• No results found

University of Groningen Developmental and pathological roles of BMP/follistatin-like 1 in the lung Tania, Navessa

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen Developmental and pathological roles of BMP/follistatin-like 1 in the lung Tania, Navessa"

Copied!
25
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

University of Groningen

Developmental and pathological roles of BMP/follistatin-like 1 in the lung

Tania, Navessa

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from

it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date:

2017

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Tania, N. (2017). Developmental and pathological roles of BMP/follistatin-like 1 in the lung. University of

Groningen.

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

1

Regulation of pulmonary inflammation by

mesenchymal cells

Hatem Alkhouri,

Wilfred J. Poppinga, Navessa P. Tania, Alaina Ammit,

Michael Schuliga

(3)

Regulation of pulmonary inflammation by mesenchymal cells

Hatem Alkhouri

1

, Wilfred J. Poppinga

2,3

, Navessa P. Tania

2,3

, Alaina Ammit

1

andMichael

Schuliga

4,5 1

Respiratory Research Group, Faculty of Pharmacy, University of Sydney, Sydney, New

South Wales, Australia

2

Department of Molecular Pharmacology, University of Groningen, Groningen, The

Netherlands

3

Groningen Research Institute of Asthma and COPD (GRIAC), University of Groningen,

Groningen, University Medical Centre Groningen, The Netherlands

4

Deptartment of Pharmacology and Therapeutics, University of Melbourne, Parkville,

Victoria, Australia

5

Lung Health Research Centre, University of Melbourne, Parkville, Victoria, Australia

(4)

6

Abstract

Pulmonary inflammation and tissue remodelling are common elements of chronic

respiratory diseases such as asthma, chronic obstructive pulmonary disease (COPD),

idiopathic pulmonary fibrosis (IPF), and pulmonary hypertension (PH). In disease,

pulmonary mesenchymal cells not only contribute to tissue remodelling, but also

have an important role in pulmonary inflammation. This review will describe the

immunomodulatory functions of pulmonary mesenchymal cells, such as airway smooth

muscle (ASM) cells and lung fibroblasts, in chronic respiratory disease. An important

theme of the review is that pulmonary mesenchymal cells not only respond to

inflammatory mediators, but produce their own mediators, whether pro-inflammatory

or pro-resolving, which influence the quantity and quality of the lung immune response.

The notion that defective pro-inflammatory or pro-resolving signalling in these cells

potentially contributes to disease progression is also discussed. Finally, the concept

of specifically targeting pulmonary mesenchymal cell immunomodulatory function to

improve therapeutic control of chronic respiratory disease is considered.

(5)

1. Introduction

Worldwide, more than 250 million people suffer from a debilitating or lethal chronic

respiratory disease,

1

such as asthma, chronic obstructive pulmonary disease (COPD),

idiopathic pulmonary fibrosis (IPF) or pulmonary hypertension (PH). Asthma,

characterized by airway inflammation, remodelling and hyper-reactivity, is one of the

most prevalent chronic respiratory diseases, causing ~1/4 of a million deaths per year

globally.

1

COPD, comprised of irreversible breakdown of lung tissue (emphysema) and

airway wall remodelling, contributes to ~3 million deaths per year, and is increasing

in incidence.

1,2

IPF, albeit less common than asthma or COPD, is a lethal interstitial

lung disease characterised by a relentlessly progressive and invasive form of lung

parenchymal fibrosis.

3

Secondary PH, a comorbidity caused primarily by hypoxia in lung

disease, features increased pulmonary vascular resistance.

4

There remains no effective

treatment for severeasthma (5-10% of asthmatics), COPD and IPF.

5

The consistent presence of inflammatory cells in the lungs of patients unequivocally

establishes pulmonary inflammation as an important component of chronic respiratory

disease. The lung inflammatory profiles of patients vary depending on the disease and

severity, and change upon exacerbation.

6–8

Airway inflammation in asthma is associated

with an increase in mast cells, eosinophils and CD4

+

T-helper-2 (Th2) lymphocytes.

However, for asthmatics with fixed airway obstruction, the inflammation is more

neutrophilic with greater CD8

+

T-helper-1 (Th1) cell involvement, akin to COPD, which

is also characterised by fixed airway obstruction.

7

Whilst IPF has a predominant Th2

cell profile, the ratio of CD8+ to CD4

+

lymphocytes increases with disease severity.

8

Like

COPD, neutrophils and macrophages are also present in lung tissue of patients with IPF.

In PH, perivascular infiltration of dendritic cells, macrophages, mast cells, T-lymphocytes

(CD4

+

and CD8

+

) and B-lymphocytes occurs.

9

In chronic respiratory disease, infiltrating

inflammatory cells produce an array of inflammatory mediators which act by autocrine

and paracrine mechanisms to not only regulate inflammatory cell function, but also

pulmonary mesenchymal cells in tissue remodelling.

In chronic respiratory disease, there is an important relationship between

inflammation and tissue remodelling. The latter describes the structural changes in

lung tissue which may contribute to respiratory dysfunction. Pulmonary mesenchymal

cells are structural cells with a well-recognised role in tissue remodelling processes

in disease. In asthma and COPD, airway smooth muscle (ASM) cell hyperplasia and

hypertrophy cause ASM enlargement, whereas airway fibroblasts contribute to sub-epithelial fibrosis in the airway wall.

10,11

In IPF, lung fibroblasts lung fibroblasts

1

have an

integral role in the progressive fibrosis which begins in the lung interstitium and invades

alveoli spaces.

12

In PH, pulmonary vascular smooth muscle cells have a prominent

role in the medial enlargement of blood vessels, which in effect reduces lumen size,

increasing vascular resistance.

13

Abnormalities of the extracellular matrix (ECM) are a

key feature of tissue remodelling in lung disease.

14

Mesenchymal cells, by the synthesis

and deposition of collagens I and III and other ECM components (e.g. fibronectin),

expand the volume of the ECM in the sub-epithelial layer of the airway wall, within ASM

(6)

6

bundles or in the lung interstitium.

15

Aside from important biomechanical contributions

in tissue remodelling, pulmonary mesenchymal cells are also potent producers of an

array of inflammatory mediators, including cytokines, chemokines and cell adhesion

molecules (CAMs).

16–19

These inflammatory mediators, as well as the ECM produced by

pulmonary mesenchymal cells, influence the type and quantity of inflammatory cells

that infiltrate airway and lung tissue in chronic respiratory disease. Furthermore, the

potential importance of inflammatory responses regulated by pulmonary mesenchymal

cells in tissue remodelling is becoming increasingly recognised. In this review, the

immunomodulatory functions of pulmonary mesenchymal cells and their potential

roles in the progression of chronic respiratory disease will be described.

2. Immunomodulatory function of pulmonary mesenchymal cells

This section will provide an overview of the types of immunomodulatory functions of

pulmonary mesenchymal cells, as summarized in

Figure 1.

Figure 1. Immunomodulatory functions of pulmonary mesenchymal cells. The solid black

arrows designate the pro-inflammatory mediators produced directly or indirectly by pulmonary

mesenchymal cells which contribute to pulmonary inflammation in disease. The black

hatched arrows represent sources of pro-inflammatory mediators which regulate pulmonary

mesenchymal cell function, including the production and expression of pro-inflammatory

mediators. The types and phenotype of the pulmonary mesenchymal and inflammatory cells

varies for disease and disease severity. Abbreviations are defined in the text.

(7)

2.1. Pro-inflammatory mediators

Pulmonary mesenchymal cells coordinate inflammatory responses by producing pro-inflammatory mediators which lead to inflammatory cell recruitment and activation.

The production of pro-inflammatory mediators by ASM cells, particularly in the context

of asthma, has been extensively studied and the subject of many reviews, including

one recent review.

20

Table 1 provides an overview of the broad range of cytokines,

chemokines and CAMs, which have been shown to be expressed by pulmonary

mesenchymal cells, primarily in in vitro cell culture studies. The ECM produced by

these cells also influences inflammatory cell recruitment. Versican and hyaluronan

for instance are ECM components produced by lung fibroblasts which regulate T-cell

trafficking and functioning in inflamed lung tissue.

21,22

Pro-inflammatory mediator

expression in pulmonary mesenchymal cells is stimulated primarily by cytokines and

growth factors produced by inflammatory cells and the epithelium.

20

The regulation

of immunomodulatory function of these cells also involves pro-resolving mediators

(section 2.2), the innate immune system (section 2.3), the plasminogen activation

system (section 2.4) and the coagulation system (section 2.5).

2.2. Pro-resolving mediators

Pulmonary mesenchymal cells may be targets for or produce mediators which have a

role in resolving inflammation. Most pro-resolving mediators with anti-inflammatory

activity, including the resolvins, protectins and lipoxins, are derived from dietary omega-3

polyunsaturated fatty acids.

53

Administration of pro-resolving lipid mediators, including

protectin D1, resolvin D1 and resolvin E1, are protective in models of lung injury and

disease.

54–57

Endogenous protectin D1 is increased in the airways in response to allergen

challenge, but less so for asthmatics than non-asthmatics.

55,58

Such observations

suggest that dys-regulated production of pro-resolving lipid mediators may contribute

to chronic respiratory disease.

59

Whilst inflammatory cells are a major source of pro-resolving lipid mediators,

55,58,59

pulmonary mesenchymal cells are a target. In cultures

of human lung fibroblasts, resolvin D1 inhibits cigarette smoke extract- and

IL-1β-induced cytokine release

60

and endotoxin-induced COX-2 expression and prostaglandin

E

2

(PGE

2

) production.

61

Current knowledge about the release of pro-resolving mediators

by pulmonary mesenchymal cells is limited, aside from the production of annexin A1,

an anti-inflammatory protein.

24

Annexin A1, like resolvin D1, is a ligand for the lipoxin A

4

receptor, ALX/FPR2. Annexin A1 expression and release is increased in lung fibroblasts

following treatment with glucocorticoids.

24

Furthermore, the silencing of annexin A1

augments TNFα-induced IL-6 release from lung fibroblasts,

24

suggesting that annexin

A1 production may be an important immunomodulatory function of pulmonary

mesenchymal cells.

2.3. Toll like receptors (TLRs)

Toll-like receptors (TLRs) activate the innate immune system in response to infection

and tissue injury. TLR ligands are: (i) derived from pathogens, including bacterial cell-surface lipopolysaccharides (LPS) and the double stranded RNA of viruses; or (ii) formed

(8)

6

Table 1. Immunomodulatory proteins expressed by airway smooth muscle (ASM) cells

and lung fibroblasts.

Type Protein Pulmonary mesenchymal cell Cytokines IL-1 ASM23, lung fibroblasts19

IL-4 Lung fibroblasts24

IL-6 ASM,25 lung fibroblasts24

IL-10 ASM23 IL-11 ASM25 IL-13 Lung fibroblasts26 GM-CSF ASM,27 lung fibroblasts19,28 LIF ASM29 OX40 ligand ASM30

Chemokines CXCL1 (Gro-α) ASM,31 lung fibroblasts32

CXCL5 (ENA-78) Lung fibroblasts32

CXCL6 (GCP-2) ASM23

CXCL8 (IL-8) ASM,33 lung fibroblasts19,26

CXCL9 (MIG) ASM34

CXCL10 (IP-10) ASM,35 lung fibroblasts36

CXCL11(ITAC) ASM34

CXCL12 (SDF-1α) ASM,34 lung fibroblasts19

CCL2 (MCP-1) ASM,37 lung fibroblasts19

CCL3 (MIP-1α) ASM38

CCL5 (RANTES) ASM,39 lung fibroblasts19,26

CCL4 (MIP-1β) ASM23

CCL7 (MCP-3) ASM,38 lung fibroblasts40

CCL8 (MCP-2) ASM37

CCL11 (Eotaxin) ASM,41 lung fibroblasts42

CCL16 (MTN-1) ASM23 CCL17 (TARC) ASM43 CCL19 (MIP-3) ASM44 CCL20 (MIP-3α) ASM45 CX3CL1 (Fractalkine) ASM46 SCF ASM47

CAMs ICAM-1 ASM,48 lung fibroblasts49

VCAM-1 ASM,19,48 lung fibroblasts49

CD40 ASM19,50 CD44 ASM51 CD90 (Thy-1) Lung fibroblasts52

Abbreviations: CD, cluster of differentiation; ENA, epithelial-derived neutrophil activating; GM-CSF, granulocyte macrophage-colony stimulating factor; ICAM, intracellular adhesion molecule;

IL, interleukin; LIF, leukaemia inhibitory factor; SCF, Stem cell factor; VCAM, vascular cell adhesion

molecule.

(9)

endogenously, such as fibrinogen

62

and annexin A2.

63

The binding of TLR ligands to their

receptors leads to the activation of nuclear factor NF-κB and/or interferon regulatory

transcription factor 3/7, which stimulates the gene expression of inflammatory

mediators. The dysregulation of TLR signalling may contribute to the development of

chronic respiratory disease. The activation of TLR4 by fibrinogen cleavage products of

coagulation proteases possibly contributes to asthma pathophysiology.

62

Pulmonary

mesenchymal cells express TLR2,

64

TLR3,

65

TLR4,

66

and TLR9

67

and their activation

stimulates IL-6, IL-8 and eotaxin production.

68,69

ASM cells release the stress-response

protein, annexin A2, which stimulates IL-6 production in both macrophages

70

and ASM

cells

63

via TLR4. In lung fibroblasts, TLR3 activation stimulates the production of RANTES,

IP-10, IL-8, type 1 IFN, TGF-β, IL-4 and IL-13,

26,71

and TLR4 regulates proliferation.

72

2.4. The plasminogen activation system

In interstitial lung tissue, the conversion of plasminogen to plasmin (“activation”), a pro-inflammatory serine protease,

73

contributes to disease.

74

Plasminogen, a plasma protein,

is relevant in lung pathology as vascular leak leads to its extravasation into inflamed

lung tissue. Both lung fibroblasts and ASM cells activate extracellular plasminogen with

subsequent effects on IL-6 and IL-8 production and cell proliferation.

63,75–77

These effects

occur at low µg/mL concentrations of plasminogen, substantially lower than that

detected in plasma. At higher concentrations of plasminogen, increased PGE

2

synthesis

and/or apoptosis are observed.

63,78

For ASM cells, plasminogen activation is mediated

by the urokinase plasminogen activator (uPA), in a manner accelerated by the annexin

A2 hetero-tetramer (AIIt),

63

an extracellular protein complex comprised of annexin A2

and S100A10 (p11). The AIIt also serves as a signal transducer for plasmin in mediating

its pro-inflammatory effects on ASM cells

77

and macrophages.

79

Whilst currently little is

known about the role of annexin A2 in respiratory disease, it is becoming increasingly

recognised for its importance in cancer.

80–84

Both uPA and annexin A2 may be novel drug

targets in the treatment of chronic respiratory disease

74

(section 5.4).

2.5. The coagulation system

The coagulation system also contributes to pulmonary inflammation in disease.

62,85

Like

plasminogen, the inactive zymogens of coagulation proteases enter inflamed lung tissue

as a consequence of vascular leak, in a process accompanied by platelet aggregation

and activation of the coagulation cascade.

86

Through the actions of thrombin, the main

activator of the coagulation system, TLR4-activating fibrinogen cleavage products are

generated. Interestingly, plasmin is also involved in the formation of fibrinogen cleavage

products,

87

suggesting that the convergence of both the coagulation and plasminogen

activation systems may play an important role in pulmonary inflammation in disease.

Thrombin and factor Xa (FXa), another coagulant, also activate PAR receptors, including

those on ASM cells and lung fibroblasts,

88,89

to elicit pro-inflammatory and remodelling

activities.

89–91

The targeting of thrombin or FXa reduces pulmonary inflammation and

tissue remodelling in murine models of lung injury and disease.

89,92,93

(10)

6

3. Pulmonary mesenchymal cells in chronic respiratory disease

3.1. Asthma

In asthma, allergen-induced airway inflammation contributes to airway

hyper-responsiveness (AHR), a process that involves spasmodic ASM contraction. Inflammation

has direct and indirect roles in AHR, causing vascular leakage, mucus hyper-secretion,

epithelial shedding, ASM thickening and sub-epithelial fibrosis.

94

Pro-inflammatory

mediators produced by ASM cells and airway fibroblasts, including IL-8, IP-10,

MIP-1α, RANTES and eotaxin, contribute to the recruitment of mast cells, lymphocytes,

eosinophils and neutrophils in asthma.

34,95–97

ASM abnormalities in asthmatics may

contribute to an increased hyper-secretory phenotype. Cytokine-induced production

of IL-8,

98

IP-10,

34

ITAC,

34

eotaxin

99

and MIP-3α

100

is greater in cultures of ASM cells

obtained from asthmatic donors than non-asthmatics donors. Furthermore, ASM cells

of asthmatics produce relatively more collagen, fibronectin and fibulin-1,

101–103

ECM

proteins which may facilitate inflammatory cell adhesion and activation. Increased

calcium handling, caused by abnormal sarco/endoplasmic reticulum calcium ATPase

(SERCA) pump function and expression, increases eotaxin expression in ASM cells of

asthmatics.

104,105

Furthermore, JNK signalling and STAT-1 activation is diminished,

106,107

whilst p65 NF-κB activation is higher

98,107

in ASM cells of asthmatic than non-asthmatic

donors. Additionally, TNF-α induced p38 mitogen-activated protein kinase (MAPK)

signalling is greater in ASM cells of donors with severe asthma, than other asthma or

control groups.

108

ASM cells of asthmatics lack expression of the full length transcription

factor CCAAT/enhancer binding protein (C/EBPα).

109

As a consequence, C/EBPβ binding

to chemokine promoters increases, causing cytokine hyper-secretion.

98

This may be

due to lack of an important anti-inflammatory protein, MAPK phosphatase 1 (MKP-1),

a critical MAPK deactivator that is explored in greater depth in section 5.3. Reduced

expression of MKP-1 was responsible for over-activation of the p38 MAPK pathway and

corticosteroid insensitivity of alveolar macrophages in severe asthma compared with

non-severe asthma.

110

Although airway fibroblast secretory function is an area of active research,

28

asthma-

associated changes in airway fibroblast inflammatory mediator production is under-explored. Whilst IL-1b-induced GM-CSF and IL-8 production is increased more in the

airway fibroblasts of asthmatics than non-asthmatics,

17

the mechanism behind this

differential cytokine production remains unknown. Interestingly, airway fibroblasts

of asthmatics in culture express lower levels of IL-13 Rα2 (a decoy receptor for IL-13

signalling) at baseline than airway fibroblasts of controls,

111

possibly augmenting IL-13-induced inflammation in asthma.

112

3.2. COPD

COPD, characterized by shortness of breath, cough and mucus hyper-secretion, is

caused primarily by tobacco exposure. Genetics/epigenetics

113,114

and the pulmonary

(11)

to ASM-related pathology, such as AHR, being far more pronounced in asthma. In COPD,

the number of airway fibroblasts with a more contractile phenotype (myofibroblasts)

is greater,

21

likely caused by increases in the expression and activity of rho-associated

coiled-coil protein kinase 1 (ROCK1).

117

Such increases will reduce airway elasticity,

as will versican, the production of which is increased in airway fibroblasts of COPD

patients.

118,119

Airway fibroblasts of COPD patients also express higher levels of IL-6 and

IL-8 than controls.

119

Intriguingly, airway fibroblasts from COPD patients, and not from

control subjects, produce prostacyclins in response to TGF-β.

120

Prostacyclins have anti-inflammatory effects in pulmonary fibrosis and PH.

121

3.3. IPF

Interstitial lung diseases (ILDs) are characterized by an abnormality in the interstitium,

the area in the lung parenchyma between the capillaries and alveolar spaces. In IPF,

a lethal form of ILD, the abnormality is a relentlessly progressive form of fibrosis that

causes irreversible damage of lung structure and function. Whilst an increased number

of inflammatory cells in the lungs of patients with IPF suggests a role of inflammation,

122

an abnormal wound-repair response of epithelial/fibroblast origin is thought to be

an underlying cause.

123

Lung fibroblasts have a pivotal role in IPF, proliferating and

differentiating into collagen producing, contractile myofibroblasts to form fibroblastic

foci. The expression of fibroblast growth factor (FGF9) is increased in IPF fibrotic foci in

situ and lung fibroblasts of IPF patients in vitro in response to TGF-β1.

124

By contrast, IFN-inducible expression of STAT1 and IP-10 is repressed in lung fibroblasts of IPF patients.

125

Human lung fibroblasts from IPF patients show constitutive activation of STAT3,

52

which

mediates oncostatin M induced fibroblast chemotaxis.

126

Oncostatin M is secreted by

inflammatory cells, such as macrophage and dendritic cells upon bacterial infection.

127

Oncostatin M is a potent mediator of pulmonary inflammation,

128

and is involved in the

induction of pulmonary eosinophilia and goblet cell hyperplasia in mice,

129,130

being up-regulated in the lung of IPF patients. Furthermore, in lung fibroblasts, eotaxin expression

is induced by oncostatin M, suggesting that lung fibroblasts play an important role in

oncostatin M-induced inflammation in IPF.

42,131

3.4. Pulmonary hypertension

PH, whether primary or secondary to an accompanying chronic respiratory disease

is characterized by vasoconstriction, in situ thrombosis and pulmonary vascular

remodelling. Hypoxia, chronic inflammation and shear stress contribute to PH

pathology.

132,133

In proximal pulmonary vessels that were previously muscularized,

medial thickening is caused by the hypertrophy, hyperplasia and ECM production of

resident pulmonary vascular smooth muscle cells. In previously non-muscular

pre-capillary arterioles, the pulmonary vascular smooth muscle cells that contribute to

medial thickening are derived from intermediate cells in blood vessels or adventitial

fibroblasts, which differentiate into pulmonary vascular smooth muscle cells. In PH, the

vascular adventitia has an important role in regulating and contributing to perivascular

inflammation.

134

Pulmonary adventitial fibroblasts, through the production and release

(12)

6

of pro-inflammatory mediators, induce the infiltration and activation of monocytes

and macrophages. Epigenetic alterations in pulmonary adventitial fibroblasts from

chronically hypoxic hypertensive calves are linked to a heightened pro-inflammatory

phenotype with the expression of IL-1β, IL-6, MCP-1, CXCL12, RANTES, CCR7, CXCR4,

GM-CSF and VCAM-1 being increased.

19

In severe PH, the epigenetic reprogramming

of human pulmonary adventitial fibroblasts to a pro-inflammatory phenotype is

associated with the decreased expression of miR-124, which regulates Notch1/PTEN/

FOXO3/p21Cip1 and p27Kip1 signalling.

135

Interestingly, aberrant PTEN phosphatase

activity may also contribute to the pro-inflammatory phenotype of pulmonary vascular

smooth muscle cells in PH. PTEN, which inhibits Akt/PI3kinase signalling, regulates

a number of cell processes including inflammation.

136

Selective deletion of the PTEN

gene in pulmonary vascular smooth muscle cells increases macrophage infiltration and

vascular remodelling in a murine model of PH.

137

4. Interactions between pulmonary mesenchymal cells and inflammatory cells

4.1. Mast cell-airway smooth muscle cell interactions

Mast cell-ASM cell interactions have an important role in asthma pathophysiology.

138

The

number of mast cells within the ASM layer of asthmatics is higher than non-asthmatics,

correlating with disease severity.

138–141

In asthma, the mast cells residing in the ASM

layer are predominantly mast cell

TC

,

35

being smaller and less granular

30

compared to

the mast cells found elsewhere in the airway wall, or within the ASM layer of non-asthmatics. Mast cell recruitment to the ASM layer requires mast cell expression of

chemokine receptors and ASM cell production of chemokines. Lung mast cells express

a wide range of chemokine receptors, including CCR3, CXCR1, 2, 3 and 4, with CXCR3

being the most highly expressed on mast cells within the ASM layer in asthma.

23,34

Important chemokines produced by ASM cells involved in mast cell recruitment include:

IL-8 (binds CXCR1);

142

IP-10 (binds CXCR3);

23,34

SDF-1α (binds CXCR4);

143

RANTES (binds

CCR1, 3 and 5);

144

and eotaxin (binds CCR3).

142

These chemokines, in conjunction with

SCF and TGF-β, are involved in the movement of mast cells to the ASM layer.

145

ASM

cells of asthmatics produce higher levels of IP-10 than ASM cells of non-asthmatics

following treatment with Th1 cytokines.

34

Under Th2 inflammatory conditions, mast cell

chemotaxis involves IL-8 and eotaxin.

142

Interestingly, the ASM cells of non-asthmatics

release factor(s) that inhibit mast cell chemotaxis under either Th1 or Th2 inflammatory

conditions.

142

CXCL1 is an inhibitory factor of mast cell migration, and is produced less

in ASM cells of asthmatics than non-asthmatics.

146

Upon ASM-mast cell contact, ASM

cells induce mast cell proliferation and maintain mast cell survival.

147

This interaction

is mediated by membrane-bound SCF expressed on ASM cells and soluble IL-6 and

CADM1 produced by mast cells.

147

In addition, numerous mast cell produced mediators

directly affect ASM cell function, a topic that has been extensively reviewed.

148–150

These mediators cause exaggerated bronchoconstriction and also modulate ASM cell

(13)

4.2. Monocyte/macrophage-fibroblast interactions

In chronic respiratory disease, blood circulating monocytes infiltrate lung tissue to

differentiate into macrophages or dendritic cells. The phagocytic and antigen presenting

functions of these cells are important in innate and adaptive immunity respectively.

Fibroblasts are ideally suited to regulate monocyte trafficking, differentiation and

functioning because of their synthetic capacity and sentinel-like positioning in interstitial

spaces. Monocytes stimulate GM-CSF production by lung fibroblasts in a manner

involving physical contact between the two cell types.

28

Lung fibroblast production of

GM-CSF, which regulates monocyte/macrophage function, is in turn increased by TNF-α

and/or IL-1β, cytokines produced by activated macrophages. In mouse models of PH,

pulmonary adventitial fibroblasts produce soluble mediators, including GM-CSF, which

influence monocyte/macrophage cell adhesion, infiltration and cytokine production.

135

Airway fibroblasts from patients with COPD express higher levels of the integrin

αvβ8, which activates TGF-β, in turn stimulating CCL2 and CCL20 production in airway

fibroblasts by an autocrine manner.

151

In a murine model of COPD, αvβ8-regulated CCL2

and CCL20 production stimulates dendritic cell migration to boost an adaptive immune

response.

151

Alternatively activated (M2) macrophages, induced by Th2 cytokines (e.g.

IL-4 and IL-13), are increasingly being recognised for their role in chronic respiratory

disease. M2 activated macrophages are the pre-dominant macrophage phenotype

present in the lungs of IPF patients

152

and are detected in higher numbers in the lung

of COPD patients who continue smoking than those who stop.

153

The M2 macrophages

have an impaired role in innate immunity, but produce a myriad of pro-inflammatory

and pro-fibrogenic mediators such as TGF-β, IL-13, CCL2, CCL17, CCL18 and CCL22.

152

Alveolar macrophages from non-IPF donors produce more CCL18 when either treated

with Th2 cytokines, co-cultured with lung fibroblasts or exposed to native collagen.

154

The latter effect of collagen occurs in a manner mediated by the β

2

-integrin.

154

As CCL18

stimulates collagen production in lung fibroblasts, an axis between M2 macrophages

and lung fibroblasts, involving a CCL18-driven positive feedback loop, may perpetuate

fibrosis in IPF.

154

5. Novel strategies to target pulmonary mesenchymal cell immunomodulatory

function

5.1. cAMP elevating agents

There is still a need to find new therapies for chronic respiratory diseases for which,

anti-inflammatory glucocorticoids alone are ineffective.

155

Roflumilast, an oral

phosphodiesterase (PDE4) inhibitor, is an anti-inflammatory drug for COPD, but has side

effects including nausea. Interestingly, both PDE4 inhibitors and β

2

-adrenergic receptor

agonists cause a rise in intracellular second messenger cyclic AMP (cAMP), but are used

pharmacologically for different targets, one inflammation, the other bronchoconstriction

(in asthma and COPD). In cultures of normal human lung fibroblasts, roflumilast, and the

β

2

-agonist, indacaterol, act synergistically to attenuate inflammatory cytokine secretion

and differentiation into a pro-fibrotic phenotype.

156

The levels of PGE

2

, an endogenous

lipid mediator that increases cAMP production, are higher in lung

157

and lung fibroblasts

(14)

6

from COPD patients.

158,159

However, in COPD, PGE

2

effects on the cAMP pathway are

reduced due to an increase in PDE4 activity.

157

As increased PDE4 activity also reduces

β

2

-agonist effectiveness, these findings imply the potential benefit of combining PDE4

inhibitors with cyclic AMP elevating agonists. Interestingly, the addition of plasmin(ogen)

to lung fibroblasts from IPF patients overcomes a similar PGE

2

resistance, by rearranging

the intracellular compartmentalization of the cAMP pathway.

160

This rearrangement

is mediated by an increased expression of the A-kinase anchoring protein AKAP9, a

scaffolding protein for protein kinase A (PKA), which amplifies the downstream pathway

of PGE

2

-cAMP-PKA.

160

Thus the anti-inflammatory and anti-fibrotic properties of PGE

2

on pulmonary mesenchymal cells may potentially be restored in chronic respiratory

diseases by approaches that rearrange cAMP compartmentalization.

160

5.2. TGF- β

1

pathways

Aberrant TGF-β

1

signalling, important in regulating pulmonary mesenchymal cells

function, contributes to pulmonary inflammation and remodelling in disease.

151,161,162

Inhibiting specific aspects of TGF-β

1

signalling may be an effective strategy to treat

chronic respiratory disease.

120

The TGF-β

1

superfamily member, activin A, is linked

with the progression of PH, stimulating pulmonary vascular smooth muscle cell

proliferation.

163

Administration of follistatin, an endogenous inhibitor of activin A

attenuates inflammation and remodelling in a murine model of pulmonary fibrosis

164

and asthma.

160

TGF-β

1

-inducible connective tissue growth factor (CTGF) is implicated in

the pathogenesis of IPF. Inhibition of CTGF reduces collagen promoter activity and its

expression in bleomycin-induced mice lung fibroblasts, suggesting CTGF neutralization

may be an option for the treatment of IPF.

165

HS 6-O-sulfotransferases 1 (HS6ST1) is

up-regulated in lung fibroblasts of IPF patients,

166

and its silencing reduces TGF-β

1

activation and subsequent collagen I and α-smooth muscle actin expression. Such data

suggests that HS6ST1 inhibition could potentially reduce TGF-β

1

-mediated lung fibrosis.

Interestingly, the IL-6 antagonist Sant7 attenuates TGF-β

1

-induced proliferation of lung

fibroblasts obtained from ILD patients,

167

suggesting that targeting IL-6 may selectively

block an important TGF-β

1

-mediated fibrotic response. Finally, inhibition of GSK-3, a

mediator of TGF-β

1

-induced pulmonary mesenchymal cell differentiation, may also be a

potential molecular target for chronic lung diseases.

168,169

5.3. MKP-1

In recent years, the important anti-inflammatory role played by the MAPK deactivator

MKP-1 in regulating inflammation in asthma has emerged. Upregulation of MKP-1 is

one of the ways in which common anti-asthma medicines, such as β

2

-agonists and

glucocorticoids, mediate their anti-inflammatory effects.

31,170,171

MKP-1 is a critical

negative feedback controller, limiting the extent and duration of pro-inflammatory

MAPK-driven cellular signalling pathways in pulmonary mesenchymal cells such as

(15)

severe asthma.

108

The concept of enhancing MKP-1 expression and/or activity to

control inflammation is currently under investigation,

174

as is the potential use of p38

MAPK inhibitors to improve corticosteroid-mediated therapeutic control of chronic

respiratory disease, especially in severe asthma.

108

Further studies are warranted.

5.4. Urokinase & Annexin A2

Urokinase and annexin A2 production by pulmonary mesenchymal cells is potentially

important in chronic respiratory disease (section 2.4). Both uPA and annexin A2 are

becoming increasingly recognised as important pathological mediators, particularly

in cancer, and their targeting by either pharmacological or antibody-based therapies

reduces tumour growth and/or metastasis in a number of pre-clinical cancer models.

80–84

Furthermore, uPA inhibitors are well tolerated in humans and have provided promising

results in recent phase I and II trials for cancer.

81

Both uPA and annexin A2 gene-deletion

reduces pulmonary inflammation in various murine models,

77,175,176

and uPA antibodies

reduce inflammation and oedema in a mouse model of acute lung injury.

177

However,

further pre-clinical characterization of these inhibitors as therapy for chronic respiratory

disease is required.

6. Conclusion

In disease, pulmonary mesenchymal cells not only respond to inflammatory mediators,

but also contribute to inflammation by producing chemokines, cytokines, CAMs

and ECM matrix which recruit and activate inflammatory cells. The dysregulation of

pulmonary mesenchymal cell immunomodulatory function is likely to contribute to the

pathogenesis of lung disease. The specific targeting of aberrant immunomodulatory

functioning in pulmonary mesenchymal cells may be a strategy to treat lung disorders

such as severe asthma, COPD, IPF and PH, for which there are no current effective

therapies.

(16)

6

References

1. Bousquet J, Khaltaev N. Global Surveillance, Prevention and Control of Chronic Respiratory Diseases - WHO - OMS -. WHO Press; 2007. http://apps.who.int/bookorders/anglais/detart1. jsp?sesslan=1&codlan=1&codcol=15&codcch=708. Accessed March 29, 2017.

2. Decramer M, Janssens W, Miravitlles M. Chronic obstructive pulmonary disease. Lancet Lond Engl. 2012;379(9823):1341-1351. doi:10.1016/S0140-6736(11)60968-9. 3. Ding Q, Luckhardt T, Hecker L, et al. New insights into the pathogenesis and treatment of idiopathic pulmonary fibrosis. Drugs. 2011;71(8):981-1001. doi:10.2165/11591490-000000000-00000. 4. Shlobin OA, Nathan SD. Pulmonary hypertension secondary to interstitial lung disease. Expert Rev Respir Med. 2011;5(2):179-189. doi:10.1586/ers.11.11. 5. Durham A, Adcock IM, Tliba O. Steroid resistance in severe asthma: current mechanisms and future treatment. Curr Pharm Des. 2011;17(7):674-684.

6. Tetley TD. Inflammatory cells and chronic obstructive pulmonary disease. Curr Drug Targets Inflamm Allergy. 2005;4(6):607-618.

7. Tubby C, Harrison T, Todd I, Fairclough L. Immunological basis of reversible and fixed airways disease. Clin Sci Lond Engl 1979. 2011;121(7):285-296. doi:10.1042/CS20110062.

8. Papiris SA, Kollintza A, Karatza M, et al. CD8+ T lymphocytes in bronchoalveolar lavage in idiopathic pulmonary fibrosis. J Inflamm Lond Engl. 2007;4:14. doi:10.1186/1476-9255-4-14.

9. Savai R, Pullamsetti SS, Kolbe J, et al. Immune and inflammatory cell involvement in the pathology of idiopathic pulmonary arterial hypertension. Am J Respir Crit Care Med. 2012;186(9):897-908. doi:10.1164/rccm.201202-0335OC.

10. Amin K, Lúdvíksdóttir D, Janson C, et al. Inflammation and structural changes in the airways of patients with atopic and nonatopic asthma. BHR Group. Am J Respir Crit Care Med. 2000;162(6):2295-2301. doi:10.1164/ajrccm.162.6.9912001.

11. Gosselink JV, Hayashi S, Elliott WM, et al. Differential Expression of Tissue Repair Genes in the Pathogenesis of Chronic Obstructive Pulmonary Disease. Am J Respir Crit Care Med. 2010;181(12):1329-1335. doi:10.1164/rccm.200812-1902OC.

12. Noble PW, Barkauskas CE, Jiang D. Pulmonary fibrosis: patterns and perpetrators. J Clin Invest. 2012;122(8):2756-2762. doi:10.1172/JCI60323.

13. Hopkins N, McLoughlin P. The structural basis of pulmonary hypertension in chronic lung disease: remodelling, rarefaction or angiogenesis? J Anat. 2002;201(4):335-348.

14. Fixman ED, Stewart A, Martin JG. Basic mechanisms of development of airway structural changes in asthma. Eur Respir J. 2007;29(2):379-389. doi:10.1183/09031936.00053506.

15. Redington AE. Airway fibrosis in asthma: mechanisms, consequences, and potential for therapeutic intervention. Monaldi Arch Chest Dis Arch Monaldi Mal Torace. 2000;55(4):317-323.

16. Koziol-White CJ, Panettieri RA. Airway smooth muscle and immunomodulation in acute exacerbations of airway disease. Immunol Rev. 2011;242(1):178-185. doi:10.1111/j.1600-065X.2011.01022.x. 17. Ward JE, Harris T, Bamford T, et al. Proliferation is not increased in airway myofibroblasts isolated

from asthmatics. Eur Respir J. 2008;32(2):362-371. doi:10.1183/09031936.00119307.

18. Flavell SJ, Hou TZ, Lax S, Filer AD, Salmon M, Buckley CD. Fibroblasts as novel therapeutic targets in chronic inflammation. Br J Pharmacol. 2008;153(Suppl 1):S241-S246. doi:10.1038/sj.bjp.0707487. 19. Li M, Riddle SR, Frid MG, et al. Emergence of fibroblasts with a proinflammatory epigenetically

altered phenotype in severe hypoxic pulmonary hypertension. J Immunol Baltim Md 1950. 2011;187(5):2711-2722. doi:10.4049/jimmunol.1100479.

20. Xia YC, Redhu NS, Moir LM, et al. Pro-inflammatory and immunomodulatory functions of airway smooth muscle: emerging concepts. Pulm Pharmacol Ther. 2013;26(1):64-74. doi:10.1016/j. pupt.2012.05.006.

21. Evanko SP, Potter-Perigo S, Bollyky PL, Nepom GT, Wight TN. Hyaluronan and versican in the control of human T-lymphocyte adhesion and migration. Matrix Biol J Int Soc Matrix Biol. 2012;31(2):90-100. doi:10.1016/j.matbio.2011.10.004.

(17)

24. Jia Y, Morand EF, Song W, Cheng Q, Stewart A, Yang YH. Regulation of lung fibroblast activation by annexin A1. J Cell Physiol. 2013;228(2):476-484. doi:10.1002/jcp.24156.

25. Elias JA, Wu Y, Zheng T, Panettieri R. Cytokine- and virus-stimulated airway smooth muscle cells produce IL-11 and other IL-6-type cytokines. Am J Physiol. 1997;273(3 Pt 1):L648-655.

26. Rudd BD, Burstein E, Duckett CS, Li X, Lukacs NW. Differential role for TLR3 in respiratory syncytial virus-induced chemokine expression. J Virol. 2005;79(6):3350-3357. doi:10.1128/JVI.79.6.3350-3357.2005.

27. Saunders MA, Mitchell JA, Seldon PM, et al. Release of granulocyte-macrophage colony stimulating factor by human cultured airway smooth muscle cells: suppression by dexamethasone. Br J Pharmacol. 1997;120(4):545-546. doi:10.1038/sj.bjp.0700998.

28. Fitzgerald SM, Chi DS, Hall HK, et al. GM-CSF induction in human lung fibroblasts by IL-1beta, TNF-alpha, and macrophage contact. J Interferon Cytokine Res Off J Int Soc Interferon Cytokine Res. 2003;23(2):57-65. doi:10.1089/107999003321455453.

29. Knight DA, Lydell CP, Zhou D, Weir TD, Robert Schellenberg R, Bai TR. Leukemia inhibitory factor (LIF) and LIF receptor in human lung. Distribution and regulation of LIF release. Am J Respir Cell Mol Biol. 1999;20(4):834-841. doi:10.1165/ajrcmb.20.4.3429.

30. Burgess JK, Carlin S, Pack RA, et al. Detection and characterization of OX40 ligand expression in human airway smooth muscle cells: a possible role in asthma? J Allergy Clin Immunol. 2004;113(4):683-689. doi:10.1016/j.jaci.2003.12.311.

31. Issa R, Xie S, Khorasani N, et al. Corticosteroid inhibition of growth-related oncogene protein-alpha via mitogen-activated kinase phosphatase-1 in airway smooth muscle cells. J Immunol Baltim Md 1950. 2007;178(11):7366-7375.

32. Brant KA, Fabisiak JP. Nickel and the Microbial Toxin, MALP-2, Stimulate Proangiogenic Mediators from Human Lung Fibroblasts via a HIF-1α and COX-2–Mediated Pathway. Toxicol Sci. 2009;107(1):227-237. doi:10.1093/toxsci/kfn208.

33. Pang L, Knox AJ. Bradykinin stimulates IL-8 production in cultured human airway smooth muscle cells: role of cyclooxygenase products. J Immunol Baltim Md 1950. 1998;161(5):2509-2515. 34. Brightling CE, Ammit AJ, Kaur D, et al. The CXCL10/CXCR3 axis mediates human lung mast cell

migration to asthmatic airway smooth muscle. Am J Respir Crit Care Med. 2005;171(10):1103-1108. doi:10.1164/rccm.200409-1220OC.

35. Hardaker EL, Bacon AM, Carlson K, et al. Regulation of TNF-alpha- and IFN-gamma-induced CXCL10 expression: participation of the airway smooth muscle in the pulmonary inflammatory response in chronic obstructive pulmonary disease. FASEB J Off Publ Fed Am Soc Exp Biol. 2004;18(1):191-193. doi:10.1096/fj.03-0170fje.

36. Dong S, Zhang X, He Y, et al. Synergy of IL-27 and TNF-α in regulating CXCL10 expression in lung fibroblasts. Am J Respir Cell Mol Biol. 2013;48(4):518-530. doi:10.1165/rcmb.2012-0340OC. 37. Wuyts WA, Vanaudenaerde BM, Dupont LJ, Demedts MG, Verleden GM. Modulation by cAMP of

IL-1beta-induced eotaxin and MCP-1 expression and release in human airway smooth muscle cells. Eur Respir J. 2003;22(2):220-226.

38. Holgate ST, Bodey KS, Janezic A, Frew AJ, Kaplan AP, Teran LM. Release of RANTES, MIP-1 alpha, and MCP-1 into asthmatic airways following endobronchial allergen challenge. Am J Respir Crit Care Med. 1997;156(5):1377-1383. doi:10.1164/ajrccm.156.5.9610064.

39. John M, Hirst SJ, Jose PJ, et al. Human airway smooth muscle cells express and release RANTES in response to T helper 1 cytokines: regulation by T helper 2 cytokines and corticosteroids. J Immunol Baltim Md 1950. 1997;158(4):1841-1847.

40. Choi ES, Jakubzick C, Carpenter KJ, et al. Enhanced monocyte chemoattractant protein-3/CC chemokine ligand-7 in usual interstitial pneumonia. Am J Respir Crit Care Med. 2004;170(5):508-515. doi:10.1164/rccm.200401-002OC.

41. Ghaffar O, Hamid Q, Renzi PM, et al. Constitutive and cytokine-stimulated expression of eotaxin by human airway smooth muscle cells. Am J Respir Crit Care Med. 1999;159(6):1933-1942. doi:10.1164/ ajrccm.159.6.9805039.

42. Fritz DK, Kerr C, Botelho F, Stampfli M, Richards CD. Oncostatin M (OSM) primes IL-13- and IL-4-induced eotaxin responses in fibroblasts: regulation of the type-II IL-4 receptor chains IL-4Ralpha and IL-13Ralpha1. Exp Cell Res. 2009;315(20):3486-3499. doi:10.1016/j.yexcr.2009.09.024.

(18)

6

muscle cells: role of IL-4 receptor genotype. Am J Physiol Lung Cell Mol Physiol. 2003;285(4):L907-914. doi:10.1152/ajplung.00120.2003.

44. Kaur D, Saunders R, Berger P, et al. Airway smooth muscle and mast cell-derived CC chemokine ligand 19 mediate airway smooth muscle migration in asthma. Am J Respir Crit Care Med. 2006;174(11):1179-1188. doi:10.1164/rccm.200603-394OC.

45. Jarai G, Sukkar M, Garrett S, et al. Effects of interleukin-1beta, interleukin-13 and transforming growth factor-beta on gene expression in human airway smooth muscle using gene microarrays. Eur J Pharmacol. 2004;497(3):255-265. doi:10.1016/j.ejphar.2004.06.055.

46. Sukkar MB, Issa R, Xie S, Oltmanns U, Newton R, Chung KF. Fractalkine/CX3CL1 production by human airway smooth muscle cells: induction by IFN-gamma and TNF-alpha and regulation by TGF-beta and corticosteroids. Am J Physiol Lung Cell Mol Physiol. 2004;287(6):L1230-1240. doi:10.1152/ ajplung.00014.2004.

47. Kassel O, Schmidlin F, Duvernelle C, Gasser B, Massard G, Frossard N. Human bronchial smooth muscle cells in culture produce stem cell factor. Eur Respir J. 1999;13(5):951-954.

48. Panettieri RA, Lazaar AL, Puré E, Albelda SM. Activation of cAMP-dependent pathways in human airway smooth muscle cells inhibits TNF-alpha-induced ICAM-1 and VCAM-1 expression and T lymphocyte adhesion. J Immunol Baltim Md 1950. 1995;154(5):2358-2365.

49. Spoelstra FM, Postma DS, Hovenga H, Noordhoek JA, Kauffman HF. Interferon-gamma and interleukin-4 differentially regulate ICAM-1 and VCAM-1 expression on human lung fibroblasts. Eur Respir J. 1999;14(4):759-766.

50. Burgess JK, Blake AE, Boustany S, et al. CD40 and OX40 ligand are increased on stimulated asthmatic airway smooth muscle. J Allergy Clin Immunol. 2005;115(2):302-308. doi:10.1016/j.jaci.2004.11.004. 51. Lazaar AL, Albelda SM, Pilewski JM, Brennan B, Puré E, Panettieri RA. T lymphocytes adhere to

airway smooth muscle cells via integrins and CD44 and induce smooth muscle cell DNA synthesis. J Exp Med. 1994;180(3):807-816.

52. Pechkovsky DV, Prêle CM, Wong J, et al. STAT3-mediated signaling dysregulates lung fibroblast-myofibroblast activation and differentiation in UIP/IPF. Am J Pathol. 2012;180(4):1398-1412. doi:10.1016/j.ajpath.2011.12.022.

53. Weylandt KH, Chiu C-Y, Gomolka B, Waechter SF, Wiedenmann B. Omega-3 fatty acids and their lipid mediators: towards an understanding of resolvin and protectin formation. Prostaglandins Other Lipid Mediat. 2012;97(3-4):73-82. doi:10.1016/j.prostaglandins.2012.01.005.

54. Aoki H, Hisada T, Ishizuka T, et al. Protective effect of resolvin E1 on the development of asthmatic airway inflammation. Biochem Biophys Res Commun. 2010;400(1):128-133. doi:10.1016/j. bbrc.2010.08.025.

55. Levy BD, Kohli P, Gotlinger K, et al. Protectin D1 is generated in asthma and dampens airway inflammation and hyperresponsiveness. J Immunol Baltim Md 1950. 2007;178(1):496-502. 56. Wang Y, Li D, Xu N, et al. Follistatin-like protein 1: a serum biochemical marker reflecting the severity

of joint damage in patients with osteoarthritis. Arthritis Res Ther. 2011;13(6):R193. doi:10.1186/ ar3522.

57. Rogerio AP, Haworth O, Croze R, et al. Resolvin D1 and aspirin-triggered resolvin D1 promote resolution of allergic airways responses. J Immunol Baltim Md 1950. 2012;189(4):1983-1991. doi:10.4049/jimmunol.1101665.

58. Miyata J, Fukunaga K, Iwamoto R, et al. Dysregulated synthesis of protectin D1 in eosinophils from patients with severe asthma. J Allergy Clin Immunol. 2013;131(2):353-360-2. doi:10.1016/j. jaci.2012.07.048.

59. Levy BD. Resolvin D1 and Resolvin E1 Promote the Resolution of Allergic Airway Inflammation via Shared and Distinct Molecular Counter-Regulatory Pathways. Front Immunol. 2012;3:390. doi:10.3389/fimmu.2012.00390.

60. Hsiao H-M, Sapinoro RE, Thatcher TH, et al. A novel anti-inflammatory and pro-resolving role for resolvin D1 in acute cigarette smoke-induced lung inflammation. PloS One. 2013;8(3):e58258. doi:10.1371/journal.pone.0058258.

(19)

through Toll-like receptor 4. Science. 2013;341(6147):792-796. doi:10.1126/science.1240342. 63. Stewart AG, Xia YC, Harris T, Royce S, Hamilton JA, Schuliga M. Plasminogen-stimulated airway

smooth muscle cell proliferation is mediated by urokinase and annexin A2, involving plasmin-activated cell signalling. Br J Pharmacol. 2013;170(7):1421-1435. doi:10.1111/bph.12422.

64. Brant KA, Fabisiak JP. Nickel alterations of TLR2-dependent chemokine profiles in lung fibroblasts are mediated by COX-2. Am J Respir Cell Mol Biol. 2008;38(5):591-599. doi:10.1165/rcmb.2007-0314OC. 65. Sugiura H, Ichikawa T, Koarai A, et al. Activation of Toll-like receptor 3 augments myofibroblast

differentiation. Am J Respir Cell Mol Biol. 2009;40(6):654-662. doi:10.1165/rcmb.2008-0371OC. 66. He Z, Zhu Y, Jiang H. Toll-like receptor 4 mediates lipopolysaccharide-induced collagen secretion

by phosphoinositide3-kinase-Akt pathway in fibroblasts during acute lung injury. J Recept Signal Transduct Res. 2009;29(2):119-125. doi:10.1080/10799890902845690.

67. Meneghin A, Choi ES, Evanoff HL, et al. TLR9 is expressed in idiopathic interstitial pneumonia and its activation promotes in vitro myofibroblast differentiation. Histochem Cell Biol. 2008;130(5):979-992. doi:10.1007/s00418-008-0466-z.

68. Sukkar MB, Xie S, Khorasani NM, et al. Toll-like receptor 2, 3, and 4 expression and function in human airway smooth muscle. J Allergy Clin Immunol. 2006;118(3):641-648. doi:10.1016/j. jaci.2006.05.013.

69. Manetsch M, Seidel P, Heintz U, et al. TLR2 ligand engagement upregulates airway smooth muscle TNFα-induced cytokine production. Am J Physiol Lung Cell Mol Physiol. 2012;302(9):L838-845. doi:10.1152/ajplung.00317.2011.

70. Swisher JFA, Burton N, Bacot SM, Vogel SN, Feldman GM. Annexin A2 tetramer activates human and murine macrophages through TLR4. Blood. 2010;115(3):549-558. doi:10.1182/ blood-2009-06-226944.

71. O’Dwyer DN, Armstrong ME, Trujillo G, et al. The Toll-like receptor 3 L412F polymorphism and disease progression in idiopathic pulmonary fibrosis. Am J Respir Crit Care Med. 2013;188(12):1442-1450. doi:10.1164/rccm.201304-0760OC.

72. He Z, Gao Y, Deng Y, et al. Lipopolysaccharide induces lung fibroblast proliferation through Toll-like receptor 4 signaling and the phosphoinositide3-kinase-Akt pathway. PloS One. 2012;7(4):e35926. doi:10.1371/journal.pone.0035926.

73. Syrovets T, Lunov O, Simmet T. Plasmin as a proinflammatory cell activator. J Leukoc Biol. 2012;92(3):509-519. doi:10.1189/jlb.0212056.

74. Schuliga M, Westall G, Xia Y, Stewart AG. The plasminogen activation system: new targets in lung inflammation and remodeling. Curr Opin Pharmacol. 2013;13(3):386-393. doi:10.1016/j. coph.2013.05.014.

75. Pendurthi UR, Ngyuen M, Andrade-Gordon P, Petersen LC, Rao LVM. Plasmin induces Cyr61 gene expression in fibroblasts via protease-activated receptor-1 and p44/42 mitogen-activated protein kinase-dependent signaling pathway. Arterioscler Thromb Vasc Biol. 2002;22(9):1421-1426. 76. Schuliga M, Harris T, Stewart AG. Plasminogen activation by airway smooth muscle is regulated by

type I collagen. Am J Respir Cell Mol Biol. 2011;44(6):831-839. doi:10.1165/rcmb.2009-0469OC. 77. Schuliga M, Langenbach S, Xia YC, et al. Plasminogen-stimulated inflammatory cytokine production

by airway smooth muscle cells is regulated by annexin A2. Am J Respir Cell Mol Biol. 2013;49(5):751-758. doi:10.1165/rcmb.2012-0404OC.

78. Bauman KA, Wettlaufer SH, Okunishi K, et al. The antifibrotic effects of plasminogen activation occur via prostaglandin E2 synthesis in humans and mice. J Clin Invest. 2010;120(6):1950-1960. doi:10.1172/JCI38369.

79. Li Q, Laumonnier Y, Syrovets T, Simmet T. Plasmin triggers cytokine induction in human monocyte-derived macrophages. Arterioscler Thromb Vasc Biol. 2007;27(6):1383-1389. doi:10.1161/ ATVBAHA.107.142901.

80. Henneke I, Greschus S, Savai R, et al. Inhibition of urokinase activity reduces primary tumor growth and metastasis formation in a murine lung carcinoma model. Am J Respir Crit Care Med. 2010;181(6):611-619. doi:10.1164/rccm.200903-0342OC.

81. Schmitt M, Harbeck N, Brünner N, et al. Cancer therapy trials employing level-of-evidence-1 disease forecast cancer biomarkers uPA and its inhibitor PAI-1. Expert Rev Mol Diagn. 2011;11(6):617-634. doi:10.1586/erm.11.47.

(20)

6

inhibits neoangiogenesis and human breast tumor growth in a xenograft model. Exp Mol Pathol. 2012;92(1):175-184. doi:10.1016/j.yexmp.2011.10.003.

83. Sharma M, Ownbey RT, Sharma MC. Breast cancer cell surface annexin II induces cell migration and neoangiogenesis via tPA dependent plasmin generation. Exp Mol Pathol. 2010;88(2):278-286. doi:10.1016/j.yexmp.2010.01.001.

84. Zheng L, Foley K, Huang L, et al. Tyrosine 23 phosphorylation-dependent cell-surface localization of annexin A2 is required for invasion and metastases of pancreatic cancer. PloS One. 2011;6(4):e19390. doi:10.1371/journal.pone.0019390.

85. Chambers RC, Scotton CJ. Coagulation cascade proteinases in lung injury and fibrosis. Proc Am Thorac Soc. 2012;9(3):96-101. doi:10.1513/pats.201201-006AW.

86. Lambrecht BN, Hammad H. Asthma and coagulation. N Engl J Med. 2013;369(20):1964-1966. doi:10.1056/NEJMcibr1311045.

87. Tchougounova E, Pejler G. Regulation of extravascular coagulation and fibrinolysis by heparin-dependent mast cell chymase. FASEB J Off Publ Fed Am Soc Exp Biol. 2001;15(14):2763-2765. doi:10.1096/fj.01-0486fje.

88. Knight DA, Lim S, Scaffidi AK, et al. Protease-activated receptors in human airways: upregulation of PAR-2 in respiratory epithelium from patients with asthma. J Allergy Clin Immunol. 2001;108(5):797-803. doi:10.1067/mai.2001.119025.

89. Scotton CJ, Krupiczojc MA, Königshoff M, et al. Increased local expression of coagulation factor X contributes to the fibrotic response in human and murine lung injury. J Clin Invest. 2009;119(9):2550-2563. doi:10.1172/JCI33288.

90. Ortiz-Stern A, Deng X, Smoktunowicz N, Mercer PF, Chambers RC. 1-dependent and PAR-independent pro-inflammatory signaling in human lung fibroblasts exposed to thrombin. J Cell Physiol. 2012;227(11):3575-3584. doi:10.1002/jcp.24061.

91. Tran T, Stewart AG. Protease-activated receptor (PAR)-independent growth and pro-inflammatory actions of thrombin on human cultured airway smooth muscle. Br J Pharmacol. 2003;138(5):865-875. doi:10.1038/sj.bjp.0705106.

92. Shinagawa K, Martin JA, Ploplis VA, Castellino FJ. Coagulation factor Xa modulates airway remodeling in a murine model of asthma. Am J Respir Crit Care Med. 2007;175(2):136-143. doi:10.1164/ rccm.200608-1097OC.

93. Zhu W, Bi M, Liu Y, et al. Thrombin promotes airway remodeling via protease-activated receptor-1 and transforming growth factor-β1 in ovalbumin-allergic rats. Inhal Toxicol. 2013;25(10):577-586. do i:10.3109/08958378.2013.813995.

94. Bousquet J, Jeffery PK, Busse WW, Johnson M, Vignola AM. Asthma. From bronchoconstriction to airways inflammation and remodeling. Am J Respir Crit Care Med. 2000;161(5):1720-1745. doi:10.1164/ajrccm.161.5.9903102.

95. Pilette C, Francis JN, Till SJ, Durham SR. CCR4 ligands are up-regulated in the airways of atopic asthmatics after segmental allergen challenge. Eur Respir J. 2004;23(6):876-884.

96. Ordoñez CL, Shaughnessy TE, Matthay MA, Fahy JV. Increased neutrophil numbers and IL-8 levels in airway secretions in acute severe asthma: Clinical and biologic significance. Am J Respir Crit Care Med. 2000;161(4 Pt 1):1185-1190. doi:10.1164/ajrccm.161.4.9812061.

97. Taha RA, Minshall EM, Miotto D, et al. Eotaxin and monocyte chemotactic protein-4 mRNA expression in small airways of asthmatic and nonasthmatic individuals. J Allergy Clin Immunol. 1999;103(3 Pt 1):476-483.

98. John AE, Zhu YM, Brightling CE, Pang L, Knox AJ. Human airway smooth muscle cells from asthmatic individuals have CXCL8 hypersecretion due to increased NF-kappa B p65, C/EBP beta, and RNA polymerase II binding to the CXCL8 promoter. J Immunol Baltim Md 1950. 2009;183(7):4682-4692. doi:10.4049/jimmunol.0803832.

99. Chan V, Burgess JK, Ratoff JC, et al. Extracellular matrix regulates enhanced eotaxin expression in asthmatic airway smooth muscle cells. Am J Respir Crit Care Med. 2006;174(4):379-385. doi:10.1164/ rccm.200509-1420OC.

(21)

101. Araujo BB, Dolhnikoff M, Silva LFF, et al. Extracellular matrix components and regulators in the airway smooth muscle in asthma. Eur Respir J. 2008;32(1):61-69. doi:10.1183/09031936.00147807. 102. Johnson PRA, Burgess JK, Underwood PA, et al. Extracellular matrix proteins modulate asthmatic

airway smooth muscle cell proliferation via an autocrine mechanism. J Allergy Clin Immunol. 2004;113(4):690-696. doi:10.1016/j.jaci.2003.12.312.

103. Lau JY, Oliver BG, Baraket M, et al. Fibulin-1 is increased in asthma--a novel mediator of airway remodeling? PloS One. 2010;5(10):e13360. doi:10.1371/journal.pone.0013360.

104. Mahn K, Hirst SJ, Ying S, et al. Diminished sarco/endoplasmic reticulum Ca2+ ATPase (SERCA) expression contributes to airway remodelling in bronchial asthma. Proc Natl Acad Sci U S A. 2009;106(26):10775-10780. doi:10.1073/pnas.0902295106.

105. Trian T, Benard G, Begueret H, et al. Bronchial smooth muscle remodeling involves calcium-dependent enhanced mitochondrial biogenesis in asthma. J Exp Med. 2007;204(13):3173-3181. doi:10.1084/jem.20070956.

106. Alrashdan YA, Alkhouri H, Chen E, et al. Asthmatic airway smooth muscle CXCL10 production: mitogen-activated protein kinase JNK involvement. Am J Physiol Lung Cell Mol Physiol. 2012;302(10):L1118-1127. doi:10.1152/ajplung.00232.2011.

107. Tan X, Alrashdan YA, Alkhouri H, Oliver BGG, Armour CL, Hughes JM. Airway smooth muscle CXCR3 ligand production: regulation by JAK-STAT1 and intracellular Ca2+. Am J Physiol - Lung Cell Mol Physiol. 2013;304(11):L790-L802. doi:10.1152/ajplung.00356.2012.

108. Chang P-J, Bhavsar PK, Michaeloudes C, Khorasani N, Chung KF. Corticosteroid insensitivity of chemokine expression in airway smooth muscle of patients with severe asthma. J Allergy Clin Immunol. 2012;130(4):877-885.e5. doi:10.1016/j.jaci.2012.07.017.

109. Roth M, Johnson PRA, Borger P, et al. Dysfunctional interaction of C/EBPalpha and the glucocorticoid receptor in asthmatic bronchial smooth-muscle cells. N Engl J Med. 2004;351(6):560-574. doi:10.1056/NEJMoa021660.

110. Hew M, Bhavsar P, Torrego A, et al. Relative corticosteroid insensitivity of peripheral blood mononuclear cells in severe asthma. Am J Respir Crit Care Med. 2006;174(2):134-141. doi:10.1164/ rccm.200512-1930OC.

111. Ingram JL, Huggins MJ, Church TD, et al. Airway fibroblasts in asthma manifest an invasive phenotype. Am J Respir Crit Care Med. 2011;183(12):1625-1632. doi:10.1164/rccm.201009-1452OC.

112. Zheng T, Liu W, Oh S-Y, et al. IL-13 receptor alpha2 selectively inhibits IL-13-induced responses in the murine lung. J Immunol Baltim Md 1950. 2008;180(1):522-529.

113. Han MK, Criner GJ. Update in chronic obstructive pulmonary disease 2012. Am J Respir Crit Care Med. 2013;188(1):29-34. doi:10.1164/rccm.201302-0319UP.

114. Qiu W, Baccarelli A, Carey VJ, et al. Variable DNA methylation is associated with chronic obstructive pulmonary disease and lung function. Am J Respir Crit Care Med. 2012;185(4):373-381. doi:10.1164/ rccm.201108-1382OC.

115. Sze MA, Dimitriu PA, Hayashi S, et al. The lung tissue microbiome in chronic obstructive pulmonary disease. Am J Respir Crit Care Med. 2012;185(10):1073-1080. doi:10.1164/rccm.201111-2075OC. 116. Jeffery PK. Remodeling and inflammation of bronchi in asthma and chronic obstructive pulmonary

disease. Proc Am Thorac Soc. 2004;1(3):176-183. doi:10.1513/pats.200402-009MS.

117. Hallgren O, Rolandsson S, Andersson-Sjöland A, et al. Enhanced ROCK1 dependent contractility in fibroblast from chronic obstructive pulmonary disease patients. J Transl Med. 2012;10:171. doi:10.1186/1479-5876-10-171.

118. Hallgren O, Nihlberg K, Dahlbäck M, et al. Altered fibroblast proteoglycan production in COPD. Respir Res. 2010;11:55. doi:10.1186/1465-9921-11-55.

119. Zhang J, Wu L, Qu J, Bai C, Merrilees MJ, Black PN. Pro-inflammatory phenotype of COPD fibroblasts not compatible with repair in COPD lung. J Cell Mol Med. 2012;16(7):1522-1532. doi:10.1111/ j.1582-4934.2011.01492.x. 120. Larsson-Callerfelt A-K, Hallgren O, Andersson-Sjöland A, et al. Defective alterations in the collagen network to prostacyclin in COPD lung fibroblasts. Respir Res. 2013;14:21. doi:10.1186/1465-9921-14-21. 121. Stitham J, Midgett C, Martin KA, Hwa J. Prostacyclin: an inflammatory paradox. Front Pharmacol. 2011;2:24. doi:10.3389/fphar.2011.00024.

(22)

6

Diseases. Semin Respir Crit Care Med. 2007;28(05):496-503. doi:10.1055/s-2007-991522.

123. Nathan N, Thouvenin G, Fauroux B, Corvol H, Clement A. Interstitial lung disease: physiopathology in the context of lung growth. Paediatr Respir Rev. 2011;12(4):216-222. doi:10.1016/j.prrv.2011.04.003. 124. Coffey E, Newman DR, Sannes PL. Expression of fibroblast growth factor 9 in normal human lung

and idiopathic pulmonary fibrosis. J Histochem Cytochem Off J Histochem Soc. 2013;61(9):671-679. doi:10.1369/0022155413497366.

125. Lindahl GE, Stock CJ, Shi-Wen X, et al. Microarray profiling reveals suppressed interferon stimulated gene program in fibroblasts from scleroderma-associated interstitial lung disease. Respir Res. 2013;14:80. doi:10.1186/1465-9921-14-80.

126. Nagahama KY, Togo S, Holz O, et al. Oncostatin M modulates fibroblast function via signal transducers and activators of transcription proteins-3. Am J Respir Cell Mol Biol. 2013;49(4):582-591. doi:10.1165/rcmb.2012-0460OC.

127. Suda T, Chida K, Todate A, et al. Oncostatin M production by human dendritic cells in response to bacterial products. Cytokine. 2002;17(6):335-340. doi:10.1006/cyto.2002.1023.

128. Botelho FM, Rangel-Moreno J, Fritz D, Randall TD, Xing Z, Richards CD. Pulmonary expression of oncostatin M (OSM) promotes inducible BALT formation independently of IL-6, despite a role for IL-6 in OSM-driven pulmonary inflammation. J Immunol Baltim Md 1950. 2013;191(3):1453-1464. doi:10.4049/jimmunol.1203318.

129. Fritz DK, Kerr C, Fattouh R, et al. A mouse model of airway disease: oncostatin M-induced pulmonary eosinophilia, goblet cell hyperplasia, and airway hyperresponsiveness are STAT6 dependent, and interstitial pulmonary fibrosis is STAT6 independent. J Immunol Baltim Md 1950. 2011;186(2):1107-1118. doi:10.4049/jimmunol.0903476.

130. Mozaffarian A, Brewer AW, Trueblood ES, et al. Mechanisms of oncostatin M-induced pulmonary inflammation and fibrosis. J Immunol Baltim Md 1950. 2008;181(10):7243-7253.

131. Smyth DC, Kerr C, Li Y, Tang D, Richards CD. Oncostatin M induction of eotaxin-1 expression requires the convergence of PI3’K and ERK1/2 MAPK signal transduction pathways. Cell Signal. 2008;20(6):1142-1150. doi:10.1016/j.cellsig.2008.02.001.

132. Pak O, Aldashev A, Welsh D, Peacock A. The effects of hypoxia on the cells of the pulmonary vasculature. Eur Respir J. 2007;30(2):364-372. doi:10.1183/09031936.00128706.

133. Crosswhite P, Sun Z. Nitric oxide, oxidative stress and inflammation in pulmonary arterial hypertension. J Hypertens. 2010;28(2):201-212. doi:10.1097/HJH.0b013e328332bcdb.

134. Stenmark KR, Frid MG, Yeager M, et al. Targeting the adventitial microenvironment in pulmonary hypertension: A potential approach to therapy that considers epigenetic change. Pulm Circ. 2012;2(1):3-14. doi:10.4103/2045-8932.94817.

135. Wang D, Zhang H, Li M, et al. MicroRNA-124 controls the proliferative, migratory, and inflammatory phenotype of pulmonary vascular fibroblasts. Circ Res. 2014;114(1):67-78. doi:10.1161/ CIRCRESAHA.114.301633.

136. Schabbauer G, Matt U, Günzl P, et al. Myeloid PTEN promotes inflammation but impairs bactericidal activities during murine pneumococcal pneumonia. J Immunol Baltim Md 1950. 2010;185(1):468-476. doi:10.4049/jimmunol.0902221.

137. Horita H, Furgeson SB, Ostriker A, et al. Selective inactivation of PTEN in smooth muscle cells synergizes with hypoxia to induce severe pulmonary hypertension. J Am Heart Assoc. 2013;2(3):e000188. doi:10.1161/JAHA.113.000188.

138. Brightling CE, Bradding P, Symon FA, Holgate ST, Wardlaw AJ, Pavord ID. Mast-cell infiltration of airway smooth muscle in asthma. N Engl J Med. 2002;346(22):1699-1705. doi:10.1056/NEJMoa012705. 139. Amin K, Janson C, Harvima I, Venge P, Nilsson G. CC chemokine receptors CCR1 and CCR4 are

expressed on airway mast cells in allergic asthma. J Allergy Clin Immunol. 2005;116(6):1383-1386. doi:10.1016/j.jaci.2005.08.053.

140. Begueret H, Berger P, Vernejoux J, Dubuisson L, Marthan R, Tunon-de-Lara JM. Inflammation of bronchial smooth muscle in allergic asthma. Thorax. 2007;62(1):8-15. doi:10.1136/thx.2006.062141. 141. Carroll NG, Mutavdzic S, James AL. Increased mast cells and neutrophils in submucosal mucous

Referenties

GERELATEERDE DOCUMENTEN

Cellular expression of Fstl1 During embryonic development, Fstl1 mRNA is detected in different organs, but the highest expression is found in the lung, predominantly in

The expression of Fstl1 mRNA and protein in lung homogenates of Fstl1-eKO mice were significantly reduced compared to age-matched controls at 1 week after birth (mRNA

We observed that the expression of Fstl1 protein was significantly increased in total lung homogenates of non-COPD ex-smokers alone (p=0.033) or combined with

Most intriguingly, this study revealed that inhibition of activin-A signalling by administration of its endogenous inhibitor follistatin, attenuates cigarette smoke-induced

Similarly 50% Fstl1 CM increased IL-6 and IL-8 mRNA by 2.0-fold expressed by bronchial epithelial cells and its potential contribution to inflammatory cytokine release

Taken together, these results demonstrate that BMP4 suppresses marker gene expression for common club cells and promotes markers for distal variant club cells in adult human

worth pursuing to trigger differentiation and maintenance of variant club cells capable of airway epithelial repair as well as suppressing the inflammatory response in the lung,

Door het remmen van de negatieve regulatie van de BMP signaaltransductieroute op differentiatie van slijmbekercellen en basale cellen zou het Toegenomen Fstl1 in