 2010) or

alternatively may neutralize a pore-lining negatively charged residue such as D993

proposed by Cui et al. (2006). Second, since the O1 → O2 transition is coupled to

ATP hydrolysis and separation of the NBD dimer, the functional role of R352 in

TMDs must depend on the status of NBDs (pre-hydrolytic versus post-hydrolytic

and/or monomeric versus dimeric). This differential role of the positive charge at

position 352 in O1 versus O2 state is reinforced by the observation that adding a

positively charged adduct at position 344 under the R352Q background increases

O1 and O2 conductance to a different degree (Fig 4E) as if the positive charged

MTSET adduct partially restore the functional role of R352 residue in a state-

dependent manner. Thus, the positive charge at position 352 (and perhaps other

positions on the same phase of TM6) makes a greater contribution to the channel

conductance in the O1 state than in the O2 state. Our recent studies on CFTR’s

TMDs suggest that TM6 may undergo a translational as well as rotational movement

Page 114: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

102

during the open-closed transition (Bai et al., 2010, 2011). The conclusion that R352

assumes different roles in O1 and O2 states suggests that the motion of TM6 is not

just confined to the opening/closing process, and must be also present upon ATP

hydrolysis (and separation of NBDs) while the gate remains open. Thus, future

studies using this mutant may also shed light on the conformational changes in

CFTR’s TMDs during gating transitions.

A few more speculations

The central theme of the energetic coupling hypothesis is that both the NBDs

and TMDs may hold certain degree of autonomy to function on its own, whereas as

an integrated protein, the conformational change of one domain facilitates the

conformational change in the other domain (i.e., dimerization of NBDs facilitates

gate opening and vice versa). This autonomy of CFTR’s TMDs is exemplified by the

observation that CFTR assumes significant spontaneous openings in the absence of

ATP (see discussion in (Chen and Hwang, 2008)) and robust ATP-independent

gating can be seen in a CFTR construct with the entire NBD2 deleted (Wang et al.,

2010). Once we allow CFTR’s NBDs and TMDs to function independently, the overall

gating motion does not have to follow the widely held view that each NBD-TMD

complex moves concurrently as a rigid body (Ivetac et al., 2007; Khare et al., 2009).

Although the exact molecular motions of CFTR gating await further studies, here we

offer one possibility: NBD dimerization drags parts of the TMDs that are in

proximity to the NBD-TMD interface whereas the part of the TMDs with a physical

Page 115: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

103

gate of the pore remains relatively static. The resulting bending of the helixes in

TMDs may induce a molecular strain. It is the relaxation of this molecular strain that

opens the gate as first proposed by Csanady et al. (2006).

Next, one may wonder why CFTR, a channel evolved from ABC transporters,

is gated by a somewhat loose coupling mechanism between the energy-harvesting

NBDs and the pore-forming TMDs? This concern over the efficiency of energy

utilization is not likely an issue for CFTR as the consumption of a few more ATP

molecules in the re-entry pathway does not affect the transport efficiency as long as

the open channel conformation is maintained. But, if the same mechanism is

applicable to an active transporter, futile cycles of ATP hydrolysis are deemed to

happen. We argue however that a slight imperfection of energy utilization, as long as

it does not pose significant evolutionary disadvantage, may not jeopardize the

survival of an organism. In fact, it has been proposed that such “molecular slip” may

even serve as a safety valve that could actually benefit organisms as several ion-

motive ATPase have been shown to exhibit non-integral stoichiometry (Nelson et al.,

2002). Moreover, a deviation from ideal energy efficiency was proposed for the CLC

transporter (Feng et al., 2010) and also demonstrated for the kinesin motor (Crevel

et al., 1999; Schnitzer et al., 2000).

If the energetic coupling hypothesis proposed here can be generalized to all

ABC proteins—a big “if” we unhesitatingly admit, inevitably one has to conclude that

there is a certain degree of energy waste. One the other hand, the autonomy of

domains underscored by this hypothesis also mean that NBDs or TMDs, capable of

Page 116: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

104

functioning as an independent unit, could shuffle in the genomic pool and appear in

other proteins that are functionally distinct from the conventional ABC transporters.

Indeed, canonical NBD domains are found in the DNA repair protein, Rad50

(Hopfner et al., 2000; Hopfner et al., 2001) and energy coupling factor (ECF)-type

transporters (Erkens et al., 2011; ter Beek et al., 2011). Such domain shuffling

phenomenon may be essential for protein diversities found in all kingdoms of life

(Doolittle, 1995). There seems no reason to think ABC proteins do not fall into this

modus operandi in evolution.

In conclusion, capitalizing on the unique gating features of R352C-CFTR, we

were able to visualize ATP hydrolysis in real time for CFTR by simply monitoring

single-channel current traces. We found a non-integral stoichiometry between the

ATP hydrolysis cycle and the gating cycle. Based on this finding, we proposed a

gating model that features an energetic coupling between CFTR’s NBDs and the

TMDs. It would be certainly interesting to see if such a mechanism can be

generalized for other ABC transporters.

Page 117: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

105

3-6. Supplementary materials

Semi-quantitative analysis of an equilibrium gating scheme for hydrolysis-deficient

CFTR channels

As mentioned in Discussion, once ATP hydrolysis is abolished (e.g., in E1371S), the

transitions between the remaining six states in Figure 3-5 obey the principle of

microscopic reversibility (scheme S1), which means the product of the forward

rates has to be equal to the product of the reverse rates in a closed loop (e.g.,

K1F·K2F·K3F·K4F = K1R·K2R·K3R·K4R). Thus, the stability of the NBD dimer in the

open state would be represented by

������ =

���∙���∙�����∙���∙�� (equation S1)

Some of the values in equation S1 can be roughly estimated:

As demonstrated in a previous study (Bompadre et al., 2007), G551D-CFTR, a

disease-associated mutation, does not respond to ATP. Since the lack of a side chain

C ATP

O2 ATP

C AD

O1

K1F

K1R

K3R

K3F

K2RK2FK4FK4R Scheme S1

C

O2

K5RK5F

Page 118: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

106

at position 551 is critical for hydrogen bond formation between the negatively

charged phosphate group of ATP and the backbone nitrogen in the signature

sequence of an NBD dimer of ABC proteins (Dawson and Locher, 2006; Oldham et al.,

2007; Ward et al., 2007), one can safely infer that this complete abolition of ATP

responsiveness by the G551D mutation is not due to a deficit in ATP binding but to

an obliteration of NBD dimerization (i.e., C ATP → C AD transition). The fact that ATP

does not elicit measurable response in the G551D mutant channel then indicates

K5F/K5R ≅ K4F/K4R in scheme S1. Thus, CFTR is not a classical ligand-gated

channel but an NBD dimerization-gated channel instead.

Since the spontaneous openings occur rarely for WT-CFTR (K5R = 0.005 s-1) and the

closing rate of the spontaneous opening (K5F) is ~ 2 s-1 (Bompadre et al., 2005a;

Bompadre et al., 2005b), K5R/K5F = 0.0025 and the value of K4R/K4F must be

equally small. In the presence of saturating [ATP], the ATP-induce channel opening

(C ATP → C AD → O1) is a much faster process that takes place within hundreds of

milliseconds in WT-CFTR. K1F (C ATP → C AD) and K2F (C AD → O1) are required to

be faster than ~1 s-1 while the backward rate K1R (C AD → C ATP) should not be >>

K2F for actualizing the ATP-induced opening rate within hundreds of milliseconds.

On the other hand, channel closing in hydrolysis-deficient mutants after washing out

ATP is extremely slow (e.g., a relaxation time constant > 50 s for E1371S-CFTR)

(Bompadre et al., 2005a; Bompadre et al., 2005b). This suggests that either K2R or

K1R /K2F is very small. To actualize a relaxation time constant of ~50 s after

washing out ATP, considering that K2F should be ~1 s-1 or faster, the value of

K2R·(K1R /K2F) could be roughly approximated as ~0.02 s-1 or smaller.

Page 119: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

107

Taken together,

������ =

���∙���∙�����∙���∙�� = ���∙������ � ∙ �

���� ∙ ����� ≤ ~0.0005 .

That is, the NBD dimer of an open state is exceedingly stable without ATP hydrolysis.

Page 120: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

108

Figure 3-S1

Figure 3-S1 The interburst dwell time histogram for R352C-CFTR. The

histogram was fit with a single exponential function and yielded a time constant of

944 ms.

60

50

40

30

20

10

0

14121086420

x103 ms

Co

un

ts

Interburst duration ( )

Page 121: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

109

Figure 3-S2

Figure 3-S2. Dwell time histogram for O1, O2 state and opening burst in

Cysless/R352C- and R352C-CFTR. (A and B) Dwell time histograms of the O1, O2

states and the opening burst (O total) for Cysless/R352C-CFTR (A) and R352C-CFTR

(B). The histograms were fit with single exponential functions and the resulting time

constant is marked next to each fitting curve.

80

60

40

20

0

8006004002000

120

80

40

0

5004003002001000

80

60

40

20

0

10008006004002000

80

60

40

20

0

5004003002001000

50

40

30

20

10

0

2000150010005000

60

40

20

0

10008006004002000

A Cysless/R352C B R352C

O1 O1

O2O2

O total O total

Co

un

ts

Co

un

ts

Co

un

ts

Co

un

ts

Co

un

ts

Co

un

ts

Dwell time (ms) Dwell time (ms)

Dwell time (ms)Dwell time (ms)

Dwell time (ms)Dwell time (ms)

= 166 msτ = 84 msτ

= 218 msτ = 43 msτ

=556 msτ = 153 msτ

Page 122: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

110

Figure 3-S3

Figure 3-S3. Gating kinetics of W401F/R352C-CFTR. (A) Representative traces

and amplitude histograms for W401F/R352C-CFTR in the presence of 2.75 mM ATP.

Consistent with the observation for R352C-CFTR, most of the opening bursts follow

the C → O1 → O2 → C gating pattern. However, more re-entry events were seen in

W401F/R352C-CFTR (Table 3-1). (B) Sample opening events expanded from the

120

80

40

0

0.30.0

A

120

80

40

0

0.40.30.20.10.0

A

120

80

40

0

0.30.0

A

160

120

80

40

0

0.30.0

A

80

60

40

20

0

8006004002000

120

80

40

0

5004003002001000

80

60

40

20

0

5004003002001000

C

O1O2

A R352C/W401F

1

2

3 4

5

B

C O1 O2

1 2 3 4 5

C

Co

un

ts

Dwell time (ms) Dwell time (ms) Dwell time (ms)

O1 O2 O total

= 94 msτ = 62 msτ = 194 msτ

200 ms

0.4

pA

1 s

C

O1O2

C

O1O2

C

O1O2

0.4

pA

C

O1O2

Page 123: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

111

traces in (A). (C) Dwell time histograms of the O1, O2 states and the opening burst

(O total). The histograms were fit with single exponential functions and the time

constants are marked.

Page 124: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

112

3-7. References

Accurso, F.J., S.M. Rowe, J.P. Clancy, M.P. Boyle, J.M. Dunitz, P.R. Durie, S.D. Sagel, D.B.

Hornick, M.W. Konstan, S.H. Donaldson, R.B. Moss, J.M. Pilewski, R.C.

Rubenstein, A.Z. Uluer, M.L. Aitken, S.D. Freedman, L.M. Rose, N. Mayer-

Hamblett, Q. Dong, J. Zha, A.J. Stone, E.R. Olson, C.L. Ordonez, P.W. Campbell,

M.A. Ashlock, and B.W. Ramsey. 2010. Effect of VX-770 in persons with cystic

fibrosis and the G551D-CFTR mutation. N Engl J Med. 363:1991-2003.

Aller, S.G., J. Yu, A. Ward, Y. Weng, S. Chittaboina, R. Zhuo, P.M. Harrell, Y.T. Trinh, Q.

Zhang, I.L. Urbatsch, and G. Chang. 2009. Structure of P-glycoprotein reveals

a molecular basis for poly-specific drug binding. Science. 323:1718-1722.

Bai, Y., M. Li, and T.C. Hwang. 2010. Dual roles of the sixth transmembrane segment

of the CFTR chloride channel in gating and permeation. J Gen Physiol.

136:293-309.

Bai, Y., M. Li, and T.C. Hwang. 2011. Structural basis for the channel function of a

degraded ABC transporter, CFTR (ABCC7). J Gen Physiol. 138:495-507.

Bompadre, S.G., T. Ai, J.H. Cho, X. Wang, Y. Sohma, M. Li, and T.C. Hwang. 2005a.

CFTR gating I: Characterization of the ATP-dependent gating of a

phosphorylation-independent CFTR channel (DeltaR-CFTR). J Gen Physiol.

125:361-375.

Bompadre, S.G., J.H. Cho, X. Wang, X. Zou, Y. Sohma, M. Li, and T.C. Hwang. 2005b.

CFTR gating II: Effects of nucleotide binding on the stability of open states. J

Gen Physiol. 125:377-394.

Bompadre, S.G., Y. Sohma, M. Li, and T.C. Hwang. 2007. G551D and G1349D, two CF-

associated mutations in the signature sequences of CFTR, exhibit distinct

gating defects. J Gen Physiol. 129:285-298.

Carson, M.R., S.M. Travis, and M.J. Welsh. 1995. The two nucleotide-binding domains

of cystic fibrosis transmembrane conductance regulator (CFTR) have distinct

functions in controlling channel activity. J Biol Chem. 270:1711-1717.

Chen, T.Y., and T.C. Hwang. 2008. CLC-0 and CFTR: chloride channels evolved from

transporters. Physiol Rev. 88:351-387.

Crevel, I., N. Carter, M. Schliwa, and R. Cross. 1999. Coupled chemical and

mechanical reaction steps in a processive Neurospora kinesin. EMBO J.

18:5863-5872.

Csanady, L. 2000. Rapid kinetic analysis of multichannel records by a simultaneous

fit to all dwell-time histograms. Biophys J. 78:785-799.

Page 125: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

113

Csanady, L., A.C. Nairn, and D.C. Gadsby. 2006. Thermodynamics of CFTR channel

gating: a spreading conformational change initiates an irreversible gating

cycle. J Gen Physiol. 128:523-533.

Csanady, L., D. Seto-Young, K.W. Chan, C. Cenciarelli, B.B. Angel, J. Qin, D.T. McLachlin,

A.N. Krutchinsky, B.T. Chait, A.C. Nairn, and D.C. Gadsby. 2005. Preferential

phosphorylation of R-domain Serine 768 dampens activation of CFTR

channels by PKA. J Gen Physiol. 125:171-186.

Csanady, L., P. Vergani, and D.C. Gadsby. 2010. Strict coupling between CFTR's

catalytic cycle and gating of its Cl- ion pore revealed by distributions of open

channel burst durations. Proc Natl Acad Sci U S A. 107:1241-1246.

Cui, G., Z.R. Zhang, A.R. O'Brien, B. Song, and N.A. McCarty. 2008. Mutations at

arginine 352 alter the pore architecture of CFTR. J Membr Biol. 222:91-106.

Dawson, R.J., and K.P. Locher. 2006. Structure of a bacterial multidrug ABC

transporter. Nature. 443:180-185.

Doolittle, R.F. 1995. The multiplicity of domains in proteins. Annual review of

biochemistry. 64:287-314.

Erkens, G.B., R.P. Berntsson, F. Fulyani, M. Majsnerowska, A. Vujicic-Zagar, J. Ter

Beek, B. Poolman, and D.J. Slotboom. 2011. The structural basis of modularity

in ECF-type ABC transporters. Nat Struct Mol Biol. 18:755-760.

Feng, L., E.B. Campbell, Y. Hsiung, and R. MacKinnon. 2010. Structure of a eukaryotic

CLC transporter defines an intermediate state in the transport cycle. Science.

330:635-641.

Gadsby, D.C., T.C. Hwang, T. Baukrowitz, G. Nagel, M. Horie, and A.C. Nairn. 1994.

Regulation of CFTR channel gating. Jpn J Physiol. 44 Suppl 2:S183-192.

Gaudet, R., and D.C. Wiley. 2001. Structure of the ABC ATPase domain of human

TAP1, the transporter associated with antigen processing. EMBO J. 20:4964-

4972.

Gunderson, K.L., and R.R. Kopito. 1994. Effects of pyrophosphate and nucleotide

analogs suggest a role for ATP hydrolysis in cystic fibrosis transmembrane

regulator channel gating. J Biol Chem. 269:19349-19353.

Gunderson, K.L., and R.R. Kopito. 1995. Conformational states of CFTR associated

with channel gating: the role ATP binding and hydrolysis. Cell. 82:231-239.

Hollenstein, K., D.C. Frei, and K.P. Locher. 2007. Structure of an ABC transporter in

complex with its binding protein. Nature. 446:213-216.

Page 126: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

114

Hopfner, K.P., A. Karcher, L. Craig, T.T. Woo, J.P. Carney, and J.A. Tainer. 2001.

Structural biochemistry and interaction architecture of the DNA double-

strand break repair Mre11 nuclease and Rad50-ATPase. Cell. 105:473-485.

Hopfner, K.P., A. Karcher, D.S. Shin, L. Craig, L.M. Arthur, J.P. Carney, and J.A. Tainer.

2000. Structural biology of Rad50 ATPase: ATP-driven conformational

control in DNA double-strand break repair and the ABC-ATPase superfamily.

Cell. 101:789-800.

Hwang, T.C., G. Nagel, A.C. Nairn, and D.C. Gadsby. 1994. Regulation of the gating of

cystic fibrosis transmembrane conductance regulator C1 channels by

phosphorylation and ATP hydrolysis. Proc Natl Acad Sci U S A. 91:4698-4702.

Ishihara, H., and M.J. Welsh. 1997. Block by MOPS reveals a conformation change in

the CFTR pore produced by ATP hydrolysis. Am J Physiol. 273:C1278-1289.

Ivetac, A., J.D. Campbell, and M.S. Sansom. 2007. Dynamics and function in a bacterial

ABC transporter: simulation studies of the BtuCDF system and its

components. Biochemistry. 46:2767-2778.

Jih, K.Y., Y. Sohma, M. Li, and T.C. Hwang. 2012. Identification of a novel post-

hydrolytic state in CFTR gating. J Gen Physiol.

Karpowich, N., O. Martsinkevich, L. Millen, Y.R. Yuan, P.L. Dai, K. MacVey, P.J. Thomas,

and J.F. Hunt. 2001. Crystal structures of the MJ1267 ATP binding cassette

reveal an induced-fit effect at the ATPase active site of an ABC transporter.

Structure. 9:571-586.

Khare, D., M.L. Oldham, C. Orelle, A.L. Davidson, and J. Chen. 2009. Alternating access

in maltose transporter mediated by rigid-body rotations. Mol Cell. 33:528-

536.

Kopeikin, Z., Y. Sohma, M. Li, and T.C. Hwang. 2010. On the mechanism of CFTR

inhibition by a thiazolidinone derivative. J Gen Physiol. 136:659-671.

Lewis, H.A., S.G. Buchanan, S.K. Burley, K. Conners, M. Dickey, M. Dorwart, R. Fowler,

X. Gao, W.B. Guggino, W.A. Hendrickson, J.F. Hunt, M.C. Kearins, D. Lorimer,

P.C. Maloney, K.W. Post, K.R. Rajashankar, M.E. Rutter, J.M. Sauder, S. Shriver,

P.H. Thibodeau, P.J. Thomas, M. Zhang, X. Zhao, and S. Emtage. 2004.

Structure of nucleotide-binding domain 1 of the cystic fibrosis

transmembrane conductance regulator. EMBO J. 23:282-293.

Mense, M., P. Vergani, D.M. White, G. Altberg, A.C. Nairn, and D.C. Gadsby. 2006. In

vivo phosphorylation of CFTR promotes formation of a nucleotide-binding

domain heterodimer. EMBO J. 25:4728-4739.

Page 127: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

115

Monod, J., J. Wyman, and J.P. Changeux. 1965. On the Nature of Allosteric Transitions:

A Plausible Model. J Mol Biol. 12:88-118.

Nelson, N., A. Sacher, and H. Nelson. 2002. The significance of molecular slips in

transport systems. Nature reviews. Molecular cell biology. 3:876-881.

Oldham, M.L., D. Khare, F.A. Quiocho, A.L. Davidson, and J. Chen. 2007. Crystal

structure of a catalytic intermediate of the maltose transporter. Nature.

450:515-521.

Ramsey, B.W., J. Davies, N.G. McElvaney, E. Tullis, S.C. Bell, P. Drevinek, M. Griese, E.F.

McKone, C.E. Wainwright, M.W. Konstan, R. Moss, F. Ratjen, I. Sermet-

Gaudelus, S.M. Rowe, Q. Dong, S. Rodriguez, K. Yen, C. Ordonez, and J.S. Elborn.

2011. A CFTR potentiator in patients with cystic fibrosis and the G551D

mutation. N Engl J Med. 365:1663-1672.

Richard, E.A., and C. Miller. 1990. Steady-state coupling of ion-channel

conformations to a transmembrane ion gradient. Science. 247:1208-1210.

Schnitzer, M.J., K. Visscher, and S.M. Block. 2000. Force production by single kinesin

motors. Nat Cell Biol. 2:718-723.

Szollosi, A., D.R. Muallem, L. Csanady, and P. Vergani. 2011. Mutant cycles at CFTR's

non-canonical ATP-binding site support little interface separation during

gating. J Gen Physiol. 137:549-562.

Szollosi, A., P. Vergani, and L. Csanady. 2010. Involvement of F1296 and N1303 of

CFTR in induced-fit conformational change in response to ATP binding at

NBD2. J Gen Physiol. 136:407-423.

ter Beek, J., R.H. Duurkens, G.B. Erkens, and D.J. Slotboom. 2011. Quaternary

structure and functional unit of energy coupling factor (ECF)-type

transporters. J Biol Chem. 286:5471-5475.

Tsai, M.F., K.Y. Jih, H. Shimizu, M. Li, and T.C. Hwang. 2010a. Optimization of the

degenerated interfacial ATP binding site improves the function of disease-

related mutant cystic fibrosis transmembrane conductance regulator (CFTR)

channels. J Biol Chem. 285:37663-37671.

Tsai, M.F., M. Li, and T.C. Hwang. 2010b. Stable ATP binding mediated by a partial

NBD dimer of the CFTR chloride channel. J Gen Physiol. 135:399-414.

Tsai, M.F., H. Shimizu, Y. Sohma, M. Li, and T.C. Hwang. 2009. State-dependent

modulation of CFTR gating by pyrophosphate. J Gen Physiol. 133:405-419.

Page 128: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

116

Vergani, P., S.W. Lockless, A.C. Nairn, and D.C. Gadsby. 2005. CFTR channel opening

by ATP-driven tight dimerization of its nucleotide-binding domains. Nature.

433:876-880.

Vergani, P., A.C. Nairn, and D.C. Gadsby. 2003. On the mechanism of MgATP-

dependent gating of CFTR Cl- channels. J Gen Physiol. 121:17-36.

Wang, W., J. Wu, K. Bernard, G. Li, G. Wang, M.O. Bevensee, and K.L. Kirk. 2010. ATP-

independent CFTR channel gating and allosteric modulation by

phosphorylation. Proc Natl Acad Sci U S A. 107:3888-3893.

Ward, A., C.L. Reyes, J. Yu, C.B. Roth, and G. Chang. 2007. Flexibility in the ABC

transporter MsbA: Alternating access with a twist. Proc Natl Acad Sci U S A.

104:19005-19010.

Yu, H., B. Burton, C.J. Huang, J. Worley, D. Cao, J.P. Johnson, Jr., A. Urrutia, J. Joubran, S.

Seepersaud, K. Sussky, B.J. Hoffman, and F. Van Goor. 2012. Ivacaftor

potentiation of multiple CFTR channels with gating mutations. Journal of

cystic fibrosis : official journal of the European Cystic Fibrosis Society.

Zhou, Z., S. Hu, and T.C. Hwang. 2001. Voltage-dependent flickery block of an open

cystic fibrosis transmembrane conductance regulator (CFTR) channel pore. J

Physiol. 532:435-448.

Zhou, Z., X. Wang, H.Y. Liu, X. Zou, M. Li, and T.C. Hwang. 2006. The two ATP binding

sites of cystic fibrosis transmembrane conductance regulator (CFTR) play

distinct roles in gating kinetics and energetics. J Gen Physiol. 128:413-422.

Page 129: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

117

CHAPTER 4

THE MOST COMMON CYSTIC FIBROSIS

ASSOCIATED MUTATION DESTABILIZES THE

DIMERIC STATE OF THE NUCLEOTIDE-BINDING

DOMAINS OF CFTR

This chapter has been modified from my manuscript published in J. Physiol.

589, 2719-2731., by Kang-Yang Jih, Min Li, Tzyh-Chang Hwang and Silvia G.

Bompadre. According to their web site,

http://www.wiley.com/bw/permis.asp?ref=0022-3751&site=1, I retain the

copyright for this work and am allowed to alter and build upon this work.

4-1. Abstract

The cystic fibrosis transmembrane conductance regulator (CFTR) is a chloride

channel that belongs to the ATP binding cassette (ABC) superfamily. The deletion of

the phenylalanine 508(ΔF508-CFTR) is the most common mutation among cystic

fibrosis (CF) patients. The mutant channels present a severe trafficking defect, and

the few channels that reach the plasma membrane are functionally impaired.

Interestingly, an ATP analog N6-(2-phenylethyl)-2’-deoxy-ATP (P-dATP) can

Page 130: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

118

increase the Po to ~0.7, implying that the gating defect of ΔF508 may involve the

ligand binding domains, such as interfering with the formation or separation of the

dimeric states of the nucleotide-binding domains (NBDs). To test this hypothesis, we

employed two approaches developed for gauging the stability of the NBD dimeric

states using the patch-clamp technique. We measured the locked-open time induced

by pyrophosphate (PPi), which reflects the stability of the full NBD dimer state, and

the ligand exchange time for ATP/P-ATP (N6-(2-phenylethyl)-ATP), which measures

the stability of the partial NBD dimer state wherein the head of NBD1 and the tail of

NBD2 remain associated. We found that both the PPi-induced locked-open time and

the ATP/P-ATP ligand exchange time of ΔF508-CFTR channels are dramatically

shortened, suggesting that the ΔF508 mutation destabilizes the full and partial NBD

dimer states. We also tested if mutations that have been shown to improve

trafficking of ΔF508-CFTR, namely the solubilizing mutation F494N/Q637R and ΔRI

(deletion of the regulatory insertion), exert any effects on these newly identified

functional defects associated with ΔF508-CFTR. Our results indicate that although

these mutations increase the membrane expression and function of ΔF508-CFTR,

they have limited impact on the stability of both full and partial NBD dimeric states

for ΔF508 channels. The structure/function insights gained from this mechanism

may provide clues for future drug design.

Page 131: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

119

4-2. Introduction

Cystic fibrosis transmembrane conductance regulator (CFTR) is the only member

in the ATP-binding cassette (ABC) transporter superfamily known to function as an ion

channel (Riordan et al., 1989; Bear et al., 1992). CFTR is mostly expressed in the apical

membrane of epithelial cells and helps maintain the fluid and electrolyte balance

across the cell membrane. Impairment of CFTR function causes cystic fibrosis (CF), the

most prevalent lethal genetic disease among Caucasians (Bobadilla et al., 2002; Cutting

et al., 2005). Although more than 1,600 mutations have been found in patients with CF

(www.genet.sickkids.on.ca/cftr/app), the deletion of phenylalanine at position 508

(ΔF508) is the most common one, associated with ~70% of CF alleles (Zielenski et al.,

1995). It is generally accepted that the ∆508 mutation causes a protein folding defect,

resulting in most channels being degraded intracellularly and very few reaching the

plasma membrane (Denning et al., 1992; Sato et al., 1996; Cheng et al., 1990; Lukacs et

al., 1993). Moreover, those few channels that actually reach the plasma membrane are

functionally impaired (Dalemans et al. 1991; Haws et al., 1996; Hwang et al., 1997;

Ostedgaard et al., 2007), with an open probability ~15 times lower than that of WT-

CFTR (Miki et al., 2010).

The structure of CFTR is similar, but more complex, than that of other members in

the ABC superfamily. In addition to the canonical domains, namely two

transmembrane domains (TMDs) and two nucleotide-binding domains (NBDs), CFTR

possesses a unique regulatory domain (R domain). Once phosphorylated by protein

kinase A (PKA) in the R domain, CFTR functions as an ATP-gated chloride channel

Page 132: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

120

(Hwang & Sheppard, 2009). Using mutant cycle analysis, Vergani and her colleagues

(Vergani et al., 2005) provided convincing evidence supporting the idea that ATP

binding induces the formation of an NBD dimer, which precedes gate opening.

Subsequent hydrolysis of the ATP molecule bound in NBD2 leads to channel closure

(Vergani et al., 2005; Mense et al., 2006, Zhou et al., 2006). Surprisingly, channel

closure does not require complete separation of the NBD dimer. It was shown recently

that closed channels enter a stable partial NBD dimer state, with the head of NBD1 and

the tail of NBD2 (i.e., ATP-binding site 1) still connected by the bound ATP, while the

ATP-binding site 2 (the head of NBD2 and the tail of NBD1) is separated (Tsai et al.,

2010b). This partial NBD dimer, which lasts for tens of seconds, allows another ATP

molecule to bind to the ATP-binding site 2 and initiate a new gating cycle (for a review

see Hwang & Sheppard, 2009).

Although the ∆F508 mutation induces severe processing and functional defects,

the crystal structures of NBD1 (Lewis et al., 2005; 2010) with the ∆F508 mutation

show little difference in protein conformation except changes in the surface

topography near the ∆F508 mutation site. It is therefore proposed that deletion of

F508 causes only a local disturbance rather than affect the overall structure of NBD1.

Because in the crystal structure of other ABC transporters, the equivalent residue of

F508 resides at the interface between the NBDs and the TMDs, it was also suggested

that ΔF508 may affect the interaction between these domains (Lewis et al., 2005;

Thibodeau et al., 2005). Indeed, biochemical studies by Serohijos et al (2008) suggest

that F508 might face the coupling helix of the fourth intracellular loop (ICL4), since

cysteines introduced at the F508 position and at several positions in the ICL4 could be

Page 133: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

121

cross-linked. Moreover, cross-linking C508 and C1068 residues inhibited channel

gating (Serohijos et al., 2008).

In contrast, our previous studies have provided evidence that the ΔF508 mutation

may affect the function of the NBDs. We found that the ATP analog, N6-(2-phenylethyl)

2’-deoxy-ATP (P-dATP), increases the Po of ΔF508 by ~15 fold, to ~0.7 (Miki et

al.,2010), a value very similar to the Po of WT-CFTR gated by P-dATP. This remarkable

increase in the Po of ΔF508-CFTR is achieved by a dramatic increase in the opening rate

and also by a prolongation of the open time. As P-dATP acts on both ATP binding sites

(i.e., NBD1 and NBD2) to rectify the gating defect associated with the ΔF508 mutation

(Miki et al., 2010), these data support the notion that the gating machinery of CFTR

might be defective in ΔF508 channels.

To further explore the possibility of a mutational defect on the function of the

NBDs, we decided to use two different functional assays we have recently established

to study the effects of the ΔF508 mutation on the stability of full and partial NBD dimer

states. First, as the lifetime of the locked-open state induced by the phosphate analog

pyrophosphate (PPi) reflects the stability of the full NBD dimer state (Tsai et al., 2009),

we can gauge the effect of the ΔF508 mutation on NBD dimers by comparing the

locked-open time between WT and ΔF508 channels. Second, the “ligand-exchange

protocol” (Tsai et al., 2010b), wherein ATP and the high-affinity analog N6-(2-

phenylethyl)-ATP (P-ATP) are suddenly exchanged, allows us to monitor how long the

nucleotides remain bound at each binding site. As dissociation of the more tightly

bound ATP in ATP-binding site 1 reports the stability of the aforementioned partial

Page 134: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

122

NBD dimer state (Tsai et al., 2010b), we can use the ligand-exchange protocol to assess

the effect of the ΔF508 mutation on the partial NBD dimer state. Since channel gating is

coupled to NBD dimer formation/separation, it is interesting to see if this basic

mechanism is perturbed in ΔF508-CFTR. In this study, results from these two assays

point to a new functional defect associated with the ΔF508 mutation. We reckon that

unraveling the molecular mechanism for this functional defect could aid our

fundamental understanding of the malfunction of the most common pathogenic

mutation associated with CF.

4-3. Material and Methods

Cell culture and transient expression system

Chinese hamster ovary (CHO) cells were grown at 37°C in Dulbecco’s Modified

Eagle’s Medium supplemented with 10% fetal bovine serum. Cells were cultured in

35-mm tissue culture dishes for one day before transfection. CFTR cDNA was

cotransfected with pEGFP-C3 (Clontech, Palo Alto, CA), encoding green fluorescence

protein, using PolyFect transfection reagent (QIAGEN). Cells expressing CFTR were

placed at 27°C for at least 2 days before electrophysiology experiments were

performed.

CFPAC-1 cells (purchased from ATCC) were grown at 37°C in Iscove’s Modified

Dulbecco’s Medium supplemented with 10% fetal bovine serum. Cells were cultured

in 35-mm tissue culture dishes for one day before virus infection.

Page 135: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

123

Mutagenesis

All mutant constructs were generated using the QuickChange XL kit (Stratagene,

La Jolla, CA) according to manufacturer’s protocols. All the CFTR cDNA were

sequenced to confirm the mutation (DNA core, University of Missouri).

Adenovirus infection

Adenoviruses expressing ΔF508 CFTR (H5.040.CMV.deltaF508) and GFP

(H5’.040CMVEGFP) were purchased from the Gene Therapy Program Vector Core in

the Department of Medicine at the University of Pennsylvania. Approximately 106

cells were co-infected with H5’.040CMVEGFP and H5.040.CMV.deltaF508 (ratio1:5)

at 1000 viral particles per cell in 2 ml medium at 37oC overnight. Cells were grown

at 27°C for 2-3 days before harvested for electrophysiological experiments.

Electrophysiological recordings

CFTR channel currents were recorded using an EPC9 or EPC10 amplifier (HEKA,

Lambrecht/Pfalz, Germany) at room temperature. The pipettes were prepared from

borosilicate capillary glass using a Flaming/Brown-type micropipette puller (P97,

Sutter Instrument Co.) and then fire-polished with a homemade microforge. The

resistance of pipettes in the bath solution was 2 to 4 MΩ. Glass chips carrying the

transfected cells were transferred to a chamber located on the stage of an inverted

microscope (IX51, Olympus). After the seal resistance was >40 GΩ, the membrane

was excised. CFTR channels were first activated by 2.75 mM ATP and 25 U/ml PKA

Page 136: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

124

until the current reached the steady state. . All test solutions contained 10 U/ml PKA

to maintain the phosphorylation level. The data were filtered at 100 Hz with an

eight-pole Bessel filter (LPF-8, Warner Instruments), and digitized to a computer at

a sampling rate of 500 Hz. The membrane potential was held at -60 mV and the

inward current was inverted for clear data presentation.

All inside-out patch experiments were performed with a fast solution exchange

perfusion system (SF-77B, Warner Instruments). The dead time of solution change

is ~30 ms (Tsai et al., 2009).

Chemicals

The pipette solution contained (in mM): 140 N-methyl-D-glucamine chloride

(NMDG-Cl), 2 MgCl2, 5 CaCl2, and 10 HEPES (pH 7.4 with NMDG). Cells were

perfused with a bath solution containing (in mM): 145 NaCl, 5 KCl, 2 MgCl2, 1 CaCl2,

5 glucose, 5 HEPES and 20 sucrose (pH 7.4 with NaOH). For inside-out configuration,

the perfusion solution contained (in mM): 150 NMDG-Cl, 2 MgCl2, 10 EGTA and 8

Tris (pH 7.4 with NMDG).

MgATP, PPi and PKA were purchased from Sigma-Aldrich. N6-(2-phenylethyl)-

ATP (P-ATP) and N6-(2-phenylethyl)-2’-deoxy ATP (P-dATP) were purchased from

Biolog Life Science Institute. PPi and MgATP were stored in 200 mM and 250 mM

stock solution respectively at -20°C. P-ATP and P-dATP were stored in 10 mM stock

at -70°C.. All chemicals were diluted to the concentration indicated in each figure

Page 137: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

125

using the inside-out perfusion solution and the pH was adjusted to 7.4 with NMDG.

For solutions containing PPi, an equal concentration of MgCl2 was added to the

solution.

Data analysis and statistics

Steady-state mean currents were calculated with Igor Pro program

(Wavemetrics, Lake Oswego, OR) after baseline subtraction. Current relaxations

were fitted with single or double exponential functions using a Levenberg-

Marquardt-based algorithm with the Igor Pro program. Ensemble currents were

generated by adding 8-15 raw traces.

Results are shown as mean ± S.E. Student’s t-test were performed for statistical

analysis using Excel (Microsoft). P<0.05 was considered significant.

4-4. Results

Effects of ΔF508 mutation on the stability of the locked-open state

To examine the effect of ΔF508 mutation on the stability of the NBD dimer, we

used the non-hydrolyzable analog PPi to lock ΔF508-CFTR channels into an open

state. Our previous studies (Tsai et al., 2009) demonstrated that PPi, by binding to

NBD2, emptied after ATP hydrolysis and dissociation of the hydrolytic products,

locks open CFTR by forming a stable NBD dimer state. The stability of this state can

Page 138: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

126

be gauged by measuring the relaxation time constant of the current decay upon

removal of ligands. Figure 4-1 A shows that application of 1 mM ATP and 4 mM PPi

to WT-CFTR increases the ATP-induced macroscopic current by 2.04 ± 0.09 fold (n =

8). Upon wash-out of the ligands, the current decays mono-exponentially with a

time constant of 31.28 ± 3.2s (n = 8). In patches containing fewer channels, stepwise

closings of the locked-open channels can be readily seen after the ligands were

removed (Figure 4-1A). Interestingly, for ΔF508-CFTR (Figure 4-1B), the same

treatment increases the macroscopic current by 2.88 ± 0.13 fold (n = 7) (Figure 4-

1C), but the locked-open time (i.e, the relaxation time constant) is ~ 10-fold shorter

than that of WT channels (3.29 ± 0.43 s, n = 5) (Figure 4-1D). This drastically

shortened locked-open time can be discerned more clearly in patches with fewer

channels. Unlike WT-CFTR, where long-lasting openings were seen after ligand

removal, ΔF508-CFTR channels closed rapidly once ATP and PPi were washed out

(Figure 4-1B).

These results suggest that deletion of the phenylalanine at position 508

significantly decreases the stability of NBD dimers. As the locked-open time for

ΔF508-CFTR is shorter than that of WT channels, it may seem odd that PPi increases

ΔF508-CFTR macroscopic currents slightly more than it does WT-CFTR currents

(Figure 4-1C). This is likely due to the much lower Po of ΔF508-CFTR with a very

long mean closed time (Dalemans et al. 1991; Haws et al., 1996; Hwang et al., 1997;

Ostedgaard et al., 2007; Miki et al., 2010). Thus, a slight prolongation of the open

time results in a significantly larger effect on Po for ΔF508-CFTR.

Page 139: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

127

Figure 4-1

Figure 4-1.Difference in locked-open time induced by PPi in WT-CFTR and

ΔF508-CFTR. Representative traces of macroscopic (left) and microscopic (right)

WT-CFTR (A) and ΔF508-CFTR (B) currents elicited by 1 mM ATP and 4 mM PPi.

The dotted line represents zero current. The bars above the traces represent the

presence of different ligands to the membrane patch (same in following figures). C.

Current increase induced by PPi for WT (n = 8) and ΔF508-CFTR (n = 7) (P < 0.01). D.

Summary of locked-open times induced by PPi for WT-CFTR (n = 7) and ΔF508-

CFTR ( n= 5).*, P < 0.01 versus WT-CFTR. . The number above each bar represents n

number for each experiment (same in following figures).

3.0

2.5

2.0

1.5

1.0

0.5

0.0

4 mM PPi1 mM ATP

4 mM PPi

1 mM ATP

1

2

3

A

B

C

D

I P

Pi/ I

AT

PR

ela

xa

tio

n t

ime

co

nsta

nt

(s)

WT∆F508

WT ∆F508

∆F508-CFTR

10

20

30

*

*

(7)

(8)

(8)

20 s

0.5

pA

20

pA

200 s

0.5

pA

5 s100 s

2 p

AWT-CFTR

4 mM PPi1 mM ATP

4 mM PPi

1 mM ATP

(5)

Page 140: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

128

Figure 4-2

Figure 4-2.Comparison of the locked-open time of E1371S- and

∆∆∆∆F508/E1371S-CFTR. Representative traces of non-hydrolytic E1371S-CFTR (A)

and ΔF508/E1371S-CFTR (B) in the presence of 1 mM ATP. C. Summary of the

locked-open times for E1371S-CFTR (n = 15) and ΔF508/E1371S-CFTR (n = 4) (P <

0.01).

100

60

20

35

30

25

20

15

10

5

200150100500

60

40

20

0

8006004002000200 s

20

pA

50 s

5 p

A

1 mM ATP 1 mM ATP

A B C∆F508/E1371S

∆F508/E1371SE1371S

Rela

xatio

n tim

e c

on

sta

nt

(s)E1371S

(15)

(4)

Page 141: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

129

To verify these results, we employed an alternative approach to produce locked-

open channels. We introduced the E1371S mutation, which is known to demolish

ATPase activity in ABC proteins including CFTR (Aleksandrovet al., 2000; Vergani et

al., 2003), into WT and ΔF508 backgrounds. For E1371S channels, the relaxation

time constant of the current decay after ATP washout is ~110 s (Figure 4-2A and C,

Zhou et al., 2006), whereas that of ΔF508/E1371S channels is only 32.45 ± 4.07 s (n

= 4) (Figure 4-2B and C). Although it is unclear why the ΔF508 mutation shows less

effect on the stability of NBD dimer under the E1371S background, the shortening of

the locked-open time seen in ΔF508/E1371S channels is consistent with the idea

that the culprit is a destabilization of the NBD dimer rather than a lower affinity or

efficacy of PPi.

Tight binding of nucleotides in NBD1 prolongs the channel locked-open time.

In a previous report (Tsai et al., 2010a), we demonstrated that the locked-open

time of WT-CFTR induced by PPi is prolonged by replacing ATP with the high affinity

ATP analog, N6-phenylethyl-ATP (P-ATP), or by introducing “gain-of-function”

mutations to the ATP-binding site 1 (mutations which increase the Po of CFTR, such

as W401F and H1348G) as the locked-open state reflects an NBD dimer with ATP-

binding site 1 occupied by ATP and ATP-binding site 2 by PPi (Tsai et al.,2009). In

Figure 4-3, we show that the gain-of-function mutations W401F and H1348G

(Figure 4-3A) and P-ATP (Figure 4-3B) also prolong the locked-open time of ΔF508-

Page 142: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

130

CFTR channels. Compared to ΔF508-CFTR, the double mutant W401F/ΔF508-CFTR

Figure 4-3

Figure 4-3.Increase of the locked-open time for ΔF508-CFTR channels by gain-

of-function mutations or the high affinity ATP analog P-ATP. A. Relaxations of

ensemble current traces for various mutations. Upper, red: ΔF508/DM locked open

with 1 mM ATP and 2 mM PPi (ensemble from 14 traces). Upper, blue: ΔF508/DM

locked open with 50 µM P-ATP and 2 mM PPi (ensemble from 14 traces). Lower, red:

ΔF508/TM locked open with 1 mM ATP and 2 mM PPi (ensemble from 8

traces).Lower, blue: ΔF508/TM locked open with 50 µM P-ATP and 2 mM PPi

(ensemble from 9 traces). B. A representative trace of ΔF508-CFTR locked open

with 50 µM P-ATP and 2 mM PPi (n = 6). C. Summary of PPi locked-open times for

each construct (ΔF508/DM: W401F/ΔF508-CFTR, ΔF508/TM:

W401F/H1348G/ΔF508-CFTR).

Page 143: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

131

(ΔF508/DM) prolonged the locked-open time by ~2 fold, and the triple mutant

W401F/H1348G/ΔF508-CFTR (ΔF508/TM) by ~4 fold. Moreover, the locked-open

time of each mutant was further prolonged when P-ATP, instead of ATP, was used as

the ligand, suggesting that the effect of P-ATP and the gain-of-function mutations are

additive. Despite these manoeuvres, the maximal locked-open time (P-ATP together

with two gain-of-function mutations) remains ~ 2/3 of that of WT-CFTR channels

locked open by ATP and PPi (20 s versus 31 s in WT-CFTR). Nonetheless, the

observation that manipulations of ligand- ATP-binding site 1 interactions result in

similar effects on the locked-open time for both WT- and ΔF508-CFTR channels

suggests a common structural basis for the locked-open state of these two channels,

namely the conformation of the NBD dimer.

ATP/P-ATP ligand exchange time in ATP-binding site 1 is accelerated in ΔF508-CFTR

The shortened locked-open time with ΔF508-CFTR channels suggests the

intriguing possibility that the binding of the ATP molecule in NBD1 may not be as

stable as in WT channels. Tsai et al., (2010b) demonstrated the existence of a partial

NBD dimer state, wherein ATP in ATP-binding site 2 has been hydrolyzed and the

hydrolytic products have been released, but one ATP remains tightly bound in ATP-

binding site 1 for tens of seconds. The stability of this partial NBD dimer can be

examined by conducting ligand exchange experiments (Tsai et al., 2010b). When the

ligand is switched from ATP to P-ATP, there is a fast current raising phase, which

represents the ATP/P-ATP exchange in the ATP-binding site 2, followed by a slow

Page 144: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

132

Figure 4-4

Figure 4. ATP/P-ATP ligand exchange for WT-CFTR and ΔF508-CFTR.

Representative current traces of WT-CFTR (A) and ΔF508-CFTR (B) upon ATP/ P-

ATP exchange. The current rising phase was fitted with a double exponential

function. C. Time constants (τ) of the slow (upper panel) and fast (lower panel)

current raising phases after the ligand is switched from ATP to P-ATP. *, P < 0.01

versus WT- CFTR.

60

50

40

30

20

10

40

30

20

10

300250200150100500

0.6

0.5

0.4

0.3

0.2

0.1

600

400

200

250200150100500

50 s

10

pA

20

0 p

A

50 s

(9)

(6)

(4)

(6)

*

50 µM P-ATP

∆F508-CFTR

50 µM P-ATP

∆F508

∆F508

WT

Tim

e c

on

sta

nt

(s)

WT

Tim

e c

on

sta

nt

(s)

WT-CFTR

2.75 mM ATP

A

slow phase

fast phase

2.75 mM ATP

B

C

Page 145: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

133

current rising phase representing the ATP/ P-ATP exchange in the ATP-binding site

1. The tightness of ligand binding in ATP-binding site 1 can be examined

quantitatively by measuring the time constant of the slow phase upon ligand

exchange in the ATP-binding site 1.

In WT-CFTR, the ligand exchange time at ATP-binding site 1 is ~ 40 s (42.18 ±

9.63 s, n = 6, Figure 4-4A, C). However, in ΔF508-CFTR, the current reached the

steady state soon after the perfusion solution was switched from ATP to P-ATP

(Figure 4-4B). As a result, the ligand exchange time for ΔF508-CFTR is dramatically

shortened (3.27 ± 0.53 s, n = 9, Figure 4-4C). Thus, the ΔF508 mutation not only

destabilizes the full NBD dimer (Figure 4-1 and Figure 4-2), it also dramatically

shortens the ligand dwell time in ATP-binding site 1 presumably by destabilizing the

partial NBD dimer state (Figure 4-4C). As expected, the time constant of the fast

rising phase is similar between WT-CFTR and ΔF508-CFTR, suggesting that the rate

of hydrolysis and sequential separation at ATP-binding site 2 is nearly identical in

both constructs.

The results described above were obtained from CHO cells transiently

expressing CFTR. However, in physiological conditions, CFTR channels are

predominantly expressed in epithelial cells. Therefore, we also carried out similar

experiments with F508-CFTR channels expressed in human CFPAC-1 epithelial

cells. The results are consistent with those observed in CHO cells (Figure 4-S1).

Page 146: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

134

Expression improving mutations improve dimer stability

ΔF508-CFTR exhibits defective processing and trafficking; thus only a small

proportion of the channels mature and traffic from the endoplasmic reticulum to the

cell surface (Denning et al., 1992; Sato et al., 1996; Cheng et al., 1990; Lukacs et al.,

1993). Several recent reports have identified mutations that help trafficking and

maturation of ΔF508 channels as well as the function (DeCarvalho et al., 2002;

Pissarra et al., 2008; Aleksandrov et al., 2010). We introduced into ∆F508-CFTR the

“solubilizing mutations”, F494N/Q637R (Pissarra et al.,2008) and the deletion of the

regulatory insertion region (ΔRI, deletion of residues 404-435) reported by

Aleksandrov et al., (2010), to test whether these trafficking improving mutations

have any effect on the gating defects described above.

We first tested the F494N/Q637R and ΔRI mutations on the WT background

and found that both the solubilizing mutations and ΔRI significantly prolong the

locked-open time and ligand exchange time. As seen in Figures 4-5A and 4-6A, in

either case, the current relaxation upon removal of ATP and PPi was significantly

slower compared with that for WT-CFTR (F494N/Q637R-CFTR: τ = 86.14 ± 12.61s,

n = 6; ΔRI-CFTR: τ = 75.33 ± 14.36 s, n = 7. Figure 4-7 summarizes the results. For

ligand exchange experiments, these mutations also significantly prolong the second

phase of current changes upon switching the ligand from ATP to P-ATP

(F494N/Q637R-CFTR: τ = 76.41 ± 12.31 s, n = 6; ΔRI-CFTR: τ = 81.78 ± 6.66 s, n = 7)

(Figures 4-5A and 4-6A, 4-7). Thus, for unknown reason, these mutations stabilize

the NBD dimer as well as the partial NBD dimer states. We next engineered these

Page 147: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

135

mutations into the ΔF508 background. We found that although the locked-open time

(F494N/Q637R/ΔF508-CFTR: τ = 5.95 ± 0.36 s, n = 8; ΔRI/ΔF508-CFTR: τ = 5.52 ±

0.45 s, n = 11) and ligand exchange time (F494N/Q637R/ΔF508-CFTR: τ = 8.44 ±

1.3 s, n = 6; ΔRI/ΔF508-CFTR: τ = 8.95 ± 1.75 s, n = 4) of F494N/Q637R/ΔF508 and

ΔF508/ΔRI channels are prolonged (Figures 4-5B, 4-6B, 4-7), they are still much

shorter than those of WT channels. As these mutations improve these gating

parameters to a similar degree for WT- and ΔF508-CFTR, their effects may not be

specific for channels carrying the ΔF508 mutation.

Besides prolonging the PPi locked-open time and the ligand exchange time,

F494N/Q637R/ΔF508 and ΔF508/ΔRI have been previously shown to improve the

function of ΔF508-CFTR (Pisarra et al, 2008; Aleksandrov et al, 2010). Here we used

the high affinity ATP analog P-dATP to boost the Po of these mutant channels. To our

surprise, we found that P-dATP still increases the current dramatically for both

compound mutants: 6.96 ± 0.17 fold increase for ΔRI/ΔF508-CFTR (n = 5), and

12.36 ± 1.21 fold increase for F494N/Q637R-CFTR (n = 8) (Figure 4-8). Previously

we found that ΔF508-CFTR channels assume a Po of ~0.7 in the presence of P-dATP,

similar to those of WT channels under the same conditions (Miki et al, 2010). While

P-dATP increased the mean macroscopic current of ΔF508-CFTR channels by 19-

fold, WT currents barely doubled in the presence of P-dATP. Thus, these

microscopic and macroscopic results allow us to accurately calculate the Po of

ΔF508 to be ~15 times smaller than the Po of WT channels at saturating [ATP] (0.03

vs 0.45; Miki et al, 2010). Since P-dATP does not alter single channel conductance

(Miki et al. 2010), the increase in the macroscopic current induced by P-dATP in

Page 148: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

136

F494N/Q637R/ΔF508 and ΔF508/ΔRI (12 and 7 fold respectively) suggests that the

Po of these mutant channels is still lower than that of WT channels when ATP is used

as the ligand. Thus, although the exact Po for these compound mutants is not

determined, we can safely conclude that these expression-improving mutations do

not completely restore the function of the ΔF508 channel in our expression system

(cf. Pissarra et al. 2008).

Page 149: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

137

Figure 4-5

Figure 4-5. Effects of “solubilizing mutations”, F494N/Q637R, on WT- and

ΔF508-CFTR channels. A. Representative current traces of F494N/Q637R-CFTR

channels locked opened by 1 mM ATP and 2 mM PPi (left) and ATP/P-ATP ligand

exchange (right). B. Representative current traces of F494N/Q637R/ ΔF508-CFTR

channels locked opened by 1 mM ATP and 2 mM PPi (left) and ATP/P-ATP ligand

exchange (right). The current decay upon ligand wash-out was fitted with a single

exponential function, while the current rising phase was fitted with a double

exponential function.

140

120

100

80

60

40

20

5004003002001000

50

40

30

20

10

5004003002001000

160

120

80

40

4003002001000

250

200

150

100

50

200150100500

20

pA

100s

40

pA

10

pA

100s

100s

2.75 mM ATP1 mM ATP

2.75 mM ATP50 µM P-ATP

50 s

50

pA

F494N/Q637R-CFTR

F494N/Q637R/∆F508-CFTR

A

B

2 mM PPi

2.75 mM ATP

1 mM ATP

2 mM PPi

2.75 mM ATP

50 µM P-ATP

Page 150: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

138

Figure 4-6

Figure 4-6. Effects of deletion of the regulatory insertion (ΔRI) on WT- and

ΔF508-CFTR channels. A. Representative current traces of ΔRI-CFTR locked open

by 1 mM ATP and 2 mM PPi (left) and ATP/P-ATP ligand exchange (right). B.

Representative current traces of ΔRI/ ΔF508-CFTR locked open by 1 mM ATP and 2

mM PPi (left) and ATP/P-ATP ligand exchange (right). The current decay upon

ligand wash-out was fitted with a single exponential function, while the current

rising phase was fitted with a double exponential function.

50

40

30

20

10

200150100500

400

300

200

100

4003002001000

80

60

40

20

250200150100500

30

25

20

15

10

5

300250200150100500

20

pA

10

0p

A1

0p

A

50 s 100s

50s

10

pA

50s

2.75 mM ATP

1 mM ATP50 µM P-ATP2.75 mM ATP

∆RI/∆F508-CFTR

∆RI-CFTRA

B

2 mM PPi

2.75 mM ATP

1 mM ATP2 mM PPi 50 µM P-ATP

2.75 mM ATP

Page 151: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

139

Figure 4-7

Figure 4-7. Summary of effects of solubilizing mutations or ΔRI on locked-open

time and the time constant of the slow phase current rise upon ATP/P-ATP ligand

exchange for different WT-CFTR constructs (A) and ΔF508-CFTR constructs (B).*,P

< 0.01 versus WT- CFTR. #,P < 0.01 versus ΔF508-CFTR. sol: solubilizing mutation,

F494N/Q637R. The number next to each bar represents n number for each

experiment.

10080604020

WT

Sol

∆RI

80604020

WT

Sol

∆RI

108642

∆F508

Sol/∆F508

∆F508/∆RI

7654321

∆F508

Sol/∆F508

∆F508/∆RI

PPi lock open time (s)

Ligand exchange time (s)

A B

*

(7)

(6) (8)

(11)

(7)

(7)*

#

#

(6)

(6)

*(7) #

# (6)

(9)

(4)

PPi lock open time (s)

Ligand exchange time (s)

Page 152: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

140

Figure 4-8

Figure 4-8. Effect of P-dATP on F494N/Q637R/ ΔF508-CFTR and ΔRI/ΔF508-

CFTR. Representative current traces of F494N/Q637R/ ΔF508-CFTR (A) and

ΔRI/ΔF508-CFTR (B) in the presence of 50 µM P-dATP . P-dATP increases the Cl-

current of F494N/Q637R/ ΔF508-CFTR by 12.36 ± 1.21 fold (n = 8) ΔRI/ΔF508-

CFTR by 6.96 ± 0.17 fold (n = 5).

70

60

50

40

30

20

10

250200150100500

40

30

20

10

250200150100500

A RI/ F508-CFTR∆ ∆

2.75 mM ATP50 M P-dATPµ

2.75 mM ATP50 M P-dATPµ

A F494N/Q637R/ F508-CFTR∆

50 s

20

pA

50 s

10

pA

Page 153: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

141

4-5. Discussion

In the current study, we used the locked-open time induced by PPi and the

ligand exchange protocol to gauge the effects of the ΔF508 mutation on the stability

of both the full and partial NBD dimer states of CFTR channels. Our results strongly

suggest that theΔF508 mutation dramatically destabilizes both. We also tested

mutations that improve the stability of both NBD dimers states as well as mutations

that had been shown to increase the surface expression of ΔF508-CFTR on these

functional parameters. All the mutants that we tested improved the function of

ΔF508-CFTR although their effects on WT channels are similar.

ΔF508 is the most common mutation found in patients with CF. Two decades of

studies have revealed several defects associated with ΔF508-CFTR: poor membrane

expression and impaired channel function. Early studies reported that the deletion

of F508 severely impeded channel maturation (Cheng et al., 1990). The majority of

ΔF508 channels are retained in the ER rather than traffic to the cell membrane

(Ward et al., 1995).This effect can be observed in Western blot, where the amount of

mature form of ΔF508-CFTR (band C) is dramatically reduced compared to WT-

CFTR (Figure 4-S2, Pissarra et al., 2008; Aleksandrov et al., 2010). The mutation also

affects the stability of the protein in the plasma membrane since ΔF508 channels are

removed from the membrane at a faster rate than WT channels (Lukacs et al., 1993).

The mechanism responsible for this reduction of the surface pool is not well

understood and it has been attributed to different processes such as recycling

deficiencies due to misfolding (Sharma et al.,2004) and accelerated endocytosis

Page 154: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

142

(Swiatecka-Urban et al., 2005), which might be related to peripheral protein quality

control (Okiyoneda et al., 2010). Thus, with less membrane insertion and more

retrieval, the channel density in the cell membrane is much less for ΔF508 than WT-

CFTR. Despite these biochemical anomalies, the crystal structure of isolated

NBD1/∆F508 showed surprisingly little difference when compared to that of WT

NBD1. Lewis et al., (2010) reported that the structural difference between WT and

ΔF508 is minor and localized to areas adjacent to ΔF508. It is therefore proposed

that enhanced dynamics of nearby residues could affect chaperone interactions with

mutant proteins (Lewis et al., 2010).

Even when those few ΔF508 channels actually traffic to the plasma membrane,

their function is gravely impaired. In several studies in native ΔF508-CFTR-

expressing tissues, including nasal epithelium, airway epithelium pancreatic ducts

and intestine, no or very little cAMP induced Cl- current was found (Riordan et al.,

1989; Knowles et al., 1983; Welsh et al., 1986; Gray et al., 1988). Our recent studies

in excised patches also revealed that ΔF508-CFTR channels exhibit a very low Po in

response to ATP (Miki et al., 2010; cf. Dalemans et al. 1991; Haws et al., 1996;

Hwang et al., 1997; Ostedgaard et al., 2007). A previous study by Schultz et al (1999)

reported a rightward shift of the ATP dose response for F508-CFTR, but we do not

observe that effect (Wang et al, 2000).We took advantage of the high affinity ATP

analog P-dATP to accurately measure the Po of ΔF508 channel expressed in CHO

cells. Our results suggest that the Po for ΔF508-CFTR is ~0.03, about 15 fold less

than that of WT channels. Interestingly, even though the chemical difference

between ATP and P-dATP is fairly minor and both nucleotides act on the nucleotide-

Page 155: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

143

binding sites, P-dATP brings the gating activity of ΔF508-CFTR to levels comparable

to those of WT channels (Miki et al., 2010). That such a minor modification of the

ligand significantly restores the function of ΔF508-CFTR suggests an underlying

molecular defect within the NBD that results in a poor response to ATP. As the

mutation does not affect the apparent affinity of the channel to ATP (Wang et al.,

2000), the simplest possibility is a slowed NBD dimer formation following ATP

binding (see below for detail). Alternatively, the F508 mutation may perturb

coupling between NBD and TMD (Serohijos et al., 2008), and/or interfere with the

function of the RI domain (Aleksandrov et al., 2010).

The hypothesis of defective coupling was proposed in Serohijos et al., (2008) by

constructing a homology model of CFTR based on the crystal structure of SAV1866,

a bacterial ABC transporter. Based on this homology model, F508 is close to the

NBD-TMD interface. Cross-linking experiments indeed showed proximity between

the F508 region of NBD1 and ICL4 as well as between NBD2 and 2nd cytoplasmic

loop (ICL2). Electrophysiological recordings further demonstrated that cross-linking

the F508 region with residues in CL4 inhibits channel activity. Therefore, the

authors suggested that the deletion of F508 could interfere with the coupling

between the NBDs and TMDs, resulting in channels with poor function.

A more recent report (Aleksandrov et al, 2010) proposed that the regulatory

insertion (RI, amino acids 404-435) may play a role in the pathogenesis of the

ΔF508 mutation as removal of the RI improves both surface expression and gating

of ΔF508-CFTR. In our study, although we confirmed that deletion of RI indeed

Page 156: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

144

somewhat improves the expression of the ΔF508 channels (Figure 4-S2), both the

PPi-induced locked-open time and the ligand exchange time of ΔRI/ΔF508-CFTR

channels are still much shorter than that of WT-CFTR, suggesting that the removal

of RI does not completely restore dimer stability. Moreover, we found that

application of P-dATP to ΔRI/ΔF508-CFTR channels increases the current by ~ 7

fold, suggesting that the Po of this compound mutant is still much lower than WT

channels (Figure4-7A).Similarly, the solubilizing mutations F494N and Q637R also

partially restore the membrane expression and function of ΔF508 channels (Figures

4-5B, 4-7B, 4-8, 4-S2 and Pisarra et al., 2008).

By employing two assays that more directly gauge the states of the NBDs, the

current study provides more insights about the basis for the functional

perturbations associated with the ΔF508 mutation. Interestingly, however, although

P-dATP brings the Po of ΔF508-CFTR back to normal levels, it failed to prolong the

PPi locked-open time and the ligand exchange time of ΔF508-CFTR to WT-CFTR

values (data not shown). If we accept the premise that these two functional assays

accurately estimate the stability of the NBD dimer states, these results directly point

to a defect in the NBDs of ΔF508-CFTR, even though little structural perturbation is

seen in the crystal structure of NBD1 carrying the F508 mutation. However, it

should be noted that our interpretation hereby is based on a crucial assumption,

that is, the gating of ΔF508-CFTR follows similar mechanisms that have been

proposed for WT-CFTR. In other words, the opening/closing of the gate for ΔF508-

CFTR, like that of WT-CFTR, is coupled to the formation and partial separation of the

NBD dimer (Tsai et al. 2009, 2010a, b). Then, our interpretation that the ΔF508

Page 157: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

145

mutation destabilizes both full and partial NBD dimers should be valid. As in an NBD

dimer, two NBDs are connected mostly by the ligand-NBD interactions and to a less

degree by the NBD-NBD interactions, we speculate that a simple explanation for the

destabilization effect observed in ΔF508 channels is that the mutation creates a

distortion which is transmitted to the dimer interface and thus weaken the dimer

stability.

Can this hypothesis also account for the decreased opening rate manifested in

ΔF508-CFTR? Results from thermodynamic studies of CFTR gating may provide a

tentative answer. Csanady et al., (2006), by carefully measuring gating parameters

at different temperatures, concluded that opening of the channel by NBD

dimerization is associated with a large increase of entropy presumably due to

dehydration at the dimer interface and dispersion of relatively ordered water

molecules to the disordered bulk. Using this energetic argument, we hypothesize

that the structural changes induced by the ΔF508 mutation at the NBD dimer

interface may hinder NBD dimerization (thus a slower opening rate); whereas P-

dATP, with a hydrophobic benzene ring added to the adenine moiety and the

removal of 2’-hydroxyl group at the ribose ring, may facilitate NBD dimerization by

lowering the energetic barrier between open and closed states. More studies are

needed to verify or disapprove this hypothesis.

Even if the aforementioned hypothesis is correct, the fact that F508 is located far

from the NBD dimer interface indicates that any changes in the NBD dimer interface

has to be allosteric in nature. Since the NBDs and other parts of the CFTR protein

Page 158: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

146

must move in a concerted way during the gating motion, it is entirely possible that

the instability of the NBD dimer states demonstrated in the current study is

secondary to alterations in those regions involved in the coupled movements.

Indeed the crystal structure of ΔF508/NBD1 (Lewis et al., 2010) does not reveal

major structural perturbation in regions that presumably form contacts with bound

ATP and NBD2. It is however important to point out that the solved crystal structure

is a monomeric form. Moreover as the crystal structure provides only a snapshot of

a protein, it is not possible to predict what conformational changes the NBDs may

undergo during a gating cycle. It is equally important to note that due to the

exponential relationship between the distribution of different kinetic states and the

changes in free energy between these states, a slight alteration in free energy due to

minor structural changes, which may not be readily seen in the crystal structure,

could be enough to lower the Po from 0.4 to 0.03 or shorten the locked open time

from 30 s to 3 s. For example, a 10-fold difference in the closing rate only reflects a

ΔΔG of ~5 kJ/ mol, a value smaller than the average strength of a hydrogen bond.

Therefore, a dramatic gating defect can occur without severe disturbance in

structure. Of note, a previous study by Roxo-Rosa et al. (2006) suggested the

possibility that the ΔF508 mutation may cause a structural instability in NBD1

which might weaken the binding energy for stable NBD dimer formation. .

Interestingly, recent studies on the defective folding mechanism and thermal

instability of isolated NBD1 constructs suggest that the ΔF508 mutation destabilizes

the native state and accelerates the formation of a partially-folded intermediate

state (Protasevich et al., 2010; Wang et al., 2010). How this structural instability of

Page 159: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

147

an isolated NBD1 relates to the instabilities revealed by our functional studies using

the whole protein remains unclear.

Irrespective of which structural mechanism may account for the gating defects

associated with ΔF508-CFTR, a better understanding of the cause underlying the

gating defect of the mutant channel may prove helpful in the development of new

CFTR potentiators (i.e. compounds that increase the activity of channels). Our latest

reports demonstrated that strengthening the ligand binding in ATP-binding site 1

either through mutations (Tsai et al, 2010a) or using ATP analogs (Miki et al., 2010)

significantly improves the activity of ΔF508-CFTR, underscoring the possibility that

targeting the NBD dimer interface could be a promising approach for drug design.

In conclusion, our study demonstrates a new functional defect associated with

ΔF508-CFTR channels. The full and partial NBD dimers of ΔF508-CFTR are less

stable than those of WT-CFTR. Elucidating the structural basis for these functional

defects of ΔF508-CFTR may open a new track for future drug design.

Page 160: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

148

4-6. Supplementary materials

Figure 4-S1

ΔF508-CFTR expressed in CFPAC-1 epithelial cells also demonstrates unstable full and

partial NBD dimers.

Figure 4-S1. Properties of ΔF508-CFTR expressed in CFPAC-1 cells. A. A representative

current trace of ΔF508-CFTR in the presence of 1 mM ATP and 2 mMPPi(n = 13). B. A

representative trace of ΔF508-CFTR locked open by 50 µM P-ATP and 2 mMPPi (n = 9).C.

Summary of the relaxation time constant (τ) for PPi-locked open ΔF508-CFTR with P-ATP or

ATP.*, P < 0.01.D. A representative current trace of ΔF508-CFTR for the ATP/P-ATP ligand

exchange experiment (Slow phase: τ = 5.93 ± 1.1 s; fast phase: τ = 0.85 ± 0.13 s (n = 9).

100

80

60

40

20

3002001000

40

30

20

10

200150100500

10

8

6

4

2

ATP P-ATP

100

80

60

40

20

20016012080400

10

pA

20

pA

50 s 100 s

40 s

20

pA

2.75 mM ATP1 mM ATP

2.75 mM ATP

Re

laxa

tio

n t

ime

co

nsta

nt

(s)

A B

C D*

2 mM PPi

2.75 mM ATP

2 mM PPi

50 µM P-ATP(9)

(13)

50 µM P-ATP

Page 161: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

149

Figure 4-S2

Western Blot analysis of WT or mutant CFTR expression:

Figure 4-S2: Western blot analysis for WT, ∆F508, ∆RI/∆F508 and ∆F508/Sol

(∆F508/F494N/Q637R) incubated at 37 °C. Mature fully-glycosylated CFTR and

core-glycosylated CFTR are labeled as Band C and Band B, respectively. N = 3.

Western Blot: Transfected cells from confluent 35-mm dishes were washed with

PBS and were lysed in 1x Laemmli sample buffer. DNA was sheared by brief

sonication. Whole cell lysates were separated on 7.5% Tris-HCl Ready gel (Bio-Rad,

Hercules, CA) and transferred onto a nitrocellulose membrane. The membrane was

blocked with 5% milk in TBST buffer (20 mM Tris, 137 mM NaCl, 0.05%Tween) at

4oC overnight and then probed with primary antibody against CFTR (clone M3A7,

Chemicon, Temecula, CA) in TBST at4oC overnight. The membrane was washed with

TBST three times and then incubated with the horseradish peroxidase-conjugated

secondary antibody (donkey anti-mouse IgG; Jackson Immuno Research

Laboratories, West Grove, PA) for 2 h at room temperature. The membrane was

Band C

Band B

WT ∆∆∆∆F508

∆∆∆∆RI/

∆∆∆∆F508

∆∆∆∆F508/

Sol

Page 162: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

150

developed with chemiluminescence reagent (Pierce, Rockford, IL) and the image

was acquired by ChemiDoc XRS+ system (BioRad, Hercules, CA).

Page 163: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

151

4-7. References

Aleksandrov AA, Chang X, Aleksandrov L & Riordan JR. (2000). The non-hydrolytic

pathway of cystic fibrosis transmembrane conductance regulator ion channel

gating. J Physiol 528 Pt 2, 259-265.

Aleksandrov AA, Kota P, Aleksandrov LA, He L, Jensen T, Cui L, Gentzsch M,

Dokholyan NV & Riordan JR. (2010). Regulatory insertion removal restores

maturation, stability and function of ∆F508 CFTR. J Mol Biol 401, 194-210.

Bear CE, Li CH, Kartner N, Bridges RJ, Jensen TJ, Ramjeesingh M & Riordan JR.

(1992). Purification and functional reconstitution of the cystic fibrosis

transmembrane conductance regulator (CFTR). Cell 68, 809-818.

Bobadilla JL, Macek M, Jr., Fine JP & Farrell PM. (2002). Cystic fibrosis: a worldwide

analysis of CFTR mutations--correlation with incidence data and application

to screening. Hum Mutat 19, 575-606.

Cheng SH, Gregory RJ, Marshall J, Paul S, Souza DW, White GA, O'Riordan CR & Smith

AE. (1990). Defective intracellular transport and processing of CFTR is the

molecular basis of most cystic fibrosis. Cell 63, 827-834.

Csanady L, Nairn AC & Gadsby DC. (2006). Thermodynamics of CFTR channel gating:

a spreading conformational change initiates an irreversible gating cycle. J Gen

Physiol 128, 523-533.

Cutting GR. (2005). Modifier genetics: cystic fibrosis. Annu Rev Genomics Hum Genet

6, 237-260.

Dalemans W, Barbry P, Champigny G, Jallat S, Dott K, Dreyer D, Crystal RG, Pavirani

A, Lecocq JP & Lazdunski M. (1991). Altered chloride ion channel kinetics

associated with the ∆F508 cystic fibrosis mutation. Nature 354, 526-528.

DeCarvalho AC, Gansheroff LJ & Teem JL. (2002). Mutations in the nucleotide-

binding domain 1 signature motif region rescue processing and functional

defects of cystic fibrosis transmembrane conductance regulator ∆508. J Biol

Chem 277, 35896-35905.

Denning GM, Anderson MP, Amara JF, Marshall J, Smith AE & Welsh MJ. (1992).

Processing of mutant cystic fibrosis transmembrane conductance regulator is

temperature-sensitive. Nature 358, 761-764.

Page 164: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

152

Gray MA, Greenwell JR & Argent BE. (1988). Secretin-regulated chloride channel on

the apical plasma membrane of pancreatic duct cells. J Membr Biol 105, 131-

142.

Haws, C. M., Neponucemo, I. B., Krouse, M. E., Wakelee, H., Law, T., Xia, Y., Nguyen, H.

& Wine, J. J. (1996). ∆F508-CFTR channels: kinetics, activation by forskolin,

and potentiation by xanthines. Am J Physiol 270, C1544-C1555.

Hwang TC & Sheppard DN. (2009). Gating of the CFTR Cl- channel by ATP-driven

nucleotide-binding domain dimerisation. J Physiol 587, 2151-2161.

Hwang, TC., Wang, F., Yang, I. C. & Reenstra, W. W. (1997). Genistein potentiates

wild-type and delta F508-CFTR channel activity. Am. J. Physiol. 273, C988-

C998.

Knowles MR, Stutts MJ, Spock A, Fischer N, Gatzy JT & Boucher RC. (1983).

Abnormal ion permeation through cystic fibrosis respiratory epithelium.

Science 221, 1067-1070.

Lewis HA, Wang C, Zhao X, Hamuro Y, Conners K, Kearins MC, Lu F, Sauder JM,

Molnar KS, Coales SJ, Maloney PC, Guggino WB, Wetmore DR, Weber PC &

Hunt JF. (2010). Structure and dynamics of NBD1 from CFTR characterized

using crystallography and hydrogen/deuterium exchange mass spectrometry.

J Mol Biol 396, 406-430.

Lewis HA, Zhao X, Wang C, Sauder JM, Rooney I, Noland BW, Lorimer D, Kearins MC,

Conners K, Condon B, Maloney PC, Guggino WB, Hunt JF & Emtage S. (2005).

Impact of the ∆F508 mutation in first nucleotide-binding domain of human

cystic fibrosis transmembrane conductance regulator on domain folding and

structure. J Biol Chem 280, 1346-1353.

Lukacs GL, Chang XB, Bear C, Kartner N, Mohamed A, Riordan JR & Grinstein S.

(1993). The ∆F508 mutation decreases the stability of cystic fibrosis

transmembrane conductance regulator in the plasma membrane.

Determination of functional half-lives on transfected cells. J Biol Chem 268,

21592-21598.

Page 165: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

153

Mense M, Vergani P, White DM, Altberg G, Nairn AC & Gadsby DC. (2006). In vivo

phosphorylation of CFTR promotes formation of a nucleotide-binding

domain heterodimer. EMBO J 25, 4728-4739.

Miki H, Zhou Z, Li M, Hwang TC & Bompadre SG. (2010). Potentiation of disease-

associated cystic fibrosis transmembrane conductance regulator mutants by

hydrolyzable ATP analogs. J Biol Chem 285, 19967-19975.

Okiyoneda T, Barriere H, Bagdany M, Rabeh WM, Du K, Hohfeld J, Young JC & Lukacs

GL. (2010). Peripheral protein quality control removes unfolded CFTR from

the plasma membrane. Science 329, 805-810.

Ostedgaard, L. S., Rogers, C. S., Dong, Q., Randak, C. O., Vermeer, D. W., Rokhlina, T.,

Karp, P. H. & Welsh, M. J. (2007). Processing and function of CFTR-∆F508 are

species-dependent. PNAS 104, 15370-15375.

Pissarra LS, Farinha CM, Xu Z, Schmidt A, Thibodeau PH, Cai Z, Thomas PJ, Sheppard

DN &Amaral MD. (2008). Solubilizing mutations used to crystallize one CFTR

domain attenuate the trafficking and channel defects caused by the major

cystic fibrosis mutation. Chem Biol 15, 62-69.

Protasevich I, Yang Z, Wang C, Atwell S, Zhao X, Emtage S, Wetmore D, Hunt JF &

Brouillette CG. (2010). Thermal unfolding studies show the disease causing

F508del mutation in CFTR thermodynamically destabilizes nucleotide-

binding domain 1. Protein Sci 19, 1917-1931.

Riordan JR, Rommens JM, Kerem B, Alon N, Rozmahel R, Grzelczak Z, Zielenski J, Lok

S, Plavsic N, Chou JL & et al. (1989). Identification of the cystic fibrosis gene:

cloning and characterization of complementary DNA. Science 245, 1066-

1073.

Roxo-Rosa M, Xu Z, Schmidt A, Neto M, Cai Z, Soares CM, Sheppard DN & Amaral MD.

(2006). Revertant mutants G550E and 4RK rescue cystic fibrosis mutants in

the first nucleotide-binding domain of CFTR by different mechanisms. Proc

Natl Acad Sci U S A 103, 17891-17896.

Sato S, Ward CL, Krouse ME, Wine JJ & Kopito RR. (1996). Glycerol reverses the

misfolding phenotype of the most common cystic fibrosis mutation. J Biol

Chem 271, 635-638.

Page 166: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

154

Schultz BD, Frizzell RA & Bridges RJ. (1999). Rescue of dysfunctional ∆F508-CFTR

chloride channel activity by IBMX. J Membr Biol 170, 51-66.

Serohijos AW, Hegedus T, Aleksandrov AA, He L, Cui L, Dokholyan NV & Riordan JR.

(2008). Phenylalanine-508 mediates a cytoplasmic-membrane domain

contact in the CFTR 3D structure crucial to assembly and channel function.

Proc Natl Acad Sci U S A 105, 3256-3261.

Sharma M, Pampinella F, Nemes C, Benharouga M, So J, Du K, Bache KG, Papsin B,

Zerangue N, Stenmark H & Lukacs GL. (2004). Misfolding diverts CFTR from

recycling to degradation: quality control at early endosomes. J Cell Biol 164,

923-933.

Swiatecka-Urban A, Brown A, Moreau-Marquis S, Renuka J, Coutermarsh B, Barnaby

R, Karlson KH, Flotte TR, Fukuda M, Langford GM & Stanton BA. (2005). The

short apical membrane half-life of rescued ∆F508-cystic fibrosis

transmembrane conductance regulator (CFTR) results from accelerated

endocytosis of ∆F508-CFTR in polarized human airway epithelial cells. J Biol

Chem 280, 36762-36772.

Thibodeau PH, Brautigam CA, Machius M & Thomas PJ. (2005). Side chain and

backbone contributions of Phe508 to CFTR folding. Nat. Struct. Mol. Biol. 12,

10-16.

Tsai MF, Jih KY, Shimizu H, Li M & Hwang TC. (2010a). Optimization of the

degenerated interfacial ATP binding site improves the function of diseases

related mutant cystic fibrosis transmembrane conductance regulator (CFTR)

channels. J Biol Chem.285, 37663-671..

Tsai MF, Li M & Hwang TC. (2010b). Stable ATP binding mediated by a partial NBD

dimer of the CFTR chloride channel. J Gen Physiol 135, 399-414.

Tsai MF, Shimizu H, Sohma Y, Li M & Hwang TC. (2009). State-dependent

modulation of CFTR gating by pyrophosphate. J Gen Physiol 133, 405-419.

Vergani P, Nairn AC & Gadsby DC. (2003). On the mechanism of MgATP-dependent

gating of CFTR Cl- channels. J Gen Physiol 121, 17-36.

Vergani P, Lockless SW, Nairn AC & Gadsby DC. (2005). CFTR channel opening by

ATP-driven tight dimerization of its nucleotide-binding domains.Nature

433,876-880.

Page 167: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

155

Wang C, Protasevich I, Yang Z, Seehausen D, Skalak T, Zhao X, Atwell S, Spencer

Emtage J, Wetmore DR, Brouillette CG & Hunt JF. (2010). Integrated

biophysical studies implicate partial unfolding of NBD1 of CFTR in the

molecular pathogenesis of F508del cystic fibrosis. Protein Sci 19, 1932-1947.

Wang F, Zeltwanger S, Hu S & Hwang TC. (2000). Deletion of phenylalanine 508

causes attenuated phosphorylation-dependent activation of CFTR chloride

channels. J Physiol 524 Pt 3, 637-648.

Ward CL, Omura S & Kopito RR. (1995). Degradation of CFTR by the ubiquitin-

proteasome pathway. Cell 83, 121-127.

Welsh MJ & Liedtke CM. (1986). Chloride and potassium channels in cystic fibrosis

airway epithelia. Nature 322, 467-470.

Zhou Z, Wang X, Liu HY, Zou X, Li M & Hwang TC. (2006). The two ATP binding sites

of cystic fibrosis transmembrane conductance regulator (CFTR) play distinct

roles in gating kinetics and energetics. J Gen Physiol 128, 413-422.

Zielenski J & Tsui LC. (1995). Cystic fibrosis: genotypic and phenotypic variations.

Annu Rev Genet 29, 777-807.

Page 168: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

156

CHAPTER 5

FUTURE DIRECTIONS

5-1 Overview

As more new evidence emerges, the boundary between ion channels and active

transporters is blurring. People start to consider, what kinds of changes, structurally

or kinetically, are necessary to convert a transporter to a channel or vice versa. As

the only ion channel in a protein family that contains a vast array of active

transporters, CFTR is no doubt an ideal candidate for studying this evolutionary

connection between channels and transporters. A lingering puzzle in the field is the

role of ATP hydrolysis in CFTR’s gating cycle. The major difference between ion

channels and active transporters is the direction of cargo flow. Active transporters

pump the cargo against its concentration gradient; therefore a free energy input is

certainly required. On the other hand, the direction of ionic flow in channels solely

depends on the external driving forces, such as concentration gradient and/or

membrane potential, rendering ATP hydrolysis in CFTR a seemingly redundant

process. But decades of studies no doubt have established the raison d’etre for the

preservation of CFTR’s capability of ATP hydrolysis during evolution: ATP

hydrolysis simply provides a shortcut for the channel to escape from the stable open

conformation and facilitates closing.

Page 169: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

157

In Chapters 2 and 3, I have illustrated the functional role of ATP hydrolysis in

CFTR gating by proposing a new CFTR gating model that features an energetic

coupling between the NBDs and the TMDs. Mechanistically, my experimental results

effectively refuted the popular strict coupling hypothesis by demonstrating the “re-

entry” pathway that allows hydrolysis of more than 1 ATP molecule for each

opening burst. Practically, the new model could pave the way for rational design of

novel CF treatments. First, as mentioned in Chapter 3, the re-entry pathway

provides an alternative route for prolonging the open time and thus can serve as a

drug target. In section 5-2, I will provide preliminary data that suggest an

involvement of the re-entry pathway in the mechanism of a novel FDA-approved

CFTR potentiater, Vx-770 (Kalydeco). Furthermore, the energetic coupling model

depicts that both NBDs and TMDs hold a certain degree of autonomy, hence granting

an opportunity of manipulating CFTR function directly by targeting the gate-forming

TMDs independently of the motion in NBDs. Such a maneuver has been proved

possible by the report by Bai et al. (2010) who showed that after modification of its

TMD by a thiol reagent, MTSET, CFTR can open frequently even in the complete

absence of ATP (Bai et al., 2010). In section 5.3, I will elaborate the feasibility of

applying this concept to CF gene therapy.

Mechanistic insights from the studies elaborated in Chapter 2 and 3 underscore

the importance of the TMDs in CFTR function. With the ongoing projects of scanning

different transmembrane helixes being carried out by my colleagues, another

interesting question arose owing to the unique feature of the R352C mutation

described in Chapter 3. Removing the positive charge at position 352 reveals two

Page 170: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

158

distinct levels of conductance (O1 and O2 states, see Chapter 3 for details). As the

transition between the O1 and O2 state is triggered by ATP hydrolysis during an

opening burst, there must be a conformational change in the TMDs that takes place

concurrently with ATP hydrolysis and such conformational change likely alters the

electrostatic effect of R352 in attracting the chloride ions thus resulting in an

ununiformed conductance levels in R352C-CFTR. Studying the role of R352 in pre-

and post-hydrolytic states can provide clues in the pore architecture and the

mechanism of ATP-induced conformational change in the TMDs. In section 5-3, I will

elaborate potential projects aimed at unveiling the pore architecture and ATP-

hydrolysis-induced conformational changes by engineering positively charged

amino acids at different positions of TM6 and TM12.

5-2 Unraveling the mechanism of CFTR potentiator, Vx-770

In recent decades, significant progress has been made in the treatment of CF.

The life expectancy for CF patients is now approaching ~40 years (Elborn et al.,

1991). However, most CF treatments only relieve the symptoms caused by CF

instead of rectifying the underlying cause; namely the malfunction of CFTR channel

(Rowe et al., 2005; Flume et al., 2009a; Flume et al., 2009b). Recently, the discovery

of a CFTR potentiator, Vx-770 (Kalydeco), via high-throughput drug screening has

successfully evolved into the first CF treatment that actually targets the fundamental

defect. Vx-770 has shown promising clinical outcomes in CF patients carrying the

G551D mutation (Accurso et al., 2010; Ramsey et al., 2011) and was recently

Page 171: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

159

approved by the FDA. The discovery of Vx-770 thus opened a new chapter for CF

treatments as we finally are able to tackle the origin of this disease.

Vx-770 can act on WT-CFTR as well as in many disease-related mutations of the

channel (Yu et al., 2012). It boosts the Po of CFTR by prolonging its open time (Van

Goor et al., 2009; Yu et al., 2012). Since the strict coupling model demands that the

formation and separation of the NBD dimer are strictly coupled to the opening and

closing of the gate respectively, the only two ways to prolong the open time is to

delay NBD dimer separation or to impede ATP hydrolysis. But neither of these two

possible mechanisms can offer an adequate explanation for the effect of Vx-770 on

G551D-CFTR, a mutation that likely prevents NBD dimerization from happening due

to its unique position in the signature sequence (Bompadre et al., 2008) and for the

same reason eliminates ATP hydrolysis (Li et al., 1996; Ramjeesingh et al., 2008). On

the other hand, the energetic coupling model proposed in Chapter 3 offers a simple

explanation. By hypothesizing that Vx-770 enhances ATP-independent gating, we

can explain the kinetic effects of Vx-770 on both WT- and G551D-CFTR. For G551D-

CFTR, since the ATP-independent openings are preserved in this mutant, our

hypothesis is that the activity of G551D-CFTR will be enhanced by Vx-770. On the

other hand, the prolongation of the open time for WT channels can be achieved

simply by promoting the re-entry pathway secondary to an enhanced ATP-

independent gating: the more stable the O2 state, the more frequently the channel

will take the re-entry pathway (see Chapter 3 for detail elaborations).

Page 172: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

160

As mentioned in Chapter 3, the C ↔ O2 transition represents the ATP-

independent opening events; a direct consequence of stabilizing the O2 state is an

increased ATP-independent activity. To test the idea that Vx-770 may affect ATP-

independent gating, I measured the ratio of ATP-independent current to ATP-

dependent current in the presence or absence of Vx-770. In WT-CFTR, I indeed

observed much higher ATP-independent activity relative to ATP-dependent current

in the presence of Vx-770 (Figure 5-1). Notably, the Po of Vx-770 treated WT-CFTR

is ~ 0.7, almost twice as high as that in the absence of Vx-770 (Van Goor et al., 2009;

Yu et al., 2012). Taking this into account, the bar chart in Figure 5-1C actually

underestimates the effect of Vx-770 on ATP-independent activity. After

normalization by their respective ATP-dependent Po, Vx-770 actually increases the

ATP-dependent Po by more than 6 fold (Figure 5-1 D).

A key concept in the energetic coupling model is that the re-entry transition is

an ATP-dependent process. The channel is more likely to go through the re-entry

pathway when the [ATP] is high. As the re-entry pathway traps the channel in the

open state, we expect that the open time of WT-CFTR would depend on the [ATP] if

the re-entry frequency is significantly increased by Vx-770. Consistent with this idea,

we found that in Vx-770 treated WT-CFTR, the mean open time is prolonged with

increasing [ATP] (Figure 5-2). Intriguingly, such a trend is not usually found in WT-

CFTR without Vx-770 treatment (Winter et al., 1994; Zeltwanger et al., 1999;

Vergani et al., 2003; Bompadre et al., 2005), probably because in the absence of Vx-

770 the re-entry frequency is too low for WT-CFTR to show the effect of [ATP] on

the open time.

Page 173: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

161

Figure 5-1

Figure 5-1. Vx-770 increases ATP-independent activity in WT-CFTR. (A and B)

Macroscopic current was induced by 2.75 mM ATP. After reaching the steady state,

ATP was washed out and the current was recorded in the absence of ATP (expanded

in red and blue box). The aforementioned protocol was performed in the absence (A)

or presence (B) of 200 nM Vx-770. (C) A bar chart summarizes the results in (A) and

(B). The ratio of ATP-independent current to ATP-dependent current was measured.

(D) The Po of ATP-independent current in the presence (right) or absence (left) of

Vx-770. The Po is calculated by normalizing the ATP-dependent current to its Po (in

the absence of Vx-770, Po = 0.4; in the presence of Vx-770, Po = 0.7).

Page 174: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

162

Figure 5-2

Figure 5-2. [ATP]-dependent mean open time of WT-CFTR in the presence of

Vx-770. (A) Representative traces of single WT-CFTR in the presence of 200 nM Vx-

770 and 10 mM (upper) or 100 µM (lower) ATP. (B) In the presence of Vx-770, the

mean open time of WT-CFTR is prolonged in higher (milimolar) [ATP] (black dots).

Such a trend is not seen in the absence of Vx-770 (blue dots). Because of the large

variation of the mean open time for WT-CFTR, the error of each data point is

approximately 50 to 100 ms, which may account for the absence of such a trend in

WT-CFTR (blue dots). In addition, it is likely that for WT-CFTR the frequency of re-

entry events is much lower in the absence of Vx-770 due to an unstable O2 state;

therefore the difference in open time at different [ATP] is obscured by the large

variation.

Page 175: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

163

Figure 5-3

Figure 5-3. Representative traces of R352C-CFTR in the presence of Vx-770.

The frequency of re-entry events in the presence of Vx-770 is significantly higher

than that in the absence of Vx-770 (Also see Table 5-1).

Page 176: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

164

The data related R352C-CFTR in Chapter 3 provides an opportunity to directly

scrutinize the effect of Vx-770 on the re-entry pathway. In Figure 5-3, we found that

the opening bursts are significantly prolonged in the presence of Vx-770.

Furthermore, the frequency of the re-entry pathway is increased when R352C-CFTR

was treated with Vx-770 (Table 5-1), again consistent with our hypothesis.

Our preliminary data suggest that Vx-770 likely promotes the re-entry pathway

and ATP-independent activity. Increased ATP-independent activity by Vx-770

indicates that the drug shifts the equilibrium between the C2 and O2 states (Figure

3-5). Since such C2 ↔ O2 transition occurs in the absence of ATP-induced

conformational change in the NBDs, Vx-770 likely achieves its effect by perturbing

the energetics of the TMDs. This idea makes sense especially when considering the

fact that Vx-770 is an extremely hydrophobic molecule that presumably could easily

partition into the lipid bilayer where the TMDs lie. This hypothesis can be tested by

adding Vx-770 in the pipette solution. If Vx-770 acts on the TMDs, it should be able

to partition into the lipid bilayer and reaches its binding site regardless it is applied

from the intracellular or extracellular side. Therefore, adding Vx-770 in the pipette

solution should potentiate CFTR channel to the same extent as when it is applied

from the cytoplasmic side. As mentioned in Chapter 1, Vx-770 is the progenitor of

medication targeting CFTR, but it benefits only a small portion of CF populations.

Thus, understanding its mechanism and site of action is pivotal for future

development of new drugs that could benefit more CF patients.

Page 177: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

165

5-3 Potential impacts of the energetic coupling model in CF gene

therapy

CF is a genetic disease that is caused by mutations of a single gene. Gene therapy

for CF has been proposed ever since the gene was cloned more than two decades

ago (Hyde et al., 1993; Yang et al., 1994; Fisher et al., 1996). One major obstacle in

CF gene therapy is the large size of the CFTR gene. The 4450 bp for the coding

sequence of CFTR is too large to fit into the adeno-associated virus (AAV) vector

with the addition of a high efficacy promoter (Flotte et al., 1993). On realizing this,

numerous attempts have been made to reduce the size of CFTR while preserving its

normal function. Removing the R domain became an option as it had been reported

that deletion of the R domain still yields a functional CFTR channel (Rich et al., 1991;

Bompadre et al., 2005). However using CFTR without the R domain for gene therapy

is potentially hazardous. Under physiological conditions, the activity of CFTR

channel mostly depends on the level of phosphorylation, removing the R domain

will therefore make the channel uncontrollable and may result in severe undesirable

outcomes.

One central feature of the energetic coupling model is a certain degree of

autonomy in both NBDs and TMDs. If the gate in TMDs indeed opens independently

of NBDs, one predicts that energetic perturbation of TMDs can alter ATP-

independent gating. Echoing this idea, a previous study by my colleague showed

that after the positively charged thiol reagent MTSET modified the cysteine

Page 178: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

166

Table 5-1

Table 5-1. Summary of different types of opening events for R352C-CFTR in

the presence of 200 nM Vx-770. The percentage of the re-entry evens (the (O1-

O2)n category) is much higher than that in the absence of Vx-770 (Tabel 3-1).

O1 - O2 O1 O2 - O1O2 (O1 - O2)n

R352C + 200 nM Vx-770

Total

2.75 mM ATP

100 M ATPµ

364 (56 %) 83 (13 %) 54 (8 %) 14 (2 %) 136 (21 %) 651 (100 %)

663 (56 %) 216 (18 %) 137 (12 %) 32 (3 %) 128 (11 %) 1176 (100 %)

Page 179: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

167

Figure 5-4

Figure 5-4. MTSET modification dramatically increases the current of

Cysless/I344C/∆NBD2-CFTR. The potentiation effect can be reversed by adding 5

mM DTT. This result suggests that modifying TMDs alone is sufficient to open CFTR

channel without NBD2.

2.75 mM ATP

1 mM MTSET

100 s

20

0 p

A

Cysless/I344C/ NBD2-CFTR∆

5 mM DTT

Page 180: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

168

engineered in the transmembrane helix 6, the mutant CFTR channel exhibit

extremely high activity even in the absence of ATP (Bai et al., 2010), suggesting the

TMDs alone is sufficient to carry out channel function. Furthermore, our preliminary

data demonstrated that the same maneuver can work in channels that have had its

entire NBD2 removed (Figure 5-4).

Since NBD2 contains ~ 250 amino acids, a CFTR construct without the entire coding

sequence for NBD2 will be small enough for making an AAV for gene delivery. The

challenge is to find a TMD mutation that imparts a significant level of ATP-

independent activity. Once such mutation is identified, the NBDs become redundant

and can be removed to shrink the size of CFTR gene. The advantage of this approach

over the deletion of R domain is that channels with its NBD2 truncated may still

retain phosphorylation-dependent regulation and hence may minimize the side

effects of hyperactive CFTR channels.

5-4 Roles of positive charge in the transmembrane domains

In Chapter 3, I introduced a mutation that reveals two distinct levels of open

state conductance (O1 and O2 respectively). As the transition between the two open

states is triggered by ATP-hydrolysis, this R352C mutation became a precise tool for

studying the gating mechanisms associated with ATP-hydrolysis. Furthermore, this

result also implies that, besides opening and closing of the gate, the TMDs undergo a

conformational change during the opening. Such a conformational change is masked

Page 181: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

169

in WT as the single channel conductance remains constant throughout the entire

opening burst and it is unveiled by the R352C mutation.

As demonstrated in Chapter 3, the positive charge in the TM6 likely plays a

critical electrostatic role in attracting intracellular chloride. After removing the

positively charged arginine, we found the conductance of the pre-hydrolytic open

state (O1 state) to be much smaller than that of the post-hydrolytic open state (O2

state); indicating that the chloride-attracting effect of the Arg352 is greater in the

O1 state than in the O2 state. Structurally, it is likely due to the rotational and/or

translational movement of the TM6 that alters the orientation and position of the

Arg352 side chain (Bai et al., 2010). Moreover, the electrostatic effect of Arg352 is

likely transferable. It had been demonstrated that swaping the R352 in TM6 and

D993 in TM9 results in a channel that behaves like WT (uniform conductance) (Cui

et al., 2008). Further supporting this notion, unpublished data from our laboratory

revealed that for R352C-CFTR, engineering a positive charge to the pore lining

residues in TM12 also (R352C/W1145R) resulted in a uniform channel conductance.

These findings indicate these TMs are physically proximal to each other so that the

placing the positive charge in either locations affects gating in a similar manner.

In the study described in Chapter 3, we also discovered that after the

Cysless/R352Q/I344C mutant channel modified by the positively charged MTSET at

the 344 position, the reduction of single channel conductance was partially

compensated and the relative difference between the O1 and O2 state also

decreased. This result suggests the charge is not only transferable between different

Page 182: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

170

TMs but also within the TM6. According to the SCAM results described by Bai et al.

(Bai et al., 2010), the 344 position is two helical turn above the 352 position and

both residues line the pore, thus our results suggests that positioning the positive

charge deeper into the pore can still attracts chloride but with less efficacy.

Furthermore, the fact that MTSET modification altered the conductance of O1 and

O2 state to a different extent suggests that the TM6 helix around the 344 position

also undergo a certain degree of conformational change.

Based on the previous study that had pinpointed the pore-lining residues in the

TM6 (Bai et al., 2010), we can engineer the positive charge at different positions of

the TM6 and measure how it would affect the O1 and O2 conductance. Such a study

not only can unravel the role of the positive charge in the TM6, but will also provide

clues regarding the pore architecture and the ATP-hydrolysis-induced

conformational change.

5-5 References

Accurso, F.J., S.M. Rowe, J.P. Clancy, M.P. Boyle, J.M. Dunitz, P.R. Durie, S.D. Sagel, D.B.

Hornick, M.W. Konstan, S.H. Donaldson, R.B. Moss, J.M. Pilewski, R.C.

Rubenstein, A.Z. Uluer, M.L. Aitken, S.D. Freedman, L.M. Rose, N. Mayer-

Hamblett, Q. Dong, J. Zha, A.J. Stone, E.R. Olson, C.L. Ordonez, P.W. Campbell,

M.A. Ashlock, and B.W. Ramsey. 2010. Effect of VX-770 in persons with cystic

fibrosis and the G551D-CFTR mutation. N Engl J Med. 363:1991-2003.

Bai, Y., M. Li, and T.C. Hwang. 2010. Dual roles of the sixth transmembrane segment

of the CFTR chloride channel in gating and permeation. J Gen Physiol.

136:293-309.

Bompadre, S.G., T. Ai, J.H. Cho, X. Wang, Y. Sohma, M. Li, and T.C. Hwang. 2005. CFTR

gating I: Characterization of the ATP-dependent gating of a phosphorylation-

independent CFTR channel (DeltaR-CFTR). J Gen Physiol. 125:361-375.

Page 183: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

171

Bompadre, S.G., M. Li, and T.C. Hwang. 2008. Mechanism of G551D-CFTR (cystic

fibrosis transmembrane conductance regulator) potentiation by a high

affinity ATP analog. J Biol Chem. 283:5364-5369.

Cui, G., Z.R. Zhang, A.R. O'Brien, B. Song, and N.A. McCarty. 2008. Mutations at

arginine 352 alter the pore architecture of CFTR. J Membr Biol. 222:91-106.

Elborn, J.S., D.J. Shale, and J.R. Britton. 1991. Cystic fibrosis: current survival and

population estimates to the year 2000. Thorax. 46:881-885.

Fisher, K.J., H. Choi, J. Burda, S.J. Chen, and J.M. Wilson. 1996. Recombinant

adenovirus deleted of all viral genes for gene therapy of cystic fibrosis.

Virology. 217:11-22.

Flotte, T.R., S.A. Afione, R. Solow, M.L. Drumm, D. Markakis, W.B. Guggino, P.L. Zeitlin,

and B.J. Carter. 1993. Expression of the cystic fibrosis transmembrane

conductance regulator from a novel adeno-associated virus promoter. J Biol

Chem. 268:3781-3790.

Flume, P.A., P.J. Mogayzel, Jr., K.A. Robinson, C.H. Goss, R.L. Rosenblatt, R.J. Kuhn, and

B.C. Marshall. 2009a. Cystic fibrosis pulmonary guidelines: treatment of

pulmonary exacerbations. American journal of respiratory and critical care

medicine. 180:802-808.

Flume, P.A., K.A. Robinson, B.P. O'Sullivan, J.D. Finder, R.L. Vender, D.B. Willey-

Courand, T.B. White, and B.C. Marshall. 2009b. Cystic fibrosis pulmonary

guidelines: airway clearance therapies. Respiratory care. 54:522-537.

Hyde, S.C., D.R. Gill, C.F. Higgins, A.E. Trezise, L.J. MacVinish, A.W. Cuthbert, R. Ratcliff,

M.J. Evans, and W.H. Colledge. 1993. Correction of the ion transport defect in

cystic fibrosis transgenic mice by gene therapy. Nature. 362:250-255.

Li, C., M. Ramjeesingh, and C.E. Bear. 1996. Purified cystic fibrosis transmembrane

conductance regulator (CFTR) does not function as an ATP channel. J Biol

Chem. 271:11623-11626.

Ramjeesingh, M., F. Ugwu, F.L. Stratford, L.J. Huan, C. Li, and C.E. Bear. 2008. The

intact CFTR protein mediates ATPase rather than adenylate kinase activity.

Biochem J. 412:315-321.

Ramsey, B.W., J. Davies, N.G. McElvaney, E. Tullis, S.C. Bell, P. Drevinek, M. Griese, E.F.

McKone, C.E. Wainwright, M.W. Konstan, R. Moss, F. Ratjen, I. Sermet-

Gaudelus, S.M. Rowe, Q. Dong, S. Rodriguez, K. Yen, C. Ordonez, and J.S. Elborn.

2011. A CFTR potentiator in patients with cystic fibrosis and the G551D

mutation. N Engl J Med. 365:1663-1672.

Page 184: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

172

Rich, D.P., R.J. Gregory, M.P. Anderson, P. Manavalan, A.E. Smith, and M.J. Welsh.

1991. Effect of deleting the R domain on CFTR-generated chloride channels.

Science. 253:205-207.

Rowe, S.M., S. Miller, and E.J. Sorscher. 2005. Cystic fibrosis. N Engl J Med. 352:1992-

2001.

Van Goor, F., S. Hadida, P.D. Grootenhuis, B. Burton, D. Cao, T. Neuberger, A. Turnbull,

A. Singh, J. Joubran, A. Hazlewood, J. Zhou, J. McCartney, V. Arumugam, C.

Decker, J. Yang, C. Young, E.R. Olson, J.J. Wine, R.A. Frizzell, M. Ashlock, and P.

Negulescu. 2009. Rescue of CF airway epithelial cell function in vitro by a

CFTR potentiator, VX-770. Proc Natl Acad Sci U S A. 106:18825-18830.

Vergani, P., A.C. Nairn, and D.C. Gadsby. 2003. On the mechanism of MgATP-

dependent gating of CFTR Cl- channels. J Gen Physiol. 121:17-36.

Winter, M.C., D.N. Sheppard, M.R. Carson, and M.J. Welsh. 1994. Effect of ATP

concentration on CFTR Cl- channels: a kinetic analysis of channel regulation.

Biophys J. 66:1398-1403.

Yang, Y., F.A. Nunes, K. Berencsi, E. Gonczol, J.F. Engelhardt, and J.M. Wilson. 1994.

Inactivation of E2a in recombinant adenoviruses improves the prospect for

gene therapy in cystic fibrosis. Nature genetics. 7:362-369.

Yu, H., B. Burton, C.J. Huang, J. Worley, D. Cao, J.P. Johnson, Jr., A. Urrutia, J. Joubran, S.

Seepersaud, K. Sussky, B.J. Hoffman, and F. Van Goor. 2012. Ivacaftor

potentiation of multiple CFTR channels with gating mutations. Journal of

cystic fibrosis : official journal of the European Cystic Fibrosis Society.

Zeltwanger, S., F. Wang, G.T. Wang, K.D. Gillis, and T.C. Hwang. 1999. Gating of cystic

fibrosis transmembrane conductance regulator chloride channels by

adenosine triphosphate hydrolysis. Quantitative analysis of a cyclic gating

scheme. J Gen Physiol. 113:541-554.

Page 185: MOLECULAR PHYSIOLOGY AND PHARMACOLOGY OF THE CFTR …

173

VITA

Kang-Yang Jih was born in Kaohsiung City, Taiwan in March 24th, 1987. He attended

National Yang-Ming University in 2005, majoring in Medicine. He enrolled in the

M.D. Ph.D. program in 2009 and pause his medical study in the same year. In 2009,

he joined the Department of Medical Pharmacology and Physiology in University of

Missouri-Columbia to pursue his Ph.D. degree. He expects to graduate in July, 2012

and return to Taiwan to complete the M.D. degree.