Sei sulla pagina 1di 12

Immunology of the lung

review

The immunology of asthma


Bart N Lambrecht1–3 & Hamida Hammad1,2

Asthma is a common disease that affects 300 million people worldwide. Given the large number of eosinophils in the airways
of people with mild asthma, and verified by data from murine models, asthma was long considered the hallmark T helper type 2
(TH2) disease of the airways. It is now known that some asthmatic inflammation is neutrophilic, controlled by the T H17 subset
of helper T cells, and that some eosinophilic inflammation is controlled by type 2 innate lymphoid cells (ILC2 cells) acting
together with basophils. Here we discuss results from in-depth molecular studies of mouse models in light of the results from the
first clinical trials targeting key cytokines in humans and describe the extraordinary heterogeneity of asthma.
© 2015 Nature America, Inc. All rights reserved.

Asthma is a chronic inflammatory disease of the conducting airways at a young age are likely to develop asthma1, and large studies have
in which many cells of the innate and adaptive immune systems act found that up to half of patients with atopic eczema developed asthma
together with epithelial cells to cause bronchial hyper-reactivity later in life2. In up to 80% of cases, patients with allergic asthma have
(BHR) (the tendency of smooth muscle cells in people with asthma concurrent allergic rhinitis. The ‘united airways disease’ hypothesis
to react to nonspecific stimuli such as cold air and exercise), mucus proposes that allergic rhinitis and asthma are manifestations of
overproduction, airway wall remodeling and airway narrowing. the same underlying disease process and that each influences the
In susceptible patients, this leads to repeated periods of shortness of severity of the other.
breath, wheezing and chest tightness. The disease is very common in Nonallergic (intrinsic) asthma often develops later in life and, per
affluent societies, in which almost 1 in 10 children and 1 in 12 adults its definition, has neither IgE reactivity to allergens in the serum
is affected, which results in substantial morbidity and annual health- nor any obvious involvement of the adaptive immune system such
care expenditure. Worldwide, up to 300 million people are affected. as type 2 helper T cells (TH2 cells). This form of the disease is more
The total cost of the disease, both in direct medical costs and indirect common in women, is often associated with chronic rhinosinusitis
costs caused by loss of productivity, is estimated to exceed $18 billion and nasal polyps, as well as obesity, and is difficult to treat, often
annually in the USA. In many patients, the disease can be controlled requiring long-term treatment with systemic steroids. The field of
by a combination of an inhaled corticosteroid (which acts to suppress clinical asthma diagnosis and treatment has undergone great concep-
the inflammation) and a short- or long-acting β2-adrenergic agonist tual shifts through the advent of genome-wide expression studies and
(which acts to open the constricting bronchial smooth muscle), but the first clinical trials using targeted therapies with biological agents.
npg

in some 5–10% of patients, the disease is refractory to corticosteroid As with many chronic inflammatory diseases, clinicians now realize
treatment and often leads to hospital admissions caused by respiratory that the division of asthma into only two clinical forms has been an
viral infection with rhinovirus. oversimplification. Different asthma phenotypes, each with a distinct
Traditionally, two forms of asthma have been defined in the clinic. pathophysiology, are now being defined as asthma endotypes. The
Most children and roughly 50% of adults have allergic asthma, in endotypes differ in terms of genetic susceptibility, environmental risk
which the disease coincides with allergic sensitization defined by factors, age of onset, clinical presentation, prognosis and response to
the presence of serum immunoglobulin E (IgE) antibodies and/or a standard and new therapies3–5. Asthma is therefore increasingly seen
positive skin-prick test to the (lipo)proteins of common inhaled or as a syndrome rather than a single disease5,6. In people with asthma
ingested allergens such as house dust mite (HDM), animal dander, who smoke, there is also considerable clinical overlap with chronic
fungal spores, plant or tree pollen, or peanuts. In children, the disease obstructive pulmonary disease, whose immunological basis has been
starts with allergic sensitization, often accompanied by eczema in the reviewed7. In this Review, we will focus on the underlying immuno-
first year of life. These children later develop allergic rhinitis and can logical basis of the various asthma endotypes, integrating results from
progress to asthma. This stepwise increase of symptoms has been human studies targeting particular pathways with results from animal
called the ‘atopic march’. In particular, children with multiple allergies studies in which a great deal of molecular detail has been gathered.

Eosinophilic asthma as a TH2 disorder


1VIB Inflammation Research Center, Ghent University, Ghent, Belgium. Since the development of the field of pulmonary immunology, asthma
2Department of Respiratory Medicine, University Hospital Ghent, Ghent,

Belgium. 3Department of Pulmonary Medicine, Erasmus MC, Rotterdam,


has been seen as the hallmark TH2 disorder of the lungs. This was
the Netherlands. Correspondence should be addressed to B.N.L. supported by the early findings that asthmatic airway inflammation
(bart.lambrecht@ugent.be) or H.H. (hamida.hammad@ugent.be). measured in bronchoalveolar lavage (BAL) fluid, bronchial biopsies or
Received 1 October; accepted 7 November; published online 18 December 2014; induced sputum is often eosinophilic in nature, regardless of whether
doi:10.1038/ni.3049 the patient is allergic or not8–11. In initial studies of people with mild

nature immunology  VOLUME 16  NUMBER 1  JANUARY 2015 45


review

Figure 1  Overview of functions of TH2 cells


and ILC2 cells in asthma. TH2 cells and ILC2
cells share many features, such as expression
of the transcription factor GATA-3, which drives
the production of TH2 cytokines, and expression Airway remodeling
Epithelial Chemokine Goblet cell
of the chemokine receptors CCR4, CCR8 and Tissue repair
damage induction metaplasia
CRTH2. Through the production of IL-5, they
control eosinophil development in the bone
marrow. Via IL-13, they can cause goblet cell ECP IL-4 Amphiregulin
metaplasia and bronchial hyperreactivity and MBP
LTC4 IL-13
Charcot-Leyden
prime the vessel wall for upregulation of the crystals LTD4
adhesion molecule VCAM-1 and ICAM-1, thus IL-4
priming eosinophil exit. TH2 cells make more IL-5, IL-13
IL-5
IL-4. Both cell types have been found to Eosinophil
TH2 & ILC2
produce IL-9 (not shown here), although this (GATA-3+)
CCR4
could be a separate population of TH9 cells.
Survival IL-25R
ECP, eosinophil cationic protein; MBP, major Activation Bronchial
basic protein; VLA-4, integrin a4b1; VCAM-1, IL-33R
MBP
hyperreactivity
vascular cell adhesion molecule 1. ECP
IL-5 IL-4 LTC4
IL-13 LTD4
TNF
to moderate asthma, increased numbers of
CD4+ cells that produce interleukin 4 (IL-4) Chemotaxis
and IL-5 were found in BAL fluid and
© 2015 Nature America, Inc. All rights reserved.

Kim Caesar/Nature Publishing Group


mucosal biopsies, correlating with the degree VLA-4
ICAM-1 VCAM-1
of airway eosinophilia12. Subsequent unsu-
VCAM-1
pervised clustering algorithms have shown
that the various asthma endotypes fall into IL-5
Vessel wall priming
TH2hi and TH2lo clusters on the basis of the for inflammatory cell
presence or absence of the cytokines IL-4, recruitment

IL-5 and IL-13 and eosinophils in blood and


Bone marrow mobilization Formation of perivascular
tissue4,13. The presence of serum IgE (atopy) of eosinophil precursors and peribronchial infiltrates
is the prototypical hallmark of adaptive TH2
immunity, driven by IL-4-induced class
switching of the immunoglobulins synthesized by B cells. Although IL-13, is able to improve lung function and reduce the frequency of
it is very easy to measure, the presence of high serum titers of IgE is a exacerbation in people with moderate to severe asthma with high
poor predictor of whether there is TH2 signature gene expression in levels of eosinophils in the blood24. In a phase 2 trial, blocking IL-13
the tissues and whether the patient will develop asthma. There may in humans with lebrikizumab (Genentech) also improved lung-
be better surrogate biomarkers of the TH2hi endotype of asthma, such function parameters, particularly in the TH2hi endotype25.
as levels of serum IL-25 and periostin that correlate well with tissue Eosinophilia in lung tissue is driven by IL-5, which supports the
eosinophilia14. development of eosinophils in the bone marrow, and by the recruit-
In mouse models of asthma driven by inhalation of the model anti- ment of eosinophils to the lung mucosa and interstitium via pro-
npg

gen ovalbumin (OVA) after intraperitoneal sensitization to OVA in duction of eotactic chemokines such as eotaxins 1, 2 and 3 (CCL11,
alum, the genetic or antibody-mediated depletion of CD4 + T cells CCL24 and CCL26, respectively). Eosinophil-derived products such
abolishes key features of asthma, whereas the adoptive transfer of TH2- as eosinophil peroxidase cause BHR directly and activate adaptive
polarized CD4+ T cells from mice with transgenic expression of an immunity through effects on dendritic cells (DCs)26–28. Eosinophils
OVA peptide–specific T cell antigen receptor (TCR) leads to the induc- can also present antigen directly to primed effector T cells29. Studies
tion of asthma features15–18. Conversely, asthma is suppressed by the of mice genetically modified to lack eosinophils have shown that
transfer of CD4+ type 1 helper T cells (TH1 cells) or administration of these cells also contribute to airway wall remodeling and subepithelial
IL-12, which induces the TH1 response19. Together with the early data membrane thickening via the release of transforming growth factor-β
from people with asthma, the case for CD4+ TH2 cells as controllers of (TGF-β), although this has not been seen in all asthma models30,31.
the disease was therefore verified (Fig. 1). Mice that lack the key TH2 Similar to neutrophils, eosinophils upon activation undergo cytolysis
cytokines IL-4, IL-5 or IL-13 all had substantial reductions in asthma and release extracellular DNA traps that contain eosinophilic granules.
features in the OVA model20. In the mouse, IL-4 is necessary for the As these granules maintain the capacity for ligand-induced secretion,
development of adaptive TH2 immunity and IgG1 and IgE antibodies the formation of DNA traps could lead to high local concentrations of
to OVA, and for priming the vessel wall for eosinophil extravasation; eosinophilic toxins such as eosinophil-derived neurotoxin, cationic
therefore, the OVA model relies heavily on IL-4 (refs. 20,21). IL-13 was proteins (eosinophil peroxidase) and major basic protein, which can
found to be necessary and sufficient for mounting BHR and for goblet damage structural cells of lungs32,33 (Fig. 1).
cell metaplasia, the process by which epithelial cells transdifferentiate The elimination of eosinophils from humans through the use of
and begin producing thick mucus containing the mucins MUC5AC antibody to IL-5 (mepolizumab; GlaxoSmithKline) led to a reduc-
and MUC5B, which clog the airway lumen22,23 (Fig. 1). tion in exacerbation frequency in a subset of patients with high lev-
In humans, treatment with an antibody to the α-chain of the recep- els of circulating eosinophils in blood and frequent exacerbations,
tor for IL-4 (dupilumab; Regeneron Pharmaceuticals), which effec- even in those receiving inhaled steroids34. As exacerbations often
tively blocks downstream signaling via the receptors for IL-4 and require systemic steroids that have serious side effects, blocking IL-5

46 VOLUME 16  NUMBER 1  JANUARY 2015  nature immunology


review

Figure 2  Relative roles of TH2 cells and


ILC2 cells in two forms of eosinophilic asthma. Allergens Pollutants, microbes,
In atopic asthma (left), eosinophilic airway glycolipids
inflammation and BHR are driven by adaptive Goblet cells
Alternatively
TH2 cells that are stimulated by DCs to produce activated
IL-5, IL-13 and IL-4, the latter driving IgE Airway macrophages
synthesis. In nonatopic or intrinsic asthma epithelium
Ym1
(right), which is not dependent on adaptive IL-33
IL-33
immunity, ILC2 cells produce IL-5 and IL-13 IL-25
DCs IL-25
TSLP IL-33
and thus cause eosinophilia and BHR. As there TSLP
NKT cells
is no specific allergen involved and as ILC2
cells produce little IL-4, there is no associated MHCII TSLPR
IgE response from B cells. Modified from ref. TCR
Mast cells Naive
185. MHCII, MHC class II; TSLPR, receptor for T cell ILC2 IL-33R
TSLP; NKT cells, natural killer T cells. IL-25R GATA-3
PGD2
RORα
IL
-9
can therefore reduce the use of oral steroids35. CRTH2
In people with asthma in whom IL-5 was TH2
blocked, there was a considerable reduction in GATA-3
the deposition of extracellular matrix compo- IL-9R
nents tenascin, lumican and procollagen III,

Kim Caesar/Nature Publishing Group


13

-5
IL
Lipoxin A4

-5

IL
L-
possibly leading to less airway remodeling36.

,I
-4
IL-9
IL
New drugs that target eosinophils currently
© 2015 Nature America, Inc. All rights reserved.

3
IL

-1
under phase 3 clinical development include

-1

IL
3
Eosinophils
an antibody to the receptor for IL-5 (ben-
Bronchial hyperreactivity
ralizumab; AstraZeneca/MedImmune) that B cells

effects depletion of eosinophils for months


after a single injection37.
Allergic eosinophilic airway inflammation Nonallergic eosinophilic airway inflammation

ILCs in eosinophilic asthma


In studies in humans, blockade of the receptor for IL-4 or of IL-5 has endogenous IL-33, which activates ILC2 cells and causes pulmonary
led to a favorable clinical response in patients with high eosinophil eosinophilia45. Subsequent data have linked ILC2 cells to the patho-
counts regardless of whether these patients were atopic or not. In genesis of allergic airway inflammation; for example, protease-
mice, administration of allergens such as HDM or Alternaria alternata containing allergens (papain and A. alternata) might activate ILC2
spores also led to airway eosinophilia in mice deficient in the RAG cells indirectly through an effect on lung epithelial cells44,46. The
recombinase, which do not produce mature T cells or B cells. Such proteolytic injury induces the release of IL-33, which in turn acti-
findings suggest that in both species, there are ways of generating type vates ILC2 cells to induce lung eosinophilia and BHR. Another way
2 cytokines and eosinophilia without involvement of the adaptive in which ILC2 cells can be activated in the lungs following allergen
immune system (Fig. 2). Innate lymphoid cells (ILCs) were initially exposure is through expression of the cysteine leukotriene receptor
classified as non-T, non-B effector cells in many disease models of CysLT1R47. LTD4, the main ligand for CysLT1R, can be induced after
TH2 immunity38–40. Published reviews have described the nomen- a single exposure to proteolytic allergens and has been found to regu-
npg

clature of the distinctive forms of ILCs, some of which produce late the activation and proliferation of ILC2 cells47. Such data indicate
interferon-γ (IFN-γ) (group 1 ILCs, which include traditional natu- that ILC2 cells can be activated very early after allergen exposure and
ral killer cells and ILC1 cells), TH2 cytokines (group 2 ILCs formerly that their production of IL-5 and IL-13 could cause some features of
known as ‘nuocytes’ or ‘natural helper cells’, now referred to as ‘ILC2 allergic asthma in a T cell–independent manner (Figs. 1 and 2). In
cells’), and IL-17 and/or IL-22, or are involved in the formation of addition to their role in allergy, ILC2 cells also contribute to BHR
lymphoid tissues (group 3 ILCs, which include ILC3 cells and lym- development in response to respiratory viruses such as influenza virus
phoid tissue–inducer cells)41. In many ways, ILC2 cells resemble TH2 and rhinovirus in adult mice and neonatal mice, respectively48,49. In
cells (Fig. 1). ILC2 cells lack antigen-specific receptors, but like TH2 the context of infection with influenza virus, ILC2 cells that produce
cells they react to the epithelium-derived cytokines IL-25, IL-33 and IL-5 also contribute to the accumulation of eosinophils in the lungs
thymic stromal lymphopoietin (TSLP)38–40. ILC2 cells that produce after viral clearance50. A role for ILC2 cells in viral infections might
TH2 cytokines (IL-13, IL-5 and IL-9) were initially described in the help to explain how viral exposure can cause or exacerbate asthma
gut of mice infected with helminths, where they contribute to tissue symptoms in some people.
eosinophilia and to mucus production, crucial for worm expulsion42. Because many of the initial reports discussed above used RAG-
ILC2 cells develop from common lymphoid progenitors in response to deficient mice, in which the contribution of ILC2 cells might be over-
IL-7 and IL-33 and depend on the transcription factors GATA-3 and rated due to their lack of a functional adaptive immune system, the
RORα. The activation of ILC2 cells is further induced by IL-25 and IL-33, relative contributions of ILC2 cells versus that of TH2 cells in asthma
produced mainly by epithelial cells in response to injury and stimula- has remained unclear. Several groups have addressed this question in
tion via pattern-recognition receptors. The administration of either mice with an intact adaptive immune system. ILC2 cells have been
cytokine to the airways of mice induces the population expansion of found to represent more than half of the cells producing TH2 cytokines
ILC2 cells that produce IL-5 and IL-13 in the lungs and BAL fluid43,44. in the lungs of mice subjected to OVA- and HDM-induced asthma51.
Intratracheal administration of chitin or infection with the intesti- Due to a lack of RORα expression in the hematopoietic compart-
nal nematode Strongyloides venezuelensis triggers the production of ment, mice that lack ILC2 cells develop less-severe lung inflammation

nature immunology  VOLUME 16  NUMBER 1  JANUARY 2015 47


review

than their wild-type littermates do after intranasal administration model, it will be essential to investigate which of the allergic asthma
of papain or HDMs52–54. The precise signals involved in the recruit- features are governed by ILC2 cells. Given the fact that depletion of
ment of ILC2 cells to the lungs after allergen exposure are not very CD4+ T cells in chronic experimental asthma affects mainly airway
clear, but microarray data suggest that the same chemokine recep- eosinophilia without affecting the development of airway remodeling,
tors that attract TH2 cells (CCR4 and CCR8), the prostaglandin D2 one possibility would be that ILC2 cells control airway remodeling18.
receptor CRTH2 and CysLTR1 might be involved47,55. Production In support of this concept, just like CD4+ TH2 cells, ILC2 cells have
of the CCR4 ligands CCL17 and CCL22, depends on signaling via been shown to produce the TGF-β-like molecule amphiregulin, which
STAT6 in epithelial cells, and STAT6-deficient mice fail to recruit contributes to airway remodeling by acting on fibroblasts and epi-
ILC2 cells to the airways56. The signals that dampen the recruitment thelial repair processes62,63. Alternatively, they might be the driving
and activation of ILC2 cells are only starting to be investigated. The force of the alternative activation of macrophages, which contribute
resolvin lipoxin A4 has been shown to suppress IL-13 production by to airway wall remodeling by promoting collagen synthesis60.
ILC2 cells, which indicates that the function of ILC2 cells in asthma Despite the rapid increase in knowledge about ILC2 cells in experi-
could be regulated by anti-inflammatory lipid mediators57 (Fig. 2). mental models of type 2 immunity, still very little is understood about
Similarly, the uptake of apoptotic cells by lung epithelial cells (a proc- the clinical relevance of ILC2 cells in human asthma. ILC2 cells may
ess dependent on the small GTPase Rac) suppresses the production of have a predominant role in a subset of people with severe, nonatopic
IL-33 and in this way suppresses activation of ILC2 cells58. Some asthma who have high eosinophil counts in the blood and lungs and
parasites have also evolved strategies to suppress the activation have a TH2 signature in the tissues, which would explain why they
of ILC2 cells. Soluble excretory and secretory products of respond favorably to blockade of the IL-4 receptor or IL-5 (Fig. 2).
Heligmosomoides polygyrus block the recruitment of ILC2 cells in This subform of the TH2hi endotype cluster of asthma often occurs
response to airway allergen exposure by blocking the epithelial with chronic rhinosinusitis and nasal polyps and does not respond
production of IL-33 (ref. 59). well to steroids. A subset of these patients also seem to have a colo-
© 2015 Nature America, Inc. All rights reserved.

Although ILC2 cells can contribute to TH2 cell–mediated lung nization of the nasal sinuses and airways with filamentous fungi that
inflammation, the method of action by which this occurs remains might represent a chronic trigger for the innate immune system64. In
unclear. Chitin is a common constituent of the wall of arthropods patients with rhinosinusitis, there is an increase in GATA-3+CRTH2+
and helminths and is a well-known trigger for the recruitment of ILC2 cells in the nasal tissues65,66. The production of IL-5 and IL-13
ILC2 cells to the lungs. After exposure to chitin, the IL-13 secreted is not as well suppressed by steroids in ILC2 cells as it is in CD4+
by ILC2 cells is crucial for inducing the alternative activation of T cells. TSLP can induce this state of steroid refractoriness in ILC2
macrophages, whereas IL-5 controls early airway eosinophilia 60. In cells by inducing phosphorylation of the transcription factor STAT5
addition to those effects on cells of the innate immune system, ILC2 and by upregulating the prosurvival factor Bcl-xL67. If ILC2 cells
cells can also directly or indirectly influence the adaptive immune are driving disease in this subset of patients, it would explain the
system through the release of cytokines. ILC2 cells may provide an steroid refractoriness. The ILC2 cells could be stimulated by chronic
early source of IL-13 that polarizes naive CD4+ T cells to become TH2 epithelial activation induced, for example, by environmental pollut-
cells. Although most TH2 priming occurs via IL-4, the experimental ants, irritants, chronic airway mycosis or repetitive viral infections
inhaled allergen papain is IL-4 independent but requires IL-13 from driving the expression of IL-25, IL-33 and TSLP (Fig. 2).
ILC2 cells to induce adaptive CD4+ T cells via migratory antigen-
presenting DCs54. ILC2 cells have also been shown to express major Neutrophilic asthma as a TH17 disorder
histocompatibility complex (MHC) class II and costimulatory Although asthma is classically associated with eosinophilia and TH2
molecules, suggesting that they may themselves act as antigen- cytokines, some patients show a neutrophil-predominant disease
presenting cells, activating CD4+ T cells during the sensitization with an absence of TH2 cytokines. In particular, patients with late-
npg

phase or during the effector phase. Despite their low expression of onset and more severe forms of asthma seem to have neutrophilic
MHC class II, human ILC2 cells are able to drive the proliferation inflammation with less reversible airway obstruction and a mixed
of HDM-specific CD4+ T cells ex vivo61. Mouse ILC2 cells are able TH1 and TH17 cytokine milieu68–70. The role of IL-17 and TH17 cells
to endocytose soluble OVA, but they are not able to induce OVA- in allergic asthma has not been fully elucidated, with IL-17 sometimes
specific naive CD4+ T cell proliferation unless antigen-derived favoring and sometimes protecting mice from the disease. These dis-
peptides are added to the coculture. Although ILC2 cells can induce crepancies can been explained by the timing of the neutralization or
some degree of T cell proliferation under some circumstances and administration of IL-17, as a protective role for IL-17A in asthma has
with soluble peptide antigen ex vivo, whether they also process and been observed only during the challenge phase71,72. Similar results
present antigens in vivo remains to be addressed. Also, gut ILC2 cells have been obtained for IL-22, another cytokine produced by TH17
have considerable expression of MHC class II, whereas expression is cells73. Another study showed that high exposure to diesel exhaust
much lower on lung ILC2 cells, which raises the question of whether particles is associated with both exacerbated asthma symptoms and
ILC2 cells are antigen-presenting cells relevant to asthma. increased serum concentrations of IL-17A in children with atopic
Such studies have begun to shed light on the relative contribution asthma74. Severe forms of asthma are characterized by increased
of ILC2 and TH2 cells in allergy. However, to pinpoint the exact role airway remodeling. In some experimental asthma models driven by
of each cell population in the disease, there is an urgent need for new HDMs or ozone, IL-17A contributes to remodeling by promoting
tools that would allow the specific depletion of ILC2 cells at different fibroblast proliferation75 and by counteracting the anti-inflammatory
stages of the response (sensitization versus effector phase). Crossing role of regulatory T cells76. In mice and humans, IL-17 can also cause
Cd4-Cre mice (which express Cre recombinase from the T cell– direct contraction of bronchial smooth muscle cells and thus cause
specific Cd4 promoter) to mice in which a loxP-flanked gene encoding BHR in the absence of neutrophilic inflammation77. It remains to
the diphtheria toxin receptor is inserted into the locus encoding the be conclusively shown where in asthmatic airways the IL-17A and
inducible costimulator ICOS produces progeny in which it is possible IL-17F are produced. In addition to conventional TCRαβ+CD4+
to deplete mice of ILC2 cells while sparing CD4+ TH2 cells61. With this T cells, TCRγδ+ T cells, invariant natural killer T cells and ILC3 cells

48 VOLUME 16  NUMBER 1  JANUARY 2015  nature immunology


review

can be a source of copious IL-17, and the relative contribution of of TH9 cells. Studies addressing the molecular pathways involved in
all of these might differ in various asthma endotypes. The cytokine IL-9 production have found that binding of the transcription factors
production by TH17 cells is notoriously resistant to inhibition by ster- Smad2, Smad3, STAT5, IRF4 and PU.1 to the Il9 promoter are essen-
oids, which explains why neutrophil-rich inflammation driven by tial for TH9 differentiation and IL-9 production98–101. This process
TH17 cells is the pathological correlate of steroid-resistant asthma. is influenced by the inhibitory SOCS proteins, and deficiency in CIS,
A clinical trial with a antibody that neutralizes the human receptor for a member of the SOCS family, causes increased TH2 differentiation
IL-17 (brodalumab; Amgen/AstraZeneca), which blocks the activity and exacerbated asthma features100. IL-9 is a mast cell growth factor
of IL-17A, IL-17F and IL-25, has shown minimal effects on outcome that promotes IL-4-driven antibody production by B cells102 and can
measures of asthma in mild to moderate disease78. It is possible that also induce goblet cell metaplasia. IL-9 has high expression in the lungs
subgroups of patients, particularly those with large numbers of spu- of patients with asthma103, and in mice, the neutralization of IL-9
tum neutrophils or with a high degree of lung function reversibility, with antibodies has been shown to ameliorate OVA- and Aspergillus
would respond more favorably70. There is also a complex interaction fumigatus–induced asthma symptoms101,104. In a chronic asthma
between TH17 cell–driven asthma and tumor-necrosis factor (TNF). model, antibody to IL-9 blocked the development of airway remod-
Lung and systemic levels of TNF are increased in patients with severe eling by reducing mast cell numbers102. Although T cells were initially
steroid-resistant asthma, although some studies have not confirmed considered as the main source of IL-9, experiments with fate-mapping
this70,79. In an adoptive transfer model of TH17 OVA-specific T cells, reporter mice have shown that following papain administration, ILC2
neutralization of TNF led to reduced neutrophilic influx in the lung cells produce greater amounts of IL-9 than do T cells105. IL-9 derived
tissue and airspaces, associated with amelioration of lung function from ILC2 cells has been shown to be an autocrine amplifier of ILC2
parameters such as lung compliance but not airway hyper-resistance 70. cell function by promoting their survival106. The relative contribu-
In clinical trials, the results of TNF blockade have also been vari- tions of IL-9 derived from TH9 cells and that derived from ILC2 cells
able79–81. It is still unclear if TNF blockade in IL-17-rich neutrophilic in allergic asthma remains to be addressed. Trials have been initiated
© 2015 Nature America, Inc. All rights reserved.

asthma would also improve steroid responsiveness. to target IL-9 in patients with asthma (MEDI-528; MedImmune), but
they have produced disappointing clinical results107.
Overlap syndromes
The view of eosinophilic asthma as an exclusive T H2 disorder and Regulatory T cells in asthma
neutrophilic asthma as an exclusive TH17 disorder is probably an For many years, it was postulated that inflammatory responses in
oversimplification seen only at the extremes of a continuous spec- asthma would develop because patients had a deficiency in natural or
trum. In many cases, there is considerable overlap in the types of induced regulatory T cells (Treg cells). In mice, the usual outcome of
cytokines found in an asthma endotype, related to severity of disease. the inhalation of harmless antigens such as OVA is tolerance, a process
This has been nicely demonstrated in a model of fungal mycosis in mediated by Foxp3+ induced Treg cells (iTreg cells)108. Indeed, mice
mice, which at extremely high challenge levels exhibit immunophe- that lack the intronic Foxp3 enhancer CNS1 show a paucity of iTreg
notype switching from a predominant TH2 response to TH1 and/or cells and develop strong TH2 responses at mucosal sites, including the
TH17 response(s)82. Some asthma features are induced by CD4+ lungs109. iTreg cells are also found in the airways of mice with asthma,
T cells that produce both TH17 cytokines and TH2 cytokines. In mice, where they have high expression of neuropilin 1 (ref. 110). Adoptive-
these IL-4+ TH17 cells induce more severe disease when transferred83. transfer studies of mice have revealed that Treg cells suppress asthma
Another study has shown that the chitinase-like protein Ym1, tradi- features through IL-10 and TGF-β, two cytokines able to suppress
tionally produced by alternatively activated macrophages elicited by pulmonary DC activation111, and through direct interactions with
IL-4 and IL-13, drives the production of IL-17 by γδ T cells and thus endothelial cells, thus preventing angiogenesis112. Another possible
contributes to lung neutrophilia and damage in a TH2 setting84. IL-4+ mechanism involves IL-35 production by ICOS+ Treg cells that have
npg

CD4+ TH17 cells have also been identified in humans, and the amount the potential to suppress IL-17-induced BHR in mice113.
of IL-17 released by such T cells in BAL fluid correlates with increased In healthy humans, immune responses to allergens can be consid-
BHR and airway obstruction85. ered active peripheral tolerance driven mainly by IL-10 and TGF-β.
It is also important to consider the contribution of the T H1 cell– The precise role of Treg cells in patients with allergic diseases has been
derived cytokine IFN-γ in eosinophilic and neutrophilic asthma. much debated. In severe asthmatics, the number of Treg cells present
Simultaneous transfer of OVA-specific TH1-polarized cells plus TH2- in the blood and sputum is lower, and their suppressive activities
polarized cells into mice is able to exacerbate bronchial hyper-reactiv- are impaired compared with those of cells from healthy subjects114.
ity in which IFN-γ acts together with IL-13 to cause smooth muscle Moreover, the percentage of Treg cells is also decreased in the BAL
contraction and activation of cells of the innate immune system and fluid of pediatric patients with asthma115. In adult patients, the data
IFN-γ itself promotes the homing of TH2 cells to the lungs86–89. IL-18 are more controversial, with some reports showing that the number of
drives the activation of TH1 cells that produce IL-13 and cause severe Treg cells in the lungs is increased116 or and other showing that number
BHR90. The airways of asthmatics contain increased amounts of IFN- is decreased117. Despite such discrepancies, all studies agree that
γ-producing CD4+ T cells, and there is a rise in the serum levels of this Treg cells are functionally impaired in people with asthma.
cytokine during acute attacks of asthma91,92. However, no studies of Interestingly, this defect in Treg cell function in allergic patients seems
humans have assessed the therapeutic effect of IFN-γ blockade. to affect only the regulation of TH2 responses. In the blood of HDM-
sensitized children, a population of Treg cells coexpressing the tran-
TH9 cells in asthma scription factors Foxp3 and GATA-3 could suppress the production
IL-9 was initially thought to be a TH2-specific cytokine induced by IL-2, TH1 cytokines but not that of TH2 cytokines by peripheral blood
IL-4 and TGF-β93. TH9 cells have been identified as a distinct helper mononuclear cells. This suggests that Treg cells in allergic patients
T cell subset94, and other studies have shown that additional stimuli might promote allergen-specific TH2 responses instead of control-
such TSLP, signaling via receptors of the Notch family or ligation of ling them. In the mouse, a similar Treg cell population unable to sup-
the receptor OX40 (refs. 95–97) also contribute to the differentiation press TH2 responses has been reported. These Treg cells express the

nature immunology  VOLUME 16  NUMBER 1  JANUARY 2015 49


review

Sensitization Challenge

Direct cDC Indirect cDC


activation activation

hi
CD11c monocytic
cells
TSLP, GM-CSF, IL-33, IL-25, IgE-mediated allergen
IL-1α, ATP, uric acid presentation
CCL2 CD11b+ cDC IL-13
CCL20 migration
β-defensins ILC2 TH2 TH2 PGD2
TH2 TH2
CCL17 Histamine
CCL22
Mast cell
Basophils
T H2
Recruitment and reactivation
TH2
Activation of cells of of TH2 cells
innate immune
system TH2
Recruitment of TH2

Kim Caesar/Nature Publishing Group


pre-cDCs
TH2

CD4 T H2
OX40L, Jagged1, IgE synthesis
IL-6, IL-23, LTC4 by B cells
Lack of IL-12 TH2
© 2015 Nature America, Inc. All rights reserved.

TFH
TH2 ILC2

Allergen presentation by
non professional APCs
TH2 and TH17 polarization

Figure 3  Epithelial cell–DC interactions during the sensitization and challenge phase of experimental asthma. Both lung cDCs and epithelial cells
express pattern-recognition receptors and can be activated directly by allergens. In response to allergens, lung epithelial cells produce chemokines
(CCL2 and CCL20) that attract immature pre-cDCs. Activated epithelial cells produce ‘instructive’ cytokines (for example, IL-1α, GM-CSF, IL-25,
IL-33 and TSLP) and danger signals (ATP and uric acid) that favor the maturation of CD11b + cDCs (which depend on the transcription factor IRF4
for maturation and migration). Activated lung CD11b + cDCs then migrate to the draining mediastinal lymph nodes, where they induce T H2 and TH17
responses. Migration is stimulated by IL-13 derived from ILC2 cells. Some helper T cells start producing IL-21 and adopt a follicular helper T cell (T FH)
fate to induce class switching to IgE in B lymphocytes. In the lymph nodes, DCs receive help from basophils to sustain T H2 responses. CD11chi DCs also
have a predominant role during the TH2 effector phase of asthma, when the lungs are repeatedly exposed to allergens (right). During allergen challenge,
poorly migratory CD11chi monocytic DCs and/or macrophages could locally restimulate effector function in lung-resident lymphocytes or they could
recruit effector TH2 cells through the production of CCL17 and CCL22. IgE-mediated allergen recognition enhances T H2 responses to inhaled
allergens, at least in humans.

inhibitory molecule TIGIT and produce fibrinogen-like protein 2, become clear that airway epithelial cells are crucial in controlling DC
which allows them to selectively spare TH2 responses while efficiently activation127. Many allergens, such as A. fumigatus spores, cockroach
npg

inhibiting TH1 and TH17 responses118. Whether this TIGIT+ popula- or HDMs, have protease activity. Proteases (such as papain or Der
tion and the GATA-3+Foxp3+ T cells identified in allergic children p 1) act on epithelial cells to decrease barrier function by cleaving
are related or not remains to be addressed. tight junction proteins and induce an innate cytokine response via
stimulation of protease-activated receptors. A. fumigatus protease
Epithelial cell–DC interactions in response to allergens activity leads to the formation of fibrinogen-cleavage products in
Whatever the endotype of asthma, the various T cells or ILCs that the fluid lining the lungs that activate Toll-like receptor 4 (TLR4)
control airway inflammation need to be activated. For T cell responses and in this way stimulate airway epithelial cells to produce IL-33
of all kinds, this involves the activation of antigen-presenting DCs that and TSLP, which activate DCs, ILC2 cells and basophils128. Cat dan-
recognize the allergen and present it to T cells in the lymph nodes der and HDM extracts contain allergens (Fel d 1 and Der p 1) that
that drain the lungs (Fig. 3). Various studies have shown that DCs are directly stimulate TLR4 on epithelial cells to produce IL-1α, IL25,
necessary and sufficient for inducing TH2 and TH17 adaptive immu- IL-33, TSLP and the cytokine GM-CSF129–131. After the inhalation
nity to inhaled allergens in mice that have not encountered the aller- of allergens, there is also the release of endogenous ‘danger signals’
gens previously119. There are different subtypes of DCs in the lungs, such as uric acid, ATP and HMGB1 (refs. 132–134). Allergens such
broadly categorized as conventional DCs (cDCs) and plasmacytoid as HDMs can also be immunogenic by triggering C-type lectin recep-
DCs (pDCs). Among cDCs, CD11b+CD172 (SIRP1α)+ cDCs that tors such as dectin-1 or dectin-2 on lung epithelial cells or DCs, which
depend on the transcription factor IRF4 are necessary and sufficient could contribute to mixed TH2 and TH17 immunity135,136. The epithe-
to induce allergic sensitization120–122, whereas CD103+XCR1+ cDCs lial cytokines induced by allergens or proteases induce TH2 immunity
that depend on the transcription factors IRF8 and BATF3 induce by activating CD11b+ cDCs (causing their migration and upregula-
tolerance to inhaled allergens123,124. pDCs also induce tolerance to tion of costimulatory molecules such as OX40L) and by activating
inhaled antigens by inducing Foxp3+ Treg cell responses to inhaled ILC2 cells and basophils that could be an important source of polar-
antigens125,126. There has been much progress in understanding of izing IL-4 and/or IL-13 for priming TH2 immunity and avoiding tol-
the process of allergic sensitization by DCs. From such studies it has erance to inhaled allergens54,130,137–139 (Fig. 3). When there is too

50 VOLUME 16  NUMBER 1  JANUARY 2015  nature immunology


review

much endotoxin in the inhaled allergen fraction, however, TH1 of mast cells have also been found between airway epithelial cells.
immunity ensues. Many of the human genome-wide association Bronchial epithelial cells exposed to IL-13 produce the mast cell
studies of asthma and atopy have found single-nucleotide polymor- growth factor SCF, which might explain the increase in intraepithe-
phisms in genes encoding molecules that control epithelial barrier lial mast cells150.
function (for example, filaggrin) and control the production or Increased numbers of basophils have been found in the blood and
responsiveness to epithelial cytokines, such as TSLP and IL1RL1 tissues of patients with asthma and mice exposed to HDMs. Basophils
(which encodes the receptor for IL-33); this indicates a crucial are an important source of IL-4 that could contribute to TH2-type
role for epithelial cells in asthma. Also, many of the environmental sensitization in trans by acting together with DCs151,152. In response to
risk factors for atopy and asthma, such as cigarette smoking, viral inhalation of papain, there is even cooperation between IL-33-activated
infections and air pollution, seem to converge on the epithelial basophils and ILC2 cells, in which basophil-derived IL-4 boosts the ter-
cell–DC interaction127. minal differentiation and activation of ILC2 cells148 (Fig. 3). However,
It has become increasingly clear that DCs and epithelial cells have although they are proposed to have antigen-presenting function,
an important role not only in sensitization to allergens but also in basophils are not proficient at stimulating naive CD4 + T cells,
ongoing asthma (Fig. 3). There is an increase in the abundance of and mice genetically modified to lack basophils (Mcpt8-Cre mice)
activated DCs in the airways of people with asthma and in mice with develop normal asthma features in response to OVA in alum153.
ongoing inflammation, and these form clusters with activated T cells Basophils could also have an important role in the effector and recall
around the airways and blood vessels in the lung140,141. Depletion response to allergens by producing lipid mediators and cytokines that
of CD11chi cells during ongoing allergen challenge suppresses the prime the vessel wall for extravasation and stimulate CD4+ effector
salient features of asthma142,143. The monocyte-derived DCs that cells directly154, and they could be involved in tissue remodeling.
accumulate in the airways have some features of macrophages, such Basophils are armed with the high-affinity IgE receptor and, just
as the expression of the Fc receptors CD64 and FcεRI, and are poorly like mast cells, could react immediately when there is reexposure to
© 2015 Nature America, Inc. All rights reserved.

migratory. They also express the chitinase-like protein Ym1, like alter- relevant allergen.
natively activated macrophages143. Their main function is the local Maybe the strongest indication of a role of the IgE in asthma is the
recruitment of effector T cells via the production of chemokines, and clinical efficacy of anti-IgE strategies in human asthma. The drug
thus they compartmentalize ongoing inflammation to the airways120. omalizumab (Xolair; Genentech/Novartis) is a humanized mono-
In ongoing disease, epithelial cells continue to fuel airway inflam- clonal antibody that binds to the constant domain of IgE, inhibit-
mation by activating incoming monocytes to adopt an immunogenic ing its interaction with the high-affinity receptor FcεRI, and has
phenotype and by producing chemokines and cytokines that activate proven efficacy in a subset of patients. Before this drug shows clini-
eosinophils, neutrophils and other cells of the innate immune system. cal effects in patients, it must be administered for a few weeks, which
Epithelial cells that go through repeated cycles of injury and repair suggests that the drug does not work by inhibiting the early degranu-
also contribute substantially to the process of airway wall remodeling lation of mast cells or basophils. Upon prolonged treatment, there is
via release of repair cytokines127. Such concepts are slowly making also a reduction in TH2 cytokine production in lung tissues, which
their way to clinical testing. AMG 157 (Amgen) is a human antibody suggests that it affects the generation of effector TH2 responses155.
to TSLP that has been shown to block the late asthmatic response and Monocytes and circulating DCs of allergic humans express FcεRI(αγ2)
bronchoconstriction in response to allergen challenge in humans, without FcεRIβ156,157. The targeting of allergens to FcεRI via IgE leads
and reduces eosinophil counts144. Human antibodies to GM-CSF and to an improvement of 1000-fold in the activation of CD4+ T cells
IL-1α have been developed for rheumatoid disease, but results for in vitro, most probably because FcεRI targets antigens to the MHC
asthma have not yet been reported. class II–rich endosomal compartment of DCs where peptides are
loaded158. Ligation of FcεRI on DCs also leads to the production of
npg

IgE effects on mast cell, basophil or DCs? CCL28, a chemokine that selectively attracts TH2 cells159. The effect of
IgE has the lowest concentration of all antibodies, and IgE is mainly triggering FcεRI on DCs has been studied in vivo only recently. DCs
known for its role in allergic disease, in which it has the potential in mice carrying a transgene expressing human FcεRI under control of
to activate mast cells and basophils when crosslinked on the high- the promoter of the gene encoding mouse integrin CD11c are at least
affinity IgE receptor FcεRI. In mast cells and basophils, this receptor 50 times more potent in presenting OVA antigen to CD4+ T cells when
complex is composed of FcεRIα and FcεRIβ and a dimer of FcεRIγ antigen-specific IgE is present160. Moreover, the presence of IgE on
chains (FcεRI(αβγ2)). The expression of FcεRI is heavily influ- DCs primes naive T cells for TH2 differentiation, and recall responses
enced by serum concentrations of IgE and also by the TH2 cytokine to OVA antigen are stronger in vivo. Although mouse cDCs do not
IL-4. The presence of antigen-specific IgE in the serum of patients express FcεRIα, mouse DCs recruited in response to viral lung infec-
with asthma leads to the immediate suggestion that mast cells tion or HDM exposure express FcεRIα in combination with FcεRIγ,
and basophils, armed with high-affinity IgE receptors, have an which leading to the expression of functional FcεRI(αγ2)120,151,161.
important role in the disease. However, studies of mast cell–deficient In mice, triggering of this receptor also leads to production of CCL28
mice have produced conflicting results. In models of mild asthma and to IL-13 production in T cells.
without the systemic adjuvant alum, mast cells have an important Although triggering via FcεRI can lead to cellular activation
proinflammatory role that is stimulated by IFN-γ145,146; in mod- dependent on the transcription factor NF-κB in DCs, it can also lead
els using alum or strong HDMs, mast cells are redundant147,148. In to the production of IL-10 and the tryptophan catabolizing enzyme
humans, the lack of clinical efficacy of antihistamines on the symp- indoleamine deoxygenase, which suggests that FcεRIα expression on
toms and signs of asthma has always been taken as evidence that mast DCs might also contribute to the inhibition of inflammation162. In
cells do not serve a big role in asthma. However, mast cells can infil- support of that, a study of mice that express human FcεRIα on DCs
trate the bronchial smooth muscle layer and contribute to BHR149, has demonstrated that ligation of IgE on DCs by relevant allergen
possibly driven by the mast cell growth factor IL-9 and release of suppresses the production of inflammatory chemokines that attract
leukotrienes. In the TH2hi endotype of asthma, increased numbers mast cells and suppresses asthma in mice163.

nature immunology  VOLUME 16  NUMBER 1  JANUARY 2015 51


review

Normal antiviral immunity to HRV Antiviral immunity to HRV in asthma

Interferon response IFN-α, IFN-β, Attachment to ICAM-1 Increased attachment


genes (ISG15) IFN-γ TLR3, TLR7, Mda5 triggering to ICAM-1

TGF-β Reduced IFN-α, Increased viral replication


Viral replication IFN-β, IFN-γ and shedding, cell death
CD103+ cDC
contained
CD8
pDC Ym1
Induction of protective CD11b+ cDC
adaptive immunity pDC
CCL2
CCL11 Alternatively TH2
IgE loaded TH17
activated
macrophage

Kim Caesar/Nature Publishing Group


TH2 TH17

Virus-induced asthma
exacerbation
TH17
TH2
Recruitment of
monocytic cells

Neutrophils
Eosinophils
© 2015 Nature America, Inc. All rights reserved.

Figure 4  Defective antiviral immunity in asthma. In the normal antiviral response to HRV (left), HRV attaches to epithelial cells by adhering to
ICAM-1. Following viral entry, TLRs and the cytosolic receptor Mda5 are triggered in lung epithelial cells to induce type I and type III interferon
responses. This leads to an antiviral state and controls viral replication, together with CD8 + T cell immunity generated by CD103+ cDCs. Other early
sources of type I interferons include pDCs and recruited monocytic cells. In asthma (right), ICAM-1 expression is upregulated, which leads to more entry
of virus. However, the induction of interferons is severely hampered. This could be intrinsic to asthma or might result from the suppressive functions
of eosinophils and alternatively activated macrophages, acting via Ym1 to suppress antiviral immunity in the epithelium. pDCs armed with IgE also fail
to produce type I interferons. The ensuing high viral load causes damage to the epithelium and thus causing locally present monocytic DCs and T H2 or
TH17 cells to cause airway inflammation rich in neutrophils and eosinophils.

Susceptibility to respiratory viral infection suppresses antiviral immunity as a secondary effect. In support
In most patients with asthma, exacerbations are caused by relatively of that proposal, triggering of FcεRI on pDCs by IgE crosslinking
mild respiratory viruses, such as human rhinovirus (HRV), respiratory interferes with the normal antiviral functions of pDCs and type I
syncytial virus or adenovirus, or by influenza virus. In childhood, interferon production178–180. When bronchial epithelial cells are
HRV is associated with persistent wheezing, which is a strong cultured with eosinophils, they are less able to produce type I and type III
risk factor for asthma later in life164,165. Experimental infection interferons in response to HRV, via eosinophil-derived TGF-β181
with HRV-16 in asthmatics causes neutrophilic inflammation (Fig. 4). The interference with antiviral immunity by type 2 immune
and steroid resistance166,167. In mice that express humanized integrin responses is probably an old adaptation response. Indeed, in the gut
ligand ICAM-1, infection with HRV can exacerbate disease168,169. immune system, a suppressive effect of helminth infection on antiviral
npg

Experimental models of infection with Sendai virus have revealed that immunity to norovirus has been described182. Alternative activation
respiratory viruses can induce an innate immune response that leads of macrophages via IL-13 caused upregulation of the chitinase-like
to the activation of natural killer T cells and alternative activation of protein Ym1, which suppresses antiviral immunity182. Chitinases and
macrophages that sustain airway pathology via IL-13 (ref. 170). chitinase-like proteins are strongly upregulated in mouse and human
In some asthma endotypes, virus-induced exacerbations are a major asthma and could be functionally linked to the diminished antiviral
problem despite adequate control of disease. The increased suscepti- state of people with asthma or to more neutrophil-rich asthma84,183.
bility to viruses could be caused by a primary defect in innate antiviral Whatever the cause of the defect, the first trials using inhaled inter-
immunity in some asthmatics, related to epithelial fragility and barrier ferons to treat viral exacerbations of asthma have been initiated 184.
disruption, or to defects in the production of TLR7 or interferon171 It is also possible to target the function of chitinase-like proteins.
(Fig. 4). In the lung, antiviral immunity relies heavily on produc-
tion of type I interferons (IFN-α and IFN-β) and type III interferons Conclusions
(IFN-λ, IL-28A, IL-28B and IL-29) and inducing interferon-stimulated Clinicians have begun to realize that asthma is a heterogeneous disease
genes, such as the gene expressing the the ubiquitin-like molecule with many endotypes. Various aspects of innate or adaptive immunity
ISG15. When exposed to HRV-16, epithelial cells from adult and to allergens, environmental triggers or viruses are involved in caus-
pediatric patients with asthma produce less IFN-β and IFN-λ and ing sensitization to allergens, symptoms of asthma, exacerbations,
have a less pronounced interferon-induced gene signature, particu- and response to therapies. There is extensive crosstalk between the
larly when the disease is poorly controlled172–174. It has been unclear airway epithelium and cells of the immune system in the initiation
whether this also causes an increased viral load in the lower respiratory and perpetuation of the disease. The results from the first intervention
tract, and whether defective interferon responses explain susceptibility trials in humans with asthma reflect the heterogeneity of this disease.
to all viruses175–177. An alternative explanation for the enhanced No single drug will be effective for all patients, but some drugs might
susceptibility to viral infection is that allergic sensitization 178 or be very effective in selected patients who are carefully identified on
the presence of TH2 cell–driven eosinophilic airway inflammation the basis of underlying immunological processes.

52 VOLUME 16  NUMBER 1  JANUARY 2015  nature immunology


review

Acknowledgments 28. Chu, D.K. et al. Indigenous enteric eosinophils control DCs to initiate
Supported by the European Union European Research Council (B.N.L.), the a primary TH2 immune response in vivo. J. Exp. Med. 211, 1657–1672
European Union Framework Programme 7 (MedALL and EUBIOPRED to B.N.L.), (2014).
29. van Rijt, L.S. et al. Airway eosinophils accumulate in the mediastinal lymph
the University of Ghent Multidisciplinary Research Platform (Group-ID, to B.N.L.)
nodes but lack antigen-presenting potential for naive T cells. J. Immunol. 171,
and Fonds Wetenschappelijk Onderzoek Vlaanderen (B.N.L. and H.H.). 3372–3378 (2003).
30. Song, D.J. et al. Anti-Siglec-F antibody reduces allergen-induced eosinophilic
COMPETING FINANCIAL INTERESTS inflammation and airway remodeling. J. Immunol. 183, 5333–5341 (2009).
The authors declare no competing financial interests. 31. Fattouh, R. et al. Eosinophils are dispensable for allergic remodeling and immunity
in a model of house dust mite-induced airway disease. Am. J. Respir. Crit. Care
Reprints and permissions information is available online at http://www.nature.com/ Med. 183, 179–188 (2011).
reprints/index.html. 32. Dworski, R., Simon, H.U., Hoskins, A. & Yousefi, S. Eosinophil and neutrophil
extracellular DNA traps in human allergic asthmatic airways. J. Allergy Clin.
1. Simpson, A. et al. Beyond atopy: multiple patterns of sensitization in relation to Immunol. 127, 1260–1266 (2011).
asthma in a birth cohort study. Am. J. Respir. Crit. Care Med. 181, 1200–1206 33. Yousefi, S., Simon, D. & Simon, H.U. Eosinophil extracellular DNA traps: molecular
(2010). mechanisms and potential roles in disease. Curr. Opin. Immunol. 24, 736–739
2. Spergel, J.M. & Paller, A.S. Atopic dermatitis and the atopic march. J. Allergy (2012).
Clin. Immunol. 112 (suppl.), S118–S127 (2003). 34. Ortega, H.G. et al. Mepolizumab treatment in patients with severe eosinophilic
3. Anderson, G.P. Endotyping asthma: new insights into key pathogenic mechanisms asthma. N. Engl. J. Med. 371, 1198–1207 (2014).
in a complex, heterogeneous disease. Lancet 372, 1107–1119 (2008). 35. Bel, E.H. et al. Oral glucocorticoid-sparing effect of mepolizumab in eosinophilic
4. Woodruff, P.G. et al. T-helper type 2-driven inflammation defines major asthma. N. Engl. J. Med. 371, 1189–1197 (2014).
subphenotypes of asthma. Am. J. Respir. Crit. Care Med. 180, 388–395 36. Flood-Page, P. et al. Anti-IL-5 treatment reduces deposition of ECM proteins in
(2009). the bronchial subepithelial basement membrane of mild atopic asthmatics.
5. Wu, W. et al. Unsupervised phenotyping of Severe Asthma Research Program J. Clin. Invest. 112, 1029–1036 (2003).
participants using expanded lung data. J. Allergy Clin. Immunol. 133, 1280–1288 37. Laviolette, M. et al. Effects of benralizumab on airway eosinophils in asthmatic
(2014). patients with sputum eosinophilia. J. Allergy Clin. Immunol. 132, 1086–1096
6. Wenzel, S.E. Asthma phenotypes: the evolution from clinical to molecular (2013).
approaches. Nat. Med. 18, 716–725 (2012). 38. Fallon, P.G. et al. Identification of an interleukin (IL)-25-dependent cell population
© 2015 Nature America, Inc. All rights reserved.

7. Brusselle, G.G., Joos, G.F. & Bracke, K.R. New insights into the immunology of that provides IL-4, IL-5, and IL-13 at the onset of helminth expulsion.
chronic obstructive pulmonary disease. Lancet 378, 1015–1026 (2011). J. Exp. Med. 203, 1105–1116 (2006).
8. De Monchy, J.G.R. et al. Bronchoalveolar eosinophilia during allergen-induced 39. Fort, M.M. et al. IL-25 induces IL-4, IL-5, and IL-13 and TH2-associated
late asthmatic reactions. Am. Rev. Respir. Dis. 131, 373–376 (1985). pathologies in vivo. Immunity 15, 985–995 (2001).
9. Humbert, M., Durham, S.R. & Ying, S. IL-4 and IL-5 mRNA and protein in 40. Kang, Z. et al. Epithelial cell-specific Act1 adaptor mediates interleukin-
bronchial biopsies from patients with atopic and non-atopic asthma: evidence 25-dependent helminth expulsion through expansion of Lin-c-Kit+ innate cell
against “intrinsic” asthma being a distinct immunopathologic entity. Am. J. population. Immunity 36, 821–833 (2012).
Respir. Crit. Care Med. 154, 1497–1504 (1996). 41. Walker, J.A., Barlow, J.L. & McKenzie, A.N. Innate lymphoid cells–how did we
10. Bentley, A.M. et al. Activated T-lymphocytes and eosinophils in the bronchial mucosa miss them? Nat. Rev. Immunol. 13, 75–87 (2013).
in isocyanate-induced asthma. J. Allergy Clin. Immunol. 89, 821–829 (1992). 42. Moro, K. et al. Innate production of TH2 cytokines by adipose tissue-associated
11. Bousquet, J. et al. Eosinophilic inflammation in asthma. N. Engl. J. Med. 323, c-Kit+Sca-1+ lymphoid cells. Nature 463, 540–544 (2010).
1033–1039 (1990). 43. Barlow, J.L. et al. Innate IL-13-producing nuocytes arise during allergic lung
12. Robinson, D.S. et al. Predominant Th2-like bronchoalveolar T lymphocyte inflammation and contribute to airways hyperreactivity. J. Allergy Clin. Immunol.
population in atopic asthma. N. Engl. J. Med. 326, 298–304 (1992). 129, 191–198 (2012).
13. Woodruff, P.G. et al. Genome-wide profiling identifies epithelial cell genes 44. Bartemes, K.R. et al. IL-33-responsive lineage- CD25+ CD44hi lymphoid cells
associated with asthma and with treatment response to corticosteroids. Proc. Natl. mediate innate type 2 immunity and allergic inflammation in the lungs.
Acad. Sci. USA 104, 15858–15863 (2007). J. Immunol. 188, 1503–1513 (2012).
14. Cheng, D. et al. Epithelial interleukin-25 is a key mediator in TH2-high, 45. Yasuda, K. et al. Contribution of IL-33-activated type II innate lymphoid cells to
corticosteroid-responsive asthma. Am. J. Respir. Crit. Care Med. 190, 639–648 pulmonary eosinophilia in intestinal nematode-infected mice. Proc. Natl. Acad.
(2014). Sci. USA 109, 3451–3456 (2012).
15. Cohn, L., Homer, R.J., Niu, N. & Bottomly, K. T helper 1 cells and interferon 46. Halim, T.Y., Krauss, R.H., Sun, A.C. & Takei, F. Lung natural helper cells are a
gamma regulate allergic airway inflammation and mucus production. J. Exp. Med. critical source of TH2 cell-type cytokines in protease allergen-induced airway
190, 1309–1318 (1999). inflammation. Immunity 36, 451–463 (2012).
16. Cohn, L. et al. TH2-induced airway mucus production is dependent on IL-4Rα, 47. Doherty, T.A. et al. Lung type 2 innate lymphoid cells express cysteinyl leukotriene
but not on eosinophils. J. Immunol. 162, 6178–6183 (1999). receptor 1, which regulates TH2 cytokine production. J. Allergy Clin. Immunol.
npg

17. Cohn, L., Homer, R.J., Marinov, A., Rankin, J. & Bottomly, K. Induction of airway 132, 205–213 (2013).
mucus production By T helper 2 (TH2) cells: a critical role for interleukin 4 in 48. Chang, Y.J. et al. Innate lymphoid cells mediate influenza-induced airway hyper-
cell recruitment but not mucus production. J. Exp. Med. 186, 1737–1747 reactivity independently of adaptive immunity. Nat. Immunol. 12, 631–638
(1997). (2011).
18. Doherty, T.A., Soroosh, P., Broide, D.H. & Croft, M. CD4+ cells are required for 49. Hong, J.Y. et al. Neonatal rhinovirus induces mucous metaplasia and airways
chronic eosinophilic lung inflammation but not airway remodeling. Am. J. Physiol. hyperresponsiveness through IL-25 and type 2 innate lymphoid cells. J. Allergy
Lung Cell. Mol. Physiol. 296, L229–L235 (2009). Clin. Immunol. 134, 429–439 (2014).
19. Gavett, S.H. et al. Interleukin 12 inhibits antigen-induced airway 50. Gorski, S.A., Hahn, Y.S. & Braciale, T.J. Group 2 innate lymphoid cell production
hyperresponsiveness, inflammation, and TH2 cytokine expression in mice. of IL-5 is regulated by NKT cells during influenza virus infection. PLoS Pathog.
J. Exp. Med. 182, 1527–1536 (1995). 9, e1003615 (2013).
20. Brusselle, G.G. et al. Attenuation of allergic airway inflammation in IL-4 deficient 51. Klein Wolterink, R.G. et al. Pulmonary innate lymphoid cells are major producers
mice. Clin. Exp. Allergy 24, 73–80 (1994). of IL-5 and IL-13 in murine models of allergic asthma. Eur. J. Immunol. 42,
21. Corry, D.B. et al. Interleukin 4, but not interleukin 5 or eosinophils, is required 1106–1116 (2012).
in a murine model of acute airway hyperreactivity. J. Exp. Med. 183, 109–117 52. Gold, M.J. et al. Group 2 innate lymphoid cells facilitate sensitization to local,
(1996). but not systemic, TH2-inducing allergen exposures. J. Allergy Clin. Immunol. 133,
22. Wills-Karp, M. et al. Interleukin-13: central mediator of allergic asthma. Science 1142–1148 (2014).
282, 2258–2261 (1998). 53. Halim, T.Y. et al. Retinoic-acid-receptor-related orphan nuclear receptor alpha is
23. Grünig, G. et al. Requirement for IL-13 independently of IL-4 in experimental required for natural helper cell development and allergic inflammation. Immunity
asthma. Science [see comments] 282, 2261–2263 (1998). 37, 463–474 (2012).
24. Wenzel, S. et al. Dupilumab in persistent asthma with elevated eosinophil levels. 54. Halim, T.Y. et al. Group 2 innate lymphoid cells are critical for the initiation of
N. Engl. J. Med. 368, 2455–2466 (2013). adaptive T helper 2 cell-mediated allergic lung inflammation. Immunity 40,
25. Corren, J. et al. Lebrikizumab treatment in adults with asthma. N. Engl. J. Med. 425–435 (2014).
365, 1088–1098 (2011). 55. Xue, L. et al. Prostaglandin D2 activates group 2 innate lymphoid cells through
26. Coyle, A.J., Perretti, F., Manzini, S. & Irvin, C.G. Cationic protein-induced sensory chemoattractant receptor-homologous molecule expressed on TH2 cells. J. Allergy
nerve activation: role of substance P in airway hyperresponsiveness and plasma Clin. Immunol. 133, 1184–1194 (2014).
protein extravasation. J. Clin. Invest. 94, 2301–2306 (1994). 56. Doherty, T.A. et al. STAT6 regulates natural helper cell proliferation during lung
27. Coyle, A.J., Ackerman, S.J., Burch, R., Proud, D. & Irvin, C.G. Human eosinophil- inflammation initiated by Alternaria. Am. J. Physiol. Lung Cell. Mol. Physiol. 303,
granule major basic protein and synthetic polycations induce airway L577–L588 (2012).
hyperresponsiveness in vivo dependent on bradykinin generation. J. Clin. Invest. 57. Barnig, C. et al. Lipoxin A4 regulates natural killer cell and type 2 innate lymphoid
95, 1735–1740 (1995). cell activation in asthma. Sci. Transl. Med. 5, 174ra126 (2013).

nature immunology  VOLUME 16  NUMBER 1  JANUARY 2015 53


review

58. Juncadella, I.J. et al. Apoptotic cell clearance by bronchial epithelial cells 89. Hessel, E.M. et al. Development of airway hyperresponsiveness is dependent on
critically influences airway inflammation. Nature 493, 547–551 (2013). interferon-γ and independent of eosinophil infiltration. Am. J. Respir. Cell Mol.
59. McSorley, H.J., Blair, N.F., Smith, K.A., McKenzie, A.N. & Maizels, R.M. Blockade Biol. 16, 325–334 (1997).
of IL-33 release and suppression of type 2 innate lymphoid cell responses by 90. Sugimoto, T. et al. Interleukin 18 acts on memory T helper cells type 1 to induce
helminth secreted products in airway allergy. Mucosal Immunol. 7, 1068–1078 airway inflammation and hyperresponsiveness in a naive host mouse. J. Exp. Med.
(2014). 199, 535–545 (2004).
60. Van Dyken, S.J. et al. Chitin activates parallel immune modules that direct distinct 91. Krug, N. et al. T-cell cytokine profile evaluated at the single cell level in BAL
inflammatory responses via innate lymphoid type 2 and γδ T cells. Immunity 40, and blood in allergic asthma. Am. J. Respir. Cell Mol. Biol. 14, 319–326
414–424 (2014). (1996).
61. Oliphant, C.J. et al. MHCII-mediated dialog between group 2 innate lymphoid 92. Corrigan, C.J. & Kay, A.B. CD4 T-lymphocyte activation in acute severe asthma.
cells and CD4+ T cells potentiates type 2 immunity and promotes parasitic Am. Rev. Respir. Dis. 141, 970–977 (1990).
helminth expulsion. Immunity 41, 283–295 (2014). 93. Schmitt, E. et al. IL-9 production of naive CD4+ T cells depends on IL-2, is
62. Zaiss, D.M. et al. Amphiregulin, a TH2 cytokine enhancing resistance to nematodes. synergistically enhanced by a combination of TGF- and IL-4, and is inhibited by
Science 314, 1746 (2006). IFN-γ. J. Immunol. 153, 3989–3996 (1994).
63. Monticelli, L.A. et al. Innate lymphoid cells promote lung-tissue homeostasis after 94. O’Garra, A., Stockinger, B. & Veldhoen, M. Differentiation of human TH17 cells
infection with influenza virus. Nat. Immunol. 12, 1045–1054 (2011). does require TGF-β!. Nat. Immunol. 9, 588–590 (2008).
64. Porter, P.C. et al. Airway surface mycosis in chronic TH2-associated airway disease. 95. Yao, W. et al. Interleukin-9 is required for allergic airway inflammation mediated
J. Allergy Clin. Immunol. 134, 325–331 (2014). by the cytokine TSLP. Immunity 38, 360–372 (2013).
65. Mjösberg, J. et al. The transcription factor GATA3 is essential for the function of 96. Elyaman, W. et al. Notch receptors and Smad3 signaling cooperate in the
human type 2 innate lymphoid cells. Immunity 37, 649–659 (2012). induction of interleukin-9-producing T cells. Immunity 36, 623–634 (2012).
66. Mjösberg, J.M. et al. Human IL-25- and IL-33-responsive type 2 innate lymphoid 97. Xiao, X. et al. OX40 signaling favors the induction of TH9 cells and airway
cells are defined by expression of CRTH2 and CD161. Nat. Immunol. 12, inflammation. Nat. Immunol. 13, 981–990 (2012).
1055–1062 (2011). 98. Chang, H.C. et al. The transcription factor PU.1 is required for the development
67. Kabata, H. et al. Thymic stromal lymphopoietin induces corticosteroid resistance of IL-9-producing T cells and allergic inflammation. Nat. Immunol. 11, 527–534
in natural helper cells during airway inflammation. Nat. Commun. 4, 2675 (2011).
(2013). 99. Liao, W. et al. Opposing actions of IL-2 and IL-21 on TH9 differentiation correlate
68. McKinley, L. et al. TH17 cells mediate steroid-resistant airway inflammation with their differential regulation of BCL6 expression. Proc. Natl. Acad. Sci. USA
and airway hyperresponsiveness in mice. J. Immunol. 181, 4089–4097 111, 3508–3513 (2014).
© 2015 Nature America, Inc. All rights reserved.

(2008). 100. Yang, X.O. et al. The signaling suppressor CIS controls proallergic T cell
69. Shaw, D.E. et al. Association between neutrophilic airway inflammation and airflow development and allergic airway inflammation. Nat. Immunol. 14, 732–740
limitation in adults with asthma. Chest 132, 1871–1875 (2007). (2013).
70. Manni, M.L. et al. The complex relationship between inflammation and lung 101. Staudt, V. et al. Interferon-regulatory factor 4 is essential for the developmental
function in severe asthma. Mucosal Immunol. 7, 1186–1198 (2014). program of T helper 9 cells. Immunity 33, 192–202 (2010).
71. Schnyder-Candrian, S. et al. Interleukin-17 is a negative regulator of established 102. Kearley, J. et al. IL-9 governs allergen-induced mast cell numbers in the lung
allergic asthma. J. Exp. Med. 203, 2715–2725 (2006). and chronic remodeling of the airways. Am. J. Respir. Crit. Care Med. 183,
72. Wakashin, H. et al. IL-23 and TH17 cells enhance TH2-cell-mediated eosinophilic 865–875 (2011).
airway inflammation in mice. Am. J. Respir. Crit. Care Med. 178, 1023–1032 103. Erpenbeck, V.J. et al. Segmental allergen challenge in patients with atopic asthma
(2008). leads to increased IL-9 expression in bronchoalveolar lavage fluid lymphocytes.
73. Besnard, A.G. et al. Dual role of IL-22 in allergic airway inflammation and its J. Allergy Clin. Immunol. 111, 1319–1327 (2003).
cross-talk with IL-17A. Am. J. Respir. Crit. Care Med. 183, 1153–1163 104. Kerzerho, J. et al. Programmed cell death ligand 2 regulates TH9 differentiation
(2011). and induction of chronic airway hyperreactivity. J. Allergy Clin. Immunol. 131,
74. Brandt, E.B. et al. Diesel exhaust particle induction of IL-17A contributes to 1048–1057 (2013).
severe asthma. J. Allergy Clin. Immunol. 132, 1194–1204 (2013). 105. Wilhelm, C. et al. An IL-9 fate reporter demonstrates the induction of an innate
75. Bellini, A. et al. Interleukin (IL)-4, IL-13, and IL-17A differentially affect the IL-9 response in lung inflammation. Nat. Immunol. 12, 1071–1077 (2011).
profibrotic and proinflammatory functions of fibrocytes from asthmatic patients. 106. Turner, J.E. et al. IL-9-mediated survival of type 2 innate lymphoid cells promotes
Mucosal Immunol. 5, 140–149 (2011). damage control in helminth-induced lung inflammation. J. Exp. Med. 210,
76. Zhao, J., Lloyd, C.M. & Noble, A. TH17 responses in chronic allergic airway 2951–2965 (2013).
inflammation abrogate regulatory T-cell-mediated tolerance and contribute to 107. Parker, J.M. et al. Safety profile and clinical activity of multiple subcutaneous
airway remodeling. Mucosal Immunol. 6, 335–346 (2012). doses of MEDI-528, a humanized anti-interleukin-9 monoclonal antibody, in two
77. Kudo, M. et al. IL-17A produced by αβ T cells drives airway hyper-responsiveness randomized phase 2a studies in subjects with asthma. BMC Pulm. Med. 11, 14
in mice and enhances mouse and human airway smooth muscle contraction. (2011).
Nat. Med. 18, 547–554 (2012). 108. Ostroukhova, M. et al. Tolerance induced by inhaled antigen involves CD4+
78. Busse, W.W. et al. Randomized, double-blind, placebo-controlled study of T cells expressing membrane-bound TGF-β and FOXP3. J. Clin. Invest. 114,
npg

brodalumab, a human anti-IL-17 receptor monoclonal antibody, in moderate to 28–38 (2004).


severe asthma. Am. J. Respir. Crit. Care Med. 188, 1294–1302 (2013). 109. Josefowicz, S.Z. et al. Extrathymically generated regulatory T cells control mucosal
79. Berry, M.A. et al. Evidence of a role of tumor necrosis factor α in refractory TH2 inflammation. Nature 482, 395–399 (2012).
asthma. N. Engl. J. Med. 354, 697–708 (2006). 110. Weiss, J.M. et al. Neuropilin 1 is expressed on thymus-derived natural regulatory
80. Wenzel, S.E. et al. A randomized, double-blind, placebo-controlled study of tumor T cells, but not mucosa-generated induced Foxp3+ T reg cells. J. Exp. Med. 209,
necrosis factor-α blockade in severe persistent asthma. Am. J. Respir. Crit. Care 1723–1742 (2012).
Med. 179, 549–558 (2009). 111. Lewkowich, I.P. et al. CD4+CD25+ T cells protect against experimentally induced
81. Morjaria, J.B. et al. The role of a soluble TNFα receptor fusion protein (etanercept) asthma and alter pulmonary dendritic cell phenotype and function. J. Exp. Med.
in corticosteroid refractory asthma: a double blind, randomised, placebo controlled 202, 1549–1561 (2005).
trial. Thorax 63, 584–591 (2008). 112. Huang, M.T. et al. Regulatory T cells negatively regulate neovasculature of airway
82. Porter, P.C. et al. Necessary and sufficient role for T helper cells to prevent remodeling via DLL4-Notch signaling. J. Immunol. 183, 4745–4754 (2009).
fungal dissemination in allergic lung disease. Infect. Immun. 79, 4459–4471 113. Whitehead, G.S. et al. IL-35 production by inducible costimulator (ICOS)-positive
(2011). regulatory T cells reverses established IL-17-dependent allergic airways disease.
83. Wang, Y.H. et al. A novel subset of CD4+ TH2 memory/effector cells that produce J. Allergy Clin. Immunol. 129, 207–215 (2012).
inflammatory IL-17 cytokine and promote the exacerbation of chronic allergic 114. Mamessier, E. et al. T-cell activation during exacerbations: a longitudinal study
asthma. J. Exp. Med. 207, 2479–2491 (2010). in refractory asthma. Allergy 63, 1202–1210 (2008).
84. Sutherland, T.E. et al. Chitinase-like proteins promote IL-17 mediated 115. Hartl, D. et al. Quantitative and functional impairment of pulmonary CD4+CD25hi
neutrophilia in a tradeoff between nematode killing and host damage. regulatory T cells in pediatric asthma. J. Allergy Clin. Immunol. 119, 1258–1266
Nat. Immunol. 15, 1116–1125 (2014). (2007).
85. Irvin, C. et al. Increased frequency of dual-positive T2/T17 cells in bronchoalveolar 116. Smyth, L.J., Eustace, A., Kolsum, U., Blaikely, J. & Singh, D. Increased airway
lavage fluid characterizes a population of patients with severe asthma. J. Allergy T regulatory cells in asthmatic subjects. Chest 138, 905–912 (2010).
Clin. Immunol. doi:10.1016/j.jaci.2014.05.038 (18 July 2014). 117. Barczyk, A. et al. Decreased percentage of CD4+Foxp3+TGF-β+ and increased
86. Randolph, D.A., Stephens, R., Carruthers, C.J. & Chaplin, D.D. Cooperation percentage of CD4+IL-17+ cells in bronchoalveolar lavage of asthmatics.
between TH1 and TH2 cells in a murine model of eosinophilic airway inflammation. J. Inflamm. (Lond.) 11, 22 (2014).
J. Clin. Invest. 104, 1021–1029 (1999). 118. Joller, N. et al. Treg cells expressing the coinhibitory molecule TIGIT selectively
87. Hansen, G., Berry, G., Dekruyff, R.H. & Umetsu, D.T. Allergen-specific TH1 cells inhibit proinflammatory TH1 and TH17 cell responses. Immunity 40, 569–581
fail to counterbalance TH2 cell-induced airway hyperreactivity but cause severe (2014).
airway inflammation. J. Clin. Invest. 103, 175–183 (1999). 119. Lambrecht, B.N. & Hammad, H. Lung dendritic cells in respiratory viral infection
88. Ford, J.G. et al. IL-13 and IFN-γ: Interactions in lung inflammation. J. Immunol. and asthma: from protection to immunopathology. Annu. Rev. Immunol. 30,
167, 1769–1777 (2001). 243–270 (2012).

54 VOLUME 16  NUMBER 1  JANUARY 2015  nature immunology


review

120. Plantinga, M. et al. Conventional and monocyte-derived CD11b+ dendritic cells 150. Dougherty, R.H. et al. Accumulation of intraepithelial mast cells with a unique
initiate and maintain T helper 2 cell-mediated immunity to house dust mite protease phenotype in TH2-high asthma. J. Allergy Clin. Immunol. 125,
allergen. Immunity 38, 322–335 (2013). 1046–1053 (2010).
121. Williams, J.W. et al. Transcription factor IRF4 drives dendritic cells to promote 151. Hammad, H. et al. Inflammatory dendritic cells–not basophils–are necessary and
TH2 differentiation. Nat. Commun. 4, 2990 (2013). sufficient for induction of TH2 immunity to inhaled house dust mite allergen.
122. Raymond, M. et al. Selective control of SIRP-α-positive airway dendritic cell J. Exp. Med. 207, 2097–2111 (2010).
trafficking through CD47 is critical for the development of TH2-mediated allergic 152. Tang, H. et al. The T helper type 2 response to cysteine proteases requires
inflammation. J. Allergy Clin. Immunol. 124, 1333–1342 (2009). dendritic cell-basophil cooperation via RAS-mediated signaling. Nat. Immunol.
123. Semmrich, M. et al. Directed antigen targeting in vivo identifies a role for CD103+ 11, 608–617 (2010).
dendritic cells in both tolerogenic and immunogenic T-cell responses. 153. Ohnmacht, C. et al. Basophils orchestrate chronic allergic dermatitis and protective
Mucosal Immunol. 5, 150–160 (2012). immunity against helminths. Immunity 33, 364–374 (2010).
124. Khare, A. et al. Cutting edge: inhaled antigen upregulates retinaldehyde 154. Wakahara, K. et al. Basophils are recruited to inflamed lungs and exacerbate
dehydrogenase in lung CD103+ but not plasmacytoid dendritic cells to induce memory TH2 responses in mice and humans. Allergy 68, 180–189 (2013).
Foxp3 de novo in CD4+ T cells and promote airway tolerance. J. Immunol. 191, 155. Holgate, S., Smith, N., Massanari, M. & Jimenez, P. Effects of omalizumab on
25–29 (2013). markers of inflammation in patients with allergic asthma. Allergy 64, 1728–1736
125. de Heer, H.J. et al. Essential role of lung plasmacytoid dendritic cells in preventing (2009).
asthmatic reactions to harmless inhaled antigen. J. Exp. Med. 200, 89–98 156. Maurer, D. et al. Peripheral blood dendritic cells express Fc ε RI as a complex
(2004). composed of Fc ε RI α- and Fc ε RI γ-chains and can use this receptor for
126. Kool, M. et al. An anti-inflammatory role for plasmacytoid dendritic cells in allergic IgE-mediated allergen presentation. J. Immunol. 157, 607–616 (1996).
airway inflammation. J. Immunol. 183, 1074–1082 (2009). 157. Bieber, T. et al. Human epidermal Langerhans cells express the high affinity
127. Lambrecht, B.N. & Hammad, H. The airway epithelium in asthma. Nat. Med. 18, receptor for immunoglobulin E (Fc ε RI). J. Exp. Med. 175, 1285–1290
684–692 (2012). (1992).
128. Millien, V.O. et al. Cleavage of fibrinogen by proteinases elicits allergic responses 158. Maurer, D. et al. Fc ε receptor I on dendritic cells delivers IgE-bound multivalent
through Toll-like receptor 4. Science 341, 792–796 (2013). antigens into a cathepsin S-dependent pathway of MHC class II presentation.
129. Hammad, H. et al. House dust mite allergen induces asthma via Toll-like J. Immunol. 161, 2731–2739 (1998).
receptor 4 triggering of airway structural cells. Nat. Med. 15, 410–416 159. Khan, S.H. & Grayson, M.H. Cross-linking IgE augments human conventional
(2009). dendritic cell production of CC chemokine ligand 28. J. Allergy Clin. Immunol.
130. Willart, M.A. et al. Interleukin-1α controls allergic sensitization to inhaled house 125, 265–267 (2010).
© 2015 Nature America, Inc. All rights reserved.

dust mite via the epithelial release of GM-CSF and IL-33. J. Exp. Med. 209, 160. Sallmann, E. et al. High-affinity IgE receptors on dendritic cells exacerbate
1505–1517 (2012). TH2-dependent inflammation. J. Immunol. 187, 164–171 (2011).
131. Herre, J. et al. Allergens as immunomodulatory proteins: the cat dander protein 161. Grayson, M.H. et al. Induction of high-affinity IgE receptor on lung dendritic cells
Fel d 1 enhances TLR activation by lipid ligands. J. Immunol. 191, 1529–1535 during viral infection leads to mucous cell metaplasia. J. Exp. Med. 204,
(2013). 2759–2769 (2007).
132. Kool, M. et al. An unexpected role for uric acid as an inducer of T helper 2 cell 162. Novak, N., Bieber, T. & Katoh, N. Engagement of Fc ε RI on human monocytes
immunity to inhaled antigens and inflammatory mediator of allergic asthma. induces the production of IL-10 and prevents their differentiation in dendritic
Immunity 34, 527–540 (2011). cells. J. Immunol. 167, 797–804 (2001).
133. Idzko, M. et al. Extracellular ATP triggers and maintains asthmatic 163. Platzer, B. et al. Dendritic cell-bound IgE functions to restrain allergic inflammation
airway inflammation by activating dendritic cells. Nat. Med. 13, 913–919 at mucosal sites. Mucosal Immunol. doi:10.1038/mi.2014.85 (17 September
(2007). 2014).
134. Ullah, M.A. et al. Receptor for advanced glycation end products and its ligand 164. Jackson, D.J. et al. Wheezing rhinovirus illnesses in early life predict asthma
high-mobility group box-1 mediate allergic airway sensitization and airway development in high-risk children. Am. J. Respir. Crit. Care Med. 178, 667–672
inflammation. J. Allergy Clin. Immunol. 134, 440–450 (2014). (2008).
135. Barrett, N.A. et al. Dectin-2 mediates TH2 immunity through the generation of 165. Calışkan, M. et al. Rhinovirus wheezing illness and genetic risk of childhood-onset
cysteinyl leukotrienes. J. Exp. Med. 208, 593–604 (2011). asthma. N. Engl. J. Med. 368, 1398–1407 (2013).
136. Norimoto, A. et al. Dectin-2 promotes house dust mite-induced TH2 and TH17 166. Papi, A. et al. Rhinovirus infection causes steroid resistance in airway epithelium
cell differentiation and allergic airway inflammation in mice. Am. J. Respir. Cell through nuclear factor κB and c-Jun N-terminal kinase activation. J. Allergy Clin.
Mol. Biol. 51, 201–209 (2014). Immunol. 132, 1075–1085 (2013).
137. Besnard, A.G. et al. IL-33-activated dendritic cells are critical for allergic airway 167. Zhu, J. et al. Airway inflammation and illness severity in response to experimental
inflammation. Eur. J. Immunol. 41, 1675–1686 (2011). rhinovirus infection in asthma. Chest 145, 1219–1229 (2014).
138. Bell, B.D. et al. The transcription factor STAT5 is critical in dendritic cells for 168. Bartlett, N.W. et al. Mouse models of rhinovirus-induced disease and exacerbation
the development of TH2 but not TH1 responses. Nat. Immunol. 14, 364–371 of allergic airway inflammation. Nat. Med. 14, 199–204 (2008).
(2013). 169. Collison, A. et al. The E3 ubiquitin ligase midline 1 promotes allergen and
139. Chu, D.K. et al. IL-33, but not thymic stromal lymphopoietin or IL-25, is central rhinovirus-induced asthma by inhibiting protein phosphatase 2A activity.
npg

to mite and peanut allergic sensitization. J. Allergy Clin. Immunol. 131, 187–200 Nat. Med. 19, 232–237 (2013).
(2013). 170. Kim, E.Y. et al. Persistent activation of an innate immune response translates
140. Huh, J.C. et al. Bidirectional interactions between antigen-bearing respiratory respiratory viral infection into chronic lung disease. Nat. Med. 14, 633–640
tract dendritic cells (DCs) and T cells precede the late phase reaction in (2008).
experimental asthma: DC activation occurs in the airway mucosa but not in the 171. Kaiko, G.E. et al. Toll-like receptor 7 gene deficiency and early-life Pneumovirus
lung parenchyma. J. Exp. Med. 198, 19–30 (2003). infection interact to predispose toward the development of asthma-like pathology
141. Thornton, E.E. et al. Spatiotemporally separated antigen uptake by alveolar in mice. J. Allergy Clin. Immunol. 131, 1331–1339 (2013).
dendritic cells and airway presentation to T cells in the lung. J. Exp. Med. 209, 172. Wark, P.A. et al. Asthmatic bronchial epithelial cells have a deficient innate
1183–1199 (2012). immune response to infection with rhinovirus. J. Exp. Med. 201, 937–947
142. van Rijt, L.S. et al. in vivo depletion of lung CD11c+ dendritic cells during allergen (2005).
challenge abrogates the characteristic features of asthma. J. Exp. Med. 201, 173. Contoli, M. et al. Role of deficient type III interferon-λ production in asthma
981–991 (2005). exacerbations. Nat. Med. 12, 1023–1026 (2006).
143. van Rijt, L.S. et al. Persistent activation of dendritic cells after resolution of 174. Edwards, M.R. et al. Impaired innate interferon induction in severe therapy
allergic airway inflammation breaks tolerance to inhaled allergens in mice. resistant atopic asthmatic children. Mucosal Immunol. 6, 797–806 (2013).
Am. J. Respir. Crit. Care Med. 184, 303–311 (2011). 175. Kennedy, J.L. et al. Comparison of viral load in individuals with and without
144. Gauvreau, G.M. et al. Effects of an anti-TSLP antibody on allergen-induced asthma during infections with rhinovirus. Am. J. Respir. Crit. Care Med. 189,
asthmatic responses. N. Engl. J. Med. 370, 2102–2110 (2014). 532–539 (2014).
145. Yu, M. et al. Identification of an IFN-γ/mast cell axis in a mouse model of chronic 176. Bochkov, Y.A. et al. Rhinovirus-induced modulation of gene expression in bronchial
asthma. J. Clin. Invest. 121, 3133–3143 (2011). epithelial cells from subjects with asthma. Mucosal Immunol. 3, 69–80
146. Heger, K. et al. A20-deficient mast cells exacerbate inflammatory responses (2010).
in vivo. PLoS Biol. 12, e1001762 (2014). 177. Spann, K.M. et al. Viral and host factors determine innate immune responses in
147. Williams, C.M. & Galli, S.J. Mast cells can amplify airway reactivity and features airway epithelial cells from children with wheeze and atopy. Thorax 69, 918–925
of chronic inflammation in an asthma model in mice. J. Exp. Med. 192, 455–462 (2014).
(2000). 178. Durrani, S.R. et al. Innate immune responses to rhinovirus are reduced by the
148. Motomura, Y. et al. Basophil-derived interleukin-4 controls the function of natural high-affinity IgE receptor in allergic asthmatic children. J. Allergy Clin. Immunol.
helper cells, a member of ILC2s, in lung inflammation. Immunity 40, 758–771 130, 489–495 (2012).
(2014). 179. Gill, M.A. et al. Counterregulation between the FcεRI pathway and antiviral
149. Brightling, C.E. et al. Mast-cell infiltration of airway smooth muscle in asthma. responses in human plasmacytoid dendritic cells. J. Immunol. 184, 5999–6006
N. Engl. J. Med. 346, 1699–1705 (2002). (2010).

nature immunology  VOLUME 16  NUMBER 1  JANUARY 2015 55


review

180. Tversky, J.R. et al. Human blood dendritic cells from allergic subjects have 183. Ober, C. et al. Effect of variation in CHI3L1 on serum YKL-40 level, risk of
impaired capacity to produce interferon-α via Toll-like receptor 9. Clin. Exp. Allergy asthma, and lung function. N. Engl. J. Med. 358, 1682–1691 (2008).
38, 781–788 (2008). 184. Djukanović, R. et al. The effect of inhaled IFN-β on worsening of asthma symptoms
181. Mathur, S.K. et al. Interaction between allergy and innate immunity: model for caused by viral infections. A randomized trial. Am. J. Respir. Crit. Care Med.
eosinophil regulation of epithelial cell interferon expression. Ann. Allergy Asthma 190, 145–154 (2014).
Immunol. 111, 25–31 (2013). 185. Brusselle, G.G., Maes, T. & Bracke, K.R. Eosinophils in the spotlight:
182. Osborne, L.C. et al. Coinfection. Virus-helminth coinfection reveals a microbiota- eosinophilic airway inflammation in nonallergic asthma. Nat. Med. 19, 977–979
independent mechanism of immunomodulation. Science 345, 578–582 (2014). (2013).
© 2015 Nature America, Inc. All rights reserved.
npg

56 VOLUME 16  NUMBER 1  JANUARY 2015  nature immunology

Potrebbero piacerti anche