Mechanisms of anion and fluid secretion by airway...

180
Mechanisms of anion and fluid secretion by airway epithelial cells Jiajie Shan Department of Physiology McGill University Montreal, Quebec, Canada December 2013 A thesis submitted to the faculty of Graduate Studies and Research in partial fulfillment of the degree of Doctorate in Philosophy Copyright © Jiajie Shan 2013

Transcript of Mechanisms of anion and fluid secretion by airway...

Mechanisms of anion and fluid secretion by airway epithelial cells

Jiajie Shan

Department of Physiology

McGill University Montreal, Quebec, Canada

December 2013

A thesis submitted to the faculty of Graduate Studies and Research in partial fulfillment of the degree of

Doctorate in Philosophy

Copyright © Jiajie Shan 2013

I

Table of contents

Abstract ………………………………………………………………………………………………………… VII

Résumé .............................................................................................................. IX

Acknowledgements ……………………………………………………………………………………... XI

Contributions of authors ..…………………………………………………………………………….. XII

Abbreviation …………………………………………………………………………………………………. XIV

Chapter 1 Introduction

1.1 The structure and function of airway epithelium …………………………………… 2

1.1.1 Ciliated cells …………………………………………………………………………….. 3

1.1.2 Basal cells ………………………………………………………………………………… 3

1.1.3 Mucous cells ……………………………………………………………………………. 3

1.1.4 Clara cells …………………………………………………………………………………. 4

1.1.5 Serous cells ………………………………………………………………………………. 4

1.1.6 Type I cells ………………………………………………………………………………… 5

1.1.7 Type II cells ……………………………………………………………………………….. 5

1.2 Junctional complexes in airways ……………………………………………………………… 5

1.3 The primary functions of airway epithelium: mucus production and

mucociliary clearance …………………………………………………………………………….. 6

1.4 Regulation of ASL ……………………………………………………………………………………. 8

1.4.1 Na+/K+-ATPase ……………………………………………………………………………. 9

1.4.2 V-ATPase ……………………………………………………………………………………. 10

II

1.4.3 NKCC ………………………………………………………………………………………….. 11

1.4.4 NBC …………………………………………………………………………………………… 11

1.4.5 The Na+-independent Cl-/HCO3- exchanger (AE) ………………………… 12

1.4.6 Pendrin (SLC26A4) …………………………………………………………………….. 12

1.4.7 The Na+/H+ exchanger (NHE) ……………………………………………………… 13

1.4.8 The H+ channel …………………………………………………………………………… 14

1.4.9 The chloride channels ………………………………………………………………… 14

1.4.10 ENaC ………………………………………………………………………………………….. 16

1.4.11 Carbonic anhydrase ……………………………………………………………………. 17

1.4.12 Regulation of ASL hydration ………………………………………………………. 18

1.4.13 pH regulation of ASL ………………………………………………………………….. 19

1.5 Models of anion transport in submucosal glands ……………………………………. 22

Chapter 2 anion transport in airway epithelial cells

2.1 Abstract …………………………………………………………………………………………………… 31

2.2 Introduction …………………………………………………………………………………………….. 32

2.3 Methods ………………………………………………………………………………………………….. 35

2.3.1 Cell culture …………………………………………………………………………………. 35

2.3.2 Solutions ……………………………………………………………………………………. 36

2.3.3 Immunoblotting …………………………………………………………………………. 37

2.3.4 Measurements of equivalent Isc and HCO3- secretion …………………. 37

2.3.5 Tracer fluxes ……………………………………………………………………………….. 38

III

2.3.6 Data analysis ………………………………………………………………………………. 39

2.4 Results ……………………………………………………………………………………………………… 39

2.4.1 HCO3- secretion accounts for the short-circuit current (Isc) …………. 39

2.4.2 HCO3- secretion is CFTR dependent ……………………………………………. 40

2.4.3 Proton secretion neutralizes some of the transported bicarbonate . 48

2.4.4 CFTR mediates apical HCO3- conductance …………………………………… 50

2.4.5 Electrogenic anion transport (Ieq) requires Na+ and CO2/HCO3- ……. 54

2.4.6 Some basolateral Cl- entry is independent of NKCC …………………….. 56

2.4.7 Evidence for basolateral Cl- loading by anion exchange ………………. 58

2.4.8 Electrogenic HCO3- secretion requires carbonic anhydrase …………. 62

2.4.9 Origin of the basal Ieq and HCO3- secretion …………………………………. 62

2.5 Discussion ………………………………………………………………………………………………… 69

2.5.1 Isc is mediated by electrogenic HCO3- transport …………………………… 69

2.5.2 Ieq depends on Na+ and CO2/HCO3- ……………………………………………… 70

2.5.3 Cl- loading depends on basolateral anion exchange mostly rather

than NKCC …………..…………………………………………………………………….. 71

2.5.4 Forskolin-stimulated HCO3- secretion requires carbonic anhydrase .. 72

2.5.5 Further evidence that CFTR mediates apical membrane HCO3-

and Cl- conductance …………………………………………………………………… 73

2.5.6 A revised anion transport model of Calu-3 cells …………………………. 75

Chapter 3 Fluid secretion by an airway epithelial cell line

IV

3.1 Abstract ………………………………………………………………………………………………….. 80

3.2 Introduction ……………………………………………………………………………………………. 81

3.3 Methods …………………………………………………………………………………………………. 84

3.3.1 Cell culture ………………………………………………………………………………… 84

3.3.2 Media ………………………………………………………………………………………… 84

3.3.3 Volume and composition of the secreted fluid ………………………….. 85

3.3.4 Data analysis ……………………………………………………………………………… 86

3.4 Results …………………………………………………………………………………………………….. 87

3.4.1 Fluid secretion is CFTR dependent ……………………………………………… 87

3.4.2 Cl- is the predominant anion in Calu-3 secretion ………………………… 87

3.4.3 HCO3-, Na+ and Cl- are all required for forskolin-stimulated

fluid secretion ……………………………………………………………………………. 89

3.4.4 Most of the fluid secretioin is independent of NKCC ………………….. 92

3.4.5 Basolateral anion exchange is important to fluid secretion ………… 92

3.4.6 Carbonic anhydrase has little effect on fluid secretion ……………….. 94

3.4.7 The time course of fluid secretion rate during 24 h stimulation …. 97

3.4.8 Evaporation does not affect the rate of fluid accumulation ………… 97

3.4.9 Osmolality determined the fluid secretion rate ………………………….. 99

3.5 Disscussion ………………………………………………………………………………………………. 101

3.5.1 Most fluid is driven by the net flux of Cl- ……………………………………… 103

3.5.2 Calu-3 secretions are only weakly alkaline ………………………………….. 104

3.5.3 A very low rate of net Cl- secretion is sufficient to

V

drive fluid secretion ……………………………………………………………………. 106

3.5.4 The osmolality determines the fluid secretion rate ……………………. 106

3.5.5 Carbonic anhydrase participates little in fluid secretion ……………… 108

Chapter 4 The role of carbonic anhydrase in anion and fluid secretion

4.1 Abstract ……………………………………………………………………………………………………. 112

4.2 Introduction ……………………………………………………………………………………………… 113

4.3 Methods …………………………………………………………………………………………………… 115

4.3.1 Cell culture …………………………………………………………………………………. 115

4.3.2 Solutions …………………………………………………………………………………….. 116

4.3.3 Measurements of equivalent Isc and HCO3- secretion ………………….. 117

4.3.4 RT-PCR ………………………………………………………………………………………… 117

4.3.5 Immunoblotting ………………………………………………………………………….. 118

4.3.6 Immunoprecipitation ………………………………………………………………….. 119

4.3.7 Fluid secretion assay …………………………………………………………………… 119

4.3.8 Data analysis ………………………………………………………………………………. 120

4.4 Results ……………………………………………………………………………………………………… 120

4.4.1 Carbonic anhydrase contributes partially to anion transport ……… 120

4.4.2 The expression of different isoforms of carbonic anhydrase ………. 121

4.4.3 CAIX is not involved in anion and fluid secretion by Calu-3 cells …. 123

4.4.4 CAII and CAXII are responsible for the forskolin-stimulated anion

secretion by Calu-3 cells …………………………………………………………….. 125

VI

4.4.5 CAII does not form a complex with CFTR ……………………………………… 128

4.5 Discussion ………………………………………………………………………………………………… 129

4.5.1 Carbonic anhydrase II is important for anion and fluid secretion .. 129

4.5.2 Further evidence that HCO3- is required for Cl- secretion ……………. 131

4.5.3 The bicarbonate transport metabolon ………………………………………… 132

4.5.4 The contributions of CAII and NBC to anion and fluid secretion …. 133

Chapter 5 General discussion

5.1 A novel model of anion transport of airway epithelial cells ………………………. 136

5.1.1 V-clamp vs. I-clamp …………………………………………………………………….. 136

5.1.2 pH-stat conditions ………………………………………………………………………. 138

5.1.3 The Calu-3 cell line ……………………………………………………………………… 138

5.1.4 Modification to existing models ………………………………………………….. 140

5.2 The determinants of fluid secretion ………………………………………………………….. 142

5.3 The role of carbonic anhydrases ……………………………………………………………….. 143

5.4 Conclusions and future direction ………………………………………………………………. 144

References ………………………………………………………………………………………………….. 146

VII

Abstract

Anion transport drives fluid into the airways and is essential for humidifying inspired

air and supplying surface liquid for mucociliary transport. Despite the importance of

airway secretion in diseases such as cystic fibrosis, the cellular mechanisms of anion

and fluid secretion remain poorly understood. To clarify these mechanisms, I studied

the Calu-3 cell line, a widely used model for serous cells in airway submucosal glands.

First, I studied anion transport using a technique which combines electrical

measurement of ion transport and pH-stat. Although a small fraction of the Cl-

secretion required basolateral entry through Na+, K+, 2Cl- cotransporter NKCC, I found

that most Cl- uptake was mediated by basolateral Cl-/HCO3- exchange. The HCO3

-

required for this exchange process was supplied by the basolateral Na+, HCO3-

cotransporter NBC and also by the hydration of CO2, which was catalyzed by carbonic

anhydrase. The data were consistent with most HCO3- and Cl- secretion occurring via

apical CFTR channels. Based on these findings, I propose a new model for

transepithelial anion secretion, which may explain early studies in the 1980’s which

indicated that most tracheal submucosal gland secretion is surprisingly insensitive to

bumetanide.

This new scheme for anion transport was further tested using fluid secretion assays

and was found to explain the ion dependencies and pharmacology of fluid secretion.

Moreover, fluid was actively secreted in the first 6 hours during the 24-h fluid

VIII

secretion assay, and secretions collected during that period had higher osmolality

than the culture medium. These results suggest that airway surface fluid is regulated

by osmolality and the water flux simply depends on the total amount of osmotically

active solute on the apical surface.

Finally, I explored the role of carbonic anhydrases because inhibitor studies indicated

that it is critically important for anion transport by Calu-3 cells. I found that several

isoforms of carbonic anhydrases are expressed in Calu-3 cells, namely CAII, CAIX and

CAXII. Inhibitor studies revealed that CAIX is not required for anion transport. CAII

is probably the isoform involved in bicarbonate secretion, although further studies

are needed to exclude the involvement of CAXII.

Taken together, these present findings support a novel cellular mechanism for anion

and fluid secretion by Calu-3 cells. This scheme may be applicable to airway

submucosal glands which control the pH, volume, and composition of airway surface

liquid and are critical for airway host defense.

IX

Résumé

Le transport d’anions entraîne le fluide dans les voies aériennes et est essentiel pour

l’humidification de l’air inspiré et pour apporter du liquide de surface nécessaire au

transport mucoscilliaire. Malgré l’importance de la sécrétion au niveau des voies

respiratoires dans le cas de maladies telles que la fibrose kystique, les mécanismes

cellulaires de la sécrétion d’anions et de fluide ne sont pas tout à fait élucidés. Afin de

clarifier les mécanismes, j’ai utilisé la lignée cellulaire Calu-3, un modèle connu de

cellules séreuses dans les glandes submucosales des voies aériennes.

En premier lieu, j’ai étudié le transport d’anions en utilisant une technique

combinant la mesure électrique de transport d’ions et le pH-stat. Bien qu’une petite

fraction de sécrétion de Cl- requiert l’entrée par le co-transporteur basolatéral Na+, K+,

2Cl- NKCC, j’ai démontré que la plupart de l’apport de Cl- est médié par l’échange

Cl-/HCO3-basolatéral. Le HCO3

- requis pour cet échange est apporté par le

co-transporteur basolatéral Na+, HCO3- NBC et également par l’hydratation de CO2,

qui est catalysée par l’anhydrase carbonique. Les données corrèlent avec la plupart

de la sécrétion de HCO3- et Cl- par les canaux CFTR apicaux. D’après ces résultats, je

propose un nouveau modèle de sécrétion transépithéliale d’anions, qui pourrait

expliquer des études précédentes datant des années 1980, qui indiquaient que la

sécrétion des glandes submucosales de trachée est majoritairement insensible au

bumétanide.

X

Ce nouveau modèle de transport d’anions a ensuite été testé par la méthode de

sécrétion de fluide et a permis d’expliquer les dépendances en ions et la

pharmacologie de la sécrétion de fluide. Aussi, le fluide est activement secrété

pendant les 6 premières heures au cours d’un essai de sécrétion de fluide sur 24

heures, et les secrétions collectées pendant cette période ont une osmolarité plus

importante que le milieu de culture. Ces résultats suggèrent que le fluide de surface

des voies aériennes est régulé par l’osmolarité et que le flux d’eau dépend

simplement de la quantité totale de soluté osmoticallement actif à la surface apicale.

Enfin, j’ai étudié le rôle des anhydrases carboniques car mes expériences d’inhibition

démontrent son importance dans le transport d’anions par les cellules Calu-3. J’ai mis

en évidence qu’il existe trois isoformes d’anhydrases carboniques exprimées dans les

cellules Calu-3 : CAII, CAIX et CAXII. Des expériences d’inhibition ont révélé que CAIX

n’est pas impliqué dans le transport d’anions. CAII est probablement l’isoforme

impliqué dans la sécrétion de bicarbonate, bien que des expériences

complémentaires soient nécessaires pour exclure l’implication de CAXII.

Pour conclure, ces résultats appuient un nouveau mécanisme cellulaire de sécrétion

d’anions et de fluide par les cellules Calu-3. Ce modèle pourrait être applicable aux

glandes submucosales des voies respiratoires qui contrôlent le pH, le volume et la

composition du liquide de surface des voies aériennes et qui sont essentielles à leur

défense.

XI

Acknowledgements

I would like to thank my supervisor Dr. John Hanrahan for his guidance, patience and

support throughout my study. I am grateful for his dedication and drive, without

which this work would not have been possible. I am also particularly thankful that I

learned a lot from his enormous knowledge throughout these years, which has

become my precious assets.

I would also like to thank my lab members, past and present. In particular, Dr. Jie Liao,

Mr. Junwei Huang, Ms. Adeline Wohlhuter, Ms. Alexandra Evagelidis and Ms. Julie

Goepp for their invaluable help throughout my study. The daily pleasant and joyous

company of my lab members eases my mind when I am stuck in the boring lab work.

I am very thankful to my committee members, Dr. John Orlowski, Dr. Gergely Lukacs

and Dr. Jean-Yves Lapointe. Their expertise provide me with insight, useful advice and

feedback for my work. I must thank them for spending their precious time in meeting

with me and discussing my work.

Finally, I would like to thank my parents for their constant support since I was born,

without them no accomplishment would be achieved.

XII

Contribution of authors

Chapter 2 contains part of a published manuscript: Shan J. et al.

Bicarbonate-dependent chloride transport drives fluid secretion by the human

airway epithelial cell line Calu-3. J Physiol 590(Pt 21): 5273-5297, 2012. For this study,

I performed Ieq and HCO3- secretion measurements in the Ussing chamber, analyzed

the data, and prepared the draft of the manuscript. Dr. Jie Liao performed Western

Blot to verify the CFTR-knockdown Calu-3 cell line, which is shown in Fig. 2.2A. Dr.

Renaud Robert performed Isc measurements shown in Fig. 2.4.

Chapter 3 also contains part of a published manuscript: Shan J. et al.

Bicarbonate-dependent chloride transport drives fluid secretion by the human

airway epithelial cell line Calu-3. J Physiol 590(Pt 21): 5273-5297, 2012. For this study,

I performed all the fluid secretion assay, analyzed the data, and prepared the draft of

the manuscript.

Chapter 4 contains a manuscript in preparation. For this study, I performed Ieq and

HCO3- secretion measurements, RT-PCR, Western Blot and co-immunoprecipitation. I

collected the data and analyzed.

Chapter 5 contains part of a published manuscript: Shan J. et al. Anion secretion by a

model epithelium: more lessons from Calu-3. Acta Physiol (Oxf) 202(3): 523-531,

2011. I mainly provided data, figure and discussion for this manuscript.

XIII

Throughout different stages of the work of this thesis, my supervisor Dr. John

Hanrahan provided guidance, instruction and discussion, including experimental

design, data analysis and the preparation of the final manuscripts.

XIV

Abbreviations

AE anion exchanger

ASL airway surface liquid

CA carbonic anhydrase

CF cystic fibrosis

CFTR cystic fibrosis transmembrane conductance regulator

Ieq equivalent short-circuit current

Isc short-circuit current

NBC sodium bicarbonate cotransporter

NKCC sodium potassium chloride cotransporter

pHi intracellular pH

1

Chapter 1 Introduction

2

1.1 The structure and function of airway epithelium

The human airways have 23 generations of dichotomous branches, which can be

divided into two parts according to their main functions, the conducting airways and

the respiratory airways. From generation 0 (trachea) to between 8 and 11, the

airways termed bronchi contain both cartilage and mucous glands. From around

generation 10 to 16, the airways lack cartilage, glands and alveoli, and are termed

conducting bronchioles. These two sections comprise the conducting airways, which

are responsible for transport of the inhaled air to the respiratory airway surfaces or

the exhaled air out of the lung. From generation 17 on, the airways become

respiratory bronchioles and have increasing numbers of alveoli. By generation 20, the

airway is full of alveoli and it ends in an alveolar sac at generation 23. These

respiratory airways are the site of gas exchange (Fischer and Widdicombe 2006,

Hollenhorst, Richter et al. 2011).

All parts of the airways are lined with an epithelium, which contains various

epithelial cell types having different morphologies and functions. Additionally, these

cell types are expressed in different proportions along the airway (Knight and Holgate

2003, Fischer and Widdicombe 2006, Hollenhorst, Richter et al. 2011, Tam,

Wadsworth et al. 2011). In the trachea, the surface epithelium is pseudostratified

and about 50 μm in height. The three main cell types are ciliated, basal and mucous

cells. With increasing airway generations, the height of the epithelium shortens

progressively. In the respiratory bronchioles, the height reaches ~10 μm and the

3

morphology of the epithelium becomes columnar.

1.1.1 Ciliated cells

Ciliated cells are the predominant cell type in the airways, accounting for ~50% of the

total cell number in the surface epithelium of all airway generations. Typically, each

cell possess up to 300 cilia on the apical surface, with numerous mitochondria

beneath them, indicating that the principal metabolic function of ciliated cells is to

provide ATP for ciliary beating and mucociliary clearance of inhaled particles out of

the lung (Knight and Holgate 2003, Fischer and Widdicombe 2006, Tam, Wadsworth

et al. 2011).

1.1.2 Basal cells

Basal cells are ubiquitously expressed along the conducting airways, though their

number decreases distally and are completely lost before the respiratory bronchioles

(Evans and Plopper 1988, Boers, Ambergen et al. 1998). These cells attach firmly to

the basement membrane and thus play a role in anchoring other superficial epithelial

cells (Evans and Plopper 1988, Evans, Cox et al. 1989, Evans, Cox et al. 1990). In

addition, basal cells act as stem cells in the lung and give rise to other epithelial cell

types, for example the ciliated cells (Hong, Reynolds et al. 2004, Hong, Reynolds et al.

2004).

1.1.3 Mucous cells

4

Mucous cells contain acidic-mucin granules and secret mucus into the airway to trap

foreign particles, which can be cleared away together with mucus by ciliated cells

(Knight and Holgate 2003, Tam, Wadsworth et al. 2011). In normal human trachea,

the estimated number of mucous cells is up to 6,800 cells/mm2 of surface epithelium

(Lumsden, McLean et al. 1984). Similar to basal cells, the number of mucous cells

declines progressively distally along the respiratory tree, and they are replaced by

Clara cells in the respiratory bronchioles (Boers, Ambergen et al. 1998). This switch

from mucous cells to Clara cells may reflect the fact that mucus is mainly a defense

against inhaled particles, which are deposited in the larger airways. Goblet cells

develop from basal cells, however in mice and perhaps humans, Clara cells can also

become mucus secreting during the mucus cell metaplasia induced by antigen

challenge (Evans, Williams et al. 2004).

1.1.4 Clara cells

Clara cells secret bronchiolar surfactants and antiproteases (Clara cell secretory

proteins, CCSPs) to protect the airway against toxic chemicals (Knight and Holgate

2003, Fischer and Widdicombe 2006, Tam, Wadsworth et al. 2011). Besides their

secretory function, they can also serve as stem cells when the number of basal cells

decreases (Hong, Reynolds et al. 2001).

1.1.5 Serous cells

Serous cells are present in the bronchioles. The morphology of serous cells resembles

5

that of goblet-shaped mucous cells however, their cellular contents differ. They are

specialized for fluid secretion into the airway to form the airway surface liquid and

also release antimicrobial factors (Rogers, Dewar et al. 1993, Fischer and

Widdicombe 2006).

1.1.6 Type I cells

Type I cells exist in the respiratory airway. Compared to other airway epithelial cells,

type I cells are flat and have very thin extensions, which cover over 98% of the

internal surface area of the lung. Thus, these cells form a thin barrier between the air

and the blood (Dobbs and Johnson 2007).

1.1.7 Type II cells

Unlike type I cells, type II cells are small cuboidal cells. These cells contain secretory

granules called lamellar bodies, and secret surfactant from these intracellular

organelles. Thus, the main functions of type II cells are the synthesis, secretion and

reuptake of the surfactant (Dobbs and Johnson 2007).

1.2 Junctional complexes in airways

Since there are various cells in the intact airway epithelium, cell-cell communication

is necessary between cells of the same cell type and between different cell types.

Therefore, paracrine purinergic signaling and junctional complexes at the contact

sites play important roles in cell-cell communication (Tam, Wadsworth et al. 2011).

6

Anchoring junctions, including desmosomes, hemidesmosomes and adherens

junctions, are vital in maintaining the structural integrity of the airway epithelium, as

they mediate direct or indirect cell-cell adhesion (Tam, Wadsworth et al. 2011).Tight

junctions, which include occludin, claudins and junctional adhesion molecule (JAM)

form zonula occludens, and are important in regulating the flow of solutes across the

epithelium by being selective for solutes of different sizes and charges (Harhaj and

Antonetti 2004).

Gap junctions consist of hexamers of connexin protein and form a pore that

transports molecules smaller than about 1 kDa between cells. They are especially

important in the bronchial and alveolar epithelium, since type I and type II cells

utilize these ‘channels’ to transit antioxidants, cytoplasmic metabolites and second

messenger signals between neighboring cells (Koval 2002, Boitano, Safdar et al.

2004).

1.3 The primary functions of airway epithelium: mucus

production and mucociliary clearance

The respiratory epithelial cells of the alveoli allow gas exchange while forming a

barrier that separates the external environment from the inner tissues and

vasculature of the lung. The conducting airways also form a barrier lined by epithelial

cells, but they are more complex and provide protection for the lung against harmful

foreign particles. Mucus-secreting goblet cells in the airway surface epithelium and

7

mucus tubules of the submucosal glands produce and secrete mucus that traps

pathogens and other particles so they can be cleared. Two important mucins in

human airways are MUC5AC and MUC5B. MUC5AC is produced mainly in the surface

epithelium and provide an acute response to direct contact with environmental

challenges, whereas MUC5B is most prominent in the submucosal glands and

involved in responses to chronic infection and inflammation (Hovenberg, Davies et al.

1996, Wickstrom, Davies et al. 1998, Thornton, Rousseau et al. 2008).

However, secreting mucus is not sufficient to protect against infection, mucus must

also be cleared by beating cilia which line the airways. To enable mucociliary

clearance, the epithelial cells secrete fluid in order to maintain a thin layer of liquid

6-7 μm in height on the epithelial surface (Tarran, Grubb et al. 2001). Most of the

secreted gel-forming mucins float on the surface of this liquid. Thus, ASL (airway

surface liquid) contains two layers, an outer mucous layer and a watery periciliary

layer between the epithelium and the mucus. The periciliary layer has lower viscosity

and enables the cilia to beat in a highly coordinated fashion that propels the mucus.

The tips of cilia penetrate into the mucous layer during the power stroke, propelling

it up the airways and out of the lung together with entrapped particles (Fischer and

Widdicombe 2006, Tam, Wadsworth et al. 2011).

In normal airways there is a fine balance between mucus production and clearance

(Evans and Koo 2009). This equilibrium is disrupted in cystic fibrosis, leading to

8

mucus plugging of the airways, infection, inflammation, epithelial damage, and

fibrosis (Boucher 2007, Boucher 2007, Chambers, Rollins et al. 2007, Mall 2008).

1.4 Regulation of ASL

The regulation of ASL hydration is a critical factor affecting mucociliary clearance, and

understanding its production is a focus of this study. Normal ASL is usually of 6-7 μm

in height, which approximates the length of the outstretched cilia and enables

efficient ciliary beating (Tarran, Grubb et al. 2001). The volume of the ASL and mucus

hydration are maintained by ion transport processes of airway epithelial cells, which

control the mass of salt (i.e. NaCl) on the airway surface, with water following

passively by osmosis (Matsui, Grubb et al. 1998). Ion transport by the surface

epithelium requires the coordinated activity of many transporters and channels,

including basolateral Na+,K+-ATPase, basolateral Na+,K+,Cl--cotransporter and apical

CFTR, apical ENaC (epithelial Na+ channel). In CF, the loss of CFTR activity reduces Cl-

secretion and may also lead to increased Na+ absorption by ENaC which prevent CF

epithelia from maintaining normal ASL volume (Boucher 2007, Boucher 2007,

Chambers, Rollins et al. 2007, Mall 2008). Consequently the height of the ASL

collapses to ~3 μm, which is not sufficient for ciliary mucociliary clearance (Tarran,

Button et al. 2005, Boucher 2007).

The pH of ASL is also important for airways defense although it is less well

understood. Firstly, pH can influence ENaC, and thus modulate sodium absorption

9

and the volume of the ASL (Awayda, Boudreaux et al. 2000). Secondly, several

proteins that actively participate in airway defense are pH-sensitive. Mucus

rheological properties are pH-dependent and the activity of some antibacterial

proteins and cationic antimicrobial peptides depend on pH (Ganz 2002). Regulation

of ASL pH is mediated by transporters, channels and enzymes, including the vacuolar

H+-ATPase, the H+ channels, and CFTR in the apical membrane, intracellular carbonic

anhydrase, and sodium bicarbonate cotransporter and anion exchangers in the

basolateral membrane etc. (Fischer and Widdicombe 2006).

Before considering current models for ion transport processes that regulate the

volume and pH of ASL, we first need to understand the structure and function of

these proteins.

1.4.1 Na+/K+-ATPase

Na+/K+-ATPase, also known as the sodium pump, is expressed ubiquitously in the

basolateral membrane of epithelia (with the exception of the choroid plexus and the

retinal pigment epithelium) (Ernst, Palacios et al. 1986, Gundersen, Orlowski et al.

1991, Crump, Askew et al. 1995). The enzyme belongs to the P-type ATPase family

because it becomes phosphorylated during the pump cycle. Na+/K+-ATPase is the

most important primary active transporter in the airway epithelium (Morth,

Pedersen et al. 2011). For each ATP hydrolysed, Na+/K+-ATPase transports 3 Na+ out

of the cell, and 2 K+ into the cell. Thus the pump generates both electrical and

10

chemical gradients that in turn drive ion movements through channels and

secondary active transporters.

Besides acting as a primary active transporter, Na+/K+-ATPase can also function as a

signaling molecule which regulates various enzymes and the formation and

maintenance of tight junctions between epithelial cells (Rajasekaran and Rajasekaran

2009).

1.4.2 V-ATPase

The vacuolar H+-ATPase (V-ATPase) is a proton pump responsible for acidifying

intracellular compartments such as endosomes and lysosomes. It also mediates

proton transport across the plasma membrane in some epithelial cells. For example,

V-ATPase in the apical membrane of intercalated cells of the renal distal tubule and

collecting duct secretes protons into the urine whereas V-ATPase in clear cells of the

epididymis acidify the epididymal fluid (Wagner, Finberg et al. 2004, Zuo, Huang et al.

2010). There is evidence for V-ATPase in airway epithelial cells, where it may be

required for sustained NADPH oxidase activity during infection (Fischer, 2011).

Bafilomycin A1 and concanamycin A are two potent macrolide inhibitors of V-ATPase.

They mainly bind to the c subunit of the V0 domain, thus preventing proton

translocation (Huss, Ingenhorst et al. 2002, Bowman, Graham et al. 2004).

11

1.4.3 NKCC

Sodium-potassium-chloride cotransporter (NKCC) is a member of the cation-chloride

cotransporter (CCC) superfamily (Haas and Forbush 2000, Russell 2000). It is

activated by cell shrinkage and participates in the regulation of cell volume. Besides

regulating the cell volume, NKCC maintains intracellular Cl- concentration ([Cl-]i)

above electrochemical equilibrium. The high [Cl-]i is used to drive net salt transport

in many epithelia. The cotransporter is expressed in the basolateral membrane of

polarized epithelial cells. Isoform 1 (NKCC1) is found in most epithelial organs except

the kidney, which express NKCC2 at the apical membrane (Gamba, Miyanoshita et al.

1994, Kaplan, Plotkin et al. 1996).

1.4.4 NBC

The sodium bicarbonate cotransporter (NBC) belongs to a superfamily of solute

carriers called the SLC4A family. NBC is mainly expressed in the basolateral

membrane of epithelial cells, and the dominant isoform in airway epithelial cells is

NBC1. The cotransporter helps regulate intracellular pH by transporting HCO3- into

the cell. The isoform in airway epithelial cells NBCe1 carries 1 Na+ and 3 HCO3- in the

same direction and therefore is electrogenic (Bernardo, Bernardo et al. 2006).

NBCe1 is inhibited by disulfonic stilbenes such as

4,4’-diisothiocyanostilbene-2,2’-disulfonic acid (DIDS) (Alpern 1985, Grassl and

Aronson 1986). The inhibition by disulfonic stilbenes occurs at the HCO3- interaction

12

site, and the amino acid motif that binds DIDS is KMIK and KLKK (Gross and Kurtz

2002). In addition to disulfonic stilbenes, NBCe1 is also inhibited competitively by

harmaline, which binds to the Na+ site (Soleimani and Aronson 1989).

1.4.5 The Na+-independent Cl-/HCO3- exchanger (AE)

Sodium independent Cl-/HCO3- exchangers also belong to the SLC4A family, and are

usually referred to simply as AEs (anion exchangers). AE isoform 2 (AE2) is widely

expressed, and in polarized epithelial cells it is usually located in the basolateral

membrane. AE2 carries out electroneutral exchange of Cl- and HCO3- in airway

epithelial cells (Dudeja, Hafez et al. 1999, Al-Bazzaz, Hafez et al. 2001), and the

directions of the fluxes depends on the chemical gradients of the two anions (Alper,

Chernova et al. 2002, Bonar and Casey 2008).

AE2 harbors covalent binding sites for disulfonic stilbenes, consequently these agents

inhibit AE2 and NBC1 unspecifically (Alper, Chernova et al. 2002).

1.4.6 Pendrin (SLC26A4)

Although pendrin, or SLC26A4, also functions as a Na+-independent Cl-/HCO3-

exchanger, it belongs to a different gene family called SLC26A, and is thus distinct

from SLC4A anion exchangers. Genetic defects in pendrin cause Pendred syndrome,

after which the exchanger was named (Wangemann 2011). Pendrin is expressed in

many different tissues, including the airways, where it localizes to the apical

13

membrane and exchanges HCO3- for other monovalent anions such as Cl-, I- or

formate (Scott, Wang et al. 1999, Scott and Karniski 2000, Pedemonte, Caci et al.

2007, Nakao, Kanaji et al. 2008).

Pendrin is predicted to have 12-15 transmembrane domains, and intracellular N- and

C- terminal domains (Mount and Romero 2004, Dossena, Rodighiero et al. 2009).

Motifs which are critical for anion transport are found within the hydrophobic core of

the transmembrane domains (Mount and Romero 2004). There is a STAS (sulfate

transporter and anti-sigma) domain at the C-terminus, which is believed to be

important for functional interactions with other proteins, including CFTR (Aravind

and Koonin 2000, Ko, Shcheynikov et al. 2002, Ko, Zeng et al. 2004).

Pendrin is inhibited by the chloride channel blocker NPPB

(5-Nitro-2-(3-phenylpropylamino) benzoic acid) and by the non-steroidal

anti-inflammatory drug niflumic acid (Dossena, Vezzoli et al. 2006, Pedemonte, Caci

et al. 2007).

1.4.7 The Na+/H+ exchanger (NHE)

NHE regulates intracellular pH and cell volume by exchanging 1 Na+ for 1 H+ and

normally transports Na+ in and H+ out. To date, six isoforms have been identified

(Fliegel 2005, Slepkov, Rainey et al. 2007). NHE1 is ubiquitously distributed and is the

only NHE isoform expressed basolaterally in airway epithelial cells (Dudeja, Hafez et

14

al. 1999, Al-Bazzaz, Hafez et al. 2001).

NHE1 is moderately sensitive to amiloride, a diuretic compound which acts mainly in

the distal nephron. To increase the potency and selectivity of this inhibitor for NHEs,

many derivatives have been synthesized such as DMA, EIPA and HMA. Some of these

novel inhibitors, notably benzoylguanidine and its derivatives, are more potent and

selective for NHE1 (Masereel, Pochet et al. 2003).

1.4.8 The H+ channel

An H+ channel in the apical membrane functions mainly in acid extrusion and is

activated by external alkaline pH and membrane depolarization (Cherny, Markin et al.

1995). So far, only one type of plasma membrane proton channel has been identified,

namely HVCN1. It has been found in many different cell types including airway

epithelial cells (DeCoursey 1991, DeCoursey and Cherny 1995, Fischer, Widdicombe

et al. 2002, Schwarzer, Machen et al. 2004, Iovannisci, Illek et al. 2010).

The most potent inhibitor of HVCN1 is Zn2+. It competitively binds to the external

surface of the proton channel, slowing the opening rate and shifting the voltage

dependence to more positive membrane potentials (Cherny and DeCoursey 1999).

1.4.9 The chloride channels

Different types of chloride channels have been found in airway epithelial cells

15

(Toczylowska-Maminska and Dolowy 2012). Among these channels, cystic fibrosis

transmembrane conductance regulator (CFTR) has been most intensively studied

because it is mutated in the common genetic disease cystic fibrosis (CF).

CFTR is apically localized in epithelia. It belongs to the ATPase-binding cassette (ABC)

transporter family, and is the only member of the family that functions as an ion

channel. It conducts both chloride and bicarbonate and has a Cl-:HCO3-

permeability ratio of 4:1 (Kim and Steward 2009). Two sets of six transmembrane

helices form transmembrane domains that contain the pore for anion flow. Each of

the membrane domains is followed by a nucleotide-binding domain (NBD), which

interacts with ATP. A central regulatory (R) domain links the N- and C-terminal

half-molecules. The R domain contains many sites for phosphorylation by protein

kinase A (PKA) and protein kinase C (PKC), and together with the two NBD s controls

the opening and closing rates of CFTR. Channel opening requires the binding of 2 ATP

molecules to the NBDs, which form a “nucleotide sandwich”, and phosphorylation of

R domain. Hydrolysis of ATP at NBD2 closes the channel gate, and dephosphorylation

of the R domain causes deactivation (Riordan 2008, Hwang and Sheppard 2009,

Lubamba, Dhooghe et al. 2012, Toczylowska-Maminska and Dolowy 2012). Besides

functioning as a chloride channel, CFTR may also exert regulatory effects on other

proteins, for example the epithelial sodium channel (ENaC) (Riordan 2008, Lubamba,

Dhooghe et al. 2012). The most potent and selective inhibitor of CFTR is the glycine

hydrazide GlyH-101

16

[N-(2-naphthalenyl)-((3,5-dibromo-2,4-dihydroxyphenyl)methylene)glycine hydrazide]

(Muanprasat, Sonawane et al. 2004, Sheppard 2004, Norimatsu, Ivetac et al. 2012). It

is voltage-dependent and therefore probably blocks the open CFTR pore. By contrast

the thiazolidinone CFTR(inh)-172

[3-[(3-trifluoromethyl)phenyl]-5-[(4-carboxyphenyl)methylene]-2-thioxo-4-thiazolidin

one] appears to inhibit within the bilayer by increasing the mean closed time and is

not voltage dependent (Kopeikin, Sohma et al. 2010).

The calcium activated chloride channel (CaCC) is another important chloride channel

which coexists with CFTR in the apical membrane of airway epithelial cells (Gabriel,

Makhlina et al. 2000). Its molecular identity is still under debate, but is probably

TMEM16A (Caputo, Caci et al. 2008, Schroeder, Cheng et al. 2008, Yang, Cho et al.

2008). Other chloride channels present in airway epithelial cells include volume

sensitive outwardly rectifying chloride channel (VSOR), bestrophins, ClC-2 channel

and outwardly rectifying chloride channel (ORCC) (Toczylowska-Maminska and

Dolowy 2012).

1.4.10 ENaC

The epithelial sodium channel ENaC is a member of the ENaC/Degenerin gene family,

which is comprised of channels with diverse functions including sodium transport

and mechanosensitivity (Kellenberger and Schild 2002). ENaC is located in the apical

membrane of epithelial cells and mediates Na+ transport across epithelia.

17

The functional unit of ENaC is probably a heterotrimer formed by α, β and γ subunits

(Kashlan and Kleyman 2011). It is interesting that in adult humans, rats and mice, all

the three ENaC subunits are highly expressed in small and medium-sized airways.

However, in the distal lung, only the α and γ subunits are expressed while the β

subunit is not found (Burch, Talbot et al. 1995, Farman, Talbot et al. 1997, Talbot,

Bosworth et al. 1999). The β- or γ-ENaC knockout mice died slightly later than the

α-knockout from severe hyperkalemia due to deficient renal K+ secretion. This

suggests that, in contrast to the kidney, Na+ transport in the lung can be maintained

efficiently by only two ENaC subunits, namely the α and β or the α and γ (Barker,

Nguyen et al. 1998, McDonald, Yang et al. 1999).

The channel blocker amiloride binds to residues in the TM2 domains of the subunits

(Kellenberger and Schild 2002, Kashlan and Kleyman 2011). These residues lie in the

outer vestibule of the channel pore, which has a diameter of ~ 5 Å, which

accommodates amiloride well (4-5 Å)

1.4.11 Carbonic anhydrase

Carbonic anhydrase is a zinc metalloenzyme which catalyzes the reversible hydration

of CO2. It is expressed ubiquitously in all living organisms, and 15 isoforms of

carbonic anhydrase have been identified in humans (Esbaugh and Tufts 2006,

Purkerson and Schwartz 2007). These isozymes can be divided into three groups, the

18

cytoplasmic CAs, the membrane-bound CAs, and CA-related proteins which have lost

their catalytic activity.

CAII is an important CA isoform, which has a high turnover rate of 106s-1 and

accounts for the majority of CA activity in different tissues (Khalifah 1971, Esbaugh

and Tufts 2006, Purkerson and Schwartz 2007). CAII is reported to bind to the

intracellular C-terminal region of AE1, AE2, NBC and NHE1 via its acidic motif in the

N-terminal region, forming a bicarbonate transport metabolon as discussed above

(McMurtrie, Cleary et al. 2004). The membrane-bound form of CA also has a high

turnover rate and may be of importance (Baird, Waheed et al. 1997, Wingo, Tu et al.

2001).

1.4.12 Regulation of ASL hydration

The hydration state of ASL is control by the mass of salt (NaCl) on the surface of

airways, with water flowing passively due to osmosis. Cl- secretion and Na+

absorption are important in regulating passive transepithelial water flow and thus

the height of the ASL. In this way, these ion transport processes are mainly

responsible for providing an optimal environment for ciliary beating (Hollenhorst,

Richter et al. 2011).

A large part of the Cl- secretion in the airways is mediated by apical CFTR (Riordan

2008, Lubamba, Dhooghe et al. 2012). In addition to CFTR, CaCC are also present on

19

the apical side and contribute to inflammatory secretion (Toczylowska-Maminska and

Dolowy 2012). NKCC and AE2 have been detected in the basolateral membrane of

airway epithelial cells, and supply Cl- from the basolateral side (Tessier, Traynor et al.

1990, Al-Bazzaz, Hafez et al. 2001). A wide variety of K+ channels are expressed in

airway epithelial cells, and several have been identified in the basolateral membrane

(Bardou, Trinh et al. 2009). Though these K+ channels are not directly involved in Cl-

secretion, they regulate the membrane potential and help to maintain the

electrochemical gradient favoring apical Cl- secretion (Bardou, Trinh et al. 2009,

Hollenhorst, Richter et al. 2011).

In terms of Na+ transport, the amiloride-sensitive ENaC contributes to apical Na+

absorption in the airways (Burch, Talbot et al. 1995), whereas on the basolateral side,

the Na+/K+ ATPase moves Na+ out of the cells and provides the driving force for apical

Na+ absorption (Hollenhorst, Richter et al. 2011). Thus, the Na+/K+ ATPase and ENaC

work in concert to mediate Na+ absorption in the airways. Besides acting as a

chloride channel, CFTR also regulates ENaC. In normal airways, CFTR inhibits ENaC

thus preventing Na+ absorption; however, in CF airways, ENaC becomes hyperactive

because of the defective CFTR (Matsui, Grubb et al. 1998, Boucher 2007, Boucher

2007). Therefore, ASL dehydrates under Na+ hyperabsorption.

1.4.13 pH regulation of ASL

Normal ASL pH is slightly acidic, averaging around 6.6-6.9 (Fischer and Widdicombe

20

2006). The ASL pH of CF patients becomes even more acidic, suggesting a breakdown

in the ASL pH regulation (Coakley, Grubb et al. 2003, Song, Salinas et al. 2006).

HCO3- is considered a weak base and may be the main determinant of the pH and

buffer capacity of the ASL (Song, Salinas et al. 2006). Under unstimulated conditions,

HCO3- has an outward net driving force across the apical membrane of airway

epithelial cells due to the negative membrane potential, and contributes to the anion

efflux through CFTR despite having lower permeability and conductance ratios

relative to Cl- (Fischer and Widdicombe 2006). The idea that CFTR alkalinizes

submucosal gland secretions is suggested by the reduced pH of CF gland secretions

(Coakley, Grubb et al. 2003, Song, Salinas et al. 2006).

NBCe1 and carbonic anhydrase are presumed to supply the HCO3- for secretion into

the lumen via CFTR (Smith and Welsh 1992, Devor, Singh et al. 1999, Devor, Bridges

et al. 2000, Krouse, Talbott et al. 2004, Ballard, Trout et al. 2006). NBCe1 supplies

HCO3- from the basolateral side whereas carbonic anhydrase catalyzes the

generation of endogenous HCO3-.

As discussed above, AE2 is also found basolaterally in airway epithelial cells (Dudeja,

Hafez et al. 1999, Al-Bazzaz, Hafez et al. 2001) but may move HCO3- out of the cell in

exchange of Cl-. Thus it would reduce the electrochemical driving force for apical

HCO3- exit via CFTR and tend to acidify the ASL (Fischer and Widdicombe 2006). It

21

should be noted however that this scheme is controversial (Garnett et al.)

Although HCO3- secretion via CFTR would cause alkalinization of the ASL, the fact

that normal ASL pH is consistently acidic (Fischer and Widdicombe 2006) indicates

that airway surface epithelial cells also secrete acid.

Other apical transporters also impact ASL pH, and some of these have been

discussed above. For example, The voltage-gated H+ channel was demonstrated in

the apical membrane of airway epithelial cells (DeCoursey 1991, DeCoursey and

Cherny 1995, Fischer, Widdicombe et al. 2002, Schwarzer, Machen et al. 2004,

Iovannisci, Illek et al. 2010). In some studies, blocking this channel by Zn2+ showed

50-70% inhibition on H+ secretion by airway epithelial cells (Fischer, Widdicombe et

al. 2002, Schwarzer, Machen et al. 2004), suggesting that in these studies the primary

acid secretion mechanism was mediated by the H+ channel. An H+,K+-ATPase is

present in the apical membrane of airway epithelial cells (Coakley, Grubb et al. 2003)

and blocking it for several hours by apical exposure to ouabain inhibits ASL

acidification, although the acute effects are small (Fischer, Widdicombe et al. 2002,

Coakley, Grubb et al. 2003, Krouse, Talbott et al. 2004, Schwarzer, Machen et al. 2004,

Song, Salinas et al. 2006). Another acid secreting transporter is the V-ATPase (Inglis,

Wilson et al. 2003). Its inhibitor bafilomycin A1 reduces ASL acidification to a variable

extent (0-60%) (Poulsen and Machen 1996, Fischer, Widdicombe et al. 2002, Inglis,

Wilson et al. 2003, Schwarzer, Machen et al. 2004, Song, Salinas et al. 2006). Though

22

NHE1 is expressed in the basolateral membrane and mainly participates in regulating

intracellular pH, it has been proposed that NHE1 could affect the apical H+ channel by

intracellular pH regulation, because the voltage-gated H+ channel can be activated by

intracellular acidification (Fischer and Widdicombe 2006). It remains unclear why

different acid transporters and channels have been identified as the primary H+

secretion pathway in airways by different investigators.

1.5 Models of anion transport in submucosal glands

Submucosal glands are important to airway defense, producing mucus and fluid in

response to neural signals (Maggi, Giachetti et al. 1995). The density of glands is

about 1 per mm2 in the trachea and becomes more abundant as the generation

grows. Each gland consists of multiple tubules that merge into a collecting duct, and

continue through a ciliated duct which eventually leads to the airway surface (Tos

1966, Meyrick, Sturgess et al. 1969). The abundant serous cells secrete water and

electrolytes, and a rich mixture of antimicrobial, anti-inflammatory, and antioxidant

compounds. CFTR is expressed in these cells according to most studies (Basbaum,

Jany et al. 1990, Engelhardt, Yankaskas et al. 1992).

To understand how serous cells regulate ASL, several groups have used submucosal

glands from human or animals and the model serous cell line Calu-3 to study anion

transport. Using Calu-3 cells, a scheme for anion transport of serous cells was

proposed which is widely accepted (Devor, Singh et al. 1999). In this cellular model,

23

cells secrete mostly HCO3- in response to elevated [cAMPi] by forskolin. HCO3

- is

taken up by cells via basolateral NBC, and then exits to the lumen through CFTR.

However, according to the model, secretion by serous cells can shift from HCO3- to Cl-

under certain conditions, for example during cell hyperpolarization, which favors the

electroneutral basolateral NKCC and thus causes the cells to secret Cl- (Fig. 1.1).

Another group made some modifications based on this model (Cuthbert and

MacVinish 2003). Though they confirmed the shift from HCO3- to Cl- secretion under

hyperpolarization, they emphasized that Cl- secretion was HCO3- dependent and the

role of carbonic anhydrase in generating HCO3-. They also suggested that carbonic

anhydrase (isoform 2) might form complexes with baolateral NBC, NHE and anion

exchanger to facilitate the transport of both HCO3- and Cl- (Fig. 1.2).

24

Figure 1.1 Model for anion transport by Calu-3 cells proposed by Devor

et al.. HCO3

- or Cl- is secreted through cAMP-stimulated apical CFTR channels according to the basolateral membrane potential (ψbl) and reversal potential of the sodium-bicarbonate cotransporter (ErevNaHCO3). Basolateral Cl- entry mediated by the sodium–potassium chloride cotransporter NKCC1 is favoured when the membrane potential is larger (i.e. more negative) than ErevNaHCO3. Basolateral HCO3

- entry via sodium-bicarbonate cotransporter NBC1 is favoured when the basolateral membrane potential is smaller (i.e. more positive) than ErevNaHCO3. 1-EBIO hyperpolarizes the basolateral membrane by activating charybdotoxin (CTX)-sensitive K+ channels. Bumetanide, ouabain and DNDS reduce secretion by inhibiting NKCC1, Na+/K+ ATPase and NBC1 respectively. Reproduced from Devor et al. 1999.

25

Figure 1.2 Model for anion transport by Calu-3 cells proposed by

Cuthbert et al.. This model is similar to the one proposed by Devor et al., except that besides NKCC1, carbonic anhydrase II (CAII) is also important to Cl- tranport. When the basolateral membrane is hyperpolarized by activating K+ channel, CAII converts CO2 to HCO3

-, which exchanges for Cl- via AE2. AE2 together with NKCC1 provides Cl- for secretion through apical CFTR. The generation of H+ by CAII conversion is extruded by NHE1 to maintain pHi. HCO3

- may activate soluble adenylyl cyclase (sAC), thus increaseing [cAMPi] and activating CFTR. Reproduced from Cuthbert et al. 2003.

26

When the above scheme was tested on submucosal glands of human and animal

models, some contradictions were found. When native glands were stimulated by

elevating [cAMPi], 1) pH of secretions from the gland was not as high as expected; 2)

the major anion in the secreted fluid was chloride rather than bicarbonate

(Jayaraman, Joo et al. 2001, Joo, Saenz et al. 2002). Bumetanide, an inhibitor of NKCC,

reduced gland secretion though not as well as did removal of HCO3-. These finding

led to the conclusion that both HCO3- and Cl- were important to gland secretion

when [cAMPi] was increased and CFTR was activated (Wine and Joo 2004) (Fig. 1.3).

All of the above studies were consistent with CFTR being the main apical exit

pathway for anion secretion, however a recent study suggested a different view

(Garnett, Hickman et al. 2011). Based on pHi measurements and fluid secretion

assays, Garnett et al. (2011) proposed that when [cAMPi] is increased, CFTR secretes

chloride while bicarbonate exits via apical pendrin in exchange for luminal chloride

(Fig. 1.4).

It has been almost 25 years since the cftr gene was cloned (Kerem, Rommens et al.

1989, Riordan, Rommens et al. 1989, Rommens, Iannuzzi et al. 1989). Nevertheless,

how anions are transported by serous cells and the role of CFTR in such process

remain controversial. The development of effective therapies for cystic fibrosis may

require an understanding of the mechanisms of anion and fluid secretion.

27

Figure 1.3 Model for anion and fluid secretion by submucosal glands

proposed by Joo et al.. Pathways that elevate intracellular Ca2+ level, such as acetylcholine (ACh), are hypothesized to activate fluid and macromolecular secretion from both serous and mucous cells. Pathways that elevate intracellular cyclic AMP level, such as VIP, are hypothesized to stimulate serous cells and mucin but not fluid secretion from mucous cells. The serous cells secrete mostly Na+, Cl- and water, and a small amount of HCO3

- comparied to Cl-. The CFTR-dependent fluid-secreting pathway is defective in CF glands. Reproduce from Joo et al., 2002.

28

Figure 1.4 Model for anion transport by Calu-3 cells proposed by

Garnett et al.. A, under non-stimulated conditions, Cl− is accumulated across the basolateral membrane by the combined action of the basolateral Na+-K+-2Cl− cotransporter, NKCC1, together with parallel activity of the NBC and AE2. Under these conditions, a basal level of CFTR activity drives a small amount of a Cl−-rich secretion with a pH of ~7.4 (25 mm HCO3

−) via electrogenic Cl− efflux through CFTR. B, elevation of cAMP inhibits basolateral AE2 and stimulates NBC. At the apical membrane, cAMP/PKA further increases CFTR activity as well as activates pendrin (SLC26A4). Stimulation of CFTR activity leads to a marked rise in net transepithelial fluid secretion driven by electrogenic Cl− efflux through CFTR, whereas the co-activation of pendrin increases the HCO3

− content of this secreted fluid to ~75 mm (pH 7.9), through coupled Cl−/HCO3

− exchange, with Cl− cycling across the apical membrane via CFTR and SLC26A4. Reproduced from Garnett et al., 2011.

29

Chapter 2 Anion transport in

airway epithelial cells

30

Previous studies by others showed that the activator of adenylyl cyclase, forskolin,

increases intracellular cAMP, stimulates PKA-phosphorylation, and activates CFTR in

Calu-3 cells. Under short-circuit current conditions, CFTR was proposed to secrete

mainly bicarbonate. However, our preliminary data revealed that the fluid secreted

by forskolin stimulated Calu-3 cell monolayers contained only ~25 mM bicarbonate.

In this chapter we investigated this and other discrepancies between our data and

previous studies. The result have led us to proposed a revised model for anion

transport by Calu-3 cells which may be applicable to airway submucosal gland serous

cells.

31

2.1 Abstract

Anion transport drives fluid onto the airway surface to supply airway surface liquid

for humidifying inspired air and also for mucociliary clearance. This anion transport is

defective in airway diseases such as cystic fibrosis. Despite the importance of anion

transport in airway epithelium, its cellular mechanisms remain poorly understood

and controversial. We studied Cl- and HCO3- fluxes across the human airway cell line

Calu-3 and a genetically matched CFTR-deficient cell line. Forskolin stimulated the

short-circuit current (Isc) across voltage-clamped monolayers, and the equivalent

short-circuit current (Ieq) calculated from the transepithelial voltage and resistance of

monolayers kept under open-circuit conditions. Isc was equivalent to the HCO3- flux

measured using the pH-stat technique. Ieq equaled the sum of net 36Cl- and HCO3-

fluxes, although Ieq and HCO3- were underestimated since both were increased by

bafilomycin and ZnCl2, inhibitors of electrogenic H+ secretion. Ieq was dependent on

the presence of Na+ and HCO3-. Forskolin-stimulated Ieq and HCO3

- secretion were

both rapidly abolished by the carbonic anhydrase inhibitor acetazolamide, indicating

that Na+,HCO3- cotransport is secondary to carbonic acid synthesis. Anion exchange is

the main mechanism of basolateral Cl- loading in these cells, consistent with the

weak (~20%) inhibition by bumetanide. Apical anion efflux was probably mediated by

CFTR channels because 1) a forskolin-stimulated current appeared after imposing a

transepithelial HCO3- gradient across basolaterally permeabilized monolayers, 2) the

HCO3- current was sensitive to CFTR inhibitors and drastically reduced in

CFTR-deficient cells, and 3) the net secretory HCO3- efflux was increased 25% in

32

bilateral Cl--free solutions, excluding a dependence on apical anion exchange. The

results suggest a model in which HCO3- recycles across the basolateral membrane in

exchange for Cl-, and the resulting HCO3--dependent supply of Cl- is secreted via CFTR

together with remaining intracellular HCO3-.

2.2 Introduction

The airways are protected by a microscopic layer of airway surface liquid (ASL), which

humidifies the air during inspiration and participates in host defense against inhaled

pathogens. Anion transport in airway epithelium is responsible for driving fluid into

the airway surface to form the ASL, and transport defects in CF impair the depth of

normal ASL (Widdicombe 2002, Tarran, Button et al. 2006, Boucher 2007), leading to

reduced mucociliary clearance and recurring infection and inflammation. Despite its

importance in innate defense and pathophysiology, the mechanisms of airway anion

transport remain poorly understood.

The Cystic Fibrosis Transmembrane conductance Regulator (CFTR) is the product of

the CF gene and plays a critical role in airway secretion. It is a

phosphorylation-regulated, non-rectifying anion channel of ~7 pS conductance which

determines the rate of secretion by many epithelia (Klyce and Wong 1977) including

those in the airways. The channel pore is permeable to Cl- and HCO3- (Gray, Pollard et

al. 1990, Poulsen, Fischer et al. 1994, Linsdell, Tabcharani et al. 1997), and secretion

of both anions is defective in CF (Widdicombe, Welsh et al. 1985, Smith and Welsh

33

1992). However, the relative contributions to apical HCO3- efflux of CFTR, or of

CFTR-regulated anion exchangers, remains controversial (Lee, Choi et al. 1999,

Ishiguro, Steward et al. 2009, Kim and Steward 2009, Garnett, Hickman et al. 2011).

Anion secretion has been studied extensively in the human airway cell line Calu-3,

which differentiates spontaneously into a predominant population of serous-like cells

which express CFTR and antimicrobial proteins, and a smaller population of

mucous-like cells which contain mucin granules. When cultured on porous supports

at the air-liquid interface, Calu-3 monolayers generate robust basal short-circuit

current (Isc) that is not explained by net 36Cl- or 22Na+ fluxes (Haws, Finkbeiner et al.

1994, Shen, Finkbeiner et al. 1994, Shan, Huang et al. 2011) and was therefore

suggested to be generated by active HCO3- secretion (Lee, Penland et al. 1998).

Forskolin stimulated both unidirectional 36Cl- fluxes across short-circuited Calu-3 cells

without inducing measureable net Cl- secretion, and the Isc was insensitive to the

NKCC1 inhibitor bumetanide (Devor, Singh et al. 1999). Based on these findings it was

suggested that the forskolin-stimulated Isc was due to electrogenic HCO3- transport

(Devor, Singh et al. 1999), and this was subsequently confirmed using the pH stat

technique (Krouse, Talbott et al. 2004).

Several cellular models have been proposed to explain transport by Calu-3

monolayers. According to one scheme, HCO3- transport occurs by Na+ coupled entry

at the basolateral membrane followed by diffusional exit through apical CFTR

34

channels (Devor, Singh et al. 1999). This scheme predicts the secretion of HCO3--rich

fluid during cAMP stimulation, and perhaps Cl- rich fluid during stimulation by

secretagogues that hyperpolarize the cell (based on unidirectional 36Cl- fluxes

measured during stimulation with the potassium channel activator 1-EBIO). Another

hyperpolarizing agonist, 7,8-benzoquinoline, was also found to stimulate Cl- secretion,

but in that study Cl- secretion was dependent on carbonic anhydrase activity and was

proposed to involve anion exchange (AE2)-mediated Cl- entry at the basolateral

membrane in parallel with NKCC (Cuthbert and MacVinish 2003). The relationship

between fluid secretion and anion transport remains uncertain. Elevated [HCO3-] was

observed in Calu-3 secretions (Irokawa, Krouse et al. 2004), however

cAMP-stimulated fluid secretion driven by a flow of bicarbonate is difficult to

reconcile with the pH of airway secretions, which are near neutrality or even slightly

acidic (Kyle, Ward et al. 1990, Coakley, Grubb et al. 2003, Fischer and Widdicombe

2006, Song, Salinas et al. 2006).

We studied HCO3- transport under open- and short-circuit conditions using an

automated pH stat apparatus and tracer fluxes. Under pH stat conditions we found

that Isc was stimulated by forskolin and was identical to the net HCO3- flux measured

simultaneously, and that the HCO3- current was probably mediated by CFTR as

previously suggested (Ballard, Trout et al. 1999, Devor, Singh et al. 1999). Unlike Isc,

the Ieq under open-circuit conditions was the sum of net 36Cl- and HCO3- fluxes, and

was dependent on both Na+ and HCO3-. Forskolin-stimulated Ieq and HCO3

- secretion

35

were both abolished by the carbonic anhydrase inhibitor acetazolamide, indicating

an important role of carbonic anhydrase in anion transport. We found that anion

exchange but not NKCC is the main mechanism of basolateral Cl- loading, which

probably explains the weak (~20%) inhibition by bumetanide. Under pH-stat

conditions, we found evidence that apical anion efflux is mediated by CFTR channels.

The results suggest a revised model for Calu-3 cell anion transport in which Cl-

transport is dependent on HCO3- via basolateral anion exchange, and most of the

HCO3- is synthesized intracellularly by carbonic anhydrase or enters the cell

basolaterally via NBCe1.

2.3 Methods

2.3.1 Cell Culture

We used two modified Calu-3 cell lines for this study, a CFTR knock-down cell line

which stably expresses a 21-mer shRNA specifically targeting CFTR mRNA, and a

control cell line expressing shRNA with four mutations that reduce its ability to

silence cftr (Sizt and Alter cell lines, respectively) (Palmer, Lee et al. 2006). Both cell

lines were cultured in EMEM containing 7% FBS. Some control experiments were also

performed using parental Calu-3 cells (HTB-55, American Type Culture Collection,

Manassas, VA) cultured in Eagle’s minimum essential medium (EMEM) containing 15%

fetal bovine serum (FBS) to allow comparison with previous studies.

Cells were seeded at 5 x 105 cells per cm2 on SnapwellsTM (1.12 cm2; Costar, Corning

36

Life Sciences, Lowell, MA) for measuring short-circuit current (Isc) and HCO3-

secretion. Fresh medium was placed on the basolateral side one day after plating,

and the apical medium was removed to establish air-liquid interface (ALI) conditions.

Any fluid that appeared spontaneously on the apical surface was removed after 3

days and cultures were maintained in a humidified, 5% CO2 incubator at 37℃ for

21-25 days. Polarization of the monolayers with respect to CFTR was confirmed by

imposing a transepithelial Cl- gradient and measuring current stimulated by forskolin

after permeabilization of the basolateral or apical membrane with nystatin (100

μg·ml-1).

2.3.2 Solutions

To measure HCO3- secretion, monolayers were mounted in modified Ussing

chambers. The apical surface was bathed with unbuffered solution containing

(mmol·l-1): 120 NaCl, 5 KCl, 1.2 MgCl2, 1.2 CaCl2, and the basolateral side was bathed

with 120 NaCl, 25 NaHCO3, 3.3 KH2PO4, 0.8 K2HPO4, 1.2 MgCl2, 1.2 CaCl2, and 10

glucose. Nominally Na+-free solution was prepared by replacing NaHCO3 and choline

chloride, and NaCl with N-methyl-D-glucamine chloride. Nominally Cl--free solution

was prepared by replacing Cl- with gluconate and increasing the [Ca2+] from 1.2 to 4

mmol·l-1 to compensate for gluconate binding. Nominally HCO3--free solution was

prepared by replacing NaHCO3 with NaHEPES. All solutions were adjusted to pH 7.4.

The apical side was stirred with 100% O2 while the basolateral side was bubbled with

95% O2/5% CO2. Both sides were gassed with 100% O2 during experiments with

37

bilateral HCO3--free solution.

Tracer fluxes were measured using the same solutions as for pH-stat experiments. In

some experiments the basolateral membrane was permeabilized by adding nystatin

from a 1,000 x stock solution in DMSO (final nystatin concentration 100 μg·ml-1). The

CFTR inhibitor GlyH-101 was added from a 1,000 x stock solution to give a final

concentration of 100 μmol·l-1 and 0.1% DMSO.

2.3.3 Immunoblotting

After SDS-PAGE on 8% gels, proteins were transferred to nitrocellulose membranes as

described previously (Luo, McDonald et al. 2009) and probed using monoclonal

antibodies: M3A7, which recognizes an epitope between amino acids 1365 and 1395

in the second nucleotide binding domain of CFTR (1:5000, gift from J.R. Riordan and

T.J. Jensen, UNC Chapel Hill, NC) (Kartner and Riordan 1998), TUB-1A2, which binds

the C-terminus of α-tubulin (1:5000, Sigma), and α5, which binds the α-subunit of

avian Na+/K+-ATPase (1:200, mAb gift from R.W. Mercer, Washington Univ., St. Louis

MO) (Takeyasu, Tamkun et al. 1988). Blots were washed, incubated with secondary

antibody conjugated to horseradish peroxidase (1:1000), visualized with enhanced

chemiluminescence (Amersham Biosciences), and analysed using ImageJ (Rasband,

2011).

2.3.4 Measurements of equivalent Isc and HCO3- secretion

38

Inserts were mounted in modified Ussing chambers (Physiologic Inst., San Diego, CA)

at 37℃ and initial studies were carried out under voltage clamp to allow comparison

of net HCO3- transport with Isc. In subsequent experiments HCO3

- secretion was

measured under open-circuit conditions, and was compared with the equivalent

short-circuit current (Ieq) calculated from Ohm’s law using the spontaneous

transepithelial potential (Vt) and transepithelial resistance (Rt). Rt was determined

from small deflections in Vt produced by bipolar current pulses (1 μA, 1 sec duration,

99.9 sec interval) delivered by voltage clamp (VCC200, Physiologic Instr.). Data were

digitized (Powerlab 8/30, AD Instruments, Montreal QC) and analyzed using Chart5

software.

HCO3- transport was measured using the pH-stat method under both open- and

short-circuit conditions. A mini-pH electrode (pHG200-8, Radiometer Analytical)

connected to the automated titration workstation (TitraLab 854, Radiometer)

delivered 1 μl aliquots of 10 mmol·l-1 HCl or 5 mmol·l-1 H2SO4 to maintain the pH

constant at 7.400 ± 0.002. The amount of acid required for this was used to calculate

the rate of HCO3- secretion. The volume of each half chamber was 4 ml. Solutions

containing 25 mmol·l-1 HCO3- were stirred vigorously with 95% O2/5% CO2.

Nominally HCO3--free solutions were bubbled with 100% O2.

2.3.5 Tracer fluxes

Calu-3 monolayers were mounted in modified Ussing chambers and equilibrated for

39

20 min. H36Cl (generous gift of W.S. Marshall, St. Francis-Xavier University, Antigonish

NS) was added to one side of the monolayer and neutralized, yielding a final Cl-

concentration of 125.3 mmol·l-1. After 20 min of mixing, duplicate 400 μl samples

were taken from the cold side at 15 min intervals and replaced with 800 μl saline.

Residual radioactivity was calculated, with correction for dilution, at the beginning of

each subsequent flux period. Samples were counted in 5 ml scintillation solution

(Tri-Carb 2810 TR, PerkinElmer, Woodbridge ON). Two 10 μl samples were taken from

the hot side at the beginning and end of the experiment and averaged to calculate

mean specific activity. Both unidirectional fluxes were measured in parallel

experiments performed at the same time using monolayers with similar Rt.

2.3.6 Data analysis

Isc or Ieq was determined at 100 sec intervals, and HCO3- net flux rate was calculated

every 5 min. Basal values were those obtained immediately before adding forskolin.

Steady-state fluxes during stimulation were calculated 30 - 60 min after forskolin

addition. Paired or unpaired student’s t-tests with p<0.05 were used for single

comparisons. One-way analysis of variance followed by the Bonferonni post-hoc test

was used for multiple comparisons

2.4 Results

2.4.1 HCO3- secretion accounts for the short-circuit current (Isc)

Electrogenic HCO3- secretion across control monolayers was measured under

40

voltage-clamp and compared with the Isc measured simultaneously. Unstimulated

HCO3- secretion and Isc were both low but increased ∼8-fold after forskolin (10

μmol·l-1) was added to the basolateral side (Fig. 2.1). The rate of HCO3- secretion

measured using pH-stat with the transepithelial potential clamped at 0 mV was equal

to the Isc within measurement error (net HCO3- secretion: 2.14 ± 0.36 μeq·cm-2·h-1; Isc:

2.30 ± 0.27 μeq·cm-2·h-1; Fig. 2.1), evidence that basal and forskolin-stimulated Isc

were both due to net HCO3- secretion. This confirms previous suggestions based on

the large discrepancy between Isc and the net 36Cl- and 22Na+ fluxes (Lee, Penland et

al. 1998, Devor, Singh et al. 1999), but differs from pH-stat results obtained during

1-EBIO stimulation, when HCO3− secretion accounted for 70% of the Isc and the

discrepancy was due to acid secretion by H+/K+-ATPase (Krouse, Talbott et al. 2004).

The effects of 1-EBIO were not examined in the present study. The pH-stat technique

required unbuffered solution on the apical side and therefore a transepithelial HCO3-

gradient; however, the basolateral-to-apical gradient of 25 mmol·l−1 HCO3- did not

cause overestimation of HCO3- secretion since the Isc was not noticeably affected

when 25 mol·l-1 HCO3-/5% CO2 was added acutely on the apical side (data not

shown), consistent with a previous report (Krouse, Talbott et al. 2004). The

stimulated Isc was similar whether symmetrical HCO3- solutions or pH-stat conditions

were used, further evidence that passive diffusion due to the HCO3- gradient

contributed little to Isc.

2.4.2 HCO3- secretion is CFTR dependent

41

To evaluate the role of CFTR during HCO3- transport, protein expression and anion

secretion were compared in control and CFTR knock-down monolayers. The

Figure 2.1 Relationship between short-circuit current (Isc) and HCO3-

secretion. Forskolin (10 μmol·l-1) and GlyH-101 (100 μmol·l-1) were added to control Calu-3 cells stably expressing shRNA that was mutated to inhibit targeting of CFTR mRNA transcripts. Note the close agreement between Isc and HCO3

- secretion and their sensitivity to the CFTR channel inhibitor GlyH-101 (n=4).

42

immunoblot in Fig. 2.2A shows CFTR expression. For comparison, Lanes 1 and 2 show

baby hamster kidney (BHK) cells stably transfected with wild-type or ΔF508-CFTR,

Lanes 3 and 4 show parental Calu-3 cells, and Lanes 5 and 6 show shRNA control and

CFTR knock-down Calu-3 cells, respectively.

BHK cells expressed mature (i.e. complex glycosylated or ‘band C' polypeptide) and

immature (core-glycosylated, band B) wild-type CFTR whereas only band B was

detected in cells expressing ΔF508 CFTR (Lanes 1 and 2, respectively). Mature and

immature CFTR glycoforms were also found in parental cells (Lanes 3 and 4), and in

cells expressing control (i.e. altered) shRNA (Lane 5). CFTR was usually not detectable

in cells that expressed shRNA targeting CFTR transcripts, indicating efficient

knock-down (Lane 6). Blots were also probed with anti-tubulin antibody to control

for variations in protein loading, and with anti-Na+/K+-ATPase antibody to assess

non-specific changes in membrane protein expression. Tubulin and Na+/K+-ATPase α

subunit levels were not affected by control or CFTR-specific shRNA. Densitometry of

CFTR expression revealed >95% reduction in band C compared with the control cell

line (Fig. 2.2A and inset), consistent with previous estimates (Palmer, Lee et al. 2006).

Band C was lower in shRNA control cells than in parental cells after normalization to

tubulin, despite the presence of four mutations in the shRNA which were intended to

reduce its binding to CFTR mRNA transcripts. Cells expressing control shRNA

43

(referred to previously as ‘alter' by (Palmer, Lee et al. 2006) were used as the control

cell line in subsequent transport experiments, bearing in mind that CFTR expression

Figure 2.2 CFTR expression and HCO3- secretion.

A, CFTR expression in stable cell lines. Lane 1, baby hamster kidney (BHK) cells transfected with wild-type CFTR expressed both mature and immature CFTR. Lane 2, BHK cells transfected with ΔF508 CFTR expressed only immature (band B) CFTR. Lanes 3,4 parental, and Lane 5 control transfected Calu-3 cells also expressed both band B and C. CFTR was not detectable in shRNA knock down cells. Comparison of μg band intensities and protein loaded indicates the following relative expression levels: BHK >> Calu-3 (Parental) > Calu-3 (control transfected) >> Calu-3 (shRNA transfected). B, The CFTR expression is compared to tubulin according to A. C, forskolin (10 μmol·l-1) stimulated Ieq and HCO3

- secretion across wild-type Calu-3

44

cells under open-circuited conditions (n=12). D, CFTR inhibitor GlyH-101 (100 μmol·l-1) nearly abolished the forskolin-stimulated Isc and HCO3

- secretion (n=6). E, the forskolin response was greatly reduced in CFTR-knockdown Calu-3 cells (n=6). F, Summary of forskolin-stimulated Ieq and HCO3

- secretion in B, C and D. ● Ieq; ○ HCO3

- secretion. Means±s.e.m.

45

is lower in these control cells than in parental Calu-3 cells that have been used in

previous studies published by others.

The effect of knocking down CFTR expression on HCO3- secretion was also

investigated under open-circuit conditions using the pH-stat technique (Fig. 2.2B-D).

In marked contrast to Isc, the Ieq calculated from transepithelial voltage and

resistance was higher than the simultaneously measured HCO3- flux; mean Ieq was

2.76 ± 0.13 μeq·cm-2·h-1 (n = 12) during forskolin stimulation whereas net HCO3-

transport was 1.14 ± 0.05 μeq·cm-2·h-1 (n = 12, Fig. 2.2B,C). This difference between

Ieq and net HCO3- flux (∼1.55 μeq·cm-2·h-1) was due to Cl- current because, as shown

below (Table 2.1), a net Cl- flux of 1.5 μeq·cm-2·h-1 in the secretory direction was

observed when unidirectional 36Cl fluxes were measured under these conditions (i.e.

open-circuited and with the same 25 mmol·l-1 HCO3- gradient that was present

during pH-stat experiments). Also, the discrepancy between Ieq and net HCO3- flux

was abolished in Cl--free solution. Ieq and net HCO3- secretion were both inhibited by

the CFTR channel blocker GlyH-101 (Fig. 2.2C); however, this inhibition developed

slowly (100 μmol·l-1; ∼20 min, see Fig. 1 and 2D in Muanprasat et al. 2004), perhaps

due to slow diffusion through the hydrophobic mucus gel on the apical surface.

Alternatively, GlyH-101 inhibition is voltage dependent therefore its affinity may

decline as CFTR channels become progressively blocked and the membrane potential

46

hyperpolarizes towards the potassium equilibrium potential EK. The stimulation of Ieq

and HCO3- secretion by forskolin was greatly reduced in CFTR knock-down cells

compared with

Table 2.1 Unidirectional 36Cl fluxes across Calu-3 cell monolayers.

Experiments were performed under open-circuit, pH-stat conditions (i.e. basolateral 25 mM HCO3

-, apical 0 mM HCO3-). The apical solution was stirred with 100% O2 and

buffered at pH 7.4 using 10 mmol·l-1 tricine.

Jsm Jms Jnet Ieq RT VT

control 1.10 ± 0.12 0.53 ± 0.11 0.58 ± 0.06 0.81 ± 0.06 375 ± 47 -8.14 ± 0.68

forskolin 3.13 ± 0.13 ***

1.63 ± 0.15 ***

1.51 ± 0.02 ***

2.75 ± 0.13 ***

289 ± 30 ***

-21.31 ± 0.97 ***

bumetanide 1.95 ± 0.08 +++

1.22 ± 0.06 +++

0.73 ± 0.05 +++

2.15 ± 0.10 +++

301 ± 32 +

-17.35 ± 0.64 +++

Jsm, Jms, Jnet , Ieq: μEquiv·cm-2·h-1, Rt: Ohms·cm2·h-1, n = 4, Mean±s.e.m. ***, P<0.001, forskolin vs. control. +++, P<0.001; +, P<0.05, bumetanide vs. forskolin

47

control cells (compare Fig. 2.2D and B), although the decrease in CFTR channel

function was somewhat less than the decrease in CFTR protein expression observed

on immunoblots (65% vs. 95% reduction). Figure 2E summarizes the mean Ieq (filled

bars) and HCO3- flux (open bars) across control vs. CFTR-deficient cells. HCO3

- was

secreted by CFTR knock-down cells during forskolin stimulation at a rate that was

similar to unstimulated control monolayers, and to stimulated control monolayers

treated with the CFTR inhibitor GlyH-101. Taken together the results confirm that

forskolin stimulates electrogenic HCO3− secretion and are consistent with CFTR

channels mediating a large part of this flux.

The stimulated Ieq under open-circuit, pH-stat conditions in Fig. 2C was 2.55

μeq·cm-2·h-1, higher than the HCO3- flux of 1.0 μeq·cm-2·h-1 measured simultaneously.

To examine if the discrepancy of ∼1.55 μeq·cm-2·h-1 is carried by Cl-,

unidirectional 36Cl- tracer fluxes were measured under the same conditions, i.e. open

circuited and with a basolateral → apical HCO3- gradient and apical O2 bubbling.

Apical pH was maintained by 25 mM HEPES during tracer experiments rather than by

titration with acid. As shown in Table 2.1, net Cl- secretion (Jnet) was observed under

basal conditions. Forskolin increased 36Cl- fluxes in both directions, but stimulation of

the secretory flow was larger and resulted in a 3-fold increase in the net flux. Thus,

48

Cl- was secreted under open circuit/pH-stat conditions (in marked contrast to Isc

conditions), and the net flux rate (1.51 ± 0.02 μeq·cm-2·h-1) accounted for the

discrepancy of 1.55 μeq·cm-2·h-1 between Ieq and net HCO3- flux measured using the

pH-stat.

2.4.3 Proton secretion neutralizes some of the transported bicarbonate

Acid secretion was also investigated because, if present, it could cause

underestimation of the HCO3- flux. Airway epithelial cells secrete H+ by multiple

mechanisms (Acevedo and Steele 1993, Fischer, Widdicombe et al. 2002, Coakley,

Grubb et al. 2003, Inglis, Wilson et al. 2003), two of which are electrically silent

(H+/K+-ATPase and Na+/H+ exchange) and therefore capable of increasing the

discrepancy between Ieq and HCO3− secretion. Two other mechanisms (electrogenic

vacuolar H+-ATPase and H+ channels) would be “invisible” in these experiments

because they would cause equivalent reductions in current and net HCO3- flux.

To assess the possible role of the non-gastric isoform of H+/K+-ATPase, the effect of

apical ouabain on Ieq and HCO3- secretion was investigated (Fig. 2.3A). Ouabain (100

μmol·l-1) did not alter basal or forskolin-stimulated Ieq or HCO3- secretion, suggesting

little role of ouabain-sensitive H+/K+-ATPase activity with an apical solution

containing 5 mmol·l-1 K+ and bubbled with 100% O2. Apical amiloride (1 mmol·l-1)

also had no effect (data not shown), evidence against significant apical proton

secretion via Na+/H+ exchange under these conditions. By contrast, pretreating cells

49

with the vacuolar H+-ATPase inhibitor bafilomycin for 30 min increased Ieq and HCO3-

secretion by ∼0.25 μeq·cm-2·h-1, and similar increases were obtained under basal and

forskolin-stimulated conditions (Fig. 2.3B and C). This implies that forskolin did not

Figure 2.3 Effect of proton transport inhibitors on Ieq and net HCO3-

secretion. A, ouabain (100 μmol·l-1) had no effect when added to the apical side to inhibit putative H+/K+-ATPase activity (n=3), B,C, pretreatment with the V-ATPase inhibitor bafilomycin (50 nmol·l-1), increased both basal and forskolin stimulated HCO3

- secretion. D, ZnCl2 (nominally 1 mmol·l-1) increased in Ieq and HCO3

- secretion slightly when added to the apical side to block putative H+ channels. ● Ieq; ○ HCO3

- secretion. Means±s.e.m.

50

cause insertion of additional proton pumps into the apical membrane (Paunescu,

Ljubojevic et al. 2010). Finally, acute exposure to ZnCl2 (1 mmol·l-1), which inhibits

acid secretion by blocking apical H+ (HVCN1) channels in airway epithelial cells

(Iovannisci, Illek et al. 2010), increased Isc and HCO3- secretion by ∼0.25 μeq·cm-2·h-1

(Fig. 2.3D), implicating proton channels in the H+ efflux. Taken together these

inhibitor studies indicate that HCO3- secretion is 30% higher across Calu-3

monolayers than the measured net flux and is partially neutralized by parallel H+

secretion. However these electrogenic H+ transport mechanisms would not increase

the discrepancy between Ieq and HCO3- net flux and are thus compatible with the

close agreement between the unidentified Ieq and 36Cl- net flux and the net HCO3-

flux and Isc under voltage clamp.

2.4.4 CFTR mediates apical HCO3- conductance

The next series of experiments used control and CFTR knock-down Calu-3 cells to

examine whether CFTR can directly mediate HCO3- flux. Monolayers were

voltage-clamped at 0 mV, the basolateral membrane was permeabilized using

nystatin (360 μg·ml-1), and an apical → basolateral gradient of 25 mM HCO3- was

imposed. Adding forskolin to control monolayers under these conditions caused a

rapid (i.e. negative) Isc, which was blocked by the inhibitor CFTRinh-172 (10 μmol·l-1)

51

(Ma, Thiagarajah et al. 2002), suggesting ion diffusion through CFTR channels (Gray,

Pollard et al. 1990, Poulsen, Fischer et al. 1994, Illek, Yankaskas et al. 1997, Linsdell,

Tabcharani et al. 1997) (Fig. 2.4A, cited with permission from (Shan, Liao et al. 2012),

Figure 2.4 CFTR-dependent HCO3- conductance at the apical membrane

of permeabilized control and CFTR knock-down (CFTR-KD) Calu-3 cells. The basolateral membrane was permeabilized using nystatin and a HCO3

- gradient was imposed across the monolayer. A, forskolin (10 μmol·l-1) stimulated a large inward current of ~8 μeq·cm-2·h-1. An inward current was stimulated by forskolin (fsk) in control cells but not CFTR-KD cells. B, summary of short-circuit currents

52

measured across control cells and CFTR-KD cells in the presence of a bicarbonate gradient. Most of the forskolin stimulated secretion (ΔIsc) was abrogated in CFTR-KD cells. Means±s.e.m.

this experiment performed by Dr. R. Robert).

Forskolin did not stimulate Isc across CFTR knock-down cells under these conditions

(Fig. 2.4A). Figure 2.4B shows the mean Isc measured across basolaterally

permeabilized monolayers. Although these results do not exclude other apical exit

mechanisms, they show that forskolin-stimulated Isc is CFTRInh-172-sensitive and

reduced by 88% in CFTR knock-down cells, consistent with the >95% decrease in

CFTR protein expression. The simplest interpretation is that CFTR channels conduct

HCO3- and mediate electrogenic HCO3

- secretion at the apical membrane under

these conditions.

To test if apical anion exchangers carry HCO3- out of the cell according to the apical

Cl- gradient, we permeabilized the basolateral membrane using nystatin in

symmetrical Cl--free solutions while supplying the monolayers with basolateral HCO3-

and CO2 by gassing basolateral solutions with 95% O2/5% CO2. The apical solutions

were unbuffered so that the HCO3- flux could be measured by pH-stat. After

stabilization, we established Cl- gradients from the apical to the basolateral side of 0,

30 or 120 mM to favor any apical anion exchange. As shown in Fig. 2.5, the Cl-

gradient caused a reversed Ieq, except when the gradient is 0 or in the CFTR

53

knock-down cells, suggesting the reversed Cl- current was mediated by Cl- diffusion

through CFTR. With the Cl- gradient set at 30 mM to support anion exchange, no

HCO3- flux could be observed until forskolin was added to stimulate CFTR, although

some HCO3- flux

Figure 2.5 CFTR mediates HCO3- flux.

Under pH-stat and Cl--free conditions, the basolateral membrane of Calu-3 cell monolayers was permeabilized using nystatin (360 μg·ml-1). A, without Cl- gradients, forskolin (10 μM) alone elicited similar increases in Ieq and HCO3- flux. B, with a 30 mM Cl- gradient from the apical to the basolateral side, no HCO3

- flux could be observed until forskolin was added. A reversed Ieq was observed when the Cl- gradient was established, and this became larger in the presence of forskolin. C, a 120 mM Cl- gradient from the apical to the basolateral side was established after basolateral permeabilization. Unlike B, HCO3

- flux could be detected without forskolin under these conditions. ● Ieq; ○ HCO3

- flux. Means±s.e.m.

54

could be detected with extreme opposing gradients of 120 mM Cl- and 24 mM HCO3-.

The results showed that under pH-stat conditions, the HCO3- flux was mediated by

CFTR rather than apical anion exchangers, at least with a moderate Cl- gradient.

2.4.5 Electrogenic anion transport (Ieq) requires Na+ and CO2/HCO3-

To examine the ionic requirements of HCO3- transport under open-circuit conditions,

Ieq and the appearance of HCO3− on the apical side were measured after bilateral

removal of CO2/HCO3-, Na+ or Cl- (Fig. 2.6A–D). The basal and forskolin-stimulated Ieq

were both abolished in nominally CO2/HCO3--free solution (Fig. 2.6B).

This dependence of Ieq on CO2/HCO3- would be consistent with Na+-nHCO3

-

(NBC)-mediated cotransport at the basolateral membrane (e.g. (Devor, Singh et al.

1999)), but also with H2CO3 synthesis and subsequent Na+/H+ exchange (Cuthbert

and MacVinish 2003). Forskolin had little effect on Ieq under nominally Na+-free

conditions but did induce electrically silent HCO3- secretion at a rate of >0.5

μeq·cm-2·h-1 in the absence of sodium (Fig. 2.6C). Importantly, transepithelial HCO3-

transport during forskolin stimulation was 25% higher under bilateral Cl--free

conditions than when Cl- was present, despite a 50% reduction in Ieq (from 2.76 ±

55

0.13 to 1.35 ± 0.32 μeq·cm-2·h-1; compare mean values indicated by open circles in

Fig. 2.6A and E). This argues strongly against Cl-/HCO3- exchange as the mechanism of

apical HCO3- exit under these conditions. Ieq was identical to net HCO3

- secretion

under symmetrical Cl--free conditions (1.32 ± 0.31 vs. 1.32 ± 0.19 μeq·cm-2·h-1,

Figure 2.6 Effects of ion substitutions. A, the normal response of Calu-3 cells to 10 μmol·l-1 forskolin in control solutions

56

(n=3). B-D, effect of 10 µmol l-1 forskolin on Calu-3 cells bathed on both sides with solutions lacking HCO3

-/CO2, Na+, or Cl-, respectively (n=6 of each). Forskolin-stimulated HCO3

- secretion was abolished in the nominal absence of HCO3

- or Na+ but was increased by 25% in Cl--free conditions, and Ieq was identical to net HCO3

- secretion. E, monolayers were bathed with apical Cl--free solution, then stimulated with 10 μmol l-1 forskolin. Note the smaller and larger HCO3

- -independent Ieq compared to when cells were bathed with symmetrical Cl- solutions. ● Ieq; ○ HCO3

- secretion (n=6 each condition; means±s.e.m.).

respectively), providing further evidence that HCO3--independent Ieq measured in

control solutions is carried by Cl-. The large transient Ieq observed immediately after

forskolin addition under control conditions(e.g. Fig. 2.6A) was eliminated under

bilateral Cl--free conditions and therefore may reflect Cl- and K+ redistribution when

CFTR channels are activated. When Cl- was removed only from the apical side to

generate a favourable Cl- gradient, forskolin caused more robust stimulation of Ieq

and a larger discrepancy between Ieq and HCO3- flux consistent with an increased Cl-

flux (Fig. 2.6E).

2.4.6 Some basolateral Cl- entry is independent of NKCC

Previous studies of NKCC1's contribution to anion secretion by Calu-3 cells have

yielded varying results, perhaps due to different experimental conditions. In the

present work, the effect of bumetanide on Isc was measured and compared with its

effects on Ieq, net HCO3- flux, tracer fluxes, and fluid transport under open-circuit

conditions. Basolateral bumetanide (20–50 μmol·l-1) did not affect Isc during

stimulation by forskolin (data not shown), in good agreement with a previous study

(Devor, Singh et al. 1999). However, under open-circuit conditions bumetanide (20

μmol·l-1) inhibited ∼20% of the basal (Fig. 2.7A) and forskolin-stimulated (Fig. 2.7B)

57

Ieq, which monitors both Cl- and HCO3- transport. Bumetanide inhibition of the Ieq

component carried by Cl- was ∼35% in control-transfected Calu-3 cells (Fig. 2.7A and

B), similar to the results with parental cells (Huang et al. 2011) and with partial

inhibition of Isc noted previously during stimulation by the hyperpolarizing

Figure 2.7 Effect of bumetanide on forskolin-stimulated Ieq. A, pretreatment with the NKCC1inhibitor bumetanide (20 μmol·l-1) partially inhibited basal Ieq (by ~20%) but did not prevent forskolin-stimulation of Ieq or HCO3

- secretion (n=6). B, basolateral bumetanide (20 μmol·l-1) caused partial inhibition (~20%) of Ieq (●) without affecting HCO3

- secretion (○) (n=6).

58

secretagogues 1-EBIO and thapsigargin (Lee et al. 1998; Krouse et al. 2004).

Net 36Cl flux under these conditions confirmed the bumetanide-sensitive component

of Ieq was mediated by Cl- transport (Table 2.1) and confirmed that a fraction of the

transepithelial Cl- transport is mediated by NKCC1 when the basolateral membrane is

hyperpolarized, either by current flowing through paracellular and transcellular shunt

pathways under open-circuit conditions (i.e. loop current), or by activation of

Ca2+-activated potassium channels under short-circuit conditions. The partial

inhibitory effect of bumetanide on Cl--dependent Ieq in Ussing chambers indicates

that substantial Cl- transport is independent of NKCC1.

2.4.7 Evidence for basolateral Cl- loading by anion exchange

Since a large portion of Cl- and HCO3- transport was not carried by NKCC, we

examined whether other basolateral transporters mediate most of the anion

transport. We found that pretreatment with DIDS (100 μmol·l-1), an inhibitor of NBC

and anion exchangers, abolished the forskolin-stimulated responses in control

transfected Calu-3 cells (Fig. 2.8A). However DIDS is not a specific inhibitor and may

inhibit basolateral NBC activity and/or anion exchangers. To examine the involvement

59

of NBCe1, we substituted Cl- with gluconate in the solutions, which should abolish

any component due to anion exchange. Applying DIDS under these conditions caused

large decreases in both Ieq and HCO3- transport (Fig. 2.8B), clearly indicating the role

of NBC in anion transport of Calu-3 cells.

60

Figure 2.8 Effects of DIDS on Calu-3 cells. A, pretreatment of DIDS (100μmol·l-1) could abolish the plateau of forskolin-stimulated Ieq and HCO3

- secretion. B, in Cl--free solutions, forskolin elicited a reduced response of Ieq, which matched HCO3

- secretion, and it could be inhibited by 100 μmol·l-1 DIDS (basolateral), an inhibitor of NBC in Cl--free solutions. ●, Ieq; ○, HCO3

- secretion (n=6 each condition; means±s.e.m.).

Next, to examine if HCO3- efflux through basolateral anion exchangers could serve as

an alternative Cl- entry pathway during anion loading and fluid secretion,

experiments were designed to monitor the ‘trans' effect of anions on the flow of

HCO3- through the basolateral membrane (Fig. 2.9). The apical membrane was

permeabilized using nystatin (100 μg·ml-1) while monolayers were bathed with

symmetrical HCO3--free solution. After the current reached a plateau, 25 mmol·l-1

NaHCO3 was added on the apical side to generate an apical → basolateral HCO3-

gradient and pH-stat was used to monitor the appearance of HCO3- on the

basolateral side.

As shown in Fig. 2.9A, apical nystatin induced a large current that was apparently due

to electrogenic Na+ absorption since it was inhibited by basolateral ouabain in control

experiments (data not shown) as has been reported for other epithelia (Lewis, Wills

et al. 1978).

After Na+ current had stabilized, imposing a transepithelial HCO3- gradient with Cl-

solution bathing the basolateral (trans) side caused Ieq to decrease and HCO3- to

appear in the basolateral compartment (Fig. 2.9A). When the same experiment was

61

performed with gluconate solution on the basolateral (trans) side, no HCO3- flux to

the basolateral side was detected by pH-stat (Fig. 2.9B). However, robust HCO3- flux

was produced with basolateral (trans) NO3- solution (Fig. 2.9C). No osmotically

induced changes in Ieq or HCO3- flux were observed in control experiments when 50

Figure 2.9 Anion exchange mediates forskolin-stimulated anion and

fluid secretion by Calu-3 cells. A and B, monolayers were incubated in HCO3

-/CO2-free or Cl-&HCO3-/CO2-free

solutions, respectively, and gassed with 100% O2. The apical membrane was permeabilized using nystatin (100 μg/ml), and HCO3

- was added to the apical side with 95% O2/5% CO2. An increase of HCO3

- transport to the basolateral side could be detected in the presence of Cl- (A), but not if Cl- was replaced by gluconate (B) (n=4 of each). C, monolayers were incubated in HCO3

-/CO2-free solutions in which Cl- was replaced by NO3

-. The apical membrane was permeabilized and HCO3-

re-introduced as described in A and B. Basolateral HCO3- flux was observed under

62

these conditions, indicating anion exchange (n=4). ● Ieq; ○ HCO3- secretion.

mmol·l-1 mannitol was added instead of 25 mmol·l-1 NaHCO3 (data not shown). These

results indicate HCO3- flow through the basolateral membrane depends on the

nature of the trans-anion and are consistent with the activity of basolateral anion

exchangers which carry HCO3-, Cl- or NO3

- but not gluconate ions.

2.4.8 Electrogenic HCO3- secretion requires carbonic anhydrase

To study the role of HCO3- synthesis in anion and fluid secretion, the effects of

acetazolamide (100 μmol·l-1) on Ieq and HCO3- transport were assayed in Ussing

chambers under pH-stat conditions. Acetazolamide abolished forskolin-stimulated

HCO3- secretion and Ieq (Fig. 2.10 A,B), suggesting a large fraction of the secreted

HCO3- arises from the intracellular hydration of CO2, as suggested previously (Krouse,

Talbott et al. 2004).

Formation of the weak acid H2CO3 and efflux of HCO3- might be expected to

stimulate Na+/H+ exchangers (NHEs) and other pHi regulatory mechanisms. However,

the NHE inhibitor amiloride (1 mmol·l-1; IC50 = 24 μmol·l-1) and the NHE1

isoform-specific inhibitor HOE-694 (20 μmol·l-1; IC50 = 0.16 μmol·l-1) (Counillon,

Scholz et al. 1993) did not affect HCO3- transport or Ieq (Fig. 2.10C). This suggests pHi

63

regulation may depend on basolateral HCO3- influx via NBCe1 and apical H+ extrusion

during forskolin stimulation.

2.4.9 Origin of the basal Ieq and HCO3- secretion

64

Figure 2.10 Effect of acetazolamide on anion by Calu-3 cells.

A, 100 µmol·l-1 acetazolamide, the inhibitor of carbonic anhydrase, greatly inhibited the forskolin-stimulated secretion (n=6). B, pretreatment with acetazolamide (100 µmol·l-1) also abolished the forskolin-stimulated secretion (n=6). C, sequential addition of the NHE inhibitors HOE694 (10 μmol·l-1) and amiloride (1 mmol·l-1) to the basolateral side did not inhibit HCO3

- secretion (n=3). ● Ieq; ○ HCO3- secretion.

Means±s.e.m.

Calu-3 monolayers have basal currents of 0.5–1.0 μeq·cm-2·h-1; however, the nature

of this constitutive transport remains uncertain. Since Calu-3 cells express purinergic

(Communi, Paindavoine et al. 1999) and adenosine receptors (Cobb, Ruiz et al. 2002)

and respond to prostaglandins (Palmer, Lee et al. 2006), we examined

pharmacologically whether these autocrine signals might be responsible for the basal

current. Ieq and HCO3- transport were unaffected by adding apyrase (10 units·ml-1) to

metabolize ATP released from the cells, 8-SPT (100 μmol·l-1) to prevent activation of

A2B adenosine receptors, or indomethicin (100 μmol·l-1) to inhibit prostaglandin

synthesis (Fig. 2.11A–C).

We then examined the role of CFTR and whether there is tonic adenylyl cyclase

activity. GlyH-101 (100 μmol·l-1), which blocks the CFTR channel with an IC50 ≈ 5

μmol·l-1 at −60 mV (Muanprasat, Sonawane et al. 2004), strongly inhibited both basal

Ieq and HCO3- transport (Fig. 2.12A).

These results suggest CFTR mediates basal secretion. As a further test for the

involvement of CFTR, the dependence of basal current on PKA was investigated.

Adding 100 μmol·l-1 Rp-cAMPS (Rp-adenosine-3’,5’-cyclic mono-phosphorothioate

65

triethylamine salt; IC50 for inhibition of PKA = 4.9 μmol·l-1), a membrane-permeant

competitive inhibitor of PKA, reduced basal Ieq and HCO3- secretion by 63% and 88%,

respectively (Fig. 2.12B), evidence that much of the basal secretion is mediated by

PKA in resting cells and can be rapidly down-regulated by a phosphatase when PKA is

Figure 2.11 Effects of autocrine signaling inhibitors on Ieq and HCO3-

66

secretion. A, the ATPase apyrase (10 units/ml) (n=3) and B, adenosine receptor antagonist 8-SPT (100 µmol·l-1) (n=3) added apically but had no effect on basal Ieq or HCO3

- secretion. C, the cyclo-oxygenase inhibitor indomethacin (100 μmol·l-1) did not inhibit Ieq or HCO3

- secretion when added bilaterally (n=3). Basal current and HCO3

- were CFTR-dependent however, since apical the CFTR channel inhibitor GlyH-101 (100 μM) abolished basal Ieq and HCO3

- secretion (n=4). ●, Ieq; ○, HCO3-

secretion.

67

Figure 2.12 Inhibitors suggest basal Ieq and HCO3- secretion depend on

CFTR and membrane-bound adenylyl cyclase activity, but not

HCO3- -stimulated adenylyl cyclase.

A, basal Ieq was abolished by the CFTR open channel blocker GlyH-101 (n=4). B, basal Ieq and net HCO3

- secretion were inhibited by 200 µmol·l-1 Rp-cAMPS, a competitive inhibitor of cAMP-dependent PKA (n=4). C, basal Ieq and HCO3

- secretion were reduced by the inhibitor of membrane-bound adenylyl cyclase, MDL-12330A (200 µmol·l-1) (n=4). D; inhibition of basal Ieq and HCO3

- secretion by 2-APB (100 μmol·l-1). E,F two chemically unrelated inhibitors of HCO3

--stimulated soluble adenylyl cyclase,

68

2-HE (20 µmol·l-1) and KH7 (50 µmol·l-1) had no effect on Ieq or HCO3- secretion (n=3

of each). G, summary of results in panels A - F. ●,■ Ieq; ○,□ HCO3- secretion.

inhibited.

The sensitivity of basal current to PKA inhibitors implies constitutive elevation of

cAMP, which may be localized near the apical membrane. There are nine

conventional membrane-bound adenylyl cyclases, and one soluble isoform which is

stimulated by HCO3- and is proposed to function as a HCO3

- sensor in various cells

(Chen, Cann et al. 2000), including Calu-3 (Wang et al. 2005). Inhibitors were used to

examine the role of these adenylyl cyclases. An antagonist of conventional

membrane-bound adenylyl cyclases, MDL-12330A (cis-N-(2-phenylcyclopentyl)

azacyclotridec-1-en-2-amine; Guellaen et al. 1977), reduced Ieq and HCO3- secretion

by 68% and 100%, respectively (200 μmol·l-1; IC50 < 10 μmol·l-1; Fig. 2.12C). Since

Ca2+-activated adenylyl cyclase might stimulate CFTR, 2-APB (100 μmol·l-1), a

non-specific inhibitor of store-operated Ca2+ entry, was also tested (Fig. 2.12D).

Remarkably, 2-APB was a potent inhibitor of basal anion secretion, reducing Ieq and

HCO3− secretion 65–90%. This suggests the membrane-bound adenylyl cyclase is

indeed activated by elevated Ca2+ near the membrane. By contrast, bilateral addition

of 2-hydroxyestradiol (2-HE; 20–50 μmol·l-1), an inhibitor of soluble adenylyl cyclase

(Fig. 2.12E), had no effect on basal Ieq or HCO3- transport. The same negative result

69

was obtained using a chemically unrelated inhibitor KH7 (50 μmol·l-1, IC50 = 3–10

μmol·l-1 in vivo; Fig. 2.12F). The effects of these inhibitors on Ieq and HCO3- transport

are summarized in Fig. 12G. Taken together they suggest the basal current in Calu-3

cells is mediated by CFTR channels, which are partially activated by PKA. This tonic

activity may be due to local production of cAMP by a membrane-bound adenylyl

cyclase in response to store-operated Ca2+ entry.

2.5 Discussion

In the current study, the human airway cell line Calu-3 was used because it forms

polarized monolayers, displays cAMP-stimulated bicarbonate and chloride secretion,

expresses the serous cell markers, and responds to gland secretagogues such as

vasoactive intestinal peptide and epinephrine (Shan, Huang et al. 2011). Some results

attained from Calu-3 cells are consistent with previous studies, while others suggest

a revised version of the current anion transport model for Calu-3 cells.

2.5.1 Isc is mediated by electrogenic HCO3- transport

The rate of HCO3- secretion measured in the present study by automated pH-stat was

equal to the simultaneously measured Isc, and both were stimulated by forskolin and

abolished by the CFTR inhibitor Inh-172. Close agreement between Isc and HCO3- flux

is consistent with previous studies of unstimulated (Lee, Penland et al. 1998) and

forskolin-stimulated Calu-3 monolayers (Devor, Singh et al. 1999). The 36Cl- net fluxes

measured under open-circuit conditions in Table 2.1 suggest that membrane

70

hyperpolarization leads to a discrepancy between Isc and HCO3- net flux, as

suggested previously during stimulation with 1-EBIO (Devor, Singh et al. 1999).

Although forskolin also induced some Cl- secretion in the present study, this is

expected since current flowing through shunt pathways would hyperpolarize the

basolateral membrane under open-circuit conditions in concert with apical

depolarization (Tamada, Hug et al. 2001). Thus, the present results extend earlier

findings to open-circuit conditions that would exist during fluid transport.

The present results with H+ transport inhibitors suggest HCO3- secretion is actually

∼30% higher than the net flux measured by pH-stat. Proton secretion by

H+/K+-ATPase has been demonstrated previously during 1-EBIO stimulation (Krouse,

Talbott et al. 2004) but was not observed in the present work. The different results

obtained using forskolin vs. 1-EBIO may be due to their opposite effects on the

membrane potential. Forskolin depolarizes the apical membrane (Tamada, Hug et al.

2001) and should favour electrogenic H+ efflux pathways whereas 1-EBIO would

cause membrane hyperpolarization and favour electroneutral acid secretion.

Electrogenic H+ secretion would be inconspicuous in the present study because it

would cause identical reductions in current and HCO3- net flux. H+/K+-ATPase activity

during 1-EBIO stimulation would consume HCO3- without affecting the current and

therefore would increase the unidentified component of the Isc, as was

demonstrated previously (Krouse, Talbott et al. 2004).

71

2.5.2 Ieq depends on Na+ and CO2/HCO3-

Basal and forskolin-stimulated Ieq were abolished in Na+ or HCO3−-free solution,

consistent with the ionic dependence of Isc reported previously (Devor, Singh et al.

1999). However, we noticed that ∼50% of the HCO3- net flux measured under

pH-stat conditions persisted despite the nominal absence of Na+, suggesting an

additional source of HCO3- besides NBCe1 (Krouse, Talbott et al. 2004). This flux may

reflect carbonic anhydrase-catalysed synthesis and apical efflux since

forskolin-stimulated HCO3- secretion was abolished by acetazolamide (Fig. 2.10A & B),

although we cannot exclude basolateral HCO3- loading through anion exchangers,

which may operate in reversed mode due to the favourable transepithelial HCO3-

gradient under pH-stat conditions.

2.5.3 Cl- loading depends on basolateral anion exchange mostly rather

than NKCC

Bumetanide (100 μM) had little effect on fluid secretion by submucosal glands in cat

trachea (Corrales, Nadel et al. 1984), which suggests the glands use mechanisms

other than NKCC1 unlike airway surface epithelial cells, which are mainly

bumetanide-sensitive. Cl- secretion by freshly isolated equine trachea declined ∼70%

when CO2 and HCO3- were removed under Isc conditions, suggesting basolateral

HCO3-/Cl- exchange (Tessier, Traynor et al. 1990). Further evidence for basolateral

anion exchange (and apical bicarbonate conductance) was obtained in human nasal

primary cells by measuring intracellular pH during extracellular Cl- replacement

72

(Paradiso, Coakley et al. 2003).

In this study, bumetanide (20-50 μM) inhibited only ~35% of the Cl- component of

the Ieq in control-transfected Calu-3 cells, indicating there is an alternative Cl- entry

mechanism besides NKCC1. We demonstrated the basolateral anion exchange by

permeabilizing the apical membrane, imposing an apical-to-basolateral HCO3-

gradient, and examining the effect of anions in the basolateral solution on the HCO3-

flux. At the presence of Cl- or NO3-, an apical to basolateral HCO3

- flux can be

detected. Though the anion exchangers mediating this basolateral Cl- loading were

not identified at the molecular level in the present study, the results suggest AE2 as

the most likely candidate since it is expressed at the basolateral membrane of Calu-3

cells (Loffing, Moyer et al. 2000). Recent studies of an AE2-deficient Calu-3 cell line

indicate that AE2 mediates essentially all Cl−/HCO3− exchange at the basolateral

membrane and plays an important role in fluid secretion; nevertheless, there is

another bumetanide-insensitive mechanism for Cl- loading at the basolateral

membrane in AE2 cells that remains to be identified (Huang, Shan et al. 2012).

2.5.4 Forskolin-stimulated HCO3- secretion requires carbonic anhydrase

Acetazolamide abolished forskolin-stimulated Isc and HCO3- secretion, as observed

previously during stimulation by other secretagogues (Krouse, Talbott et al. 2004).

The simplest explanation for this inhibition is that secreted HCO3- must be

synthesized in the epithelial cells by carbonic anhydrase to sustain electrogenic

73

secretion, even when NBCe1 is available to mediate basolateral HCO3- loading.

Acetazolamide inhibition was surprisingly rapid, perhaps because carbonic anhydrase

activity and the supply of intracellular bicarbonate are limiting when intracellular

pCO2 and [HCO3-] are low due to HCO3

-/CO2-free solution on the apical side.

Alternatively, during forskolin stimulation carbonic anhydrase may form a

bicarbonate transport metabolon with other transporters such as NBCe1 and AE2, so

that inhibiting the enzyme with acetazolamide also rapidly inhibits those transporters.

It is interesting to consider the role of carbonic anhydrase activity in intracellular pH

regulation. Forskolin-stimulated apical HCO3- efflux effectively converts the weak

intracellular acid H2CO3 to a strong acid H+. Cells must neutralize this acid load to

sustain a high rate of HCO3- secretion during steady-state forskolin stimulation.

Basolateral HCO3- entry via NBCe1 probably mediates most of this pHi regulation, as

it occurs during recovery from ammonium-induced acid loads (Inglis, Wilson et al.

2003). Thus, use of basolateral HCO3- entry to neutralize intracellular H+ may

generate CO2 which can be reused for carbonic anhydrase-catalysed HCO3- synthesis,

and this may be especially important when intracellular CO2 and HCO3- are low and

rate limiting. Future studies should examine whether the sensitivity to acetazolamide

is a universal feature of HCO3- secretion or a consequence of using pH-stat conditions

that are admittedly non-physiological.

2.5.5 Further evidence that CFTR mediates apical membrane HCO3− and

Cl− conductance

74

When the basolateral membrane was permeabilized using nystatin and an

apical-to-basolateral HCO3- gradient was imposed, subsequent addition of forskolin

produced a (reversed) Isc across control monolayers that was sensitive to the

inhibitor CFTRinh-172. Although GlyH-101 has also been shown to inhibit the Cl-

channel/exchanger Slc26a9 (Bertrand, Zhang et al. 2009), the anion exchangers

Slc26a3, -a6 and -a11 (Stewart, Shmukler et al. 2011), and mitochondrial function

(Kelly, Trudel et al. 2010), similar results were obtained using CFTR knock-down cells.

Together these results demonstrate the HCO3- conductance of CFTR under these

conditions, consistent with previous studies of Calu-3 monolayers (Illek, Yankaskas et

al. 1997, Tamada, Hug et al. 2001, Krouse, Talbott et al. 2004) and with the single

channel permeability ratio, which ranges between 0.13 and 0.26 (Gray, Pollard et al.

1990, Poulsen, Fischer et al. 1994, Linsdell, Tabcharani et al. 1997). Various factors

could influence permeability of the CFTR pore to Cl- and HCO3-. The pore is partially

blocked by intracellular anions and permeability is enhanced when extracellular Cl- is

replaced with [HCO3-] (Li, Holstead et al. 2011). It may also be modulated by the

WNK1-OSR1/SPAK kinase pathway during stimulated secretion (Park, Nam et al.

2010).

Forskolin-stimulated Isc was less affected by shRNA than CFTR protein expression,

consistent with studies using a different CFTR-deficient Calu-3 cell line (MacVinish,

Cope et al. 2007). Indeed, basal Ieq and HCO3− flux were similar in control and

CFTR-deficient monolayers even though pharmacological studies indicated that basal

75

anion secretion is mediated by CFTR. This result would be explained if some band C

protein that had been lost in knock down cells was not at the cell surface, or if the

decline in conductance caused by silencing CFTR was partly offset by an increase in

anion driving force as would be expected if CFTR was rate-limiting.

On the other hand, when the apical anion exchange is considered, our data suggest

that CFTR that mediates most HCO3- flux under pH-stat conditions. Nevertheless we

cannot exclude the possibility that apical anion exchangers participate in

physiological HCO3- secretion. The pH-stat constantly neutralizes any secreted HCO3

-

and prevents the accumulation of apical HCO3-, thus a basolateral-to-apical 25 mM

HCO3- gradient always exists under these conditions and HCO3

- may readily exit the

cell through CFTR. However, in the airways, HCO3- does accumulate in the thin film of

ASL. To overcome the less favorable transepithelial HCO3- gradient and drive more

HCO3- secretion, airway epithelial cells may then utilize the apical Cl- gradient and the

apical exchangers to transport HCO3-.

2.5.6 A revised anion transport model of Calu-3 cells

The present results suggest a revised model of anion secretion by Calu-3 cells (Fig.

2.13). 50–70% of the basolateral Cl- loading occurs by anion exchange, which is

increased during forskolin-stimulated secretion. This scheme differs from a model in

which NKCC mediates basolateral Cl- entry (Devor, Singh et al. 1999). We hypothesize

that basolateral Cl- loading via basolateral anion exchange increases during forskolin

76

stimulation, in contrast to another recent model for Calu-3 in which forskolin

stimulation was proposed to inhibit this basolateral anion exchange (Garnett,

Hickman et al. 2011).

Figure 2.13 Model for anion transport by Calu-3 cells proposed by Shan

et al.. In resting cells, an apical Ca2+ microdomain produced by store-operated Ca2+ entry causes partial activation of CFTR through stimulation of membrane-bound adenylyl cyclase and local elevation of [cAMP]. Secretagogues that further elevate [cAMPi] stimulate apical Cl− and HCO3

− efflux, creating an acid load that may further increase CFTR open probability (Chen et al. 2009). Regulation of pHi by HCO3

− that enters via NBCe1 provides CO2 for bicarbonate synthesis by carbonic anhydrase, which sustains HCO3

− secretion. Most bicarbonate entering via NBCe1 is recycled basolaterally by anion exchange during Cl− loading. Not shown in this scheme are other anion exchangers and Na+/H+ exchangers that are active under other conditions; i.e. Na+/H+ exchange is likely to mediate pHi regulation during stimulation by secretagogues that hyperpolarize the basolateral membrane.

77

Forskolin stimulates a large transient current which is abolished in Cl--free solution

and therefore probably mediated by apical Cl- and basolateral K+ efflux. While the

decline in intracellular Cl- activity would favour basolateral anion exchange, cell

shrinkage produced by the loss of these solutes may trigger a volume regulatory

increase and stimulation of Cl- entry through NKCC1 (Jiang, Chernova et al. 1997). In

this scheme for Calu-3 cells, most apical HCO3- efflux is sustained by intracellular

HCO3- synthesis, which is catalysed by carbonic anhydrase. The acid load produced by

forskolin-stimulated HCO3- efflux may be neutralized by HCO3

- entry through NBCe1,

whereas H+ extrusion via electroneutral NHE1 (Cuthbert and MacVinish 2003) and

H+/K+-ATPase (Krouse, Talbott et al. 2004) mediate pHi regulation during the

response to hyperpolarizing secretagogues.

Calu-3 monolayers had significant basal Ieq and HCO3- secretion. Basal activity was

sensitive to inhibitors of CFTR (GlyH-101), PKA catalytic subunit (Rp-cAMPS) and

membrane-bound adenylyl cyclase (MDL-12330A). Adenylyl cyclase may be

constitutively activated by Ca2+ entry through the store-operated Ca2+ entry channel

Orai1. A similar mechanism might explain the high basal activity of CFTR noted in

other tissues such as the sweat duct (Quinton 1983) and small airways (Wang, Lytle

78

et al. 2005). The present results do not allow discrimination of conductances that are

strictly dependent on CFTR activity (Bertrand, Zhang et al. 2009).

Chapter 3 Fluid secretion by an

airway epithelial cell line

79

In the previous chapter, we proposed a revised cellular scheme for anion transport by

Calu-3 cells and showed the importance of HCO3- for the short-circuit current and

transepithelial Cl- flux. However, preliminary studies of fluid secretion showed that

the concentration of HCO3- was surprisingly low, ~25 mM. We therefore proceeded

to study the relationships between different anions and fluid secretion, and

investigated how fluid secretion is regulated.

80

3.1 Abstract

Fluid secretion is reduced in cystic fibrosis, a disease caused by CFTR dysfuncion. In

normal airways, CFTR mediates the secretion of Cl- and HCO3- however the precise

roles of these anions in fluid transport by airway epithelial cells is not well

understood. In this study, we collected fluid produced by genetically matched

CFTR-deficient and CFTR-expressing cell lines derived from a human airway epithelial

cell line (Calu-3), and measured the volume, pH, anion composition and the

osmolality of the samples. Forskolin elicited CFTR-dependent fluid secretion.

Compared to the CFTR-expressing cell line, the volume of fluid secretion was greatly

reduced and the pH was more acidic in the CFTR-deficient cell line, suggesting less Cl-

and HCO3- transport, consistent with the results in the previous chapter. Fluid

secretion depended on the presence of all three ions (Na+, Cl- and HCO3-). Cl- was the

predominant anion in the secretions and therefore drove most of the fluid transport.

Consistent with the partial inhibition of Cl- transport by bumetanide (see Ieq

measurements in previous chapter), the volume of fluid produced was inhibited only

~20% by bumetanide, indicating a minimal role for NKCC1 and the presence of an

alternative Cl- transport mechanism at the basolateral membrane. Moreover adding

NO3-, which is not carried by NKCC1, restored most of the fluid secretion in nominally

81

Cl--free medium, suggesting an alternative mechanism, e.g. anion exchange, exists at

the basolateral membrane for Cl- loading. These results support a model proposed in

the previous chapter, in which most HCO3- taken up by NBCe1 is recycled

basolaterally by exchange for Cl-, and the resulting HCO3--dependent Cl- transport

generates most of the osmotic driving force for fluid secretion. Consistent with this

interpretation, additing acetazolamide in this fluid secretion assay, an inhibitor of

carbonic anhydrase that abolished Ieq and HCO3- secretion, caused only ~20%

inhibition of fluid transport. Time course studies revealed that most fluid was

produced during the first 6 h in this 24 h assay. On the other hand, we found that

evaporation did not affect the accuracy of the volume measurements. These data

prompted us to propose a novel but simple model for the regulation of ASL volume,

in which fluid secretion or absorption is mainly determined by osmolality created by

secreted ions.

3.2 Introduction

The airways are covered by a thin film of liquid which humidifies inspired air and

enables the cilia to beat and clear mucus containing pathogens and other inhaled

particles from the lungs. Most airway surface liquid (ASL) is produced by submucosal

glands, which secrete a complex mixture of ions, fluid, mucins and antimicrobial

factors when stimulated by agonists that mobilize intracellular Ca2+ and/or [cAMP]

(Wine and Joo 2004). The depth of ASL is 5–30 μm in healthy airways (Widdicombe

2002, Boucher 2003), but is reduced in cystic fibrosis (CF) due to abnormal

82

transepithelial ion and water transport (Tarran, Trout et al. 2006), and this diminishes

mucociliary clearance and leads to recurring infections and chronic inflammation.

Despite the important role of ASL in airway host defense and pathophysiology, the

relationship between anion and fluid secretion by airway epithelia remains poorly

understood.

Isosmotic fluid secretion by most epithelia is driven by the active transport of anions,

which are followed passively by cations and osmotically-obliged water (Ballard and

Inglis 2004). While NKCC1 is the main route for basolateral Cl- uptake during

secretion across many epithelia, in some tissues such as airway submucosal glands,

fluid secretion is insensitive to the NKCC1 inhibitor bumetanide (Corrales, Nadel et al.

1984). Basolateral anion exchange has been proposed as an alternative mechanism

for basolateral Cl- entry in salivary glands (Case, Hunter et al. 1984, Novak and Young

1986, Pirani, Evans et al. 1987, Turner and George 1988) and intact trachea, where

most secretion is from the glands (Tessier, Traynor et al. 1990). Removing exogenous

CO2 and HCO3- inhibits fluid secretion by individual airway submucosal glands ∼50%,

and similar results are obtained during stimulation by vasoactive intestinal peptide

(VIP) and/or acetylcholine (Joo, Saenz et al. 2002, Choi, Joo et al. 2007). This

dependence of fluid secretion on exogenous CO2/HCO3- would be greater if not for

the metabolic production of CO2 and HCO3− by the epithelium.

Previous studies have suggested various cellular models for anion and fluid secretion

83

in the airways. Although Cl- and HCO3- can permeate CFTR (Gray, Pollard et al. 1990,

Poulsen, Fischer et al. 1994, Linsdell, Tabcharani et al. 1997), apical anion exchange

has recently been proposed to mediate most HCO3− secretion by Calu-3 cells

(Garnett, Hickman et al. 2011). This model is similar to those that have been

proposed for duodenum (Simpson, Gawenis et al. 2005) and pancreatic duct (Lee,

Choi et al. 1999), which produce strongly alkaline secretions that are driven by HCO3−.

However, the pH of submucosal gland secretions is near neutrality in 5% CO2 (pH

6.9–7.1) (Song, Salinas et al. 2006) similar to the ASL pH, which ranges between 6.2

and 7.2 (Fischer and Widdicombe 2006). The low HCO3− concentration of airway

submucosal gland secretions and ASL (~20 mmol·l−1) is difficult to reconcile with the

high rates of HCO3− secretion in other tissues where HCO3

− drives fluid secretion.

Although HCO3− would contribute 20 mosmol·l−1 to the osmotic pressure of gland

secretions (i.e. < 7% of the total solute concentration of 300 mosmol·l−1), removing

exogenous CO2/HCO3− reduces the volume of secretions by ≥ 50% (Joo, Irokawa et al.

2002). These differences indicate that although HCO3- may not be at a high

concentration in the ASL, it is nevertheless vital for fluid secretion and the

maintenance of ASL volume.

In this study, we sampled fluid produced by polarized Calu-3 monolayers at different

time points and measured its volume, pH, Cl- concentration and osmolality. By

comparing CFTR-expressing and CFTR-deficient cell lines, we found that forskolin

elicits CFTR-dependent fluid secretion. Fluid transport by Calu-3 monolayers was

84

dependent on Na+, Cl- and HCO3- and was abolished by removal of any of these ions.

Cl- was the predominant anion in the secreted fluid, whereas the concentration of

HCO3- was relatively low. These and other unexpected results prompted us to

investigate the role of anions in fluid transport and propose a novel, but simple,

mechanism for the regulation of ASL volume.

3.3 Methods

3.3.1 Cell Culture

The CFTR-deficient cell line and the control CFTR-expressing cell lines called Sizt and

Alter, respectively, were obtained from Dr. Scott O’Grady at the Univ. of Minnesota

(Palmer, Lee et al. 2006). Both cell lines were cultured in Eagle's minimum essential

medium (EMEM; Wisent Bioproducts, St. Bruno, QC) containing 7% fetal bovine

serum (FBS). Cells were seeded at 5 × 105 cells per cm2 on Transwells (4.67 cm2,

Corning) for fluid secretion assays. Fresh medium was placed on the basolateral side

one day after plating and the apical medium was removed to establish an air–liquid

interface (ALI). Any fluid that appeared spontaneously on the apical surface was

removed after 3 days. Cultures were maintained in a humidified 5% CO2 incubator at

37°C and studied after 21–25 days.

3.3.2 Media

To maintain viability during long fluid secretion experiments, EMEM medium

containing essential amino acids was placed on the basolateral side. The salts of

85

amino acids (l-histidine·HCl, l-lysine·HCl) and vitamins (choline chloride,

pyridoxine·HCl, thiamine·HCl) contributed 0.6 mmol·l−1 Cl− to the nominally Cl−-free

medium. Cl− contamination probably did not affect the results however, since adding

this concentration of Cl− during pH-stat measurements had no detectable effect on

Ieq, which is a measure of the net Cl− + HCO3− flux under open-circuit conditions.

Monolayers bathed with basolateral 25 mmol·l−1 HCO3− medium were kept in 5%

CO2/95% air, which was nominally saturated with H2O during fluid secretion

measurements unless otherwise noted. Control experiments performed with

Transwells mounted on an orbital shaker in a 5% CO2/95% O2 atmosphere to provide

better oxygenation yielded only slightly higher fluid secretion rates (data not shown).

Humidified air (0.035% CO2) was used when measuring fluid secretion with

nominally HCO3−-free basolateral medium.

3.3.3 Volume and composition of the secreted fluid

Fluid was aspirated from the apical surface and fresh media containing replacement

ions, activators or inhibitors were added on the basolateral side to begin fluid

secretion assays. In some experiments, agents were added to the apical side in a

small volume of Krebs–Henseleit solution (mmol·l−1: 120 NaCl, 25 NaHCO3, 3.3

KH2PO4, 0.8 K2HPO4, 1.2 MgCl2, 1.2 CaCl2 and 10 mannitol), which was subtracted

from the volume measured at the end of the experiment when calculating secretion

rates. Apical fluid was collected at 24 h intervals using a pipettor and the

time-averaged fluid secretion rate was calculated after subtracting the volume of

86

vehicle if added on the apical side. Although cumulative volume increased linearly for

at least 3 days, the instantaneous secretion rate declined over the course of each 24

h collection period, therefore total volume secreted under these conditions is

sub-maximal. A low rate of evaporation was observed from siliconized Transwells in

control measurements (∼0.27 μl·cm−2·h−1) despite being kept in a humidified

incubator. However, when cultures were coated with a film of water-saturated

hexadecane (a solvent having low vapor pressure, which eliminates evaporation), the

volume of secretions that could be retrieved from cell monolayers was not altered.

High osmotic permeability apparently enables sufficient H2O diffusion through the

monolayer to compensate for evaporative H2O loss from the surface of uncoated

cultures so that the steady-state apical fluid volume is determined simply by the

quantity of solute on the surface. Correcting for evaporation would cause

overestimation of active fluid transport, therefore no correction was applied. The Cl−

concentrations of secretions and media were measured using an ADVIA 1650

biochemistry system (Bayer) after 10-fold dilution in distilled H2O. pH was measured

using a micro-electrode (9826BN, Orion) which was either kept inside the 5% CO2

incubator to reduce equilibration time or immersed in the sample while bubbling

with 5% CO2. The Henderson–Hasselbalch equation was used to calculate the

equilibrium HCO3− concentration:

where pK = 6.09 and PCO2 = %CO2/100·(760-47). No corrections were made for small

errors in water vapor pressure (slightly less than 47 mmHg), laboratory elevation

87

above sea level (∼31 m) or daily fluctuations in barometric pressure. The incubator

was set using the built-in infrared controller and confirmed independently within

±0.01% using an external, non-dispersive infrared (NDIR) sensor (GMM221, Vaisala,

Lake Villa, IL, USA).

3.3.4 Data analysis

Paired or unpaired Student's t tests with P < 0.05 were used for single comparisons,

whereas one-way analysis of variance followed by the Bonferonni post hoc test was

used for multiple comparisons.

3.4 Results

3.4.1 Fluid secretion is CFTR dependent

Dehydration is a hallmark of CF airway secretions, and suggests a role of CFTR in fluid

secretion. To test if fluid transport by Calu-3 cells is CFTR dependent, apical fluid was

removed at time 0 h and secretions were collected from control and CFTR

knock-down cultures at 24 h intervals under control conditions and during forskolin

stimulation (Fig. 3.1A). Little fluid was produced by control monolayers that were left

untreated or exposed only to vehicle (0.1% basolateral DMSO). However, fluid

secretion increased to > 40 μl·day−1 during stimulation with forskolin (10 μmol·l−1),

and a similar volume was produced over the course of several days, leading to the

cumulative secretion of ∼130 μl.

88

Figure 3.1B shows the time-averaged fluid secretion rate measured daily for 3 days

under each condition. GlyH-101 (100 μmol·l−1) nearly abolished forskolin-stimulated

fluid secretion when added to the apical side in 400 μl PBS, suggesting CFTR channel

activity is required for fluid transport.

3.4.2 Cl− is the predominant anion in Calu-3 secretions

89

Figure 3.1 CFTR-dependent fluid secretion by control and CFTR knock

down Calu-3 cells. A, Fluid secretion under different conditions: ■ control Calu-3 cells stimulated with 10 μmol·l-1 forskolin; ▲ untreated, ▼ 0.1 % DMSO, or ● forskolin + GlyH-101 (100 μmol·l-1) treated control Calu-3 cells, and ◆ CFTR knock down Calu-3 cells treated with forskolin (n=6 each condition). B, summary of results in A. Means±s.e.m.

90

To investigate if Cl- and/or HCO3- is essential to fluid secretion, the pH and the [Cl-] of

apical fluid on control and CFTR knock-down monolayers was measured. Surprisingly,

the pH of the fluid was in the range 7.2–7.6 when equilibrated with 5% CO2 (Table

3.1). The Cl− concentration in fluid secreted by control shRNA transfected, CFTR

knock-down and parental Calu-3 cell lines was ∼ 120 mmol·l−1. Thus, Cl− was the

predominant anion and was 2- to 4-fold higher than the [HCO3−] calculated using the

Henderson–Hasselbalch equation. [HCO3−] was similar in secretions collected after

several hours or several days of forskolin stimulation (data not shown), and the

osmotic pressure of secretions at the end of collection periods was equal to that of

the basolateral medium within measurement error (Table 3.1), suggesting most fluid

is osmotically driven by transepithelial Cl− rather than by HCO3− secretion despite the

close correspondence between Isc and HCO3− secretion (Fig. 2.1), and the absence of

detectable 36Cl− net flux under Isc conditions (Devor, Singh et al. 1999).

3.4.3 HCO3−, Na+ and Cl− are all required for forskolin-stimulated fluid

secretion

The low [HCO3−] and high [Cl−] of Calu-3 secretions prompted us to study the ionic

requirements for fluid secretion and Ieq. Fluid transport was reduced from 45.1 ± 7.9

μl·cm−2·day−1 to 0.8 ± 0.4 μl·cm−2·day−1 when monolayers were bathed with

nominally HCO3−-free medium (Fig. 3.2), and a similar decrease to 1.0 ± 0.8

μl·cm−2·day−1 was observed in nominally Na+-free medium. These effects of removing

HCO3− and Na+ were anticipated since Ieq also depended on these ions (e.g. Fig. 2.6B

91

Table 3.1 Composition of fluid secreted by Calu-3 monolayers.

Fluid was collected daily during forskolin (10 μmol l-1) stimulation and pooled for analysis.

Control transfected CFTR knockdown Culture Medium

Measured [Cl-] 120 ± 5 123 ± 6 135 ± 16

Measured pH 7.55 ± 0.04 7.28 ± 0.02 7.43 ± 0.03

Calculated [HCO3-] 31.35 ± 2.1 16.08 ± 0.52*** 24.4 ± 2.2

Osmotic pressure 309 ± 16 301 ± 6 310 ± 16

mmol·l-1 or mosmol·l-1; n = 3 - 12 each, Mean ± s.e.m. ***, P<0.001, CFTR knockdown vs. control transfected.

92

Figure 3.2 Effects of ion substitutions on fluid secretion. A, fluid secretion was abolished in the nominal absence of Na+, HCO3

-/CO2 or Cl-. B, summary of results in A. ■ normal medium; ▲ Na+-free medium; ▼ Cl--free medium; ◆ HCO3

-/CO2-free medium (n=6 each condition; means±s.e.m.).

93

and C). More surprising was the effect of replacing Cl− with gluconate, which nearly

abolished fluid secretion (from 45.1 ± 7.9 μl·cm−2·day−1 to 2.0 ± 0.7 μl·cm−2·day−1; Fig.

3.2). This contrasts with the effect on Ieq, 60% of which persisted in nominally

Cl−-free solution (Fig. 2.6D). Thus, forskolin-stimulated HCO3− secretion depends on

the simultaneous presence of Na+ and HCO3− but not Cl−, whereas fluid secretion

requires all three ions (Fig. 3.2).

Ion substitution effects and the high [Cl−] of secretions suggested that most fluid is

osmotically driven by transepithelial Cl− rather than HCO3− transport.

3.4.4 Most of the fluid secretion is independent of NKCC

Bumetanide (20 μmol·l−1) reduced fluid secretion from 33.5 ± 6.98 to 26.6 ± 1.39

μl·cm−2·day−1; however, the remaining 70–80% of the fluid secretion was insensitive

to this inhibitor despite the high [Cl−] of the secreted fluid (Fig. 3.3). Inhibition of fluid

transport was not further increased by raising bumetanide concentration to 50

μmol·l−1 (data not shown). Partial inhibition of fluid transport by bumetanide

paralleled its effect on Cl−-dependent Ieq in Ussing chambers, reinforcing the

conclusion that substantial Cl− and fluid transport is independent of NKCC1.

3.4.5 Basolateral anion exchange is important to fluid secretion

In the previous chapter, we have shown that basolateral anion exchange is important

to Cl- loading and AE2 is a candidate to be the mediator of the anion exchange. Since

94

Figure 3.3 Effects of bumetanide on forskolin-stimulated fluid

secretion. forskolin-stimulated fluid secretion was inhibited ~15% by bumetanide (▲) when compared to forskolin alone (■) (n=6 each condition).

95

NO3− substituted for Cl− during basolateral anion exchange, we wondered if it could

also support fluid transport. Forskolin stimulated robust fluid secretion when

monolayers were bathed with nominally Cl−-free NO3− medium on the basolateral

side (Fig. 3.4). The rate of fluid transport from basolateral NO3− solution was > 60% of

that measured with Cl− solution during the first 24 h stimulation, then gradually

declined, presumably because Cl− is needed for other cellular processes in order to

maintain long-term viability. Nevertheless, NO3− did sustain fluid secretion for many

hours, evidence that its basolateral entry is supported by HCO3− efflux in apically

permeabilized monolayers. HCO3− recycling through basolateral anion exchangers

can apparently load Calu-3 cells with either Cl− or NO3−, and both these anions

support fluid secretion because both are permeant through apical CFTR channels

(Linsdell, Tabcharani et al. 1997).

3.4.6 Carbonic anhydrase has little effect on fluid secretion

In chapter 2, we found that carbonic anhydrase is vital to anion transport by Calu-3

cells, since inhibiting it with acetazolamide abolished forskolin-stimulated

transepithelial anion transport as measured by Ieq. Thus, we speculated that carbonic

anhydrase may also play an important role in fluid transport, and that treatment with

acetazolamide should greatly reduce fluid secretion by stimulated Calu-3 cells.

However, application of acetazolamide (100 μM) only reduced fluid secretion by 27%

(Fig. 3.5), regardless of whether cells were pretreated with acetazolamide or if it was

added during forskolin stimulation.

96

Figure 3.4 Anion exchange mediates forskolin-stimulated fluid secretion

by Calu-3 cells. Fluid secretion was largely maintained for the first 24 h when Cl- was replaced with NO3

- but was nearly abolished after gluconate replacement. ■ normal medium; ▲ low-Cl- gluconate medium; ▼ low-Cl- nitrate medium.

97

Figure 3.5 Effects of acetazolamide on fluid secretion by Calu-3 cells. ■, without acetazolamide; ▲, with acetazolamide (n=6 of each). Acetazolamide caused only a small reduction in fluid secretion by Calu-3 cells.

98

3.4.7 The time course of fluid secretion rate during 24 h stimulation

To see if the fluid was secreted constantly during 24 h stimulation, we collected fluid

and measured the volume every hour, and then put the fluid back onto the apical

surface of Calu-3 monolayers for the next measurement to simulate a normal

secretion assay.

Fig. 3.6 shows that after forskolin stimulation, the control CFTR-expressing Calu-3

monolayers actively secreted during the first 6 h at an average rate of ~5 μl·cm-2·h-1,

and nearly 75% of the total daily amount was secreted during this initial period. After

the first 6 h, fluid secretion rate declined by ~90% to 0.5 μl·cm-2·h-1and fluid

accumulated slowly on the monolayers.

3.4.8 Evaporation does not affect the rate of fluid accumulation

It is believed that evaporation reduces the fluid volume under these conditions and

affects the accuracy of the volume measurements, consequently others have

corrected for evaporation (Garnett et al., 2011). We determined the evaporation rate

by sealing the Transwell with silicone polymer, and adding 200 μl medium onto the

apical compartment of the well. We observed a low rate of evaporation at ~0.3

μl·cm-2·h-1 despite keeing the Transwells in an atmosphere that was > 90% saturated

with H2O.

To test the impact of this evaporation on measurements of fluid secretion, we

99

Figure 3.6 Time course of fluid secretion during 24 h fluid secretion

assay. Calu-3 cell monolayers were stimulated by forskolin (10 μM). Most fluid was secreted during the first 6 h after stimulation. After the rapid secretion phase, there was a slow secretion phase, in which the secretion rate was almost one tenth of the one in the rapid phase (n=6 of each).

100

utilized water-saturated hexadecane, a solvent which has a very low vapor pressure

that should prevent evaporation. It was shown to be effective when the medium on

the siliconized Transwells was covered by a thin film of water-saturated hexadecane

(Fig.3.7A). However, when Calu-3 monolayers were coated with hexadecane, the

volume of secreted fluid that could be retrieved from the monolayers was not altered

(Fig. 3.7B). This rather surprising result suggests that the high osmotic permeability

and the osmotic driving force created by secreted ions is sufficient for H2O diffusion

through the monolayer to fully compensate for evaporative H2O losses. Therefore,

the apical fluid volume is determined simply by the quantity of solute on the surface.

The common practice of correcting for evaporation, which ignores this passive water

flux, is expected to cause overestimation of active fluid transport, therefore no

correction was applied to data from the fluid secretion assays.

3.4.9 Osmolality determined the fluid secretion rate

To explore the hypothesis that total solute content determines fluid flux, we applied

apical solutions with different osmolalities to unstimulated and stimulated Calu-3

monolayers.

Without forskolin stimulation, monolayers at the air-liquid interface (control) and in

solutions with 290 mOsm showed similar fluid secretion rates, ~10 μl·cm-2·h-1.

Solutions at 500 mOsm caused an increase in water flux at a rate of 60 μl·cm-2·h-1,

whereas exposure to 55 mOsm solution led to absorption at 12 μl·cm-2·h-1 rather

101

Figure 3.7 Evaporation did not affect fluid accumulation. A, 200 μl culture medium was added into the apical compartments of Transwells, which were sealed with silicon basolaterally. The culture medium was either covered with a thin film of hexadecane or left uncovered. After 24 h in the incubator, without hexadecane, the volume of the culture medium was reduced to less than 150 μl due to evaporation (n=3 of each). B, Calu-3 cells were cultured on Transwells and stimulated using 10 μM forskolin. There was no difference in the volume of fluid collected between groups with and without hexadecane (n=6 of each).

102

than secretion (Fig. 3.8A). During forskolin stimulation, addition of 290 mOsm

solution apically caused a significant reduction in the fluid secretion rate from ~40 to

~28 μl·cm-2·h-1, compared to control (Fig. 3.8B), presumably because adding solution

on the apical side caused dilution of the osmolytes secreted by the monolayer and

reduced the driving force that was generated initially by forskolin stimulated of salt

secretion.

To further investigate how osmolality regulates the fluid secretion rate, we studied

the time course of osmolality changes in secretions. We collected secreted fluid and

basolateral medium every hour for the first 5 h during a 24 h fluid secretion assay

and measured the osmolality. We found that the osmolality of apical fluid was always

higher than in the basolateral medium during the first 5 h of stimulated fluid

secretion by Calu-3 monolayers (Fig. 3.8C). After the initial active secretion phase,

the osmolality of the fluid and the medium became similar, consistent with the

observed slowing of the secretion rate in Fig. 3.6.

These results confirmed that we could elicit net fluid flux without stimulating the

monolayers, by imposing an osmotic gradient, and that the rate of active fluid

secretion was correlated with the transepithelial osmotic gradient generated by salt

transport.

3.5 Discussion

103

Figure 3.8 Osmolality determine the fluid secretion rate.

A, without stimulation by forskolin, Calu-3 cell monolayers were bathed apically with solutions at different osmolality values. Different fluid secretion (absorption) rates were observed accordingly to the osmolality (n=6 of each). B, with stimulation by forskolin, apical solution at 290 mOsm greatly reduced the fluid secretion rate, when compared to control (without apical solution)(n=6 of each). C, a time course of the fluid and the culture medium osmolality. Fluid secretion and the culture medium were collected every hour during the first 5 h and one final time at 24 h. the osmolality was measured. During the rapid secretion phase, the osmolality of the fluid was generally higher than the one of the medium. However, the osmolality of the fluid and the medium eventually became the same at the end of 24 h fluid secretion assay (n=6 of each).

104

In the present study, we compared fluid secretion by CFTR-expressing and

CFTR-deficient cell lines, investigated the ionic requirements for fluid secretion, and

studied aspects of fluid transport regulation. The results are consistent with those in

the previous chapter, and with the anion transport model we have proposed for

Calu-3 cells.

3.5.1 Most fluid is driven by the net flux of Cl−

The [Cl-] of secretions was ∼120 mmol·l−1, or about 4-fold higher than the [HCO3−].

This implies that most fluid secretion is driven by transepithelial Cl− rather than

HCO3− flux during forskolin stimulation despite the fact that active Cl− transport is not

detected under Isc conditions (Devor, Singh et al. 1999) or open circuit conditions

(this thesis), and that Isc can be fully explained by the net HCO3− flux. Using the [Cl-]

of secretions, we calculate that during a 24 h fluid secretion assay, the Cl- secretion

rate is ~0.2 μeq·cm−2·h−1, a number much smaller than the Cl- component of the Ieq

(1.5 μeq·cm−2·h−1). This calculation clearly shows that a very low rate of Cl− transport

would be sufficient to sustain fluid secretion at the observed rate. Such low rates are

difficult to detect using radiotracer fluxes because they are the difference between

two large unidirectional fluxes, and because there is only one useful radioisotope of

Cl- thus unidirectional fluxes must be measured across different monolayers, which

introduces variability larger than the net flux.

A related finding was that much of the Cl− transport which drives fluid secretion is

105

independent of NKCC1. This conclusion was based on anion selectivity and

pharmacological data. Robust fluid secretion was observed for 24 h when the

basolateral side was bathed with nominally Cl−-free solution containing NO3−, which

does not bind to the highly selective anion site on NKCC transporters and therefore is

not transported (Kinne, Kinne-Saffran et al. 1986). Further evidence for

NKCC1-independent Cl− entry is the relative insensitivity of fluid secretion to

bumetanide. These results suggest other basolateral Cl− entry pathways such as

anion exchangers are important for fluid secretion. NO3− is carried by many anion

exchangers including AE2 (Humphreys, Jiang et al. 1994).

Although HCO3− is secreted by Calu-3 cells under open-circuit conditions, its

concentration in the secreted fluid is low relative to Cl−. Thus HCO3- seems to be

involved in Cl- loading via basolateral anion exchange rather than driving the fluid

secretion. The high concentration of Cl− in secretions suggests that it together with

electrically coupled Na+ flux provides most of the osmotic driving force for fluid

secretion.

3.5.2 Calu-3 secretions are only weakly alkaline

The normal pH of ASL and gland secretions is in the range 6.2–7.2 (Fischer and

Widdicombe 2006). Weakly buffered solutions placed on cultures of surface

epithelium become acidic, and pH falls more rapidly on CF than non-CF monolayers,

consistent with a defect in parallel HCO3− secretion (Coakley, Grubb et al. 2003).

106

Calculations based on the pH of gland fluid collected in the presence vs. absence of

HCO3− (Song, Salinas et al. 2006) suggests that these secretions normally contain

about 14 mmol·l−1 HCO3−. Moreover, secretions from pig glands are 0.2–0.4 units

more acidic when secreted in the presence of CFTR inhibitors, mimicking the low pH

seen with CF glands (Song, Salinas et al. 2006).

HCO3−-driven fluid transport was expected to generate strongly alkaline secretions

having pH > 8; however, the maximal pH during forskolin stimulation in the present

experiments was ∼7.6. The calculated [HCO3−] in secretions was slightly higher than

in the basolateral medium (33 vs. 25 mmol·l−1), and higher apical [HCO3−] was

achieved by parental Calu-3 monolayers, which have significantly higher CFTR

expression (∼50 mm [HCO3−]; D. Kim, unpublished observations, Fig. 2A);

nevertheless, secretions were always much less alkaline than expected if the fluid

were driven significantly by HCO3− transport. The values obtained with parental cells

were generally consistent with previous studies in a virtual gland preparation (74 mM

HCO3− (Irokawa, Krouse et al. 2004)), and in cyclophilin B shRNA control Calu-3 cells

cultured on transwells (60 mM (Garnett, Hickman et al. 2011)). Silencing CFTR

reduced the [HCO3−] of secretions by ∼50% in the present study, but had no effect

on the pH and HCO3− concentration in a previous study (Garnett et al. 2011), perhaps

because the knock-down cells used in that study had higher residual CFTR expression

(28% vs. < 5% in the present work). The present results are consistent with the

decreased [HCO3−] reported for CF gland secretions (Song, Salinas et al. 2006),

107

although the starting value in glands was lower (i.e. 3.5 mmol·l−1 vs. 13.8 mmol·l−1

[HCO3−] in CF vs. non-CF glands, respectively).

In the present work, the total volume secreted and Cl− content were both reduced

10-fold by silencing CFTR expression and the [HCO3−] was reduced by 2-fold. Similar

changes in CF airways may contribute to disease symptoms by altering mucus release

or rheology, innate immunity or ciliary beating (Shan, Huang et al. 2011).

3.5.3 A very low rate of net Cl− secretion is sufficient to drive fluid

secretion

HCO3− secretion accounted for ∼40% of the Ieq under pH-stat conditions. The

remainder (1.5 μeq·cm−2·h−1) was carried by Cl−, as revealed by measuring 36Cl−

fluxes under the same conditions. Nevertheless, Cl− net flux was negligible during

previous short-circuit current experiments (Devor et al. 1999) and under open-circuit

conditions in the present study (Table 1), making it difficult to understand how Calu-3

monolayers could produce a Cl−-rich fluid. However, it is important to note that a net

Cl− flux of only 0.2 μeq·cm−2·h−1 would be sufficient to account for the Cl− content of

secretions in the present, and such low rates cannot be determined reliably in Ussing

chambers.

3.5.4 The osmolality determines the fluid secretion rate

The optimal height of ASL is ~7 μm, which is also the height of outstretched cilia. The

108

height of ASL is fine tuned in the airways, but it is not clear how airway epithelia

sense the volume and adjust the fluid secretion rate accordingly. Some studies

hypothesized that cilia of airway epithelial cells might act in a mechanosensory

capacity to signal ASL volume status to the underlying epithelia (Tarran, Grubb et al.

2001, Chambers, Rollins et al. 2007). However, a role for cilia in mechanically sensing

ASL volume has not been demonstrated. Other studies have suggested that soluble

stimulatory molecules such as adenosine and channel-activating proteases or sodium

channel regulators such as SPLUNC1 in the ASL may function as chemosensors that

regulate the balance between secretion and absorption (Garcia-Caballero,

Rasmussen et al. 2009).

Although Calu-3 does not maintain a constant ASL volume equivalent to surface

airway cells, our studies suggest a simple mechanism for regulating this balance in

which osmotic driving force created by secreted solutes determines fluid secretion

rate and steady-state volume. Firstly, our data from the evaporation test suggest that

the osmotic driving force created by forskolin stimulated ion secretion drives a

passive flux of solute-free water, and hypertonicity caused by evaporation has a

similar effect, drawing water from the basolateral side into the apical compartment

to compensate for that lost by evaporation.

Secondly, incubating Calu-3 monolayers with solutions having high or low osmolality

caused water flows without forskolin stimulation and rapidly generated absorption or

109

secretion. However adding isotonic solution on the apical side reduced the active

fluid secretion rate, presumably by diluting the local osmotic driving force created by

forskolin stimulated salt secretion. Thirdly, during the initial fast rate of secretion, the

osmolality of secretions was always slightly higher than the basolateral medium,

again suggesting that the osmotic driving force due to secretion drives fluid transport.

Later during the slow secretion phase, when some fluid had accumulated in the

apical compartment, the osmotic driving force was reduced and fluid was secreted at

a slower rate.

All of the present data suggest that osmolality determines fluid secretion rate and

thus regulates the ASL volume. In this simple view, osmotic pressure is both the

sensor and driving force for water flux, and luminal dessication caused by inhaled air

and sodium absorption by ENaC ensures that there is never too much ASL.

3.5.5 Carbonic anhydrase participates little in fluid secretion

In the previous chapter, we found that acetazolamide abolishes forskolin-stimulated

anion transport, indicating an important role of carbonic anhydrase in anion

transport. We therefore anticipated that carbonic anhydrase would also participate

actively in fluid secretion during forskolin stimulation. However, acetazolamide

caused only 27% inhibition of fluid transport when added before forskolin or acutely

during forskolin stimulation. We investigate the discrepancy between the impact of

carbonic anhydrase inhibitor on anion transport and fluid secretion in the next

110

chapter.

111

Chapter 4 The role of carbonic anhydrase in anion and fluid secretion

112

In previous chapters we found that pharmacological inhibition of carbonic anhydrase

dramatically reduced transepithelial anion transport but had only a modest effect on

the rate of fluid secretion. Therefore in this chapter we investigate further the role of

carbonic anhydrases in anion and fluid secretion by Calu-3 cells.

113

4.1 Abstract

Carbonic anhydrase is a ubiquitous metalloenzyme which catalyzes the hydration of

carbon dioxide to carbonic acid. The role of this enzyme in bicarbonate production

and transepithelial secretion has been demonstrated in many tissues, including the

airways. However, our results in previous chapters revealed a discrepancy between

the effect of a carbonic anhydrase inhibitor on anion transport and its effect on fluid

secretion, prompting us to study its role further. Without any contribution of NBC to

basolateral bicarbonate entry, forskolin still increased the equivalent short-circuit

current (Ieq) across Calu-3 monolayers although less than when NBC is operating.

Since we showed that bicarbonate is essential for Ieq this suggests that carbonic

anhydrase contributes significantly to transepithelial bicarbonate transport.

Expression of the carbonic anhydrase isozymes CAII, CAIX and CAXII was

demonstrated at the mRNA level by PCR and at the protein level by immunoblotting.

Attempts to co-immunoprecipitate CAII with CFTR were not successful, which argues

against a physical complex between these proteins (i.e. a metabolon), at least under

these conditions. Acetazolamide inhibited only the forskolin stimulated component

of Ieq attributable to bicarbonate, it did not affect the component mediated by

chloride flux through CFTR, therefore it did not act by blocking CFTR channels. CAIX is

often upregulated in cancer cells and was robustly expressed in Calu-3, which is a

tumor cell line, however the relatively CAIX-selective inhibitor

4-(2-aminoethyl)-benzenesulfonamide had no effect on forskolin-stimulated Ieq or

bicarbonate secretion, suggesting CAIX plays little role in anion and fluid secretion.

114

Taken together, the results suggest that bicarbonate production by the CAII isoform

may be important for anion and fluid secretion across Calu-3 cells.

4.2 Introduction

Carbonic anhydrase is a zinc metalloenzyme which catalyzes the reversible hydration

of CO2 to carbonic acid, which then dissociates spontaneously to protons and

bicarbonate ions. Carbonic anhydrase (CA) is expressed ubiquitously in all living

organisms, and 15 distinct isoforms have been identified in humans (Esbaugh and

Tufts 2006). The isozymes can be divided into three groups, the cytoplasmic CAs,

membrane-bound CAs, and the CA-related proteins which have lost their catalytic

activity. CAII is the most prevalent isoform and belongs to the cytoplasmic group. It

has a high enzymatic turnover rate of 106 s-1 and is thought to mediate much of the

CA activity in various tissues including airways (Khalifah 1971, Esbaugh and Tufts

2006, Purkerson and Schwartz 2007). On the other hand, some membrane-bound

CAs become over-expressed and may have roles in pathological states, for example

cancers. Examples of these isoforms include CAIX and CAXII, which are associated

with cancer progression and metastasis (McDonald, Winum et al. 2012).

Previous studies suggested that carbonic anhydrase is important for anion transport

by airway epithelial cells (Welsh and Smith 2001, Cuthbert, Supuran et al. 2003, Yue,

Lau et al. 2008). When CFTR was activated directly or indirectly, 35-50% of the anion

secretion was dependent on carbonic anhydrase based on the inhibition caused by

115

acetazolamide. The contribution of carbonic anhydrase is presumed to be due to its

production of endogenous bicarbonate. The CAII isoform may bind to the

intracellular C-terminal region of various transport proteins through an acidic motif

in its N-terminus, forming a functional complex that facilitates bicarbonate transport

called a metabolon (McMurtrie, Cleary et al. 2004). These partners include AE1, AE2,

NBC1, NBC3 and NHE1, some of which are essential for anion and fluid secretion by

Calu-3 cells as discussed in previous chapters. It should be noted however that the

idea of carbonic anhydrase forming metabolons remains controversial (Piermarini,

Kim et al. 2007).

CAIX and CAXII are over-expressed in tumors due to activation of a hypoxia

responsive element in their promoter, thus they have become important therapeutic

targets for the treatment of cancer (McDonald, Winum et al. 2012). Expression of

cancer-related carbonic anhydrases is restricted in normal tissues, and genetic

disruption of CAIX in mice results in a mild phenotype (Gut, Parkkila et al. 2002). It

seems that CAIX and CAXII play only minor, if any, role in bicarbonate generation in

normal cells when CAII is present. Interestingly, CAIX expression is upregulated

~2-fold in mice that are null for CAII (Pan, Leppilampi et al. 2006), suggesting that

CAIX compensates for the loss of functional CAII in this setting. Calu-3 is a widely

used model for airway epithelial cells and it is also an adenocarcinoma cell line,

therefore CAIX and CAXII could play important roles in the generation of endogenous

bicarbonate and the secretion of anions and/or fluid in this system.

116

CAII is the most widely expressed and usually the most abundant carbonic anhydrase

isoform, however carbonic anhydrases have not been characterized in airway

epithelial cells. In the present study we identified CA isoforms in Calu-3 cells and

studied their potential roles in anion and fluid secretion. We found that carbonic

anhydrase activity was able to partially sustain forskolin-stimulated Ieq

independently of basolateral NBC. Though CAII, CAIX and CAXII were the most

prominent isoforms, the CAIX-selective inhibitor 4-(2-aminoethyl)-

benzenesulfonamide had a negligible impact compared to the broad-spectrum CA

inhibitor acetazolamide. On the other hand, CAXII-specific inhibitor U-104

(1-(4-fluorophenyl)-3-(4-sulfamoylphenyl)urea) partially inhibited the

forskolin-stimulated Ieq and HCO3- secretion. These results suggest that CAII, but not

CAIX, is important for anion and fluid secretion across Calu-3 cells, presumably by

contributing some endogenous bicarbonate.

4.3 Methods

4.3.1 Cell Culture

Two Calu-3 cell lines provided by Dr. Scott O’Grady were used, a parental line with

normal CFTR expression and a CFTR-deficient cell line in which CFTR was knocked

down by stable expression of shRNA that targets CFTR mRNA transcripts. Both cell

lines were cultured in EMEM containing 7% FBS. Some control experiments were also

performed using parental Calu-3 cells (HTB-55, American Type Culture Collection,

117

Manassas, VA) cultured in Eagle’s minimum essential medium (EMEM) containing 15%

fetal bovine serum (FBS) to allow comparison with previous studies.

Cells were seeded at 5 x 105 cells per cm2 on SnapwellsTM (1.12 cm-2; Costar, Corning

Life Sciences, Lowell, MA) when measuring Ieq and HCO3- secretion, or on Transwells

(4.67 cm2, Corning) for experiments involving RT-PCR, immunoblotting,

co-immunoprecipitation and fluid secretion assays. Fresh medium was placed on the

basolateral side one day after plating and the apical medium was removed to

establish air-liquid interface (ALI) conditions. Any fluid that appeared spontaneously

on the apical surface after 3 days was removed. Cultures were maintained in a

humidified, 5% CO2 incubator at 37℃ for 21-25 days.

4.3.2 Solutions

To study the contribution of carbonic anhydrase to forskolin-stimulated Ieq and

HCO3- secretion, monolayers in the Ussing chamber were bathed apically with acidic

solution (pH was clamped at 6.0 by an automated titration workstation; TitraLab 854,

Radiometer) containing (mmol·l-1): 120 NaCl, 5 KCl, 25 NaHCO3, 1.2 MgCl2, 1.2 CaCl2.

The apical solution was constantly gassed with 95% O2/5% CO2. The basolateral side

was bathed with 120 NaCl, 3.3 KH2PO4, 0.8 K2HPO4, 1.2 MgCl2, 1.2 CaCl2, 10 HEPES

and 10 glucose, gassed with 100% O2. The nominally HCO3--free solution blocked

activity of basolateral NBCe1, while the apical solution provided CO2 to support the

activity of intracellular carbonic anhydrase. The pH of the apical solution was

118

clamped at 6 to minimize bicarbonate accumulation and avoid significant buffer

capacity that would preclude pH-stat measurements of HCO3- secretion.

4.3.3 Measurements of equivalent Isc and HCO3- secretion

Inserts were mounted in modified Ussing chambers (Physiologic Inst., San Diego, CA)

at 37℃. HCO3- secretion was measured under open-circuit conditions, and was

compared with the equivalent short-circuit current (Ieq) calculated from Ohm’s law

using the spontaneous transepithelial potential (Vt) and transepithelial resistance

(Rt). Rt was determined from the small deflections in Vt produced by bipolar current

pulses (1 μA, 1 sec duration, 99.9 sec interval) which were delivered by the voltage

clamp (VCC200, Physiologic Instr.). Data were digitized (Powerlab 8/30, AD

Instruments, Montreal QC) and analyzed using Chart5 software.

HCO3- transport was measured using the pH-stat method under open-circuit

conditions. A mini-pH electrode (pHG200-8, Radiometer Analytical) connected to an

automated titration workstation (TitraLab 854, Radiometer) delivered 1 μl aliquots of

10 mmol l-1 HCl to the apical solution to maintain the pH constant at 6.000 ± 0.002.

The amount of acid required for this was used to calculate the rate of HCO3-

secretion. The volume of each half chamber was 4 ml. Solutions containing 25 mmol

l-1 HCO3- were stirred vigorously with 95% O2/5% CO2. Nominally HCO3

--free

solutions were bubbled with 100% O2.

119

4.3.4 RT-PCR

Total RNA was isolated using the RNAqueous-4PCR extraction kit (Ambion) according

to the manufacturer's instructions. Samples were treated with DNase (Ambion) and

reverse transcription was performed using Superscript III (Invitrogen) and 2 μg of

RNA primed with 200 ng random hexamers and 50 μmol l−1 oligo(dT). mRNAs for

each CA isoform were detected as described previously (Tarun, Bryant et al. 2003).

Isoform-specific primers were used and RT-PCR was performed using a PTC-100

Thermal controller (MJ Research) with 30 ng of each primer in 20 μl PCR reaction

buffer. Amplification was carried out using the following thermal cycling conditions:

initial denaturation (2 min at 95°C), 15 cycles of touchdown PCR (95°C for 45 s, 75°C

(–1℃ for each subsequent cycle) for 1 min) and 30 cycles of two-step PCR (95°C for

45 s, 60°C for 1 min). For analysis of the PCR products, 10 µl of the PCR reactions

were loaded onto a 2% TAE (Tris acetate EDTA) agarose gel and subjected to

electrophoresis. After staining with ethidium bromide (0.5 µg/ml), the gel was

photographed. We used GAPDH as an endogenous control to normalize the results.

4.3.5 Immunoblotting

After SDS-PAGE on 12% gels, proteins were transferred to nitrocellulose membranes

and probed using antibodies against: CAII (1:50000, ab6621, Abcam), CAIX (1:2000,

ab107257, Abcam), Na+/K+-ATPase (1:200, mAb gift from R.W. Mercer, Washington

Univ., St. Louis MO; (Takeyasu, Tamkun et al. 1988)), β-actin (1:500, SC-1615, Santa

Cruz). Blots were washed, incubated with secondary antibody conjugated to

120

horseradish peroxidase (1:1000), and visualized with enhanced chemiluminescence

(Amersham Biosciences).

4.3.6 Immunoprecipitation

Calu-3 cell monolayers seeded on Transwells were treated with 10 μM forskolin or

DMSO for 24 h before protein crosslinking by Dithiobis(succinimidyl propionate)

(DSP). After crosslinking, cells were rinsed with PBS and lysed. CFTR was

immunoprecipitated on TrueBlot™ anti-mouse Ig IP beads (00-8811, eBioscience)

using the monoclonal antibody M3A7 (1:1000, gift from J.R. Riordan and T.J. Jensen,

UNC Chapel Hill, NC (Kartner and Riordan 1998)) by incubation for 1 h at 4℃. Beads

were then carefully washed to retrieve samples, whichwere subjected to 8%

SDS-PAGE. After transfer to nitrocellulose membrane, protein complexes were

probed using anti-CAII antibody (1:10000, ab6621, Abcam). Blots were washed,

incubated with secondary antibody conjugated to horseradish peroxidase (1:1000),

and visualized using enhanced chemiluminescence (Amersham Biosciences).

4.3.7 Fluid secretion assay

At the beginning of each assay, any fluid in the apical compartment of the Transwells

was removed. To eliminate the effects of serum and maintain viability during long

fluid secretion experiments, EMEM medium without serum but containing essential

amino acids was placed on the basolateral side. Calu-3 cell monolayers bathed with

basolateral 25 mmol·l−1 HCO3− medium were kept in 5% CO2–95% air, which was

121

nominally saturated with H2O during fluid secretion measurements. Chemicals were

added to the basolateral medium to avoid disrupting the air-liquid interface. After 24

h treatment, fluid secreted by cell monolayers was collected and the volume was

measured.

4.3.8 Data analysis

Isc or Ieq was determined at 100 sec intervals, and HCO3- net flux rate was calculated

every 5 min. Basal values were those obtained immediately before adding forskolin.

Paired or unpaired student’s t-tests with p<0.05 were used for single comparisons.

4.4 Results

4.4.1 Carbonic anhydrase contributes partially to anion transport

To investigate the effects of carbonic anhydrase on anion transport by airway

epithelial cells, we bathed the basolateral side of Calu-3 cell monolayers with

nominally HCO3-- (and CO2-) free solution gassed with 100% O2 to prevent

basolateral NBC from transporting bicarbonate into the cell. However, this should not

prevent the intracellular production of bicarbonate by carbonic anhydrase if CO2 is

supplied apically.

The apical solution was gassed with 95% O2/5% CO2 and clamped at pH 6 so that

pCO2 could be supplied to the cells while bicarbonate secretion is measured using

the pH-stat technique. Under these conditions, the apical solution should still

122

provide CO2 for carbonic anhydrase-catalyzed synthesis of HCO3- but would have

minimal buffer , enabling the measurement of HCO3- secretion by the monolayer.

Basal Ieq and HCO3- secretion under these conditions were similar to those measured

normal solutions (Fig. 4.1, compare with Fig. 2.2B). Applying forskolin again elicited a

transient peak Ieq followed by a plateau (Fig. 4.1), although the peak and plateau

were both reduced somewhat, probably due to the absence of basolateral HCO3- and

thus NBC activity. Based on comparison of the ΔIeq under these conditions with

control ΔIeq, we estimated the contribution of carbonic anhydrase towards

forskolin-stimulated anion current to be 35-40%, within the range reported

previously by other groups (Cuthbert, Supuran et al. 2003, Yue, Lau et al. 2008).

Interestingly, very little acid had to be added to the apical side to maintain pH

constant under these conditions, which implies that little of the forskolin-stimulated

current was carried by HCO3- secretion. The simplest interpretation is that

bicarbonate generated endogenously by carbonic anhydrase was exchanged for Cl- at

the basolateral membrane, and that Ieq was mainly due to net transepithelial Cl-

secretion under these conditions.

4.4.2 The expression of different isoforms of carbonic anhydrase

Two isoforms of carbonic anhydrase, CAIX and CAXII are commonly over-expressed in

cancer cells (McDonald, Winum et al. 2012). As Calu-3 is an adenocarcinoma cell line,

123

Figure 4.1 Carbonic anhydrase contributes partially to anion transport

by Calu-3 cells. Calu-3 cell monolayers were bathed basolaterally with nominally HCO3

-/CO2-free solution gassed with 100% O2. The apical unbuffered solution was gassed with 95% O2/5% CO2 and clamped at pH 6. 10 μM forskolin was added after basal Ieq and HCO3

- became stable (n=4). ●, Ieq; ○, HCO3- secretion.

124

it would not be surprising if it also had up-regulated CAIX and CAXII expression. This

raises the question “Is endogenous bicarbonate synthesis mediated by CAIX and/or

CAXII, in addition to other CA isoforms such as CAII?”

To answer this question, we began by determining the expression of the CA isoform

genes using the reverse transcriptase polymerase chain reaction (RT-PCR). The

RT-PCR revealed that CAII, CAIX and CAXII were highly expressed at the mRNA level

(Fig. 4.2A). Unfortunately, an antibody against CAXII is not available and we could not

confirm its expression at the protein level, however expression of CAII and CAIX

proteins was confirmed by immunoblotting (Fig. 4.2B, C). Both carbonic anhydrases

were exuberantly expressed in wild-type Calu-3 cells and CFTR-deficient cells stably

expressing shRNA against CFTR. Carbonic anhydrase expression was not altered by

raising intracellular cAMP with forskolin.

4.4.3 CAIX is not involved in anion and fluid secretion by Calu-3 cells

Our focus shifted to CAIX because of its up-regulation in cancer cells. We

hypothesized that CAIX might be a source of bicarbonate for chloride exchange and

that the 35-40% contribution carbonic anhydrase contribution to anion transport was

the combined efforts of CAIX and CAII.

To test this, we used 4-(2-aminoethyl)-benzenesulfonamide, a potent inhibitor which

is selective for CAIX over CAII (Svastova, Hulikova et al. 2004) in contrast to the

125

Figure 4.2 The expression of different isoforms of carbonic anhydrases

in Calu-3 cells.

A, mRNA expression of different carbonic anhydrases in Calu-3 cells. Lanes from the left to the right were ladder, CAI, CAII, CAIII, CAIX, CAXa, CAXb, CAXI, CAXII, CAIX, CAXII, CAXIII, CAXIX and GAPDH. The left graph was Calu-3 and the right one was negative control. B, CAII expression in Calu-3 cells. Lane 1 & 2, wild-type Calu-3 cells without forskolin stimulation; Lane 3 & 4, wild-type Calu-3 cells with 24 h forskolin stimulation; Lane 5 & 6, CFTR-deficient Calu-3 cells without forskolin stimulation; Lane 7 & 8, CFTR-deficient Calu-3 cells with 24 h forskolin stimulation. C, CAIX expression in Calu-3 cells. Lane 1 & 2, wild-type Calu-3 cells without forskolin stimulation; Lane 3 & 4, wild-type Calu-3 cells with 24 h forskolin stimulation; Lane 5 & 6, CFTR-deficient Calu-3 cells without forskolin stimulation; Lane 7 & 8, CFTR-deficient Calu-3 cells with 24 h forskolin stimulation.

126

non-specific CA inhibitor acetazolamide. However, during forskolin stimulation,

adding 4-(2-aminoethyl)- benzenesulfonamide at concentrations up to 100 µM did

not alter Ieq or HCO3- secretion (Fig. 4.3). Fluid secretion assays with

4-(2-aminoethyl)-benzenesulfonamide were also consistent with the findings from

Ussing chambers (data not shown). The volume of fluid secreted by Calu-3

monolayers was not affected by the CAIX inhibitor. These results suggest that CAIX

is expressed but plays little if any role in anion and fluid secretion by Calu-3 cells.

4.4.4 CAII and CAXII are responsible for the forskolin-stimulated anion

secretion in Calu-3 cells

As CAIX was excluded, only two candidates were left, one was CAII, the other was

CAXII. I used the CAXII-specific inhibitor U-104 to test if the enzyme was involved in

the forskolin-stimulated response in Calu-3 cells.

As Fig. 4.4 shows, applying U-104 (10 μM, a concentration to inhibit CAXII) after

forskolin partially inhibited the Ieq and HCO3- secretion, indicating CAXII contributed

to the anion transport.

Though I didn’t use any CAII-selective inhibitor to test the involvement of the enzyme,

it is clear that CAII also contributed to the anion transport in Calu-3 cell when

stimulated, because only two candidates were possibly responsible and CAXII

contributed partially, CAII must be the contributor to the remaining Ieq and HCO3-

127

Figure 4.3 CAIX is not involved in anion transport by Calu-3 cells.

The inhibitor of CAIX, 4-(2-aminoethyl)-benzenesulfonamide, did not affect forskolin-stimulated anion secretion by Calu-3 cells (n=3). ● , Ieq; ○ , HCO3

- secretion.

128

Figure 4.4 CAXII contributes partially to the anion transport by Calu-3

cells.

The inhibitor of CAXII U-104 (10 μM) partially inhibited the forskolin-stimulated anion secretion by Calu-3 cells (n=3). ●, Ieq; ○, HCO3

- secretion.

129

secretion.

4.4.5 CAII does not form a complex with CFTR

In the previous two chapters, we discovered a discrepancy between the role of CA in

anion transport and fluid secretion. Inhibiting CA with acetazolamide abolished

forskolin-stimulated Ieq and HCO3- secretion, but in the fluid secretion assay it caused

only ~30% inhibition. Enzyme inhibition by acetazolamide might be expected to

cause slow changes in bicarbonate transport and pH, however its effects were very

rapid, therefore we initially suspected that acetazolamide might abolish Ieq by

blocking CFTR. However the non-HCO3- component of Ieq was not inhibited by

acetazolamide, although it was presumably mediated by CFTR. This suggests that

acetazolamide is not a blocker of the CFTR pore.

Since carbonic anhydrase is thought to form a metabolon with acid/base transporters

which facilitates bicarbonate flux (McMurtrie, Cleary et al. 2004), we wondered if

acetazolamide might act by disrupting a complex between CFTR and CAII, the

predominant carbonic anhydrase in most cells.

To test this, we immunoprecipitated CFTR, separated the proteins by SDS-PAGE, and

probed the resulting blots with antibody against CAII to see if it was co-precipitated.

However no CAII was detected in the CFTR precipitates (Fig. 4.4). This suggests that

CAII and CFTR do not form a stable complex and makes it unlikely that acetazolamide

130

inhibits CFTR function by disrupting its association with CAII.

4.5 Discussion

Results in the current chapter support the involvement of carbonic anhydrase in

anion and fluid secretion by Calu-3 cells, with the former being more strongly

affected.

4.5.1 Carbonate anhydrase II is important for anion and fluid secretion

The role of CAII is suggested by several lines of evidence: 1) acetazolamide-sensitive

carbonic anhydrase maintained basal secretion when the basolateral HCO3- and NBC

was unable to supply HCO3- (Fig. 4.1), 2) CO2 was equally effective in supporting

basal secretion when supplied from the basolateral or apical side, compatible with its

hydration by a soluble carbonic anhydrase that is distributed throughout the cell, and

3) the data suggest that carbonic anhydrase contributes 30-40% of the

forskolin-stimulated anion and fluid secretion by Calu-3 cells through its ability to

generate endogenous bicarbonate.

Three carbonic anhydrases were identified in Calu-3 cells, the ubiquitously expressed

soluble CAII, and the membrane-bound forms CAIX and CAXII. Although all three

could potentially play a role in bicarbonate handling, the insensitivity of HCO3-

secretion to an inhibitor of CAIX suggested it plays little role in anion and fluid

secretion. Only the other hand, CAII and CAXII were showed to contribute to the

131

Figure 4.4 Co-immunoprecipitation between CAII and CFTR.

Wild-type Calu-3 cell protein (400 μg) was immunoprecipitated with CFTR antibody, and then blotted with CAII antibody. Lane 1, sample protein immunoprecipitated with CFTR antibody; Lane 2, sample protein without CFTR antibody; Lane 3, total lysate blotted against CFTR and CAII antibody.

132

anion secretion by Calu-3 cells. However the fact that CAIX and CAXII are normally

expressed at low levels and only become upregulated in adenocarcinomas suggests

that CAII may be the one that supports secretion in normal airway epithelial cells.

4.5.2 Further evidence that HCO3- is required for Cl- secretion

We showed in previous chapters that Cl- is the main anion secreted by Calu-3 cell

monolayers under physiological (open-circuit) conditions. This prominence is most

obvious when the activity of NBC is blocked by removing HCO3- from the basolateral

solution. Without functional NBC so that carbonic anhydrase supplied all the

bicarbonate to the cell, transepithelial bicarbonate secretion was reduced to 30-40%

that measured under normal conditions, and the bicarbonate generated by carbonic

anhydrase was insufficient to drive Cl- and HCO3- secretion via CFTR. Thus it seems

that airway epithelial cells have a mechanisms that preferably exchanges HCO3- for

Cl- at the basolateral membrane, so that Cl- becomes the main anion secreted under

these conditions (Fig. 4.1). The mechanisms were not investigated in detail in this

study, however we speculate that the selectivity of CFTR for Cl- over HCO3- and the

relative affinities of some transporters, e.g. AE2, for these anions may account for

this preferential exchange at the basolateral membrane. Moreover, as we found in

the previous chapters, bicarbonate supports chloride secretion but does not drive

fluid transport significantly in Calu-3 cells. Thus it is that airway epithelial cells would

humidify the airway lumen by secreting Cl- and fluid (rather than HCO3-) and recycle

HCO3- at the basolateral membrane when there is insufficient HCO3

- available to

133

support transepithelial flux of both anions.

4.5.3 The bicarbonate transport metabolon

CAII is thought to form complexes with other transport proteins that mediate

bicarbonate transport. These transporters include AE1, AE2, NBC1, NBC3 and NHE1

(McMurtrie, Cleary et al. 2004). Bicarbonate transport metabolons may exist in

Calu-3 cells since they express AE2 and NBC1, however our experiments did not

provide any evidence for such an association with CFTR.

Calu-3 cells need a supply of HCO3- to maintain HCO3

- secretion and also to support

basolateral Cl- entry by anion exchange during forskolin-stimulated Cl- secretion. CAII

may function in de novo HCO3- synthesis by generating HCO3

- from intracellular CO2.

We speculate that any HCO3- entering on NBCe1 which is not recycled by basolateral

anion exchange for Cl- becomes neutralized by the intracellular acid load that is

generated when HCO3- exits through CFTR. The resulting dehydration would produce

CO2 near CFTR which could then be rehydrated locally by carbonic anhydrase.

Alternatively, since CO2 diffuses rapidly, the basolateral supply of HCO3- that

supports Cl- entry via anion exchange may depend on a basolateral CAII-AE

metabolon that increases the efficiency of basolateral HCO3- extrusion.

We used a membrane-permeant crosslinking reagent in our attempt to capture

proteins complexed with CFTR, however we cannot exclude the possibility that CFTR

134

and CAII were weakly associated and the metabolon was disrupted during the

immunoprecipation, or during long-term experiments such as the fluid secretion

assay, which requires 24 h. Moreover, as shown in Chapter 3, active fluid secretion

was rapid initially but slowed dramatically after the first 6 h. The complex might be

disrupted at the beginning of the assay but remain undetected functionally because

CAII and NBC working independently can compensate during a 24 h assay.

Bicarbonate metabolons might also be inconspicuous if one of the interacting

proteins was inhibited before the supply of HCO3- became rate-limiting. This could

occur when Calu-3 monolayers are pre-treated with acetazolamide; i.e. if inhibiting

CAII beforehand prevented the formation of metabolons, this could potentially

explain why acetazolamide caused less dramatic inhibition of the forskolin-stimulated

secretion than during acute acetazolamide exposure.

4.5.4 The contributions of CAII and NBC to anion and fluid secretion

In our revised model for anion and fluid secretion by airway epithelial cells, CAII and

NBC both supply bicarbonate. When they are combined, some HCO3- is exchanged

for Cl- via basolateral AE and both anions are secreted through CFTR. Basolateral

NKCC also supplies Cl- which becomes predominant during stimulation by

secretagogues that hyperpolarize the cell (Devor, Singh et al. 1999), however during

forskolin (cAMP) stimulation it seems to play a minor role compared with AE2.

135

Chapter 5 General discussion

136

The first clear description of CF appeared in 1938 when Dorothy Anderson published

her paper entitled “Cystic fibrosis of the pancreas and its relation to celiac disease ”

(Lubamba, Dhooghe et al. 2012). Chronic respiratory infection was recognized as one

of the major symptoms, thus antibiotics were introduced for the treatment of CF in

the 1940s. As stated above, salty tasting skin is one of the hallmarks of CF and results

from increased secretion of Cl- and Na+ by the defective sweat glands, therefore the

diagnostic pilocarpine sweat test was brought into use in 1959 (Gibson and Cooke

1959). In the 1980s, the understanding of CF pathophysiology progressed rapidly

with reports of altered Na+ transport across CF respiratory epithelia (Knowles, Gatzy

et al. 1981), and impermeability of the CF sweat gland to Cl- (Quinton 1983, Quinton

and Bijman 1983). Another milestone in the CF field was reached in 1989 when the

CF gene was cloned (Kerem, Rommens et al. 1989, Riordan, Rommens et al. 1989,

Rommens, Iannuzzi et al. 1989). During the following decades, enormous progress

towards understanding of the molecular biology of the CF gene and its protein

product CFTR has firmly established that loss of CFTR function and/or expression is

the basic defect in CF. Nevertheless, the mechanisms of transepithelial ion transport

in the airways and the precise role of CFTR in CF pathogenesis has remained poorly

understood. The major goal of this thesis was to characterize the ion transport and

the related fluid secretion in airway epithelial cells.

In chapter 2, I focused on transepithelial anion transport by the polarized airway

epithelial cell line Calu-3. The airway epithelial monolayer forms a barrier with

137

transporters and channels on both sides (i.e. in the apical and basolateral

membranes). The results are consistent with the basolateral transporters NKCC, NBC

and AE mediating uptake of Cl- and HCO3- and are secreted mainly through apical

CFTR channels. In chapter 3, I investigated fluid secretion by airway epithelial cells,

which is driven by ions, or more accurately, by the osmolality gradient generated by

salt secretion. The results in chapter 3 confirmed the model for anion transport that I

proposed in chapter 2. In chapter 4 carbonic anhydrase was the main focus, and was

pursued due to the dramatic effect of inhibiting this enzyme on bicarbonate

transport and the important role of endogenous bicarbonate generation suggested in

chapter 2 and 3. I attempted to characterize its different isoforms and their

functions.

5.1 A novel model of anion transport of airway

epithelial cells

The results led to a novel model for the transepithelial anion transport in airway

epithelial cells. Before going into details of the model, I should discuss some of the

assumptions and limitations of the experimental design.

5.1.1 V-clamp vs. I-clamp

Short-circuit current (Isc) measurements are the most widely used method for

monitoring transepithelial ion transport (Ussing and Zerahn 1999). In this technique,

the transepithelial voltage is clamped at 0 mV by injecting current in the opposite

138

direction to that generated by the cells (V-clamp). There are no transepithelial

electrical or chemical gradients under classical short-circuit current conditions, thus

Isc provides a rigorous measure of net electrogenic ion transport across the

epithelium under these conditions.

Under I-clamp, voltage is measured and the current is clamped at 0 µA/cm2,

conditions that are referred to as open-circuit. Changes in transepithelial ion

transport under I-clamp are reflected by changes in transepithelial voltage. For

comparison with previous studies, the equivalent short-circuit current (Ieq) was also

calculated from the transepithelial voltage and resistance.

Ieq measures net electrogenic ion transport, but in the presence of a transepithelial

electrochemical gradient and therefore is not a measure of active transport.

Nevertheless, I-clamp has some advantages over V-clamp. I-clamp is more relevant to

physiological transport since it does not perturb the transepithelial and

transmembrane electrochemical gradients, which are important features of polarized

epithelia. Second, since no current is passed through the epithelium from an external

source, I-clamp does cause membrane dielectric breakdown or other damage to the

monolayer.

I used open-circuit conditions and measured V and R so that Ieq could be calculated.

This is appropriate for studying bicarbonate transport since rigorous short-circuit

139

conditions with symmetrical solutions are not possible under pH-stat conditions

anyway; CO2 and bicarbonate are present only on the basolateral side while the

apical solution must remain unbuffered to detect the secretion of base (see below).

Also, I found that the titration workstation worked best under open-circuit conditions,

presumably because no current flowed through reference electrode during pH

measurements under I-clamp.

5.1.2 pH-stat conditions

To directly measure HCO3- secretion, I used the pH-stat technique. A major

advantage of this method is that it provides an accurate measure of net bicarbonate

flux. One limitation is the possibility that the HCO3- gradient from the basolateral to

the apical side may lead to an artificially elevated rate of HCO3- secretion and

perhaps other artifacts caused by an electrochemical gradient for HCO3- which is

essentially infinite. For example the large outward gradient for HCO3- at the apical

membrane under these conditions might favor HCO3- apical exit via CFTR channels

rather than transporters or exchangers, although I obtained no evidence for this.

5.1.3 The Calu-3 cell line

In all the studies described in this thesis, I used the Calu-3 cell line as a model for

airway epithelial cells. Calu-3 is an adenocarcinoma cell line which was developed by

Jorgen Fogh (Fogh, Fogh et al. 1977). When Calu-3 cells were cultured on porous

supports at the air-liquid interface they formed tight junctions, had transepithelial

140

resistance >100 Ω·cm2, and generated short-circuit current (Isc) that was stimulated

by isoproterenol (Shen, Finkbeiner et al. 1994). Isc was also increased by

calcium-mediated secretagogues (bradykinin, histamine methacholine, Ca2+

ionophores). Those responses were abolished when the basolateral membrane was

permeabilized using the pore-forming antifungal agent nystatin, indicating that the

monolayers were polarized and that calcium mobilizing agonists acted primarily at

the basolateral membrane, presumably by activating K+ channels (Shen, Finkbeiner et

al. 1994).

Calu-3 cells express high levels of CFTR (Haws, Finkbeiner et al. 1994) and also

produce the serous cell markers lysozyme (Duszyk 2001) and lactoferrin (Dubin,

Robinson et al. 2004), however they lack the dense granules which characterize

serous cells in vivo. Moreover, secretagogues that normally stimulate lysozyme and

lactoferrin release by native glands (e.g methacholine, adrenaline, neutrophil

elastase, lipopolysaccharide) are not effective on Calu-3 monolayers (Dubin,

Robinson et al. 2004). Like most epithelial cell lines, Calu-3 cells also secrete some

mucus despite being a serous cell model. About 25% of the cells in a typical air-liquid

interface culture develop translucent mucin granules of 1–2 μm diameter (Shen,

Finkbeiner et al. 1994, Kreda, Okada et al. 2007) which contain the mucins MUC5AC

and MUC5B like those in submucosal glands (Kreda, Okada et al. 2007). Secretion of

antimicrobial factors and mucins suggests a mixed serous-mucous phenotype,

however this is not necessarily abnormal as serous cells in submucosal glands also

141

produce some mucin. Indeed, about one-third of the serous tubules in native glands

also express significant amounts of MUC7, which is colocalized with lysozyme

(Sharma, Dudus et al. 1998).

Another feature of Calu-3 cells which sometimes causes concern is their abnormal

karyotype. According to American Type Culture Collection (Manassas, VA, USA),

Calu-3 cells lack normal chromosomes 1, 13, 15 and 17, and are triploid for some

other (unidentified) chromosomes. Genes that are relevant for anion transport and

host defense would normally be situated on those missing chromosomes. For

example PDZD1/CAP70/NHERF3 is on chromosome 1 at 1q21, and the NADPH

oxidases DUOX1 and DUOX2 are on chromosome 15 at 15q21. However, it is clear

that at least some of the genes have been translocated to other chromosomes and

are still expressed. The loss of chromosome 17 should eliminate sodium-hydrogen

exchanger regulator factor 1 (NHERF1) at 17q25.1, yet NHERF1 is readily

immunoprecipitated from Calu-3 cells (Liedtke, Raghuram et al. 2004) and its role in

Calu-3 cells has been demonstrated using RNA interference. Thus although Calu-3

cells have chromosomal abnormalities, their consequences are less severe than

expected. In summary, the Calu-3 cell line is an imperfect, but still useful model for

airway epithelia.

5.1.4 Modifications to existing models

This section considers how existing models for Calu-3 transport might be revised to

142

accommodate the results from this study, which were obtained under conditions that

have not been used previously to study Calu-3 (i.e. open circuit, pH-stat).

The [Cl−] of fluid secreted by Calu-3 cells was ∼120 mmol·l−1, much higher than the

[HCO3−]. This implies that Calu-3 cells secrete mostly Cl− rather than HCO3

− during

forskolin stimulation despite the fact that active Cl− transport was not detected in a

previous study (Devor, Singh et al. 1999). Another interesting finding in this thesis is

that most of the Cl− transport was independent of NKCC1. This conclusion was based

on both anion selectivity and pharmacological data. Robust fluid secretion was

observed for 24 h when the basolateral side was bathed with nominally Cl−-free

solution containing NO3−, which does not bind to the highly selective anion site on

NKCC transporters and therefore is not transported (Kinne, Kinne-Saffran et al. 1986).

Further evidence for NKCC1-independent Cl− entry is the relative insensitivity of fluid

secretion to bumetanide. These results suggest that other basolateral Cl− entry

pathways such as anion exchangers are more important. The fact that NO3− is carried

by many anion exchangers including AE2 (Humphreys, Jiang et al. 1994) led to the

investigation of these exchangers in terms of their contribution to Cl- transport.

Basolateral anion exchange was demonstrated by permeabilizing the apical

membrane, imposing an apical-to-basolateral HCO3− gradient, and examining the

effect of anions in the basolateral solution (i.e. on the “trans” side) on the HCO3− flux

as measured by basolateral pH-stat. Our data strongly suggest that some HCO3− that

143

enters the cell via NBCe1 returns to the basolateral side by exchanging for Cl− (or

NO3−). The anion exchangers that mediate this basolateral Cl− loading were not

identified at the molecular level in the present study; however, AE2 is the most likely

candidate since it is expressed at the basolateral membrane of Calu-3 cells (Loffing,

Moyer et al. 2000). Recent studies of an AE2-deficient Calu-3 cell line generated in

this lab indicate that AE2 mediates essentially all Cl−/HCO3− exchange at the

basolateral membrane and plays an important role in fluid secretion (Huang, Shan et

al. 2012).

A recent study suggested that HCO3- is secreted apically via pendrin but not CFTR

(Garnett, Hickman et al. 2011). The results in this thesis provide no evidence for

significant bicarbonate transport via pendrin when functional CFTR is present.

However, as discussed above, our results do not preclude a role of pendrin under

other experimental conditions since the large bicarbonate gradient used here favors

bicarbonate secretion and might drive an abnormally large flux through CFTR,

masking the role of pendrin.

5.2 The determinants of fluid secretion

The [Cl-] of secretions was ∼120 mmol·l−1, or about 4-fold higher than the [HCO3−].

This implies that most fluid secretion is driven by transepithelial transport of Cl−

rather than HCO3−. However, HCO3

- plays a critical role in secretion as it is needed by

the basolateral anion exchanger for Cl- loading. The pharmacology results suggest

144

this is the predominant mechanism of Cl- uptake, with NKCC supplying a smaller

fraction of the Cl-.

Our further studies on fluid secretion indicate that osmolality provides the driving

force for fluid secretion as expected and that under physiological conditions, Cl-

secretion followed by passive cation transport establishes the osmotic gradient. Net

water flux can also be achieved artificially by adjusting osmolality with mannitol,

which supports the development of inhaled mannitol as a treatment for increasing

fluid secretion in CF airways.

5.3 The role of carbonic anhydrases

I found three carbonic anhydrase isoforms (II, IX, XII) in Calu-3 cells at the mRNA level,

and the expression of CAII and CAIX was further demonstrated at the protein level.

Expression of these isoforms is not surprising since CAII is ubiquitous while the other

two are cancer-associated, and Calu-3 is an adenocarcinoma cell line.

The fact that the broad inhibitor of carbonic anhydrases, acetazolamide, had a strong

inhibitory effect on anion transport by Calu-3 suggests carbonic anhydrase plays an

important role in this process. The lack of effect of

4-(2-aminoethyl)-benzenesulfonamide, a specific inhibitor of CAIX, suggests that this

isoform is not involved in anion transport.

145

Treating cells with acetazolamide had a strong inhibitory effect whether it was added

before or during forskolin stimulation, however the possibility of a metabolon with

NBC was not explored systematically. If a metabolon does exist in Calu-3 cells, CAII

would be the most likely candidate carbonic anhydrase.

Though CAXII also contributes to the anion secretion in Calu-3 cells, its effects are

probably due to the upregulation of CAXII in adenocarcinomas. In normal tissue

without such an upregulation, CAII may be the major player.

5.4 Conclusions and future direction

In my thesis, I successfully characterized transepithelial anion transport in airway

epithelial cells using the Calu-3 cell line as a model. I propose a revised model for

anion and fluid secretion by Calu-3 which is likely to be relevant to submucosal

glands and distinct from the mechanism present in the surface epithelium. This

model may explain the results in an early study by others showing that exposure to a

high concentration of bumetanide on the basolateral side (100 µM) has almost no

effect on fluid secretion by cat trachea during phenylephrine stimulation of the

submucosal glands (Corrales, Nadel et al. 1984). That convincing study is rarely cited

because the results could not be reconciled with the paradigm that NKCC drives fluid

transport in secretory epithelia.

There are still important questions which need to be addressed: Does pendrin

146

transport HCO3- under physiological and/or pathological conditions? Does the model

I have proposed for Calu-3 cells also apply to primary submucosal gland cultures?

What is the role of HCO3- in airway secretions beside supporting Cl- loading by

basolateral anion exchange? Answering these questions will provide a better

understanding of CF and may lead to better treatment of the disease.

147

References Acevedo, M. and L. W. Steele (1993). "Na(+)-H+ exchanger in isolated epithelial tracheal cells from sheep. Involvement in tracheal proton secretion." Exp Physiol 78(3): 383-394.

Al-Bazzaz, F. J., N. Hafez, S. Tyagi, C. A. Gailey, M. Toofanfard, W. A. Alrefai, T. M. Nazir, K. Ramaswamy and P. K. Dudeja (2001). "Detection of Cl--HCO3- and Na+-H+ exchangers in human airways epithelium." JOP 2(4 Suppl): 285-290.

Alper, S. L., M. N. Chernova and A. K. Stewart (2002). "How pH regulates a pH regulator: a regulatory hot spot in the N-terminal cytoplasmic domain of the AE2 anion exchanger." Cell Biochem Biophys 36(2-3): 123-136.

Alpern, R. J. (1985). "Mechanism of basolateral membrane H+/OH-/HCO-3 transport in the rat proximal convoluted tubule. A sodium-coupled electrogenic process." J Gen Physiol 86(5): 613-636.

Aravind, L. and E. V. Koonin (2000). "The STAS domain - a link between anion transporters and antisigma-factor antagonists." Curr Biol 10(2): R53-55.

Awayda, M. S., M. J. Boudreaux, R. L. Reger and L. L. Hamm (2000). "Regulation of the epithelial Na(+) channel by extracellular acidification." Am J Physiol Cell Physiol 279(6): C1896-1905.

Baird, T. T., Jr., A. Waheed, T. Okuyama, W. S. Sly and C. A. Fierke (1997). "Catalysis and inhibition of human carbonic anhydrase IV." Biochemistry 36(9): 2669-2678.

Ballard, S. T. and S. K. Inglis (2004). "Liquid secretion properties of airway submucosal glands." J Physiol 556(Pt 1): 1-10.

Ballard, S. T., L. Trout, Z. Bebok, E. J. Sorscher and A. Crews (1999). "CFTR involvement in chloride, bicarbonate, and liquid secretion by airway submucosal glands." Am J Physiol 277(4 Pt 1): L694-699.

Ballard, S. T., L. Trout, J. Garrison and S. K. Inglis (2006). "Ionic mechanism of forskolin-induced liquid secretion by porcine bronchi." Am J Physiol Lung Cell Mol Physiol 290(1): L97-104.

Bardou, O., N. T. Trinh and E. Brochiero (2009). "Molecular diversity and function of K+ channels in airway and alveolar epithelial cells." Am J Physiol Lung Cell Mol Physiol 296(2): L145-155.

148

Barker, P. M., M. S. Nguyen, J. T. Gatzy, B. Grubb, H. Norman, E. Hummler, B. Rossier, R. C. Boucher and B. Koller (1998). "Role of gammaENaC subunit in lung liquid clearance and electrolyte balance in newborn mice. Insights into perinatal adaptation and pseudohypoaldosteronism." J Clin Invest 102(8): 1634-1640.

Basbaum, C. B., B. Jany and W. E. Finkbeiner (1990). "The serous cell." Annu Rev Physiol 52: 97-113.

Bernardo, A. A., C. M. Bernardo, D. J. Espiritu and J. A. Arruda (2006). "The sodium bicarbonate cotransporter: structure, function, and regulation." Semin Nephrol 26(5): 352-360.

Bertrand, C. A., R. Zhang, J. M. Pilewski and R. A. Frizzell (2009). "SLC26A9 is a constitutively active, CFTR-regulated anion conductance in human bronchial epithelia." J Gen Physiol 133(4): 421-438.

Boers, J. E., A. W. Ambergen and F. B. Thunnissen (1998). "Number and proliferation of basal and parabasal cells in normal human airway epithelium." Am J Respir Crit Care Med 157(6 Pt 1): 2000-2006.

Boitano, S., Z. Safdar, D. G. Welsh, J. Bhattacharya and M. Koval (2004). "Cell-cell interactions in regulating lung function." Am J Physiol Lung Cell Mol Physiol 287(3): L455-459.

Bonar, P. T. and J. R. Casey (2008). "Plasma membrane Cl(-)/HCO(3)(-) exchangers: structure, mechanism and physiology." Channels (Austin) 2(5): 337-345.

Boucher, R. C. (2003). "Regulation of airway surface liquid volume by human airway epithelia." Pflugers Arch 445(4): 495-498.

Boucher, R. C. (2007). "Airway surface dehydration in cystic fibrosis: pathogenesis and therapy." Annu Rev Med 58: 157-170.

Boucher, R. C. (2007). "Evidence for airway surface dehydration as the initiating event in CF airway disease." J Intern Med 261(1): 5-16.

Bowman, E. J., L. A. Graham, T. H. Stevens and B. J. Bowman (2004). "The bafilomycin/concanamycin binding site in subunit c of the V-ATPases from Neurospora crassa and Saccharomyces cerevisiae." J Biol Chem 279(32): 33131-33138.

Burch, L. H., C. R. Talbot, M. R. Knowles, C. M. Canessa, B. C. Rossier and R. C. Boucher (1995). "Relative expression of the human epithelial Na+ channel subunits in

149

normal and cystic fibrosis airways." Am J Physiol 269(2 Pt 1): C511-518.

Caputo, A., E. Caci, L. Ferrera, N. Pedemonte, C. Barsanti, E. Sondo, U. Pfeffer, R. Ravazzolo, O. Zegarra-Moran and L. J. Galietta (2008). "TMEM16A, a membrane protein associated with calcium-dependent chloride channel activity." Science 322(5901): 590-594.

Case, R. M., M. Hunter, I. Novak and J. A. Young (1984). "The anionic basis of fluid secretion by the rabbit mandibular salivary gland." J Physiol 349: 619-630.

Chambers, L. A., B. M. Rollins and R. Tarran (2007). "Liquid movement across the surface epithelium of large airways." Respir Physiol Neurobiol 159(3): 256-270.

Chen, Y., M. J. Cann, T. N. Litvin, V. Iourgenko, M. L. Sinclair, L. R. Levin and J. Buck (2000). "Soluble adenylyl cyclase as an evolutionarily conserved bicarbonate sensor." Science 289(5479): 625-628.

Cherny, V. V. and T. E. DeCoursey (1999). "pH-dependent inhibition of voltage-gated H(+) currents in rat alveolar epithelial cells by Zn(2+) and other divalent cations." J Gen Physiol 114(6): 819-838.

Cherny, V. V., V. S. Markin and T. E. DeCoursey (1995). "The voltage-activated hydrogen ion conductance in rat alveolar epithelial cells is determined by the pH gradient." J Gen Physiol 105(6): 861-896.

Choi, J. Y., N. S. Joo, M. E. Krouse, J. V. Wu, R. C. Robbins, J. P. Ianowski, J. W. Hanrahan and J. J. Wine (2007). "Synergistic airway gland mucus secretion in response to vasoactive intestinal peptide and carbachol is lost in cystic fibrosis." J Clin Invest 117(10): 3118-3127.

Coakley, R. D., B. R. Grubb, A. M. Paradiso, J. T. Gatzy, L. G. Johnson, S. M. Kreda, W. K. O'Neal and R. C. Boucher (2003). "Abnormal surface liquid pH regulation by cultured cystic fibrosis bronchial epithelium." Proc Natl Acad Sci U S A 100(26): 16083-16088.

Cobb, B. R., F. Ruiz, C. M. King, J. Fortenberry, H. Greer, T. Kovacs, E. J. Sorscher and J. P. Clancy (2002). "A(2) adenosine receptors regulate CFTR through PKA and PLA(2)." Am J Physiol Lung Cell Mol Physiol 282(1): L12-25.

Communi, D., P. Paindavoine, G. A. Place, M. Parmentier and J. M. Boeynaems (1999). "Expression of P2Y receptors in cell lines derived from the human lung." Br J Pharmacol 127(2): 562-568.

Corrales, R. J., J. A. Nadel and J. H. Widdicombe (1984). "Source of the fluid

150

component of secretions from tracheal submucosal glands in cats." J Appl Physiol Respir Environ Exerc Physiol 56(4): 1076-1082.

Counillon, L., W. Scholz, H. J. Lang and J. Pouyssegur (1993). "Pharmacological characterization of stably transfected Na+/H+ antiporter isoforms using amiloride analogs and a new inhibitor exhibiting anti-ischemic properties." Mol Pharmacol 44(5): 1041-1045.

Crump, R. G., G. R. Askew, S. E. Wert, J. B. Lingrel and C. H. Joiner (1995). "In situ localization of sodium-potassium ATPase mRNA in developing mouse lung epithelium." Am J Physiol 269(3 Pt 1): L299-308.

Cuthbert, A. W. and L. J. MacVinish (2003). "Mechanisms of anion secretion in Calu-3 human airway epithelial cells by 7,8-benzoquinoline." Br J Pharmacol 140(1): 81-90.

Cuthbert, A. W., C. T. Supuran and L. J. MacVinish (2003). "Bicarbonate-dependent chloride secretion in Calu-3 epithelia in response to 7,8-benzoquinoline." J Physiol 551(Pt 1): 79-92.

DeCoursey, T. E. (1991). "Hydrogen ion currents in rat alveolar epithelial cells." Biophys J 60(5): 1243-1253.

DeCoursey, T. E. and V. V. Cherny (1995). "Voltage-activated proton currents in membrane patches of rat alveolar epithelial cells." J Physiol 489 ( Pt 2): 299-307.

Devor, D. C., R. J. Bridges and J. M. Pilewski (2000). "Pharmacological modulation of ion transport across wild-type and DeltaF508 CFTR-expressing human bronchial epithelia." Am J Physiol Cell Physiol 279(2): C461-479.

Devor, D. C., A. K. Singh, L. C. Lambert, A. DeLuca, R. A. Frizzell and R. J. Bridges (1999). "Bicarbonate and chloride secretion in Calu-3 human airway epithelial cells." J Gen Physiol 113(5): 743-760.

Dobbs, L. G. and M. D. Johnson (2007). "Alveolar epithelial transport in the adult lung." Respir Physiol Neurobiol 159(3): 283-300.

Dossena, S., S. Rodighiero, V. Vezzoli, C. Nofziger, E. Salvioni, M. Boccazzi, E. Grabmayer, G. Botta, G. Meyer, L. Fugazzola, P. Beck-Peccoz and M. Paulmichl (2009). "Functional characterization of wild-type and mutated pendrin (SLC26A4), the anion transporter involved in Pendred syndrome." J Mol Endocrinol 43(3): 93-103.

Dossena, S., V. Vezzoli, N. Cerutti, C. Bazzini, M. Tosco, C. Sironi, S. Rodighiero, G. Meyer, U. Fascio, J. Furst, M. Ritter, L. Fugazzola, L. Persani, P. Zorowka, C. Storelli, P.

151

Beck-Peccoz, G. Botta and M. Paulmichl (2006). "Functional characterization of wild-type and a mutated form of SLC26A4 identified in a patient with Pendred syndrome." Cell Physiol Biochem 17(5-6): 245-256.

Dubin, R. F., S. K. Robinson and J. H. Widdicombe (2004). "Secretion of lactoferrin and lysozyme by cultures of human airway epithelium." Am J Physiol Lung Cell Mol Physiol 286(4): L750-755.

Dudeja, P. K., N. Hafez, S. Tyagi, C. A. Gailey, M. Toofanfard, W. A. Alrefai, T. M. Nazir, K. Ramaswamy and F. J. Al-Bazzaz (1999). "Expression of the Na+/H+ and Cl-/HCO-3 exchanger isoforms in proximal and distal human airways." Am J Physiol 276(6 Pt 1): L971-978.

Duszyk, M. (2001). "CFTR and lysozyme secretion in human airway epithelial cells." Pflugers Arch 443 Suppl 1: S45-49.

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." Nat Genet 2(3): 240-248.

Ernst, S. A., J. R. Palacios, 2nd and G. J. Siegel (1986). "Immunocytochemical localization of Na+,K+-ATPase catalytic polypeptide in mouse choroid plexus." J Histochem Cytochem 34(2): 189-195.

Esbaugh, A. J. and B. L. Tufts (2006). "The structure and function of carbonic anhydrase isozymes in the respiratory system of vertebrates." Respir Physiol Neurobiol 154(1-2): 185-198.

Evans, C. M. and J. S. Koo (2009). "Airway mucus: the good, the bad, the sticky." Pharmacol Ther 121(3): 332-348.

Evans, C. M., O. W. Williams, M. J. Tuvim, R. Nigam, G. P. Mixides, M. R. Blackburn, F. J. DeMayo, A. R. Burns, C. Smith, S. D. Reynolds, B. R. Stripp and B. F. Dickey (2004). "Mucin is produced by clara cells in the proximal airways of antigen-challenged mice." Am J Respir Cell Mol Biol 31(4): 382-394.

Evans, M. J., R. A. Cox, S. G. Shami and C. G. Plopper (1990). "Junctional adhesion mechanisms in airway basal cells." Am J Respir Cell Mol Biol 3(4): 341-347.

Evans, M. J., R. A. Cox, S. G. Shami, B. Wilson and C. G. Plopper (1989). "The role of basal cells in attachment of columnar cells to the basal lamina of the trachea." Am J Respir Cell Mol Biol 1(6): 463-469.

152

Evans, M. J. and C. G. Plopper (1988). "The role of basal cells in adhesion of columnar epithelium to airway basement membrane." Am Rev Respir Dis 138(2): 481-483.

Farman, N., C. R. Talbot, R. Boucher, M. Fay, C. Canessa, B. Rossier and J. P. Bonvalet (1997). "Noncoordinated expression of alpha-, beta-, and gamma-subunit mRNAs of epithelial Na+ channel along rat respiratory tract." Am J Physiol 272(1 Pt 1): C131-141.

Fischer, H. and J. H. Widdicombe (2006). "Mechanisms of acid and base secretion by the airway epithelium." J Membr Biol 211(3): 139-150.

Fischer, H., J. H. Widdicombe and B. Illek (2002). "Acid secretion and proton conductance in human airway epithelium." Am J Physiol Cell Physiol 282(4): C736-743.

Fliegel, L. (2005). "The Na+/H+ exchanger isoform 1." Int J Biochem Cell Biol 37(1): 33-37.

Fogh, J., J. M. Fogh and T. Orfeo (1977). "One hundred and twenty-seven cultured human tumor cell lines producing tumors in nude mice." J Natl Cancer Inst 59(1): 221-226.

Gabriel, S. E., M. Makhlina, E. Martsen, E. J. Thomas, M. I. Lethem and R. C. Boucher (2000). "Permeabilization via the P2X7 purinoreceptor reveals the presence of a Ca2+-activated Cl- conductance in the apical membrane of murine tracheal epithelial cells." J Biol Chem 275(45): 35028-35033.

Gamba, G., A. Miyanoshita, M. Lombardi, J. Lytton, W. S. Lee, M. A. Hediger and S. C. Hebert (1994). "Molecular cloning, primary structure, and characterization of two members of the mammalian electroneutral sodium-(potassium)-chloride cotransporter family expressed in kidney." J Biol Chem 269(26): 17713-17722.

Ganz, T. (2002). "Antimicrobial polypeptides in host defense of the respiratory tract." J Clin Invest 109(6): 693-697.

Garcia-Caballero, A., J. E. Rasmussen, E. Gaillard, M. J. Watson, J. C. Olsen, S. H. Donaldson, M. J. Stutts and R. Tarran (2009). "SPLUNC1 regulates airway surface liquid volume by protecting ENaC from proteolytic cleavage." Proc Natl Acad Sci U S A 106(27): 11412-11417.

Garnett, J. P., E. Hickman, R. Burrows, P. Hegyi, L. Tiszlavicz, A. W. Cuthbert, P. Fong and M. A. Gray (2011). "Novel role for pendrin in orchestrating bicarbonate secretion in cystic fibrosis transmembrane conductance regulator (CFTR)-expressing airway

153

serous cells." J Biol Chem 286(47): 41069-41082.

Gibson, L. E. and R. E. Cooke (1959). "A test for concentration of electrolytes in sweat in cystic fibrosis of the pancreas utilizing pilocarpine by iontophoresis." Pediatrics 23(3): 545-549.

Grassl, S. M. and P. S. Aronson (1986). "Na+/Hco3-Co-Transport in Basolateral Membrane-Vesicles Isolated from Rabbit Renal-Cortex." Journal of Biological Chemistry 261(19): 8778-8783.

Gray, M. A., C. E. Pollard, A. Harris, L. Coleman, J. R. Greenwell and B. E. Argent (1990). "Anion selectivity and block of the small-conductance chloride channel on pancreatic duct cells." Am J Physiol 259(5 Pt 1): C752-761.

Gross, E. and I. Kurtz (2002). "Structural determinants and significance of regulation of electrogenic Na(+)-HCO(3)(-) cotransporter stoichiometry." Am J Physiol Renal Physiol 283(5): F876-887.

Gundersen, D., J. Orlowski and E. Rodriguez-Boulan (1991). "Apical polarity of Na,K-ATPase in retinal pigment epithelium is linked to a reversal of the ankyrin-fodrin submembrane cytoskeleton." J Cell Biol 112(5): 863-872.

Gut, M. O., S. Parkkila, Z. Vernerova, E. Rohde, J. Zavada, M. Hocker, J. Pastorek, T. Karttunen, A. Gibadulinova, Z. Zavadova, K. P. Knobeloch, B. Wiedenmann, J. Svoboda, I. Horak and S. Pastorekova (2002). "Gastric hyperplasia in mice with targeted disruption of the carbonic anhydrase gene Car9." Gastroenterology 123(6): 1889-1903.

Haas, M. and B. Forbush, 3rd (2000). "The Na-K-Cl cotransporter of secretory epithelia." Annu Rev Physiol 62: 515-534.

Harhaj, N. S. and D. A. Antonetti (2004). "Regulation of tight junctions and loss of barrier function in pathophysiology." Int J Biochem Cell Biol 36(7): 1206-1237.

Haws, C., W. E. Finkbeiner, J. H. Widdicombe and J. J. Wine (1994). "CFTR in Calu-3 human airway cells: channel properties and role in cAMP-activated Cl- conductance." Am J Physiol 266(5 Pt 1): L502-512.

Hollenhorst, M. I., K. Richter and M. Fronius (2011). "Ion transport by pulmonary epithelia." J Biomed Biotechnol 2011: 174306.

Hong, K. U., S. D. Reynolds, A. Giangreco, C. M. Hurley and B. R. Stripp (2001). "Clara cell secretory protein-expressing cells of the airway neuroepithelial body

154

microenvironment include a label-retaining subset and are critical for epithelial renewal after progenitor cell depletion." Am J Respir Cell Mol Biol 24(6): 671-681.

Hong, K. U., S. D. Reynolds, S. Watkins, E. Fuchs and B. R. Stripp (2004). "Basal cells are a multipotent progenitor capable of renewing the bronchial epithelium." Am J Pathol 164(2): 577-588.

Hong, K. U., S. D. Reynolds, S. Watkins, E. Fuchs and B. R. Stripp (2004). "In vivo differentiation potential of tracheal basal cells: evidence for multipotent and unipotent subpopulations." Am J Physiol Lung Cell Mol Physiol 286(4): L643-649.

Hovenberg, H. W., J. R. Davies and I. Carlstedt (1996). "Different mucins are produced by the surface epithelium and the submucosa in human trachea: identification of MUC5AC as a major mucin from the goblet cells." Biochem J 318 ( Pt 1): 319-324.

Huang, J., J. Shan, D. Kim, J. Liao, A. Evagelidis, S. L. Alper and J. W. Hanrahan (2012). "Basolateral chloride loading by the anion exchanger type 2: role in fluid secretion by the human airway epithelial cell line Calu-3." J Physiol 590(Pt 21): 5299-5316.

Humphreys, B. D., L. Jiang, M. N. Chernova and S. L. Alper (1994). "Functional characterization and regulation by pH of murine AE2 anion exchanger expressed in Xenopus oocytes." Am J Physiol 267(5 Pt 1): C1295-1307.

Huss, M., G. Ingenhorst, S. Konig, M. Gassel, S. Drose, A. Zeeck, K. Altendorf and H. Wieczorek (2002). "Concanamycin A, the specific inhibitor of V-ATPases, binds to the V(o) subunit c." J Biol Chem 277(43): 40544-40548.

Hwang, T. C. and D. N. Sheppard (2009). "Gating of the CFTR Cl- channel by ATP-driven nucleotide-binding domain dimerisation." J Physiol 587(Pt 10): 2151-2161.

Illek, B., J. R. Yankaskas and T. E. Machen (1997). "cAMP and genistein stimulate HCO3- conductance through CFTR in human airway epithelia." Am J Physiol 272(4 Pt 1): L752-761.

Inglis, S. K., S. M. Wilson and R. E. Olver (2003). "Secretion of acid and base equivalents by intact distal airways." Am J Physiol Lung Cell Mol Physiol 284(5): L855-862.

Iovannisci, D., B. Illek and H. Fischer (2010). "Function of the HVCN1 proton channel in airway epithelia and a naturally occurring mutation, M91T." J Gen Physiol 136(1): 35-46.

155

Irokawa, T., M. E. Krouse, N. S. Joo, J. V. Wu and J. J. Wine (2004). "A "virtual gland" method for quantifying epithelial fluid secretion." Am J Physiol Lung Cell Mol Physiol 287(4): L784-793.

Ishiguro, H., M. C. Steward, S. Naruse, S. B. Ko, H. Goto, R. M. Case, T. Kondo and A. Yamamoto (2009). "CFTR functions as a bicarbonate channel in pancreatic duct cells." J Gen Physiol 133(3): 315-326.

Jayaraman, S., N. S. Joo, B. Reitz, J. J. Wine and A. S. Verkman (2001). "Submucosal gland secretions in airways from cystic fibrosis patients have normal [Na(+)] and pH but elevated viscosity." Proc Natl Acad Sci U S A 98(14): 8119-8123.

Jiang, L., M. N. Chernova and S. L. Alper (1997). "Secondary regulatory volume increase conferred on Xenopus oocytes by expression of AE2 anion exchanger." Am J Physiol 272(1 Pt 1): C191-202.

Joo, N. S., T. Irokawa, J. V. Wu, R. C. Robbins, R. I. Whyte and J. J. Wine (2002). "Absent secretion to vasoactive intestinal peptide in cystic fibrosis airway glands." J Biol Chem 277(52): 50710-50715.

Joo, N. S., Y. Saenz, M. E. Krouse and J. J. Wine (2002). "Mucus secretion from single submucosal glands of pig. Stimulation by carbachol and vasoactive intestinal peptide." J Biol Chem 277(31): 28167-28175.

Kaplan, M. R., M. D. Plotkin, W. S. Lee, Z. C. Xu, J. Lytton and S. C. Hebert (1996). "Apical localization of the Na-K-Cl cotransporter, rBSC1, on rat thick ascending limbs." Kidney Int 49(1): 40-47.

Kartner, N. and J. R. Riordan (1998). "Characterization of polyclonal and monoclonal antibodies to cystic fibrosis transmembrane conductance regulator." Methods Enzymol 292: 629-652.

Kashlan, O. B. and T. R. Kleyman (2011). "ENaC structure and function in the wake of a resolved structure of a family member." Am J Physiol Renal Physiol 301(4): F684-696.

Kellenberger, S. and L. Schild (2002). "Epithelial sodium channel/degenerin family of ion channels: a variety of functions for a shared structure." Physiol Rev 82(3): 735-767.

Kelly, M., S. Trudel, F. Brouillard, F. Bouillaud, J. Colas, T. Nguyen-Khoa, M. Ollero, A. Edelman and J. Fritsch (2010). "Cystic fibrosis transmembrane regulator inhibitors CFTR(inh)-172 and GlyH-101 target mitochondrial functions, independently of

156

chloride channel inhibition." J Pharmacol Exp Ther 333(1): 60-69.

Kerem, B., J. M. Rommens, J. A. Buchanan, D. Markiewicz, T. K. Cox, A. Chakravarti, M. Buchwald and L. C. Tsui (1989). "Identification of the cystic fibrosis gene: genetic analysis." Science 245(4922): 1073-1080.

Khalifah, R. G. (1971). "The carbon dioxide hydration activity of carbonic anhydrase. I. Stop-flow kinetic studies on the native human isoenzymes B and C." J Biol Chem 246(8): 2561-2573.

Kim, D. and M. C. Steward (2009). "The role of CFTR in bicarbonate secretion by pancreatic duct and airway epithelia." J Med Invest 56 Suppl: 336-342.

Kinne, R., E. Kinne-Saffran, B. Scholermann and H. Schutz (1986). "The anion specificity of the sodium-potassium-chloride cotransporter in rabbit kidney outer medulla: studies on medullary plasma membranes." Pflugers Arch 407 Suppl 2: S168-173.

Klyce, S. D. and R. K. Wong (1977). "Site and mode of adrenaline action on chloride transport across the rabbit corneal epithelium." J Physiol 266(3): 777-799.

Knight, D. A. and S. T. Holgate (2003). "The airway epithelium: structural and functional properties in health and disease." Respirology 8(4): 432-446.

Knowles, M., J. Gatzy and R. Boucher (1981). "Increased bioelectric potential difference across respiratory epithelia in cystic fibrosis." N Engl J Med 305(25): 1489-1495.

Ko, S. B., N. Shcheynikov, J. Y. Choi, X. Luo, K. Ishibashi, P. J. Thomas, J. Y. Kim, K. H. Kim, M. G. Lee, S. Naruse and S. Muallem (2002). "A molecular mechanism for aberrant CFTR-dependent HCO(3)(-) transport in cystic fibrosis." EMBO J 21(21): 5662-5672.

Ko, S. B., W. Zeng, M. R. Dorwart, X. Luo, K. H. Kim, L. Millen, H. Goto, S. Naruse, A. Soyombo, P. J. Thomas and S. Muallem (2004). "Gating of CFTR by the STAS domain of SLC26 transporters." Nat Cell Biol 6(4): 343-350.

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(6): 659-671.

Koval, M. (2002). "Sharing signals: connecting lung epithelial cells with gap junction channels." Am J Physiol Lung Cell Mol Physiol 283(5): L875-893.

157

Kreda, S. M., S. F. Okada, C. A. van Heusden, W. O'Neal, S. Gabriel, L. Abdullah, C. W. Davis, R. C. Boucher and E. R. Lazarowski (2007). "Coordinated release of nucleotides and mucin from human airway epithelial Calu-3 cells." J Physiol 584(Pt 1): 245-259.

Krouse, M. E., J. F. Talbott, M. M. Lee, N. S. Joo and J. J. Wine (2004). "Acid and base secretion in the Calu-3 model of human serous cells." Am J Physiol Lung Cell Mol Physiol 287(6): L1274-1283.

Kyle, H., J. P. Ward and J. G. Widdicombe (1990). "Control of pH of airway surface liquid of the ferret trachea in vitro." J Appl Physiol (1985) 68(1): 135-140.

Lee, M. C., C. M. Penland, J. H. Widdicombe and J. J. Wine (1998). "Evidence that Calu-3 human airway cells secrete bicarbonate." Am J Physiol 274(3 Pt 1): L450-453.

Lee, M. G., J. Y. Choi, X. Luo, E. Strickland, P. J. Thomas and S. Muallem (1999). "Cystic fibrosis transmembrane conductance regulator regulates luminal Cl-/HCO3- exchange in mouse submandibular and pancreatic ducts." J Biol Chem 274(21): 14670-14677.

Lewis, S. A., N. K. Wills and D. C. Eaton (1978). "Basolateral membrane potential of a tight epithelium: ionic diffusion and electrogenic pumps." J Membr Biol 41(2): 117-148.

Li, M. S., R. G. Holstead, W. Wang and P. Linsdell (2011). "Regulation of CFTR chloride channel macroscopic conductance by extracellular bicarbonate." Am J Physiol Cell Physiol 300(1): C65-74.

Liedtke, C. M., V. Raghuram, C. C. Yun and X. Wang (2004). "Role of a PDZ1 domain of NHERF1 in the binding of airway epithelial RACK1 to NHERF1." Am J Physiol Cell Physiol 286(5): C1037-1044.

Linsdell, P., J. A. Tabcharani, J. M. Rommens, Y. X. Hou, X. B. Chang, L. C. Tsui, J. R. Riordan and J. W. Hanrahan (1997). "Permeability of wild-type and mutant cystic fibrosis transmembrane conductance regulator chloride channels to polyatomic anions." J Gen Physiol 110(4): 355-364.

Loffing, J., B. D. Moyer, D. Reynolds, B. E. Shmukler, S. L. Alper and B. A. Stanton (2000). "Functional and molecular characterization of an anion exchanger in airway serous epithelial cells." Am J Physiol Cell Physiol 279(4): C1016-1023.

Lubamba, B., B. Dhooghe, S. Noel and T. Leal (2012). "Cystic fibrosis: insight into CFTR pathophysiology and pharmacotherapy." Clin Biochem 45(15): 1132-1144.

Lumsden, A. B., A. McLean and D. Lamb (1984). "Goblet and Clara cells of human

158

distal airways: evidence for smoking induced changes in their numbers." Thorax 39(11): 844-849.

Luo, Y., K. McDonald and J. W. Hanrahan (2009). "Trafficking of immature DeltaF508-CFTR to the plasma membrane and its detection by biotinylation." Biochem J 419(1): 211-219, 212 p following 219.

Ma, T., J. R. Thiagarajah, H. Yang, N. D. Sonawane, C. Folli, L. J. Galietta and A. S. Verkman (2002). "Thiazolidinone CFTR inhibitor identified by high-throughput screening blocks cholera toxin-induced intestinal fluid secretion." J Clin Invest 110(11): 1651-1658.

MacVinish, L. J., G. Cope, A. Ropenga and A. W. Cuthbert (2007). "Chloride transporting capability of Calu-3 epithelia following persistent knockdown of the cystic fibrosis transmembrane conductance regulator, CFTR." Br J Pharmacol 150(8): 1055-1065.

Maggi, C. A., A. Giachetti, R. D. Dey and S. I. Said (1995). "Neuropeptides as regulators of airway function: vasoactive intestinal peptide and the tachykinins." Physiol Rev 75(2): 277-322.

Mall, M. A. (2008). "Role of cilia, mucus, and airway surface liquid in mucociliary dysfunction: lessons from mouse models." J Aerosol Med Pulm Drug Deliv 21(1): 13-24.

Masereel, B., L. Pochet and D. Laeckmann (2003). "An overview of inhibitors of Na(+)/H(+) exchanger." Eur J Med Chem 38(6): 547-554.

Matsui, H., B. R. Grubb, R. Tarran, S. H. Randell, J. T. Gatzy, C. W. Davis and R. C. Boucher (1998). "Evidence for periciliary liquid layer depletion, not abnormal ion composition, in the pathogenesis of cystic fibrosis airways disease." Cell 95(7): 1005-1015.

McDonald, F. J., B. Yang, R. F. Hrstka, H. A. Drummond, D. E. Tarr, P. B. McCray, Jr., J. B. Stokes, M. J. Welsh and R. A. Williamson (1999). "Disruption of the beta subunit of the epithelial Na+ channel in mice: hyperkalemia and neonatal death associated with a pseudohypoaldosteronism phenotype." Proc Natl Acad Sci U S A 96(4): 1727-1731.

McDonald, P. C., J. Y. Winum, C. T. Supuran and S. Dedhar (2012). "Recent developments in targeting carbonic anhydrase IX for cancer therapeutics." Oncotarget 3(1): 84-97.

McMurtrie, H. L., H. J. Cleary, B. V. Alvarez, F. B. Loiselle, D. Sterling, P. E. Morgan, D. E.

159

Johnson and J. R. Casey (2004). "The bicarbonate transport metabolon." J Enzyme Inhib Med Chem 19(3): 231-236.

Meyrick, B., J. M. Sturgess and L. Reid (1969). "A reconstruction of the duct system and secretory tubules of the human bronchial submucosal gland." Thorax 24(6): 729-736.

Morth, J. P., B. P. Pedersen, M. J. Buch-Pedersen, J. P. Andersen, B. Vilsen, M. G. Palmgren and P. Nissen (2011). "A structural overview of the plasma membrane Na+,K+-ATPase and H+-ATPase ion pumps." Nat Rev Mol Cell Biol 12(1): 60-70.

Mount, D. B. and M. F. Romero (2004). "The SLC26 gene family of multifunctional anion exchangers." Pflugers Arch 447(5): 710-721.

Muanprasat, C., N. D. Sonawane, D. Salinas, A. Taddei, L. J. Galietta and A. S. Verkman (2004). "Discovery of glycine hydrazide pore-occluding CFTR inhibitors: mechanism, structure-activity analysis, and in vivo efficacy." J Gen Physiol 124(2): 125-137.

Nakao, I., S. Kanaji, S. Ohta, H. Matsushita, K. Arima, N. Yuyama, M. Yamaya, K. Nakayama, H. Kubo, M. Watanabe, H. Sagara, K. Sugiyama, H. Tanaka, S. Toda, H. Hayashi, H. Inoue, T. Hoshino, A. Shiraki, M. Inoue, K. Suzuki, H. Aizawa, S. Okinami, H. Nagai, M. Hasegawa, T. Fukuda, E. D. Green and K. Izuhara (2008). "Identification of pendrin as a common mediator for mucus production in bronchial asthma and chronic obstructive pulmonary disease." J Immunol 180(9): 6262-6269.

Norimatsu, Y., A. Ivetac, C. Alexander, N. O'Donnell, L. Frye, M. S. Sansom and D. C. Dawson (2012). "Locating a plausible binding site for an open-channel blocker, GlyH-101, in the pore of the cystic fibrosis transmembrane conductance regulator." Mol Pharmacol 82(6): 1042-1055.

Novak, I. and J. A. Young (1986). "Two independent anion transport systems in rabbit mandibular salivary glands." Pflugers Arch 407(6): 649-656.

Palmer, M. L., S. Y. Lee, D. Carlson, S. Fahrenkrug and S. M. O'Grady (2006). "Stable knockdown of CFTR establishes a role for the channel in P2Y receptor-stimulated anion secretion." J Cell Physiol 206(3): 759-770.

Pan, P., M. Leppilampi, S. Pastorekova, J. Pastorek, A. Waheed, W. S. Sly and S. Parkkila (2006). "Carbonic anhydrase gene expression in CA II-deficient (Car2-/-) and CA IX-deficient (Car9-/-) mice." J Physiol 571(Pt 2): 319-327.

Paradiso, A. M., R. D. Coakley and R. C. Boucher (2003). "Polarized distribution of HCO3- transport in human normal and cystic fibrosis nasal epithelia." J Physiol 548(Pt

160

1): 203-218.

Park, H. W., J. H. Nam, J. Y. Kim, W. Namkung, J. S. Yoon, J. S. Lee, K. S. Kim, V. Venglovecz, M. A. Gray, K. H. Kim and M. G. Lee (2010). "Dynamic regulation of CFTR bicarbonate permeability by [Cl-]i and its role in pancreatic bicarbonate secretion." Gastroenterology 139(2): 620-631.

Paunescu, T. G., M. Ljubojevic, L. M. Russo, C. Winter, M. M. McLaughlin, C. A. Wagner, S. Breton and D. Brown (2010). "cAMP stimulates apical V-ATPase accumulation, microvillar elongation, and proton extrusion in kidney collecting duct A-intercalated cells." Am J Physiol Renal Physiol 298(3): F643-654.

Pedemonte, N., E. Caci, E. Sondo, A. Caputo, K. Rhoden, U. Pfeffer, M. Di Candia, R. Bandettini, R. Ravazzolo, O. Zegarra-Moran and L. J. Galietta (2007). "Thiocyanate transport in resting and IL-4-stimulated human bronchial epithelial cells: role of pendrin and anion channels." J Immunol 178(8): 5144-5153.

Piermarini, P. M., E. Y. Kim and W. F. Boron (2007). "Evidence against a direct interaction between intracellular carbonic anhydrase II and pure C-terminal domains of SLC4 bicarbonate transporters." J Biol Chem 282(2): 1409-1421.

Pirani, D., L. A. Evans, D. I. Cook and J. A. Young (1987). "Intracellular pH in the rat mandibular salivary gland: the role of Na-H and Cl-HCO3 antiports in secretion." Pflugers Arch 408(2): 178-184.

Poulsen, J. H., H. Fischer, B. Illek and T. E. Machen (1994). "Bicarbonate conductance and pH regulatory capability of cystic fibrosis transmembrane conductance regulator." Proc Natl Acad Sci U S A 91(12): 5340-5344.

Poulsen, J. H. and T. E. Machen (1996). "HCO3-dependent pHi regulation in tracheal epithelial cells." Pflugers Arch 432(3): 546-554.

Purkerson, J. M. and G. J. Schwartz (2007). "The role of carbonic anhydrases in renal physiology." Kidney Int 71(2): 103-115.

Quinton, P. M. (1983). "Chloride impermeability in cystic fibrosis." Nature 301(5899): 421-422.

Quinton, P. M. and J. Bijman (1983). "Higher bioelectric potentials due to decreased chloride absorption in the sweat glands of patients with cystic fibrosis." N Engl J Med 308(20): 1185-1189.

Rajasekaran, S. A. and A. K. Rajasekaran (2009). "Na,K-ATPase and epithelial tight

161

junctions." Front Biosci 14: 2130-2148.

Riordan, J. R. (2008). "CFTR function and prospects for therapy." Annu Rev Biochem 77: 701-726.

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(4922): 1066-1073.

Rogers, A. V., A. Dewar, B. Corrin and P. K. Jeffery (1993). "Identification of serous-like cells in the surface epithelium of human bronchioles." Eur Respir J 6(4): 498-504.

Rommens, J. M., M. C. Iannuzzi, B. Kerem, M. L. Drumm, G. Melmer, M. Dean, R. Rozmahel, J. L. Cole, D. Kennedy, N. Hidaka and et al. (1989). "Identification of the cystic fibrosis gene: chromosome walking and jumping." Science 245(4922): 1059-1065.

Russell, J. M. (2000). "Sodium-potassium-chloride cotransport." Physiol Rev 80(1): 211-276.

Schroeder, B. C., T. Cheng, Y. N. Jan and L. Y. Jan (2008). "Expression cloning of TMEM16A as a calcium-activated chloride channel subunit." Cell 134(6): 1019-1029.

Schwarzer, C., T. E. Machen, B. Illek and H. Fischer (2004). "NADPH oxidase-dependent acid production in airway epithelial cells." J Biol Chem 279(35): 36454-36461.

Scott, D. A. and L. P. Karniski (2000). "Human pendrin expressed in Xenopus laevis oocytes mediates chloride/formate exchange." Am J Physiol Cell Physiol 278(1): C207-211.

Scott, D. A., R. Wang, T. M. Kreman, V. C. Sheffield and L. P. Karniski (1999). "The Pendred syndrome gene encodes a chloride-iodide transport protein." Nat Genet 21(4): 440-443.

Shan, J., J. Huang, J. Liao, R. Robert and J. W. Hanrahan (2011). "Anion secretion by a model epithelium: more lessons from Calu-3." Acta Physiol (Oxf) 202(3): 523-531.

Shan, J., J. Liao, J. Huang, R. Robert, M. L. Palmer, S. C. Fahrenkrug, S. M. O'Grady and J. W. Hanrahan (2012). "Bicarbonate-dependent chloride transport drives fluid secretion by the human airway epithelial cell line Calu-3." J Physiol 590(Pt 21): 5273-5297.

162

Sharma, P., L. Dudus, P. A. Nielsen, H. Clausen, J. R. Yankaskas, M. A. Hollingsworth and J. F. Engelhardt (1998). "MUC5B and MUC7 are differentially expressed in mucous and serous cells of submucosal glands in human bronchial airways." Am J Respir Cell Mol Biol 19(1): 30-37.

Shen, B. Q., W. E. Finkbeiner, J. J. Wine, R. J. Mrsny and J. H. Widdicombe (1994). "Calu-3: a human airway epithelial cell line that shows cAMP-dependent Cl- secretion." Am J Physiol 266(5 Pt 1): L493-501.

Sheppard, D. N. (2004). "CFTR channel pharmacology: novel pore blockers identified by high-throughput screening." J Gen Physiol 124(2): 109-113.

Simpson, J. E., L. R. Gawenis, N. M. Walker, K. T. Boyle and L. L. Clarke (2005). "Chloride conductance of CFTR facilitates basal Cl-/HCO3- exchange in the villous epithelium of intact murine duodenum." Am J Physiol Gastrointest Liver Physiol 288(6): G1241-1251.

Slepkov, E. R., J. K. Rainey, B. D. Sykes and L. Fliegel (2007). "Structural and functional analysis of the Na+/H+ exchanger." Biochem J 401(3): 623-633.

Smith, J. J. and M. J. Welsh (1992). "cAMP stimulates bicarbonate secretion across normal, but not cystic fibrosis airway epithelia." J Clin Invest 89(4): 1148-1153.

Soleimani, M. and P. S. Aronson (1989). "Ionic mechanism of Na+-HCO3- cotransport in rabbit renal basolateral membrane vesicles." J Biol Chem 264(31): 18302-18308.

Song, Y., D. Salinas, D. W. Nielson and A. S. Verkman (2006). "Hyperacidity of secreted fluid from submucosal glands in early cystic fibrosis." Am J Physiol Cell Physiol 290(3): C741-749.

Stewart, A. K., B. E. Shmukler, D. H. Vandorpe, F. Reimold, J. F. Heneghan, M. Nakakuki, A. Akhavein, S. Ko, H. Ishiguro and S. L. Alper (2011). "SLC26 anion exchangers of guinea pig pancreatic duct: molecular cloning and functional characterization." Am J Physiol Cell Physiol 301(2): C289-303.

Svastova, E., A. Hulikova, M. Rafajova, M. Zat'ovicova, A. Gibadulinova, A. Casini, A. Cecchi, A. Scozzafava, C. T. Supuran, J. Pastorek and S. Pastorekova (2004). "Hypoxia activates the capacity of tumor-associated carbonic anhydrase IX to acidify extracellular pH." FEBS Lett 577(3): 439-445.

Takeyasu, K., M. M. Tamkun, K. J. Renaud and D. M. Fambrough (1988). "Ouabain-sensitive (Na+ + K+)-ATPase activity expressed in mouse L cells by transfection with DNA encoding the alpha-subunit of an avian sodium pump." J Biol

163

Chem 263(9): 4347-4354.

Talbot, C. L., D. G. Bosworth, E. L. Briley, D. A. Fenstermacher, R. C. Boucher, S. E. Gabriel and P. M. Barker (1999). "Quantitation and localization of ENaC subunit expression in fetal, newborn, and adult mouse lung." Am J Respir Cell Mol Biol 20(3): 398-406.

Tam, A., S. Wadsworth, D. Dorscheid, S. F. Man and D. D. Sin (2011). "The airway epithelium: more than just a structural barrier." Ther Adv Respir Dis 5(4): 255-273.

Tamada, T., M. J. Hug, R. A. Frizzell and R. J. Bridges (2001). "Microelectrode and impedance analysis of anion secretion in Calu-3 cells." JOP 2(4 Suppl): 219-228.

Tarran, R., B. Button and R. C. Boucher (2006). "Regulation of normal and cystic fibrosis airway surface liquid volume by phasic shear stress." Annu Rev Physiol 68: 543-561.

Tarran, R., B. Button, M. Picher, A. M. Paradiso, C. M. Ribeiro, E. R. Lazarowski, L. Zhang, P. L. Collins, R. J. Pickles, J. J. Fredberg and R. C. Boucher (2005). "Normal and cystic fibrosis airway surface liquid homeostasis. The effects of phasic shear stress and viral infections." J Biol Chem 280(42): 35751-35759.

Tarran, R., B. R. Grubb, J. T. Gatzy, C. W. Davis and R. C. Boucher (2001). "The relative roles of passive surface forces and active ion transport in the modulation of airway surface liquid volume and composition." J Gen Physiol 118(2): 223-236.

Tarran, R., L. Trout, S. H. Donaldson and R. C. Boucher (2006). "Soluble mediators, not cilia, determine airway surface liquid volume in normal and cystic fibrosis superficial airway epithelia." J Gen Physiol 127(5): 591-604.

Tarun, A. S., B. Bryant, W. Zhai, C. Solomon and D. Shusterman (2003). "Gene expression for carbonic anhydrase isoenzymes in human nasal mucosa." Chem Senses 28(7): 621-629.

Tessier, G. J., T. R. Traynor, M. S. Kannan and S. M. O'Grady (1990). "Mechanisms of sodium and chloride transport across equine tracheal epithelium." Am J Physiol 259(6 Pt 1): L459-467.

Thornton, D. J., K. Rousseau and M. A. McGuckin (2008). "Structure and function of the polymeric mucins in airways mucus." Annu Rev Physiol 70: 459-486.

Toczylowska-Maminska, R. and K. Dolowy (2012). "Ion transporting proteins of human bronchial epithelium." J Cell Biochem 113(2): 426-432.

164

Tos, M. (1966). "Development of the tracheal glands in man. Number, density, structure, shape, and distribution of mucous glands elucidated by quantitative studies of whole mounts." Acta Pathol Microbiol Scand 68: Suppl 185:183+.

Turner, R. J. and J. N. George (1988). "Cl(-)-HCO3- exchange is present with Na+-K+-Cl- cotransport in rabbit parotid acinar basolateral membranes." Am J Physiol 254(3 Pt 1): C391-396.

Ussing, H. H. and K. Zerahn (1999). "Active transport of sodium as the source of electric current in the short-circuited isolated frog skin. Reprinted from Acta. Physiol. Scand. 23: 110-127, 1951." J Am Soc Nephrol 10(9): 2056-2065.

Wagner, C. A., K. E. Finberg, S. Breton, V. Marshansky, D. Brown and J. P. Geibel (2004). "Renal vacuolar H+-ATPase." Physiol Rev 84(4): 1263-1314.

Wang, X., C. Lytle and P. M. Quinton (2005). "Predominant constitutive CFTR conductance in small airways." Respir Res 6: 7.

Wangemann, P. (2011). "The role of pendrin in the development of the murine inner ear." Cell Physiol Biochem 28(3): 527-534.

Welsh, M. J. and J. J. Smith (2001). "cAMP stimulation of HCO3- secretion across airway epithelia." JOP 2(4 Suppl): 291-293.

Wickstrom, C., J. R. Davies, G. V. Eriksen, E. C. Veerman and I. Carlstedt (1998). "MUC5B is a major gel-forming, oligomeric mucin from human salivary gland, respiratory tract and endocervix: identification of glycoforms and C-terminal cleavage." Biochem J 334 ( Pt 3): 685-693.

Widdicombe, J. H. (2002). "Regulation of the depth and composition of airway surface liquid." J Anat 201(4): 313-318.

Widdicombe, J. H., M. J. Welsh and W. E. Finkbeiner (1985). "Cystic fibrosis decreases the apical membrane chloride permeability of monolayers cultured from cells of tracheal epithelium." Proc Natl Acad Sci U S A 82(18): 6167-6171.

Wine, J. J. and N. S. Joo (2004). "Submucosal glands and airway defense." Proc Am Thorac Soc 1(1): 47-53.

Wingo, T., C. Tu, P. J. Laipis and D. N. Silverman (2001). "The catalytic properties of human carbonic anhydrase IX." Biochem Biophys Res Commun 288(3): 666-669.

165

Yang, Y. D., H. Cho, J. Y. Koo, M. H. Tak, Y. Cho, W. S. Shim, S. P. Park, J. Lee, B. Lee, B. M. Kim, R. Raouf, Y. K. Shin and U. Oh (2008). "TMEM16A confers receptor-activated calcium-dependent chloride conductance." Nature 455(7217): 1210-1215.

Yue, G. G., C. B. Lau, K. P. Fung, P. C. Leung and W. H. Ko (2008). "Effects of Cordyceps sinensis, Cordyceps militaris and their isolated compounds on ion transport in Calu-3 human airway epithelial cells." J Ethnopharmacol 117(1): 92-101.

Zuo, W. L., J. H. Huang, J. J. Shan, S. Li, P. Y. Wong and W. L. Zhou (2010). "Functional studies of acid transporter in cultured rat epididymal cell." Fertil Steril 93(8): 2744-2749.