Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2017 Jun 1.
Published in final edited form as: J Allergy Clin Immunol. 2016 Jun;137(6):1631–1645. doi: 10.1016/j.jaci.2016.04.009

Fifty years later: Emerging functions of IgE antibodies in host defense, immune regulation, and allergic diseases

Hans C Oettgen 1
PMCID: PMC4898788  NIHMSID: NIHMS785274  PMID: 27263999

Abstract

Fifty years ago, after a long search, IgE emerged as the circulating factor responsible for triggering allergic reactions. Its extremely low levels in plasma, which are present at logs less than the other immunoglobulin isotypes, created significant hurdles for scientists working to reveal its identity. We now know that IgE levels are invariably increased in patients affected by atopic conditions and that this provides the critical link between the antigen recognition role of the adaptive immune system and the effector functions of mast cells and basophils at mucosal and cutaneous sites of environmental exposure. This review discusses the established mechanisms of action of IgE in pathologic immediate hypersensitivity, as well as its multifaceted roles in protective immunity, control of mast cell homeostasis, and its more recently revealed immunomodulatory functions.

Keywords: IgE, mast cells, anaphylaxis, humanized mice, tryptase, peanut allergy


The ability of a circulating factor to transfer allergen-specific immediate hypersensitivity was recognized early in the 20th century when Prausnitz and Kustner described a component of the γ-globulin fraction of plasma, then called reagin and now recognized as IgE, that was capable of passing skin test responsiveness from a sensitized subject to a naive host in the passive cutaneous anaphylaxis assay. As implied in the terms immediate and hypersensitivity, IgE has unique properties among immunoglobulin isotypes in its abilities both to induce extremely rapid pathologic responses, including potentially fatal anaphylaxis, and to act as a highly sensitive immunologic amplifier capable of triggering reactions after the interaction of minute quantities of antigen with just a few IgE molecules. These functions have rendered IgE an attractive target for pharmacologic intervention and IgE blockade. Many aspects of IgE immunobiology stand out as unique, including its regulation, its specific cellular receptors, the effector cell lineages mediating its functions, and its immunoregulatory properties, all of which are discussed in this review.

GENERATION OF IgE+ B CELLS: IgE ISOTYPE SWITCHING

The assembly of a functional IgE gene involves a sequence of DNA recombination events within the immunoglobulin heavy chain (IgH) locus, which spans 1250 kb in human subjectse.1 In pro-B cells in the bone marrow, transcription through an assortment of genomic heavy chain variable region gene (VH), diversity segment (D), and heavy chain joining segment (JH) exons triggers a recombination-activating gene 1– and 2–mediated process leading to their recombination to generate a diverse repertoire of VHDJH cassettes, each encoding a VH domain of fixed antigen specificity. Because this VHDJH cassette is situated just upstream of the Cµ and Cδ exons, B cells emerging from the bone marrow produce µ and δ-heavy chain transcripts and are both IgM+ and IgD+. Later in B-cell life, on exposure to cytokine and T-cell stimuli, B cells can undergo immunoglobulin class-switch recombination (CSR) in which a second somatic rearrangement results in the juxtaposition of VHDJH cassettes with one of a series of CH gene segments (Cγ, Cε, or Cα), each containing the CH exons encoding constant region domains for their respective isotypes (Fig 1). Switched B cells retain the antigen specificity dictated by their original VHDJH cassette but acquire the specific biological effector functions conferred by new Fc regions. Much of what we now know about CSR in general was learned from careful study of the specific process of IgE switching.

FIG 1.

FIG 1

IgE CSR. Before switching, the IgH locus in a B cell is in its germline configuration, with exons encoding the heavy chain constant region domains distributed over 150 kb of genomic DNA. Stimulation with IL-4 initiates ε-germline transcription through the Sε region. Clustering of Gs results in a very tight interaction between the transcribed RNA and DNA template, leaving a single nontemplate DNA strand. Secondary structures arising in the single strand cause stalling of RNAse polymerase II (Pol II), which results in recruitment of AID. AID catalyzes dC→dU conversions. The resultant dU nucleotides are deaminated by uracil N-glycosylase (UNG), generating a basic sites, which are substrates for apurinic/apyrimidinic endonuclease 1 (APE1). Single-stranded DNA breaks introduced at high density by this enzyme ultimately lead to DSBs. Breaks in Sµ and Sε are then annealed by means of classical nonhomologous end-joining (C-NHEJ), a DNA repair process involving the enzymes Ku70, Ku80, and DNA-dependent protein kinase, catalytic subunit. The products of this reaction are an episomal switch excision circle along with a functional IgE gene in which the VDJ cassette encoding the heavy chain variable regions is juxtaposed to the Cε exons encoding the constant domains.

Molecular genetic mechanism of IgE isotype switching

The process of CSR involves the sequential steps of (1) transcriptional activation of one of the CH loci, (2) chemical modification of nucleotides in the ε-switch region (Sε), (3) introduction of double-stranded DNA breaks (DSBs) in switch regions upstream of µ and the activated CH locus, and (4) a DNA repair process leading to annealing of the VDJ and CH regions (Fig 1).2,3 In some situations switching can be sequential. For example, B cells initially switching from IgM to IgG can later undergo a second CSR from γ to ε or α.

Cytokine- and receptor-mediated regulation of IgE CSR

Each of the CH gene segments (except Cδ) is an autonomous transcriptional unit 1 to 10 kb in length with its own cytokine-regulated promoter. The Iε promoter controls transcription at the ε-locus and contains binding sites for signal transducer and activator of transcription (STAT) 6, nuclear factor κB (NF-κB), Pax5, E2A, NFIL3, AP-1, C/EBP, and PU.1. The promoter is activated by IL-4 and/or IL-13 binding to receptors on B cells, leading to activation of the transcription factor STAT6. Simultaneous engagement of CD40 on B cells by its ligand, CD40L (CD154), which is transiently expressed on activated helper T cells, contributes a key second signal, activating NF-κB in a signal transduction pathway involving intracellular proteins from the TNF receptor-associated factor family of TNF receptor-associated factors.4,5 STAT6 and NF-κB sites are adjacent to each other, and the 2 transcription factors act synergistically to drive transcription.6 CD40L is encoded on the X chromosome, and boys with X-linked immunodeficiency with hyper-IgM syndrome have mutations in this gene.711 Additional TNF-type receptor-ligand pairs are able to provide similar stimulatory signals to those delivered by CD40/CD40L ligation.

One TNF family member, B cell-activating factor of the TNF family (BAFF), which is expressed on monocytes and dendritic cells (DCs), binds to transmembrane activator and CAML interactor (TACI) on cytokine-stimulated B cells, inducing isotype switching, even in the absence of T cells bearing CD40L.12,13 Although BAFF can drive IgE switching and respiratory epithelium produces BAFF, with increases in bronchoalveolar lavage fluid of segmental allergen-challenged subjects, its physiologic relevance in IgE regulation remains to be clarified.14,15 Consistent with the existence of T cell–independent mechanisms, it has been observed that IgE CSR occurs in the airway mucosa of patients with respiratory allergy.16 McCoy et al17 showed that IgE can be produced, even in mice and human subjects with no T cells or MHC II, and that the levels of this “natural” IgE increase with age. Such mice do not mount effective antigen-specific IgE responses, and natural IgE shows no evidence of somatic hypermutation.

Regulation of germline transcription: The Iε promoter

Transcripts arising from activation of the Iε promoter are referred to as ε-germline RNA and encompass the small Iε exon, as well as exons Cε1 to Cε4.18,19 These RNAs are spliced, capped, and transported to the cytosol but do not give rise to proteins because multiple stop codons present in the first exon, Iε, and have been referred to as sterile.20 Despite this lack of a functional protein product, production of germline εRNA is indispensable to IgE switching. Targeted deletion of either the I exon or its promoter in CH loci completely ablates switching, and insertion of a constitutively activated promoter can drive class-switching.2123

The critical function of germline transcription is to induce structural changes, such as stretches of single-stranded DNA, in S-regions, which lead to recruitment of key enzymes mediating chemical DNA modification, breakage, and repair. Switch regions contain evolutionarily conserved repeats of A/T-GC-A/T sequences, the preferred substrate for the enzyme activation-induced cytidine deaminase (AID). Among heavy chain switch regions, Sε contains the fewest of these repeats, perhaps rendering it a less efficient switch target than the other isotypes.24 RNA transcribed through switch regions contains abundant Gs. The G-rich Sε RNA transcripts bind more tightly to the C-rich template DNA than does its DNA complement. As a result, the nontemplate DNA is left free in a single-stranded form that tends to form secondary structures known as an R-loops, as well as stem loops, G quartets, and others. These structures cause stalling of RNA polymerase II, which is linked to AID in complex with Spt5, a protein recruited to the stalled RNAse polymerase II (Fig 1).25

Induction of DNA breaks and their repair in CSR

AID catalyzes dC→dU conversions, giving rise to dU:dG mismatches.26,27 A rare autosomal form of hyper-IgM syndrome (HIGM2), which is associated with marked lymphoid hypertrophy, is caused by mutations in AID.28 Subsequent sequential actions on these mismatched sites by the enzymes uracil N-glycosylase and then apurinic/apyrimidinic endonuclease 1 generate single-stranded DNA breaks, which at high density give rise to DSBs. Corresponding DSBs at Sµ and Sε are annealed by the DNA repair process of classical nonhomologous end-joining, and the Cµ exons are brought together to create a functional IgE gene. Consistent with this pathway, B cells lacking Ku70, Ku80, and DNA-dependent protein kinase, catalytic subunit, all of which are involved in nonhomologous end-joining, cannot execute isotype switching normally.29,30 For this final repair step to be directed to the correct DNA regions, the Sµ and acceptor Sε regions must be brought into physical proximity, a process mediated by chromosome looping coordinated by proteins interacting with the Eµ enhancer and the IgH 3′ regulatory region (the Ca enhancer).31

IgE MEMORY

In most antibody responses the processes of antigen-driven B-cell expansion, affinity maturation, isotype switch recombination, and generation of long-term B-cell memory occur in the germinal centers of secondary lymphoid tissues. For instance, high-affinity IgG-committed B cells arise in germinal centers after their IgM+ progenitors are driven by cytokine signals and costimulatory molecules from follicular helper T cells to switch to IgG (µ-γ switch), followed by affinity maturation and generation of long-lived memory B cells. The process is different for IgE. IgE+ B cells are short-lived in germinal centers, exhibiting both a tendency toward rapid transition to plasma cells and a susceptibility to apoptotic cell death. These properties might reflect a special fate of B cells expressing transmembrane IgE.32

The generation of high-affinity IgE responses and long-term memory for IgE occurs through unique mechanisms. The current understanding of IgE responses is in flux, but there is accumulating evidence that affinity maturation of IgE requires a step-wise process in which B cells sequentially undergo µ-γ and then γ-ε switches. Such a mechanism is suggested by the fact that high-affinity IgE B-cell clones tend to have hybrid switch sequences (Sµ-S-γ-Sε), which is consistent with their prior existence as IgG clones, and mice lacking the Cγ locus do not exhibit affinity maturation of their IgE responses.33 These observations indicate that IgE memory might reside mostly in that intermediate IgG+ B-cell stage.34,35

However, there is some conflicting evidence supporting the existence of IgE+ B-cell memory. Talay et al,36 using a transgenic mouse model in which cells expressing membrane IgE transcripts also produce a green fluorescent protein, found that IgE memory could develop through an IgE+ germinal center intermediate and ultimately reside in IgE+ B cells. Using a similar approach, Yang et al37 found evidence supporting a germinal center pathway for IgE+ cell formation but observed that the IgE+ cells exhibited a unique fate, rapidly upregulating the transcription factor Blimp-1 and transitioning to plasma cells. In another IgE transgenic reporter model, He et al38 found that IgE+ B cells do not contribute to long-lived memory. Addressing the pathways of IgE affinity maturation and memory in human subjects has been much more of a challenge. Recent very elegant work by Looney et al39 at Stanford involving the sequencing of more than 15 million IgH regions in healthy and allergic subjects suggests that most IgE+ cells arise from high-affinity antigen experienced (somatically hypermutated) IgG+ precursors.

IgE STRUCTURE

IgE is the least abundant antibody isotype in plasma, present at levels (about 100 ng/mL) several logs lower than circulating IgG antibodies (5–10 mg/mL). Like other immunoglobulins, IgE antibodies are tetramers, comprised of 2 ε-heavy and 2 light chains (κ or λ) linked by numerous intrachain disulfide bonds (Fig 2). Variable (V) sequences at the N-termini of the heavy (VH) and light (VL) chains create unique antigen-specific binding sites. The C-terminal regions of the ε-heavy chains are made up of 4 Cε domains (compared with 3 for γ-heavy chains), each encoded by one of the Cε1 to Cε4 exons located near the 3′ end of the heavy chain locus (IgH). The IgE Cε2–4 Fc domains confer its isotype-specific functions, including binding to its receptors, FcεRI and CD23. Unlike Fcγ and Fcµ, Fcε does not activate complement.

FIG 2.

FIG 2

Structures of the mature ε-heavy chain gene and IgE protein. The ε-heavy chain gene generated in the IgH locus through deletional CSR (Fig 1) contains a VDJ cassette encoding the heavy chain variable regions juxtaposed to the Cε exons encoding the constant domains. A hybrid Sµ/Sε sequence remains between the VH and Cε exons, a by product of the isotype-switching process. The secreted IgE protein has 4 constant regions, 1 more of these domains than IgG. Intrachain disulfide bonds are contained within each of the immunoglobulin domains. IgE molecules are heavily glycosylated (7 N-linked sites as indicated by ovals). The N394 oligosaccharide is essential for IgE-FcεRI binding. The transmembrane form of IgE, which contains M1 and M2 exons encoded by ε-mRNA splice isoforms, is expressed at the surface of IgE+ B cells.

Contact with the high-affinity IgE receptor α-chain is mediated by Cε3.40 Cε2 is in a position comparable to the flexible “hinge” region contained in Cγ of the γ-heavy chains, and in its receptor-bound configuration, there can be a sharp turn in the molecule in the Cε2–3 region, with the pair of Cε2 domains folding back over Cε3–4.41 In IgE+ B cells hydrophobic sequences encoded by M1 and M2 exons present in mRNA splice variants encode a transmembrane form of IgE. IgE antibodies are more heavily glycosylated than other immunoglobulin isotypes with 7N-linked glycosylation consensus sequences (N-X-S/T) on each ε-heavy chain, one of which (at N394 in Cε3) is required for IgE binding to its high-affinity receptor, FcεRI.42,43 As a result of their heavy glycosylation, IgE antibodies have an affinity for galectins, lectin-type proteins that can interact with both free and cell-bound IgE.

In contrast to IgG antibodies, which have a half-life of about 3 weeks, IgE is very short-lived in plasma (half-life, <1 day), but receptor-bound IgE can remain fixed to mast cells in tissues for weeks or months. This long tissue half-life has significant clinical implications. For example, solid organ transplant recipients can exhibit peanut reactions mediated by mast cell-fixed donor IgE in the organ.44,45

IgE RECEPTORS

The biological functions of IgE antibodies are mediated by their dual interactions with specific antigens and 2 structurally very different receptors, FcεRI and CD23, each present on a broad range of effector cells.

FcεRI structure

The high-affinity IgE receptor FcεRI is a multimeric protein expressed in 2 isoforms, a tetrameric αβγ2 receptor present on mast cells and basophils and a trimeric αγ2 receptor expressed by eosinophils, platelets, monocytes, DCs, and Langerhans cells, although at 10- to 100-fold lower levels (Fig 3). The α-chain contains 2 extracellular immunoglobulin-like domains and mediates binding to the Cε3 domain of the ε-heavy chain of IgE. The β-subunit has 4 transmembrane-spanning domains with both N- and C-termini on the cytosolic side of the plasma membrane. The γ-chains are transmembrane dimeric disulfide-linked proteins. The β- and γ-chains have intracellular immunoreceptor tyrosine-based activation motifs (ITAMs), 18 amino acid tyrosine-containing sequences that provide docking sites for Src homology 2 (SH2) domain- containing proteins when phosphorylated, including the proximal signaling protein tyrosine kinase of FcεRI signaling, spleen tyrosine kinase (SYK). IgE antibodies play an important role in regulating the density of their own high-affinity receptor, stabilizing FcεRI at the cell surface.4752

FIG 3.

FIG 3

IgE receptors. The high-affinity IgE receptor FcεRI is expressed in its tetrameric form (αβγ2) on mast cells and basophils. In human subjects a trimeric form (αγ2) is found on a number of lineages, including various types of professional APCs. CD23, the low-affinity IgE receptor, is broadly distributed and is a type II transmembrane protein (N-terminus intracellular) assembled as a multimer with α-helical coiled-coil stalks terminating in IgE-binding C-type lectin heads. Protease-sensitive sites in the stalks can be cleaved by endogenous proteases (including ADAM 10) or exogenous proteases (including the Der p 1 protease of dust mites).

Dehlink et al53 have described a circulating soluble form of FcεRI-α. In cultured cells crosslinking of surface FcεRI triggers generation of the soluble α-chain. The physiologic functions of the soluble form of FcεRI-α are under investigation, but the observation that recombinant soluble FcεRI-α can block mast cell and basophil activation, as well as passive cutaneous anaphylaxis, suggests functions in downregulation of allergic responses.

FcεRI activation on mast cells and basophils: Immediate hypersensitivity

FcεRI has high affinity for IgE (Kd, 10−9 mol/L), and therefore mast cell and basophil FcεRI are fully occupied under physiologic conditions, and once bound, IgE remains permanently attached until internalized.54 In the classic immediate hypersensitivity reaction, allergen-induced cross-linking of IgE bound through FcεRI triggers a cascade of signaling events, leading to mediator release and gene transcription.55 Cross-linking of neighboring FcεRI receptors leads to aggregation and transphosphorylation of cytosolic ITAMs on the FcεRI β- and γ-chains by constitutively receptor-associated Lyn tyrosine kinase. These ITAM phosphotyrosines provide docking sites for the SH2-containing SYK protein tyrosine kinase. Activated SYK phosphorylates tyrosines in a number of adapter molecules, including LAT1/2, SLP-76, and Grb2, leading to the assembly of a supramolecular plasma membrane-localized signaling complex. The coordinate initiation of downstream signaling pathways from this complex ultimately leads to increases in cytosolic calcium, activation of gene transcription (IL-4, TNF, and IL-6), induction of synthesis of prostaglandins and cysteinyl leukotrienes and granule fusion with the plasma membrane, leading to release of preformed mediators of hypersensitivity, including histamine, proteoglycans, and proteases. These mediators rapidly induce vasodilation, plasma extravasation, tissue edema, mucus production, and smooth muscle constriction. In many subjects immediate responses are followed by delayed “late-phase” reactions. These manifest as repeated onset of airflow obstruction, gastrointestinal symptoms, skin inflammation, or anaphylaxis 8 to 24 hours after allergen challenge and after the initial response has completely subsided. Passively transferred IgE antibodies confer both acute and late-phase sensitivity to allergen challenge, and interference with IgE signaling, mast cell activation, or inhibition of the mast cell mediators blocks both responses.56,57

The nanomolar affinity of FcεRI for IgE together with the potency of the vasoactive mediators produced by mast cells and basophils make a highly bioamplified response system to allergens, the sensitivity of which is subject to regulation. Modulation of the cell-surface density of FcεRI by ambient IgE levels is a strong determinant of signaling threshold, and the effects of omalizumab in downregulating FcεRI likely account for much of the clinical benefit of IgE blockade. The sensitivity of FcεRI signaling is also diminished after repeated activation events, a phenomenon that might be the result of SYK downregulation.58 Concurrent signaling by IgG antibodies interacting with the same allergens as FcεRI can also attenuate the strength of the IgE signal. These antibodies interact with the inhibitory IgG receptor FcγRIIb, setting off negative signaling pathways originating from receptor-associated protein tyrosine phosphatases and inositol phosphatases. The reduction in allergen sensitivity occurring after subcutaneous immunotherapy in aeroallergen-sensitive patients and also in subjects with food allergy completing oral immunotherapy (both of which retain significant IgE levels) might be mediated in part by the strong inhibitory IgG responses induced by these therapies.59,60

Blocking immediate hypersensitivity: Inhibition of FcεRI/IgE interaction

Effective pharmacologic blockade of the binding of IgE to FcεRI was achieved with the introduction of omalizumab, a humanized mAb recognizing the Cε3 domain of free but not receptor-bound IgE antibodies.61 Omalizumab effectively blocks both acute IgE-mediated responses to inhaled and ingested allergens, as well as late-phase responses to inhaled allergens. 62,63 In seminal clinical trials led by Soler in Europe64 and Busse in the United States,65 IgE blockade by omalizumab was shown to decrease the frequency of flares in steroid-requiring asthmatic patients but did not lead to substantial improvement in lung function or day-to-day symptoms. This suggests that although immediate hypersensitivity reactions are blunted by IgE inhibition, TH2 cell-driven eosinophil influx, mucus metaplasia, and airway remodeling, aspects of asthma previously shown to arise independently of IgE in animal models, are unaffected.66 The beneficial effect of omalizumab in preventing acute responses to allergen inhalation might be mediated in large part by the downregulation of IgE receptors in the setting of low ambient IgE levels, with a consequent decrease in responsiveness of allergen-challenged mast cells and basophils. 6769 The effects of omalizumab on FcεRI are complex, however, as evidenced by the findings of Zaidi et al70 that β-chain levels decrease while SYK levels and the efficiency of signaling by individual FcεRI molecules are increased after omalizumab treatment.

A number of additional strategies for IgE inhibition have been explored or are under development. IgE synthesis is not affected by omalizumab, a limitation that led to the development of quilizumab, a humanized mAb that targets the M1-prime segment of membrane-expressed IgE present on IgE-switched B cells. Unfortunately, although quilizumab reduced circulating IgE levels to greater than 40%, it did not significantly affect asthma flares, lung function, or quality of life in adults with poorly controlled asthma.71 This result is consistent with the past observations of Casale et al72 that the effectiveness of IgE blockade in reducing clinical symptoms correlates with reduction of IgE to very low levels. Because omalizumab does not displace IgE bound to FcεRI at current treatment doses, the onset of clinical benefit is delayed for weeks or months until preexisting cell-associated IgE is internalized. This limitation has prodded the development of disruptive agents that dislodge IgE from its receptor. Among these is a designed ankyrin repeat protein, DARPin E2_79, which efficiently dissociates FcεRI-IgE complexes and blocks IgE-mediated mast cell activation in culture and in vivo, as developed by Jardetzky et al.73,74 Low-molecular-weight compounds, including DNA aptamers and peptides, have also been studied as potential blockers of IgE binding to FcεRI, but a range of issues, including their susceptibility to hydrolysis, bioavailability, and pharmacodynamics, have prevented their translation to clinical use.7579

Antigen-independent signaling through FcεRI: Cytokinergic IgEs and histamine-releasing factor

Although classical dogma has held that signaling through FcεRI is dependent on antigen/IgE-mediated receptor aggregation, evidence accumulating over the past 15 years has demonstrated that the binding of IgE to FcεRI can deliver activating signals, even in the absence of allergen. Experiments with cultured bone marrow mast cells revealed that IgE acts through FcεRI in the absence of antigen to exert a survival-enhancing effect, protecting these cells from apoptosis after growth factor withdrawal.80,81 Indeed, there is in vivo evidence that mast cell survival is regulated by IgE. Tissue mast cell expansion in parasitized mice and in animals exposed to allergens depends on the presence of IgE antibodies.8284 Thus in addition to their role in allergen-triggered mast cell activation, IgE antibodies are key regulators of mast cell homeostasis.

Since the initial observation that IgE antibodies control mast cell survival, a number of additional responses have been found to be induced in cultured mast cells by IgE in the absence of antigen, including cytokine production, histamine release, leukotriene synthesis, and calcium flux.8587 There is indirect evidence that the baseline cytokine profile of cutaneous mast cells is influenced by IgE levels, even in the absence of antigen.88 Not all IgE antibodies are equally capable of inducing antigen-dependent mast cell activation. Those with the greatest activity have been referred to as “cytokinergic.”89 Detailed analyses of the structure of one of the most cytokinergic IgE antibodies, SPE-7, which is specific for the hapten dinitrophenyl, have revealed possible mechanisms underlying antigen-independent IgE/FcεRI aggregation. Bax et al90 provided evidence for homotypic interactions between the variable domains (Fv) of free SPE-7 and their counterparts on FcεRI-bound SPE-7.90 Previously, James et al91 reported that SPE-7 exists in conformational isomers, some of which bind to autoantigens, including the protein thioredoxin. Taken together, these observations suggest that FcεRI activation by IgE in the absence of nominal antigen can be mediated by a tendency to self-associate, as well as to recognize autoantigens, interactions that, even at low affinity, lead to sufficient FcεRI crosslinking to initiate detectable signaling in this very highly amplified pathway. It is possible that such nonspecific IgE interactions with autologous proteins drive the mast cell activation underlying idiopathic urticaria. Patients affected by this disorder do not have allergen-specific IgE antibodies but exhibit dramatic and often rapid improvement after treatment with omalizumab.92

An additional mechanism for antigen-independent activation of IgE signaling might involve histamine-releasing factors (HRFs), which have been studied for more than 30 years. In 1995, MacDonald et al93 reported the molecular structure of HRF, which turned out to be a secreted protein variously known as translationally controlled tumor protein, fortilin, and other names. In addition to playing key roles in cell-cycle progression and proliferation as an intracellular protein, the secreted form of HRF is found in bronchoalveolar and nasal lavage fluids and binds to a subset of IgE and IgG molecules.94,95 The physiologic contributions of HRF to IgE-driven FcεRI signaling in vivo are currently under investigation.

CD23, the low-affinity IgE receptor: Structure

Commonly referred to as the low-affinity IgE receptor, CD23 actually binds IgE tightly with a KA of 108 .96 RNA splice variants encode 2 isoforms of the protein CD23a, which is present predominantly on B cells, and CD23b, which is expressed on a wide range of cells, including monocytes, DCs, Langerhans cells, eosinophils, and gastrointestinal and respiratory epithelial cells. Unlike FcεRI or any of the Fcγ receptors, CD23 is a type II transmembrane protein (N-terminus intracellular) with an IgE-binding C-type lectin domain, making it the only immunoglobulin receptor that is not a member of the immunoglobulin superfamily.97 The IgE-binding domains of CD23 are connected to its transmembrane segments through α-helical coiled-coil stalks (Fig 3). These mediate multimerization, and only oligomeric CD23 binds IgE.98 In addition to serving as a receptor for IgE, CD23 in human subjects binds to a second ligand, the B-cell surface molecule CD21 (also known as CR2 and the EBV receptor).99 Like FcεRI, surface levels of CD23 are regulated by IgE itself.100 Occupancy by IgE protects sensitive sites proximal to its C-type lectin heads from cleavage by a variety of proteases, including allergens (like the Der p 1 protease of dust mites) and the endogenous a disintegrin and metalloproteinase (ADAM) proteases, especially ADAM10. Released oligomeric CD23 heads, referred to as soluble CD23 (sCD23), retain their IgE-binding capacity.

CD23 functions: Transepithelial allergen transport, facilitated antigen presentation, and regulation of IgE synthesis

CD23 on intestinal epithelial cells mediates the transport of food allergen-IgE complexes from the gut lumen into the mucosa.101 CD23 on antigen-presenting cells (APCs) has been shown to mediate “facilitated antigen presentation,” enhancing the uptake of antigen-IgE complexes for processing and presentation to T cells.102104 Gustavsson et al105 and Martin et al106 have described a mechanism whereby circulating B cells can transport IgE-antigen complexes to the spleen, where B-cell exosomes, generated in an ADAM10- and CD23-mediated process, transfer antigen to DCs for uptake, processing, and presentation. Ligation of CD23 on B cells by activating antibodies inhibits IgE synthesis, 107 and transgenic mice overexpressing CD23 have suppressed IgE responses.108,109 Conversely, CD23-deficient mice have higher and more sustained specific IgE titers after immunization.110112 As predicted by these observations, anti-CD23 therapy (lumiliximab) results in decreased IgE levels.113,114 Unlike transmembrane CD23, soluble CD23 (sCD23) fragments, which are generated by means of proteolytic cleavage of the IgE-binding heads, can have an opposing effect, enhancing IgE production.115

IgE IN HOST DEFENSE

The evolutionary persistence of the IgE isotype, along with a committed network of receptors and effector cell lineages, points to an important adaptive advantage. A clue regarding the forces driving the selection of this system is provided by the presence of high IgE levels in subjects residing in helminth-endemic regions, an association that has long suggested that IgE might be important in controlling host-parasite interactions. In areas with a high prevalence of Schistosoma mansoni, total IgE levels correlate with the subject’s resistance to reinfection, and studies done more than 30 years ago by Joseph et al116 showed that IgE can opsonize S mansoni for killing by eosinophils, providing direct evidence for a protective function. Animal models have provided an additional direct evidence for a protective function of IgE. IgE enhances granuloma formation in the liver and promotes clearance of adult worms during primary infection with S mansoni in mice.117 IgE antibodies also enhance production of parasite-specific IgG1 antibodies, which have been implicated in anti-schistosomal immunity. IgE facilitates the elimination of another helminth, Trichinella spiralis, from the intestine and drives the destruction of tissue cysts, which are heavily coated with IgE antibodies.82

One might predict that the high IgE levels present in parasite-endemic regions would be associated with increased rates of atopic disease. However, the data on this point are complicated. A number of studies have pointed to the opposite effect, namely a decrease in allergic conditions in parasitized subjects.118,119 In the setting of chronic filarial infestation, this has been associated mechanistically with an IL-10-driven state of TH1 and TH2 suppression. In contrast, investigations in Ascaris or Toxocara species-infected subjects point to a parasite-induced increase in allergen-specific TH2 responses and allergic symptoms, including wheezing.120122 Cross-reactivity between parasitic antigens and allergens might play a significant role, as has been suggested by the observation that IgE formed in response to the tropomyosin of Ascaris lumbricoides recognizes the homologous proteins from the dust mite or cockroach. 123

The abundance of mast cells in the skin along with the itch response elicited by arthropods, such as ticks, has suggested a role for IgE in protection from these ectoparasites. Such a defense mechanism might also indirectly protect against the numerous pathogens for which ticks serve as vectors. Brown et al124 first demonstrated that antibodies to the tick conferred a basophil-mediated protective response in guinea pigs. Subsequent studies with genetically manipulated mice confirmed that both mast cells and basophils contribute to tick immunity in an IgE-dependent mechanism.125127

Even a single exposure to the lone star tick, Amblyomma americanum, can induce a strong IgE response to tick antigens including a carbohydrate determinant, galactose-α-1,3-galatose, which is also present in certain meats. This response is responsible for a unique form of delayed food-induced anaphylaxis first described by Commins et al,128 which is most common in the southern and eastern United States and occurs 3 to 6 hours after ingestion of mammalian food products, including beef and pork. IgE antibodies with the same galactose-α-1,3-galatose specificity are also responsible for more classical systemic anaphylaxis induced by oligosaccharides on the Fab portion of cetuximab, a chimeric mouse/human mAb against the epidermal growth factor receptor used for the treatment of various metastatic cancers.129

Proteases from IgE-activated mast cells inactivate reptile and hymenoptera venoms

In some subjects exposure to Hymenoptera venoms elicits strong IgE responses that can even provoke life-threatening anaphylactic reactions on re-exposure. Emerging data suggest that this might represent a pathological overreaction of a defense mechanism originally evolved to inactivate toxic venom constituents and thereby protect the host. A beneficial function for mast cell activation in the inactivation of venoms was first suggested by the observations of Higginbotham130 in 1965 that mast cells prevented the formation of local skin lesions in mice injected with snake venom. An elegant series of experiments performed over the past decade by the group of Stephen Galli has given rise to an entirely new paradigm in which IgE antibodies might mediate resistance to both reptile and arthropod venoms.131134 These studies showed that exposure of mice to snake or bee venoms induces the production of IgE antibodies that mediate resistance to challenge with normally lethal doses. The failure of mast cell-deficient KitW-sh/KitW-sh or IgE-deficient (Igh7−/−) mice to acquire such protection implicates the IgE-mast cell pathway in this adaptive response. Independent studies by Palm et al135 at Yale showed that acquired resistance to bee venom phospholipase A2 was also mediated by an IgE mechanism. Although these findings are currently restricted to mouse models, they prod us to rethink the dogma that IgE responses to venom are invariably pathologic in human subjects.

IMMUNE REGULATION BY IgE

Both FcεRI and CD23 are expressed on APCs and can facilitate the uptake of allergen bound to IgE.104 The trimeric form (αγ2) of FcεRI is expressed by monocytes, as well as conventional and plasmacytoid DCs, in many tissues of human subjects (but not mice) and can mediate antigen uptake for presentation.136,137 FcεRI on dermal DCs and Langerhans cells in the skin of patients with atopic dermatitis is markedly upregulated during allergic flares.138 FcεRI signaling in cultured monocytes and DCs leads to activation of the NF-κB pathway and production of proinflammatory mediators, including TNF-α, IL-6, and CCL-28.139142 FcεRI crosslinking has also been observed to elicit the production of anti-inflammatory factors, including IL-10 and indoleamine 2,3-dioxygenase, and to suppress T-cell proliferation in vitro.143145 An immunosuppressive effect of IgE signaling in the setting of viral infections is suggested by the FcεRI-mediated attenuation of type I interferon production in response to the exposure of plasmacytoid DCs to influenza virus, rhinovirus, or Toll-like receptor ligands.146,147 Asthmatic patients have been reported to have impaired rhinovirus-induced IFN-α and IFN-β responses in bronchoalveolar lavage cells,148 and it has been speculated that the decrease in asthma flares during respiratory virus seasons observed after omalizumab treatment149 might relate to the inhibition of IgE-mediated suppression of innate viral immunity through this type I interferon pathway.

The absence of FcεRI on rodent APCs has made it a challenge to assess the physiologic role of IgE/FcεRI on APCs with the usual mouse models of allergic diseases and TH2 induction. Significant advances have been made in this regard by the generation of transgenic mice constitutively expressing FcεRI under control of the CD11c promoter active in DCs.150 Sallmann et al150 found that these mice have an enhanced pulmonary late-phase response (lung infiltration with inflammatory cells and antigen-specific T cells) to ovalbumin (OVA) inhalation after intraperitoneal or epicutaneous OVA sensitization. In contrast, a study by Platzer et al151 using the same animals revealed a less severe phenotype with decreased inflammation, airway responsiveness, and mast cell progenitor recruitment and attenuated TH2 responses. These investigators also observed blunted responses in an OVA food allergy model and decreased total IgE levels, the latter observation consistent with prior findings of Greer et al152 showing that FcεRI on DCs promotes IgE clearance by internalization and transport to an endolysosomal compartment. The conflicting findings in these reports reveal the need for additional investigation on the functions of DC FcεRI.

B cells, Langerhans cells, follicular DCs, T cells, and eosinophils express CD23.153 The binding of allergen by specific IgE bound to CD23 on cultured APCs facilitates its uptake. 154,155 The relevance of CD23-mediated antigen uptake to immune responses in vivo has been shown in mouse models in which wild-type but not CD23−/− animals immunized intravenously produce stronger IgG responses when specific IgE is provided along with antigen at the time of immunization.102,103 In this system CD23−/− mice acquire responsiveness to IgE after reconstitution with B cells from CD23+ donors.105,156 Observations by Gustavsson et al105 suggest a process whereby allergen-specific IgE-allergen complexes bound to B cells in the periphery through CD23 are transported to B-cell follicles in the spleen, where they are made available to splenic DCs.

Adjuvant effects of IgE-activated mast cells and basophils on allergic inflammation and TH2 responses

Mast cells activated by IgE-FcεRI produce an array of cytokines that exert proinflammatory and immunomodulatory effects.157 Mast cell development and survival are dependent on the cell-surface receptor protein tyrosine kinase c-Kit. Kit-mutant mice, which lack mast cells, have provided an excellent tool for analysis of the immunomodulatory effects of mast cells in vivo. Studies by Stephen Galli’s group comparing the responses of c-Kit-deficient KitW-sh/W-sh and/or KitW/W-v mice that lack mast cells with those of control animals and mutant mice reconstituted with normal or cytokine-deficient mast cells first showed that mast cell-derived TNF-α can amplify the development of allergic airway and skin inflammation in adjuvant-free systems of allergen senstization.157159 These observations ran counter to those of some prior investigations that did not show a mast cell role in the induction of allergic inflammation but in which sensitization was achieved by using artificial adjuvants, such as alum, a discrepancy that suggested that mast cells themselves act to stimulate emerging immune responses, substituting for artificial adjuvants.160

We have observed a similar adjuvant function of IgE-activated mast cells in a mouse model of peanut allergy, showing that the efficient induction of peanut-specific TH2 cells and IgE production in atopy-prone IL4raF709 mice is dependent on the presence of IL-4-producing mast cells and IgE antibodies.84 Analyses of IL-4-producing lineages using 4get reporter mice have revealed that mast cells are a major source of IL-4 in the intestine,161 and Finkelman et al162 have shown that enteral challenge of food-sensitized mice leads to a rapid IgE-dependent increase in plasma IL-4 levels. Mast cells, although not professional APCs, express MHC II and OX40L on activation, and mouse and human mast cells are able to present antigen in vitro.163,164 IgE-activated mast cells have also been shown to reduce IL-12 production by DCs, conferring a TH2-inducing phenotype.165,166

In addition to promoting TH2 responses, there is evidence that IgE-activated mast cells suppress the generation of regulatory T (Treg) cells and might divert them toward a TH2 or TH17 phenotype. Among the cytokines produced by IgE-activated mast cells, IL-1, IL-4, IL-6, and TNF-α are known to inhibit Treg cell development and function.167,168 In the IL4raF709 murine model of peanut allergy mentioned above, IgE-mediated mast cell activation has been shown to favor a reprogramming of the Treg phenotype with induction of GATA3 expression and production of IL-4.169 Mast cell-derived IL-6 destabilizes Treg cell expression of Foxp3, resulting in a shift to IL-17 production.170 The interaction of mast cell OX40L with OX40 on Treg cells inhibits Treg cell suppressive functions. 170,171

Basophils have also been implicated as IgE-triggered inducers of allergic inflammation and TH2 responses. Twenty-five years ago, Seder et al 172 identified FcεRI+ cells with ultrastructural features characteristic of basophils in the bone marrow and circulation as the major nonlymphocytic source of IL-4 in parasitized mice, a finding subsequently confirmed by the same group and others in IL-4 reporter mice.173,174 Allergic skin inflammation elicited by means of passive immunization with hapten-specific IgE, followed by intradermal challenge, was used by Karasuyama et al175 to implicate basophils in the induction of IgE-mediated allergic skin inflammation. Subsequent investigations using a genetic strategy to deplete basophils (transgenic expression of the diphtheria toxin receptor under control of the basophil-specific mcpt8 promoter) also blocked the induction of IgE-driven skin inflammation. 176 We and others 177179 have found that the same basophil depletion strategies attenuate TH2 sensitization and allergic airway inflammation induced by allergen inhalation, and Noti et al180 have reported that the induction of food allergy by means of epicutaneous antigen application is dependent on IgE antibodies and basophils.

Limitations inherent in the approaches used to deplete mast cells or basophils have occasionally given rise to controversy. 161 Some of the strategies used to remove mast cells either do not completely eliminate all mast cell subsets or have unwanted off-target effects on basophils and other lineages. 181183 Conversely, the common approaches applied for basophil depletion, MAR-1 treatment (anti-FcεRI antibody), and Mcpt8cre transgenes might have off-target effects on mast cells. 184,185 In the most convincing studies complementation of a lost phenotype in mast cell- or basophil-deficient recipients by means of transfer of the corresponding highly purified cell population has been used to confirm the physiologic relevance of the targeted lineage. The aggregate findings to date obtained by using a range of model systems support key roles for both IgE-activated mast cells and basophils in induction of TH2 responses, and it is likely that they have complementary and synergistic functions.

IgE and asthma: Epidemiologic evidence for a role in driving allergic inflammation

A function for IgE antibodies in the induction of allergic inflammation in human subjects is strongly suggested by epidemiologic data. There is a close correlation between “allergic sensitization,” the production of aeroallergen-specific IgE antibodies detected by prick skin testing or direct IgE measurement, and development of asthma. 186,187 In wheezing toddlers greater levels of allergen-specific IgE are associated with persistence of wheezing to school age.188 The German Multicenter Allergy Birth Cohort Study of 1314 children revealed the importance of IgE-mediated sensitivity in disease progression.189 During the preschool years, wheeze symptoms were similar between allergic and nonallergic children. However, by adolescence, 90% of the nonallergic children had no wheezing and normal lung function, whereas more than half of allergic wheezers had active asthma with impaired lung function.

CONCLUSION

In addition to bringing to a close the quest for the elusive reagin, the discovery of IgE 50 years ago ushered in a fruitful era of investigation into its genetics, structure, and functions. In many cases research on IgE has had a truly far-reaching effect on our understanding of fundamental immunologic processes, including the mechanisms of immunoglobulin class-switching and the functions and interactions of immunoglobulin receptors. The coming decades will likely witness further advances in our understanding of IgE biology along with the introduction of next-generation anti-IgE therapies and innovative strategies to manipulate the IgE axis to modulate allergic disease.

FIG 4.

FIG 4

Proposed adjuvant and immunoregulatory functions of IgE and FcεRI. Mast cells and basophils residing in mucosal and skin sites produce IL-4 in response to antigen-induced IgE-FcεRI signaling. IL-4 promotes the induction of TH2 cells and sustains their local survival. These provide the IL-4 and cognate T-B interactions critical for driving IgE class-switching in mucosal B cells. Mast cells suppress Treg cell expansion and function, possibly through cytokines, including IL-4 and IL-6. Trimeric FcεRI present on APCs facilitates antigen uptake for presentation to local T cells.

Acknowledgments

Supported by National Institute of Allergy and Infectious Diseases grants 1R01AI119918-01 and 5T32AI007512-28.

I thank Drs Oliver Burton and Amanda Stranks for their critical review of the manuscript.

Abbreviations used

ADAM

A disintegrin and metalloproteinase

AID

Activation-induced cytidine deaminase

APC

Antigen-presenting cell

BAFF

B cell–activating factor of the TNF family

CSR

Class-switch recombination

DC

Dendritic cell

DSB

Double-stranded DNA break

HRF

Histamine-releasing factor

IgH

Immunoglobulin heavy chain

ITAM

Immunoreceptor tyrosine-based activation motif

JH

Heavy chain joining segment

NF-κB

Nuclear factor κB

OVA

Ovalbumin

ε-Switch region

STAT

Signal transducer and activator of transcription

SYK

Spleen tyrosine kinase

Treg

Regulatory T

VH

Heavy chain variable region gene

GLOSSARY

ANTIGEN-PRESENTING CELLS

Cells that take up antigens and process them into peptides for display on MHC proteins on their surfaces for presentation to T-cell receptors on T cells specific for the same antigens

CD11c

A cell-surface molecule with a broad expression found on immune cells

CD40

A costimulatory protein found on B cells and antigen-presenting cells (APCs) that is required for their activation. The binding of its ligand, CD40L, on helper T cells activates B cells and APCs and induces a variety of downstream effects

CLASS-SWITCH RECOMBINATION

A mechanism that changes a B cell’s production of immunoglobulin from one type to another in which the constant region of the heavy chain is changed but the variable region of the heavy chain stays the same

FcεRI

The high-affinity receptor for the Fc region of IgE, which was constitutively expressed on mast cells, basophils, eosinophils, platelets, monocytes, dendritic cells, and Langerhans cells

FOLLICULAR HELPER T (TFH) CELLS

Nonpolarized effector T cells specialized in homing to the B-cell areas of secondary lymphoid tissue through interactions mediated by the chemokine receptor CXCR5 and its ligand, CXCL13

HAPTENS

Small molecules that elicit an immune response only when covalently bound to a large carrier, typically a protein antigen

IL-10

A cytokine produced primarily by monocytes and, to a lesser extent, by lymphocytes, which has pleiotropic effects in immunoregulation and inflammation by limiting inflammation and thereby preventing damage to the host

IL-12

A cytokine produced by dendritic cells, macrophages, and human B-lymphoblastoid cells in response to antigen stimulation. IL-12 is involved in the differentiation of naive T cells in to TH1 cells. It stimulates the production of IFN-γ and TNF-α from T cells and natural killer cells and reduces IL-4-mediated suppression of IFN-γ

Mcpt8DTR

A human diphtheria toxin receptor under the control of the mast cell protease 8 (Mcpt8) promoter, which is active only in basophils. Mice transgenic for Mcpt8DTR have diphtheria toxin only on their basophils and, when treated with the toxin, will be specifically depleted of basophils

MHC II

A complex that presents peptides derived from extracellular antigens to T-cell receptors

N-LINKED GLYCOSYLATION

Attachment of the oligosaccharide known as glycan to a nitrogen atom that is required for the structure and function of some eukaryotic proteins

NUCLEAR FACTOR κB (NF-κB)

A protein complex that controls transcription of DNA and plays a key role in regulating the immune response to infection. NF-κB is found in almost all animal cell types and is involved in cellular responses to a variety of stimuli

OPSONIZE

The process by which a pathogen is labeled and made more susceptible to phagocytosis

OVALBUMIN

The main protein found in egg white, which is a well-characterized allergen used in immunologic studies

OX40-OX40L

Members of the TNF superfamily expressed on a variety of cells, including activated CD4+ and CD8+ T cells. The OX40-OX40L complex has been shown to regulate cytokine production from T cells, antigen-presenting cells, natural killer cells, and natural killer T cells and modulate cytokine receptor signaling. This complex plays a central role in the development of multiple inflammatory and autoimmune diseases, making them ideal therapeutic candidates

RECOMBINATION-ACTIVATING GENES (RAGs)

Genes that encode enzymes that play an important role in the rearrangement and recombination of the genes of immunoglobulin and T-cell receptor molecules. RAG-1 and RAG-2 cellular expression is restricted to developing lymphocytes and generation of mature B and T lymphocytes

RNA POLYMERASE II

An enzyme found in eukaryotic cells that catalyzes the transcription of DNA to synthesize precursors of mRNA and most small nuclear RNA and microRNA

SOMATIC HYPERMUTATION

A cellular mechanism affecting the variable regions of immunoglobulin genes of immune cells, which diversifies and increases the affinity of antibodies

SPLEEN TYROSINE KINASE

A nonreceptor cytoplasmictyrosine kinase composed of a dual SH2 domain separated by a linker domain that plays a crucial role in immune receptor signaling

Src HOMOLOGY 2 DOMAIN (SH2 DOMAIN)

A sequence-specific phosphotyrosine-binding module commonly found in adapter proteins that aids in the signal transduction of receptor tyrosine kinase pathways, which allow proteins containing those domains to dock to phosphorylated tyrosine residues on other proteins

TRANSMEMBRANE ACTIVATOR AND CAML INTERACTOR (TACI [TNFRSF13B GENE])

A protein found on the surfaces of B cells that is known to promote cell signaling, plays a role in B-cell survival and maturation, and is involved in class-switch recombination and antibody production

TNF-α

Secreted by macrophages, mast cells, and many other cell types, this cytokine’s primary role is the regulation of immune cells. Moreover, it is involved in the regulation of a wide spectrum of biological processes, including cell proliferation, differentiation, apoptosis, lipid metabolism, and coagulation

TOLL-LIKE RECEPTOR LIGANDS

Ligands binding to a class of receptors expressed on macrophages and dendritic cells that recognize conserved microbial particles that can activate an immune response

V(D)J RECOMBINATION

The somatic assembly of component gene segments that encode antigen recognition sites of receptors expressed on B and T lymphocytes

Footnotes

Disclosure of potential conflict of interest: H. C. Oettgen has received a grant from the National Institutes of Health and has had consultant arrangements with Genentech.

REFERENCES

  • 1.Chaudhuri J, Alt FW. Class-switch recombination: interplay of transcription, DNA deamination and DNA repair. Nat Rev Immunol. 2004;4:541–552. doi: 10.1038/nri1395. [DOI] [PubMed] [Google Scholar]
  • 2.Tong P, Wesemann DR. Molecular mechanisms of IgE class switch recombination. Curr Top Microbiol Immunol. 2015;388:21–37. doi: 10.1007/978-3-319-13725-4_2. [DOI] [PubMed] [Google Scholar]
  • 3.Stavnezer J, Schrader CE. IgH chain class switch recombination: mechanism and regulation. J Immunol. 2014;193:5370–5378. doi: 10.4049/jimmunol.1401849. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Cheng G, Cleary AM, Ye Z, Hong DI, Lederman S, Baltimore D. Involvement of CRAF1, a relative of TRAF, in CD40 signaling. Science. 1995;267:1494–1498. doi: 10.1126/science.7533327. [DOI] [PubMed] [Google Scholar]
  • 5.Ishida T, Mizushima S, Azuma S, Kobayashi N, Tojo T, Suzuki K, et al. Identification of TRAF6, a novel tumor necrosis factor receptor-associated protein that mediates signaling from an amino-terminal domain of the CD40 cytoplasmic region. J Biol Chem. 1996;271:28745–28748. doi: 10.1074/jbc.271.46.28745. [DOI] [PubMed] [Google Scholar]
  • 6.Shen CH, Stavnezer J. Activation of the mouse Ig germline epsilon promoter by IL-4 is dependent on AP1 transcription factors. J Immunol. 2001;166:411–423. doi: 10.4049/jimmunol.166.1.411. [DOI] [PubMed] [Google Scholar]
  • 7.Allen RC, Armitage RJ, Conley ME, Rosenblatt H, Jenkins NA, Copeland NG, et al. CD40 ligand gene defects responsible for X-linked hyper-IgM syndrome. Science. 1993;259:990–993. doi: 10.1126/science.7679801. [DOI] [PubMed] [Google Scholar]
  • 8.Aruffo A, Farrington M, Hollenbaugh D, Li X, Milatovich A, Nonoyama S, et al. The CD40 ligand, gp39, is defective in activated T cells from patients with X-linked hyper-IgM syndrome. Cell. 1993;72:291–300. doi: 10.1016/0092-8674(93)90668-g. [DOI] [PubMed] [Google Scholar]
  • 9.DiSanto JP, Bonnefoy JY, Gauchat JF, Fischer A, de Saint Basile G. CD40 ligand mutations in x-linked immunodeficiency with hyper-IgM. Nature. 1993;361:541–543. doi: 10.1038/361541a0. [DOI] [PubMed] [Google Scholar]
  • 10.Fuleihan R, Ramesh N, Loh R, Jabara H, Rosen RS, Chatila T, et al. Defective expression of the CD40 ligand in X chromosome-linked immunoglobulin deficiency with normal or elevated IgM. Proc Natl Acad Sci USA. 1993;90:2170–2173. doi: 10.1073/pnas.90.6.2170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Korthauer U, Graf D, Mages HW, Briere F, Padayachee M, Malcolm S, et al. Defective expression of T-cell CD40 ligand causes X-linked immunodeficiency with hyper-IgM. Nature. 1993;361:539–541. doi: 10.1038/361539a0. [DOI] [PubMed] [Google Scholar]
  • 12.Castigli E, Geha RS. TACI, isotype switching, CVID and IgAD. Immunol Res. 2007;38:102–111. doi: 10.1007/s12026-007-8000-2. [DOI] [PubMed] [Google Scholar]
  • 13.Castigli E, Wilson SA, Scott S, Dedeoglu F, Xu S, Lam KP, et al. TACI and BAFF-R mediate isotype switching in B cells. J Exp Med. 2005;201:35–39. doi: 10.1084/jem.20032000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Kato A, Truong-Tran AQ, Scott AL, Matsumoto K, Schleimer RP. Airway epithelial cells produce B cell-activating factor of TNF family by an IFN-beta-dependent mechanism. J Immunol. 2006;177:7164–7172. doi: 10.4049/jimmunol.177.10.7164. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Kato A, Xiao H, Chustz RT, Liu MC, Schleimer RP. Local release of B cell-activating factor of the TNF family after segmental allergen challenge of allergic subjects. J Allergy Clin Immunol. 2009;123:369–375. doi: 10.1016/j.jaci.2008.11.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Cameron L, Gounni AS, Frenkiel S, Lavigne F, Vercelli D, Hamid Q. S epsilon S mu and S epsilon S gamma switch circles in human nasal mucosa following ex vivo allergen challenge: evidence for direct as well as sequential class switch recombination. J Immunol. 2003;171:3816–3822. doi: 10.4049/jimmunol.171.7.3816. [DOI] [PubMed] [Google Scholar]
  • 17.McCoy KD, Harris NL, Diener P, Hatak S, Odermatt B, Hangartner L, et al. Natural IgE production in the absence of MHC Class II cognate help. Immunity. 2006;24:329–339. doi: 10.1016/j.immuni.2006.01.013. [DOI] [PubMed] [Google Scholar]
  • 18.Del Prete G, Maggi E, Parronchi P, Chretien I, Tiri A, Macchia D, et al. IL-4 is an essential factor for the IgE synthesis induced in vitro by human T cells clones and their supernatants. J Immunol. 1988;140:4193–4198. [PubMed] [Google Scholar]
  • 19.Vercelli D, Jabara H, Arai K-I, Geha R. Induction of human IgE synthesis requires interleukin 4 and T/B interactions involving the T cell receptor/CD3 complex and MHC class II antigens. J Exp Med. 1989;169:1295–1307. doi: 10.1084/jem.169.4.1295. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Gauchat J-F, Lebman D, Coffman R, Gascan H, de Vries J. Structure and expression of germline e transcripts in human B cells induced by interleukin 4 to switch to IgE production. J Exp Med. 1990;172:463–473. doi: 10.1084/jem.172.2.463. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Jung S, Rajewsky K, Radbruch A. Shutdown of class switch recombination by deletion of a switch region control element. Science. 1993;259:984–987. doi: 10.1126/science.8438159. [DOI] [PubMed] [Google Scholar]
  • 22.Zhang J, Bottaro A, Li S, Stewart V, Alt FW. A selective defect in IgG2b switching as a result of targeted mutation of the I gamma 2b promoter and exon. EMBO J. 1993;12:3529–3537. doi: 10.1002/j.1460-2075.1993.tb06027.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Seidl KJ, Bottaro A, Vo A, Zhang J, Davidson L, Alt FW. An expressed neo(r) cassette provides required functions of the 1 gamma2b exon for class switching. Int Immunol. 1998;10:1683–1692. doi: 10.1093/intimm/10.11.1683. [DOI] [PubMed] [Google Scholar]
  • 24.Hackney JA, Misaghi S, Senger K, Garris C, Sun Y, Lorenzo MN, et al. DNA targets of AID evolutionary link between antibody somatic hypermutation and class switch recombination. Adv Immunol. 2009;101:163–189. doi: 10.1016/S0065-2776(08)01005-5. [DOI] [PubMed] [Google Scholar]
  • 25.Pavri R, Gazumyan A, Jankovic M, Di Virgilio M, Klein I, Ansarah-Sobrinho C, et al. Activation-induced cytidine deaminase targets DNA at sites of RNA polymerase II stalling by interaction with Spt5. Cell. 2010;143:122–133. doi: 10.1016/j.cell.2010.09.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Longerich S, Basu U, Alt F, Storb U. AID in somatic hypermutation and class switch recombination. Curr Opin Immunol. 2006;18:164–174. doi: 10.1016/j.coi.2006.01.008. [DOI] [PubMed] [Google Scholar]
  • 27.Soulas-Sprauel P, Rivera-Munoz P, Malivert L, Le Guyader G, Abramowski V, Revy P, et al. V(D)J and immunoglobulin class switch recombinations: a paradigm to study the regulation of DNA end-joining. Oncogene. 2007;26:7780–7791. doi: 10.1038/sj.onc.1210875. [DOI] [PubMed] [Google Scholar]
  • 28.Revy P, Muto T, Levy Y, Geissmann F, Plebani A, Sanal O, et al. Activation-induced cytidine deaminase (AID) deficiency causes the autosomal recessive form of the hyper-IgM syndrome (HIGM2) Cell. 2000;102:565–575. doi: 10.1016/s0092-8674(00)00079-9. [DOI] [PubMed] [Google Scholar]
  • 29.Manis JP, Dudley D, Kaylor L, Alt FW. IgH class switch recombination to IgG1 in DNA-PKcs-deficient B cells. Immunity. 2002;16:607–617. doi: 10.1016/s1074-7613(02)00306-0. [DOI] [PubMed] [Google Scholar]
  • 30.Manis JP, Gu Y, Lansford R, Sonoda E, Ferrini R, Davidson L, et al. Ku70 is required for late B cell development and immunoglobulin heavy chain class switching. J Exp Med. 1998;187:2081–2089. doi: 10.1084/jem.187.12.2081. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Wuerffel R, Wang L, Grigera F, Manis J, Selsing E, Perlot T, et al. SS synapsis during class switch recombination is promoted by distantly located transcriptional elements and activation-induced deaminase. Immunity. 2007;27:711–722. doi: 10.1016/j.immuni.2007.09.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Laffleur B, Duchez S, Tarte K, Denis-Lagache N, Peron S, Carrion C, et al. Self-restrained B cells arise following membrane IgE expression. Cell Rep. 2015 doi: 10.1016/j.celrep.2015.01.023. [Epub ahead of print] [DOI] [PubMed] [Google Scholar]
  • 33.Xiong H, Dolpady J, Wabl M, Curotto de Lafaille MA, Lafaille JJ. Sequential class switching is required for the generation of high affinity IgE antibodies. J Exp Med. 2012;209:353–364. doi: 10.1084/jem.20111941. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Akdis M, Akdis CA. IgE class switching and cellular memory. Nat Immunol. 2012;13:312–314. doi: 10.1038/ni.2266. [DOI] [PubMed] [Google Scholar]
  • 35.Wu LC, Zarrin AA. The production and regulation of IgE by the immune system. Nat Rev Immunol. 2014;14:247–259. doi: 10.1038/nri3632. [DOI] [PubMed] [Google Scholar]
  • 36.Talay O, Yan D, Brightbill HD, Straney EE, Zhou M, Ladi E, et al. IgE(+) memory B cells and plasma cells generated through a germinal-center pathway. Nat Immunol. 2012;13:396–404. doi: 10.1038/ni.2256. [DOI] [PubMed] [Google Scholar]
  • 37.Yang Z, Sullivan BM, Allen CD. Fluorescent in vivo detection reveals that IgE(+) B cells are restrained by an intrinsic cell fate predisposition. Immunity. 2012;36:857–872. doi: 10.1016/j.immuni.2012.02.009. [DOI] [PubMed] [Google Scholar]
  • 38.He JS, Meyer-Hermann M, Xiangying D, Zuan LY, Jones LA, Ramakrishna L, et al. The distinctive germinal center phase of IgE+ B lymphocytes limits their contribution to the classical memory response. J Exp Med. 2013;210:2755–2771. doi: 10.1084/jem.20131539. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Looney TJ, Lee JY, Roskin KM, Hoh RA, King J, Glanville J, et al. Human B-cell isotype switching origins of IgE. J Allergy Clin Immunol. 2016;137 doi: 10.1016/j.jaci.2015.07.014. 579-86.e7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Garman SC, Wurzburg BA, Tarchevskaya SS, Kinet JP, Jardetzky TS. Structure of the Fc fragment of human IgE bound to its high-affinity receptor Fc epsilonRI alpha. Nature. 2000;406:259–266. doi: 10.1038/35018500. [DOI] [PubMed] [Google Scholar]
  • 41.Wan T, Beavil RL, Fabiane SM, Beavil AJ, Sohi MK, Keown M, et al. The crystal structure of IgE Fc reveals an asymmetrically bent conformation. Nat Immunol. 2002;3:681–686. doi: 10.1038/ni811. [DOI] [PubMed] [Google Scholar]
  • 42.Arnold JN, Wormald MR, Sim RB, Rudd PM, Dwek RA. The impact of glycosylation on the biological function and structure of human immunoglobulins. Annu Rev Immunol. 2007;25:21–50. doi: 10.1146/annurev.immunol.25.022106.141702. [DOI] [PubMed] [Google Scholar]
  • 43.Shade KT, Platzer B, Washburn N, Mani V, Bartsch YC, Conroy M, et al. A single glycan on IgE is indispensable for initiation of anaphylaxis. J Exp Med. 2015;212:457–467. doi: 10.1084/jem.20142182. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Legendre C, Caillat-Zucman S, Samuel D, Morelon S, Bismuth H, Bach JF, et al. Transfer of symptomatic peanut allergy to the recipient of a combined liver-and-kidney transplant. N Engl J Med. 1997;337:822–824. doi: 10.1056/NEJM199709183371204. [DOI] [PubMed] [Google Scholar]
  • 45.Castells M, Boyce J. Transfer of peanut allergy by a liver allograft. N Engl J Med. 1998;338:202–203. doi: 10.1056/NEJM199801153380319. [DOI] [PubMed] [Google Scholar]
  • 46.Kinet JP. The high-affinity IgE receptor (Fc epsilon RI): from physiology to pathology. Annu Rev Immunol. 1999;17:931–972. doi: 10.1146/annurev.immunol.17.1.931. [DOI] [PubMed] [Google Scholar]
  • 47.Malveaux FJ, Conroy MC, Adkinson NF, Jr, Lichtenstein LM. IgE receptors on human basophils. Relationship to serum IgE concentration. J Clin Invest. 1978;62:176–181. doi: 10.1172/JCI109103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Furuichi K, Rivera J, Isersky C. The receptor for immunoglobulin E on rat basophilic leukemia cells: effect of ligand binding on receptor expression. Proc Natl Acad Sci USA. 1985;82:1522–1525. doi: 10.1073/pnas.82.5.1522. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Lantz CS, Yamaguchi M, Oettgen HC, Katona IM, Miyajima I, Kinet JP, et al. IgE regulates mouse basophil Fc epsilon RI expression in vivo. J Immunol. 1997;158:2517–2521. [PubMed] [Google Scholar]
  • 50.Yamaguchi M, Lantz CS, Oettgen HC, Katona IM, Fleming T, Miyajima I, et al. IgE enhances mouse mast cell Fc(epsilon)RI expression in vitro and in vivo: evidence for a novel amplification mechanism in IgE-dependent reactions. J Exp Med. 1997;185:663–672. doi: 10.1084/jem.185.4.663. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Yamaguchi M, Sayama K, Yano K, Lantz CS, Noben-Trauth N, Ra C, et al. IgE enhances Fc epsilon receptor I expression and IgE-dependent release of histamine and lipid mediators from human umbilical cord blood-derived mast cells: synergistic effect of IL-4 and IgE on human mast cell Fc epsilon receptor I expression and mediator release. J Immunol. 1999;162:5455–5465. [PubMed] [Google Scholar]
  • 52.Saini SS, Klion AD, Holland SM, Hamilton RG, Bochner BS, Macglashan DW., Jr The relationship between serum IgE and surface levels of FcepsilonR on human leukocytes in various diseases: correlation of expression with FcepsilonRI on basophils but not on monocytes or eosinophils. J Allergy Clin Immunol. 2000;106:514–520. doi: 10.1067/mai.2000.108431. [DOI] [PubMed] [Google Scholar]
  • 53.Dehlink E, Platzer B, Baker AH, Larosa J, Pardo M, Dwyer P, et al. A soluble form of the high affinity IgE receptor, FcepsilonRI, circulates in human serum. PLoS One. 2011;6:e19098. doi: 10.1371/journal.pone.0019098. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Kulczycki A, Jr, Isersky C, Metzger H. The interaction of IgE with rat basophilic leukemia cells. I. Evidence for specific binding of IgE. J Exp Med. 1974;139:600–616. doi: 10.1084/jem.139.3.600. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Gilfillan AM, Rivera J. The tyrosine kinase network regulating mast cell activation. Immunol Rev. 2009;228:149–169. doi: 10.1111/j.1600-065X.2008.00742.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Shampain MP, Behrens BL, Larsen GL, Henson PM. An animal model of late pulmonary responses to Alternaria challenge. Am Rev Respir Dis. 1982;126:493–498. doi: 10.1164/arrd.1982.126.3.493. [DOI] [PubMed] [Google Scholar]
  • 57.Cockcroft DW, Murdock KY. Comparative effects of inhaled salbutamol, sodium cromoglycate and beclomethasone dipropionate on allergen-induced early astmatic responses, late asthmatic responses and increased bronchial responsiveness to histamine. J Allergy Clin Immunol. 1987;79:734–740. doi: 10.1016/0091-6749(87)90204-1. [DOI] [PubMed] [Google Scholar]
  • 58.Macglashan D, Miura K. Loss of syk kinase during IgE-mediated stimulation of human basophils. J Allergy Clin Immunol. 2004;114:1317–1324. doi: 10.1016/j.jaci.2004.08.037. [DOI] [PubMed] [Google Scholar]
  • 59.Burton OT, Logsdon SL, Zhou JS, Medina-Tamayo J, Abdel-Gadir A, Noval Rivas M, et al. Oral immunotherapy induces IgG antibodies that act through FcγRIIb to suppress IgE-mediated hypersensitivity. J Allergy Clin Immunol. 2014;134 doi: 10.1016/j.jaci.2014.05.042. 1301-7.e6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Daeron M, Latour S, Malbec O, Espinosa E, Pina P, Pasmans S, et al. The same tyrosine-based inhibition motif, in the intracytoplasmic domain of Fc gamma RIIB, regulates negatively BCR-, TCR-, and FcR-dependent cell activation. Immunity. 1995;3:635–646. doi: 10.1016/1074-7613(95)90134-5. [DOI] [PubMed] [Google Scholar]
  • 61.Logsdon SL, Oettgen HC. Anti-IgE therapy: clinical utility and mechanistic insights. Curr Top Microbiol Immunol. 2015;388:39–61. doi: 10.1007/978-3-319-13725-4_3. [DOI] [PubMed] [Google Scholar]
  • 62.Fahy JV, Fleming HE, Wong HH, Liu JT, Su JQ, Reimann J, et al. The effect of an anti-IgE monoclonal antibody on the early- and late-phase responses to allergen inhalation in asthmatic subjects. Am J Respir Crit Care Med. 1997;155:1828–1834. doi: 10.1164/ajrccm.155.6.9196082. [DOI] [PubMed] [Google Scholar]
  • 63.Boulet LP, Chapman KR, Cote J, Kalra S, Bhagat R, Swystun VA, et al. Inhibitory effects of an anti-IgE antibody E25 on allergen-induced early asthmatic response. Am J Respir Crit Care Med. 1997;155:1835–1840. doi: 10.1164/ajrccm.155.6.9196083. [DOI] [PubMed] [Google Scholar]
  • 64.Soler M, Matz J, Townley R, Buhl R, O’Brien J, Fox H, et al. The anti-IgE antibody omalizumab reduces exacerbations and steroid requirement in allergic asthmatics. Eur Respir J. 2001;18:254–261. doi: 10.1183/09031936.01.00092101. [DOI] [PubMed] [Google Scholar]
  • 65.Busse W, Corren J, Lanier BQ, McAlary M, Fowler-Taylor A, Cioppa GD, et al. Omalizumab, anti-IgE recombinant humanized monoclonal antibody, for the treatment of severe allergic asthma. J Allergy Clin Immunol. 2001;108:184–190. doi: 10.1067/mai.2001.117880. [DOI] [PubMed] [Google Scholar]
  • 66.Mehlhop PD, van de Rijn M, Goldberg AB, Brewer JP, Kurup VP, Martin TR, et al. Allergen-induced bronchial hyperreactivity and eosinophilic inflammation occur in the absence of IgE in a mouse model of asthma. Proc Natl Acad Sci USA. 1997;94:1344–1349. doi: 10.1073/pnas.94.4.1344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Macglashan DW, Jr, Bochner BS, Adelman DC, Jardieu PM, Togias A, McKenzie-White J, et al. Down-regulation of Fc(epsilon)RI expression on human basophils during in vivo treatment of atopic patients with anti-IgE antibody. J Immunol. 1997;158:1438–1445. [PubMed] [Google Scholar]
  • 68.Beck LA, Marcotte GV, MacGlashan D, Togias A, Saini S. Omalizumab-induced reductions in mast cell Fce psilon RI expression and function. J Allergy Clin Immunol. 2004;114:527–530. doi: 10.1016/j.jaci.2004.06.032. [DOI] [PubMed] [Google Scholar]
  • 69.Prussin C, Griffith DT, Boesel KM, Lin H, Foster B, Casale TB. Omalizumab treatment downregulates dendritic cell FcepsilonRI expression. J Allergy Clin Immunol. 2003;112:1147–1154. doi: 10.1016/j.jaci.2003.10.003. [DOI] [PubMed] [Google Scholar]
  • 70.Zaidi AK, Saini SS, Macglashan DW., Jr Regulation of Syk kinase and FcRbeta expression in human basophils during treatment with omalizumab. J Allergy Clin Immunol. 2010;125 doi: 10.1016/j.jaci.2009.12.996. 902-8.e7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Harris JM, Maciuca R, Bradley MS, Cabanski CR, Scheerens H, Lim J, et al. A randomized trial of the efficacy and safety of quilizumab in adults with inadequately controlled allergic asthma. Respir Res. 2016;17:29. doi: 10.1186/s12931-016-0347-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Casale TB, Condemi J, LaForce C, Nayak A, Rowe M, Watrous M, et al. Effect of omalizumab on symptoms of seasonal allergic rhinitis: a randomized controlled trial. JAMA. 2001;286:2956–2967. doi: 10.1001/jama.286.23.2956. [DOI] [PubMed] [Google Scholar]
  • 73.Eggel A, Baravalle G, Hobi G, Kim B, Buschor P, Forrer P, et al. Accelerated dissociation of IgE-FcεRI complexes by disruptive inhibitors actively desensitizes allergic effector cells. J Allergy Clin Immunol. 2014;133 doi: 10.1016/j.jaci.2014.02.005. 1709-19.e8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Kim B, Eggel A, Tarchevskaya SS, Vogel M, Prinz H, Jardetzky TS. Accelerated disassembly of IgE-receptor complexes by a disruptive macromolecular inhibitor. Nature. 2012;491:613–617. doi: 10.1038/nature11546. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Zhou JS, Sandomenico A, Severino V, Burton OT, Darling A, Oettgen HC, et al. An IgE receptor mimetic peptide (PepE) protects mice from IgE mediated anaphylaxis. Mol Biosyst. 2013;9:2853–2859. doi: 10.1039/c3mb70286c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Helm B, Kebo D, Vercelli D, Glovsky MM, Gould H, Ishizaka K, et al. Blocking of passive sensitization of human mast cells and basophil granulocytes with IgE antibodies by a recombinant human epsilon-chain fragment of 76 amino acids. Proc Natl Acad Sci USA. 1989;86:9465–9469. doi: 10.1073/pnas.86.23.9465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Nakamura GR, Reynolds ME, Chen YM, Starovasnik MA, Lowman HB. Stable “zeta” peptides that act as potent antagonists of the high-affinity IgE receptor. Proc Natl Acad Sci USA. 2002;99:1303–1308. doi: 10.1073/pnas.022635599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Sandomenico A, Monti SM, Marasco D, Dathan N, Palumbo R, Saviano M, et al. IgE-binding properties and selectivity of peptide mimics of the FcvarepsilonRI binding site. Mol Immunol. 2009;46:3300–3309. doi: 10.1016/j.molimm.2009.07.025. [DOI] [PubMed] [Google Scholar]
  • 79.Wiegand TW, Williams PB, Dreskin SC, Jouvin MH, Kinet JP, Tasset D. High-affinity oligonucleotide ligands to human IgE inhibit binding to Fc epsilon receptor I. J Immunol. 1996;157:221–230. [PubMed] [Google Scholar]
  • 80.Asai K, Kitaura J, Kawakami Y, Yamagata N, Tsai M, Carbone DP, et al. Regulation of mast cell survival by IgE. Immunity. 2001;14:791–800. doi: 10.1016/s1074-7613(01)00157-1. [DOI] [PubMed] [Google Scholar]
  • 81.Kalesnikoff J, Huber M, Lam V, Damen JE, Zhang J, Siraganian RP, et al. Monomeric IgE stimulates signaling pathways in mast cells that lead to cytokine production and cell survival. Immunity. 2001;14:801–811. doi: 10.1016/s1074-7613(01)00159-5. [DOI] [PubMed] [Google Scholar]
  • 82.Gurish MF, Bryce PJ, Tao H, Kisselgof AB, Thornton EM, Miller HR, et al. IgE enhances parasite clearance and regulates mast cell responses in mice infected with Trichinella spiralis . J Immunol. 2004;172:1139–1145. doi: 10.4049/jimmunol.172.2.1139. [DOI] [PubMed] [Google Scholar]
  • 83.Mathias CB, Freyschmidt EJ, Caplan B, Jones T, Poddighe D, Xing W, et al. IgE influences the number and function of mature mast cells, but not progenitor recruitment in allergic pulmonary inflammation. J Immunol. 2009;182:2416–2424. doi: 10.4049/jimmunol.0801569. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Burton OT, Noval Rivas M, Zhou JS, Logsdon SL, Darling AR, Koleoglou KJ, et al. Immunoglobulin E signal inhibition during allergen ingestion leads to reversal of established food allergy and induction of regulatory T cells. Immunity. 2014;41:141–151. doi: 10.1016/j.immuni.2014.05.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Kitaura J, Song J, Tsai M, Asai K, Maeda-Yamamoto M, Mocsai A, et al. Evidence that IgE molecules mediate a spectrum of effects on mast cell survival and activation via aggregation of the FcepsilonRI. Proc Natl Acad Sci USA. 2003;100:12911–12916. doi: 10.1073/pnas.1735525100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Tanaka S, Takasu Y, Mikura S, Satoh N, Ichikawa A. Antigen-independent induction of histamine synthesis by immunoglobulin E in mouse bone marrow-derived mast cells. J Exp Med. 2002;196:229–235. doi: 10.1084/jem.20012037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Pandey V, Mihara S, Fensome-Green A, Bolsover S, Cockcroft S. Monomeric IgE stimulates NFAT translocation into the nucleus, a rise in cytosol Ca2+, degranulation, and membrane ruffling in the cultured rat basophilic leukemia-2H3 mast cell line. J Immunol. 2004;172:4048–4058. doi: 10.4049/jimmunol.172.7.4048. [DOI] [PubMed] [Google Scholar]
  • 88.Bryce PJ, Miller ML, Miyajima I, Tsai M, Galli SJ, Oettgen HC. Immune sensitization in the skin is enhanced by antigen-independent effects of IgE. Immunity. 2004;20:381–392. doi: 10.1016/s1074-7613(04)00080-9. [DOI] [PubMed] [Google Scholar]
  • 89.Kawakami T, Kitaura J. Mast cell survival and activation by IgE in the absence of antigen: a consideration of the biologic mechanisms and relevance. J Immunol. 2005;175:4167–4173. doi: 10.4049/jimmunol.175.7.4167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Bax HJ, Bowen H, Dodev TS, Sutton BJ, Gould HJ. Mechanism of the antigen-independent cytokinergic SPE-7 IgE activation of human mast cells in vitro. Sci Rep. 2015;5:9538. doi: 10.1038/srep09538. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.James LC, Roversi P, Tawfik DS. Antibody multispecificity mediated by conformational diversity. Science. 2003;299:1362–1367. doi: 10.1126/science.1079731. [DOI] [PubMed] [Google Scholar]
  • 92.Maurer M, Rosen K, Hsieh HJ, Saini S, Grattan C, Gimenez-Arnau A, et al. Omalizumab for the treatment of chronic idiopathic or spontaneous urticaria. N Engl J Med. 2013;368:924–935. doi: 10.1056/NEJMoa1215372. [DOI] [PubMed] [Google Scholar]
  • 93.MacDonald SM, Rafnar T, Langdon J, Lichtenstein LM. Molecular identification of an IgE-dependent histamine-releasing factor. Science. 1995;269:688–690. doi: 10.1126/science.7542803. [DOI] [PubMed] [Google Scholar]
  • 94.Kashiwakura JC, Ando T, Matsumoto K, Kimura M, Kitaura J, Matho MH, et al. Histamine-releasing factor has a proinflammatory role in mouse models of asthma and allergy. J Clin Invest. 2012;122:218–228. doi: 10.1172/JCI59072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Kawakami T, Kashiwakura J, Kawakami Y. Histamine-releasing factor and immunoglobulins in asthma and allergy. Allergy Asthma Immunol Res. 2014;6:6–12. doi: 10.4168/aair.2014.6.1.6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Lee WT, Conrad DH. Murine B cell hybridomas bearing ligand-inducible Fc receptors for IgE. J Immunol. 1986;136:4573–4580. [PubMed] [Google Scholar]
  • 97.Gould HJ, Sutton BJ. IgE in allergy and asthma today. Nat Rev Immunol. 2008;8:205–217. doi: 10.1038/nri2273. [DOI] [PubMed] [Google Scholar]
  • 98.Dierks SE, Bartlett WC, Edmeades RL, Gould HJ, Rao M, Conrad DH. The oligomeric nature of the murine Fc epsilon RII/CD23. Implications for function. J Immunol. 1993;150:2372–2382. [PubMed] [Google Scholar]
  • 99.Aubry J-P, Pochon S, Graber P, Jansen K, Bonnefoy J-Y. CD21 is a ligand for CD23 and regulates IgE production. Nature. 1992;358:505–507. doi: 10.1038/358505a0. [DOI] [PubMed] [Google Scholar]
  • 100.Kisselgof AB, Oettgen HC. The expression of murine B cell CD23, in vivo, is regulated by its ligand, IgE. Int Immunol. 1998;10:1377–1384. doi: 10.1093/intimm/10.9.1377. [DOI] [PubMed] [Google Scholar]
  • 101.Yang PC, Berin MC, Yu LC, Conrad DH, Perdue MH. Enhanced intestinal transepithelial antigen transport in allergic rats is mediated by IgE and CD23 (FcepsilonRII) J Clin Invest. 2000;106:879–886. doi: 10.1172/JCI9258. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Heyman B, Liu T, Gustavsson S. In vivo enhancement of the specific antibody response via the low-affinity receptor for IgE. Eur J Immunol. 1993;23:1739–1742. doi: 10.1002/eji.1830230754. [DOI] [PubMed] [Google Scholar]
  • 103.Gustavsson S, Hjulstrom S, Liu T, Heyman B. CD23/IgE-mediated regulation of the specific antibody response in vivo. J Immunol. 1994;152:4793–4800. [PubMed] [Google Scholar]
  • 104.Rosenwasser LJ. Mechanisms of IgE Inflammation. Curr Allergy Asthma Rep. 2011;11:178–183. doi: 10.1007/s11882-011-0179-6. [DOI] [PubMed] [Google Scholar]
  • 105.Gustavsson S, Wernersson S, Heyman B. Restoration of the antibody response to IgE/antigen complexes in CD23- deficient mice by CD23+ spleen or bone marrow cells. J Immunol. 2000;164:3990–3995. doi: 10.4049/jimmunol.164.8.3990. [DOI] [PubMed] [Google Scholar]
  • 106.Martin RK, Brooks KB, Henningsson F, Heyman B, Conrad DH. Antigen transfer from exosomes to dendritic cells as an explanation for the immune enhancement seen by IgE immune complexes. PLoS One. 2014;9:e110609. doi: 10.1371/journal.pone.0110609. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Sherr E, Macy E, Kimata H, Gilly M, Saxon A. Binding the low affinity Fc epsilon R on B cells suppresses ongoing human IgE synthesis. J Immunol. 1989;142:481–489. [PubMed] [Google Scholar]
  • 108.Payet ME, Woodward EC, Conrad DH. Humoral response suppression observed with CD23 transgenics. J Immunol. 1999;163:217–223. [PubMed] [Google Scholar]
  • 109.Payet-Jamroz M, Helm SL, Wu J, Kilmon M, Fakher M, Basalp A, et al. Suppression of IgE responses in CD23-transgenic animals is due to expression of CD23 on nonlymphoid cells. J Immunol. 2001;166:4863–4869. doi: 10.4049/jimmunol.166.8.4863. [DOI] [PubMed] [Google Scholar]
  • 110.Yu P, Kosco-Vilbois M, Richards M, Kohler G, Lamers MC. Negative feedback regulation of IgE synthesis by murine CD23. Nature. 1994;369:753–756. doi: 10.1038/369753a0. [DOI] [PubMed] [Google Scholar]
  • 111.Cernadas M, De Sanctis GT, Krinzman SJ, Mark DA, Donovan CE, Listman JA, et al. CD23 and allergic pulmonary inflammation: potential role as an inhibitor. Am J Respir Cell Mol Biol. 1999;20:1–8. doi: 10.1165/ajrcmb.20.1.3299. [DOI] [PubMed] [Google Scholar]
  • 112.Haczku A, Takeda K, Hamelmann E, Loader J, Joetham A, Redai I, et al. CD23 exhibits negative regulatory effects on allergic sensitization and airway hyperresponsiveness. Am J Respir Crit Care Med. 2000;161:952–960. doi: 10.1164/ajrccm.161.3.9905046. [DOI] [PubMed] [Google Scholar]
  • 113.Rosenwasser LJ, Meng J. Anti-CD23. Clin Rev Allergy Immunol. 2005;29:61–72. doi: 10.1385/CRIAI:29:1:061. [DOI] [PubMed] [Google Scholar]
  • 114.Rosenwasser LJ, Busse WW, Lizambri RG, Olejnik TA, Totoritis MC. Allergic asthma and an anti-CD23 mAb (IDEC-152): results of a phase I, single-dose, dose-escalating clinical trial. J Allergy Clin Immunol. 2003;112:563–570. doi: 10.1016/s0091-6749(03)01861-x. [DOI] [PubMed] [Google Scholar]
  • 115.Saxon A, Kurbe LM, Behle K, Max EE, Zhang K. Inhibition of human IgE production via Fc epsilon R-II stimulation results from a decrease in the mRNA for secreted but not membrane epsilon H chains. J Immunol. 1991;147:4000–4006. [PubMed] [Google Scholar]
  • 116.Joseph M, Auriault C, Capron A, Vorng H, Viens P. A new function for platelets: IgE-dependent killing of schistosomes. Nature. 1983;303:810–812. doi: 10.1038/303810a0. [DOI] [PubMed] [Google Scholar]
  • 117.King CL, Xianli J, Malhotra I, Liu S, Mahmoud AA, Oettgen HC. Mice with a targeted deletion of the IgE gene have increased worm burdens and reduced granulomatous inflammation following primary infection with Schistosoma mansoni . J Immunol. 1997;158:294–300. [PubMed] [Google Scholar]
  • 118.van den Biggelaar AH, van Ree R, Rodrigues LC, Lell B, Deelder AM, Kremsner PG, et al. Decreased atopy in children infected with Schistosoma haematobium: a role for parasite-induced interleukin-10. Lancet. 2000;356:1723–1727. doi: 10.1016/S0140-6736(00)03206-2. [DOI] [PubMed] [Google Scholar]
  • 119.Cooper PJ, Chico ME, Rodrigues LC, Ordonez M, Strachan D, Griffin GE, et al. Reduced risk of atopy among school-age children infected with geohelminth parasites in a rural area of the tropics. J Allergy Clin Immunol. 2003;111:995–1000. doi: 10.1067/mai.2003.1348. [DOI] [PubMed] [Google Scholar]
  • 120.Hunninghake GM, Soto-Quiros ME, Avila L, Ly NP, Liang C, Sylvia JS, et al. Sensitization to Ascaris lumbricoides and severity of childhood asthma in Costa Rica. J Allergy Clin Immunol. 2007;119:654–661. doi: 10.1016/j.jaci.2006.12.609. [DOI] [PubMed] [Google Scholar]
  • 121.Dold S, Heinrich J, Wichmann HE, Wjst M. Ascaris-specific IgE and allergic sensitization in a cohort of school children in the former East Germany. J Allergy Clin Immunol. 1998;102:414–420. doi: 10.1016/s0091-6749(98)70129-0. [DOI] [PubMed] [Google Scholar]
  • 122.Alshishtawy MM, Abdella AM, Gelber LE, Chapman MD. Asthma in Tanta, Egypt: serologic analysis of total and specific IgE antibody levels and their relationship to parasite infection. Int Arch Allergy Appl Immunol. 1991;96:348–354. doi: 10.1159/000235520. [DOI] [PubMed] [Google Scholar]
  • 123.Sereda MJ, Hartmann S, Lucius R. Helminths and allergy: the example of tropomyosin. Trends Parasitol. 2008;24:272–278. doi: 10.1016/j.pt.2008.03.006. [DOI] [PubMed] [Google Scholar]
  • 124.Brown SJ, Galli SJ, Gleich GJ, Askenase PW. Ablation of immunity to Amblyomma americanum by anti-basophil serum: cooperation between basophils and eosinophils in expression of immunity to ectoparasites (ticks) in guinea pigs. J Immunol. 1982;129:790–796. [PubMed] [Google Scholar]
  • 125.Matsuda H, Watanabe N, Kiso Y, Hirota S, Ushio H, Kannan Y, et al. Necessity of IgE antibodies and mast cells for manifestation of resistance against larval Haemaphysalis longicornis ticks in mice. J Immunol. 1990;144:259–262. [PubMed] [Google Scholar]
  • 126.Wada T, Ishiwata K, Koseki H, Ishikura T, Ugajin T, Ohnuma N, et al. Selective ablation of basophils in mice reveals their nonredundant role in acquired immunity against ticks. J Clin Invest. 2010;120:2867–2875. doi: 10.1172/JCI42680. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Steeves EB, Allen JR. Basophils in skin reactions of mast cell-deficient mice infested with Dermacentor variabilis . Int J Parasitol. 1990;20:655–667. doi: 10.1016/0020-7519(90)90124-6. [DOI] [PubMed] [Google Scholar]
  • 128.Commins SP, James HR, Kelly LA, Pochan SL, Workman LJ, Perzanowski MS, et al. The relevance of tick bites to the production of IgE antibodies to the mammalian oligosaccharide galactose-alpha-1,3-galactose. J Allergy Clin Immunol. 2011;127 doi: 10.1016/j.jaci.2011.02.019. 1286-93.e6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Chung CH, Mirakhur B, Chan E, Le QT, Berlin J, Morse M, et al. Cetuximab-induced anaphylaxis and IgE specific for galactose-alpha-1,3-galactose. N Engl J Med. 2008;358:1109–1117. doi: 10.1056/NEJMoa074943. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Higginbotham RD. Mast cells and local resistance to Russell’s viper venom. J Immunol. 1965;95:867–875. [PubMed] [Google Scholar]
  • 131.Metz M, Piliponsky AM, Chen CC, Lammel V, Abrink M, Pejler G, et al. Mast cells can enhance resistance to snake and honeybee venoms. Science. 2006;313:526–530. doi: 10.1126/science.1128877. [DOI] [PubMed] [Google Scholar]
  • 132.Akahoshi M, Song CH, Piliponsky AM, Metz M, Guzzetta A, Abrink M, et al. Mast cell chymase reduces the toxicity of Gila monster venom, scorpion venom, and vasoactive intestinal polypeptide in mice. J Clin Invest. 2011;121:4180–4191. doi: 10.1172/JCI46139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Marichal T, Starkl P, Reber LL, Kalesnikoff J, Oettgen HC, Tsai M, et al. A beneficial role for immunoglobulin E in host defense against honeybee venom. Immunity. 2013;39:963–975. doi: 10.1016/j.immuni.2013.10.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Galli SJ. The mast cell-IgE paradox: from homeostasis to anaphylaxis. Am J Pathol. 2016;186:212–224. doi: 10.1016/j.ajpath.2015.07.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Palm NW, Rosenstein RK, Yu S, Schenten DD, Florsheim E, Medzhitov R. Bee venom phospholipase A2 induces a primary type 2 response that is dependent on the receptor ST2 and confers protective immunity. Immunity. 2013;39:976–985. doi: 10.1016/j.immuni.2013.10.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Bieber T, de la Salle H, Wollenberg A, Hakimi J, Chizzonite R, Ring J, et al. Human epidermal Langerhans cells express the high affinity receptor for immunoglobulin E (Fc epsilon RI) J Exp Med. 1992;175:1285–1290. doi: 10.1084/jem.175.5.1285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Maurer D, Ebner C, Reininger B, Fiebiger E, Kraft D, Kinet JP, et al. The high affinity IgE receptor (Fc epsilon RI) mediates IgE-dependent allergen presentation. J Immunol. 1995;154:6285–6290. [PubMed] [Google Scholar]
  • 138.Novak N, Gros E, Bieber T, Allam JP. Human skin and oral mucosal dendritic cells as ‘good guys’ and ‘bad guys’ in allergic immune responses. Clin Exp Immunol. 2010;161:28–33. doi: 10.1111/j.1365-2249.2010.04162.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Von Bubnoff D, Matz H, Cazenave JP, Hanau D, Bieber T, De La Salle H. Kinetics of gene induction after FcepsilonRI ligation of atopic monocytes identified by suppression subtractive hybridization. J Immunol. 2002;169:6170–6177. doi: 10.4049/jimmunol.169.11.6170. [DOI] [PubMed] [Google Scholar]
  • 140.von Bubnoff D, Novak N, Kraft S, Bieber T. The central role of FcepsilonRI in allergy. Clin Exp Dermatol. 2003;28:184–187. doi: 10.1046/j.1365-2230.2003.01209.x. [DOI] [PubMed] [Google Scholar]
  • 141.Schroeder JT, Chichester KL, Bieneman AP. Toll-like receptor 9 suppression in plasmacytoid dendritic cells after IgE-dependent activation is mediated by autocrine TNF-alpha. J Allergy Clin Immunol. 2008;121:486–491. doi: 10.1016/j.jaci.2007.09.049. [DOI] [PubMed] [Google Scholar]
  • 142.Khan SH, Grayson MH. Cross-linking IgE augments human conventional dendritic cell production of CC chemokine ligand 28. J Allergy Clin Immunol. 2010;125:265–267. doi: 10.1016/j.jaci.2009.09.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.von Bubnoff D, Matz H, Frahnert C, Rao ML, Hanau D, de la Salle H, et al. FcepsilonRI induces the tryptophan degradation pathway involved in regulating T cell responses. J Immunol. 2002;169:1810–1816. doi: 10.4049/jimmunol.169.4.1810. [DOI] [PubMed] [Google Scholar]
  • 144.Le T, Tversky J, Chichester KL, Bieneman AP, Huang SK, Wood RA, et al. Interferons modulate Fc epsilon RI-dependent production of autoregulatory IL-10 by circulating human monocytoid dendritic cells. J Allergy Clin Immunol. 2009;123:217–223. doi: 10.1016/j.jaci.2008.09.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Rowe RK, Gill MA. Asthma: the interplay between viral infections and allergic diseases. Immunol Allergy Clin North Am. 2015;35:115–127. doi: 10.1016/j.iac.2014.09.012. [DOI] [PubMed] [Google Scholar]
  • 146.Durrani SR, Montville DJ, Pratt AS, Sahu S, DeVries MK, Rajamanickam V, et al. Innate immune responses to rhinovirus are reduced by the high-affinity IgE receptor in allergic asthmatic children. J Allergy Clin Immunol. 2012;130:489–495. doi: 10.1016/j.jaci.2012.05.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Gill MA, Bajwa G, George TA, Dong CC, Dougherty II, Jiang N, et al. Counter-regulation between the FcepsilonRI pathway and antiviral responses in human plasmacytoid dendritic cells. J Immunol. 2010;184:5999–6006. doi: 10.4049/jimmunol.0901194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Sykes A, Edwards MR, Macintyre J, del Rosario A, Bakhsoliani E, Trujillo-Torralbo MB, et al. Rhinovirus 16-induced IFN-α and IFN-β are deficient in bronchoalveolar lavage cells in asthmatic patients. J Allergy Clin Immunol. 2012;129 doi: 10.1016/j.jaci.2012.03.044. 1506-14.e6. [DOI] [PubMed] [Google Scholar]
  • 149.Busse WW, Morgan WJ, Gergen PJ, Mitchell HE, Gern JE, Liu AH, et al. Randomized trial of omalizumab (anti-IgE) for asthma in inner-city children. N Engl J Med. 2011;364:1005–1015. doi: 10.1056/NEJMoa1009705. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Sallmann E, Reininger B, Brandt S, Duschek N, Hoflehner E, Garner-Spitzer E, et al. High-affinity IgE receptors on dendritic cells exacerbate Th2-dependent inflammation. J Immunol. 2011;187:164–171. doi: 10.4049/jimmunol.1003392. [DOI] [PubMed] [Google Scholar]
  • 151.Platzer B, Baker K, Vera MP, Singer K, Panduro M, Lexmond WS, et al. Dendritic cell-bound IgE functions to restrain allergic inflammation at mucosal sites. Mucosal Immunol. 2015;8:516–532. doi: 10.1038/mi.2014.85. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Greer AM, Wu N, Putnam AL, Woodruff PG, Wolters P, Kinet JP, et al. Serum IgE clearance is facilitated by human FcepsilonRI internalization. J Clin Invest. 2014;124:1187–1198. doi: 10.1172/JCI68964. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Delespesse G, Sarfati M, Hofstetter H, Suter U, Nakajima T, Peleman R, et al. Structure, function and clinical relevance of the low affinity receptor for IgE. Immunol Invest. 1988;17:363–387. doi: 10.3109/08820138809049845. [DOI] [PubMed] [Google Scholar]
  • 154.Kehry MR, Hudak SA. Characterization of B-cell populations bearing Fce receptor II. Cell Immunol. 1989;118:504–515. doi: 10.1016/0008-8749(89)90397-3. [DOI] [PubMed] [Google Scholar]
  • 155.Pirron U, Schlunck T, Prinz JC, Rieber EP. IgE-dependent antigen focusing by human B lymphocytes is mediated by the low-affinity receptor for IgE. Eur J Immunol. 1990;20:1547–1551. doi: 10.1002/eji.1830200721. [DOI] [PubMed] [Google Scholar]
  • 156.Fujiwara H, Kikutani H, Suematsu S, Naka T, Yoshida K, Yoshida K, et al. The absence of IgE antibody-mediated augmentation of immune responses in CD23-deficient mice. Proc Natl Acad Sci USA. 1994;91:6835–6839. doi: 10.1073/pnas.91.15.6835. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Galli SJ, Tsai M. IgE and mast cells in allergic disease. Nat Med. 2012;18:693–704. doi: 10.1038/nm.2755. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Williams CM, Galli SJ. Mast cells can amplify airway reactivity and features of chronic inflammation in an asthma model in mice. J Exp Med. 2000;192:455–462. doi: 10.1084/jem.192.3.455. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Nakae S, Ho LH, Yu M, Monteforte R, Iikura M, Suto H, et al. Mast cell-derived TNF contributes to airway hyperreactivity, inflammation, and TH2 cytokine production in an asthma model in mice. J Allergy Clin Immunol. 2007;120:48–55. doi: 10.1016/j.jaci.2007.02.046. [DOI] [PubMed] [Google Scholar]
  • 160.Takeda K, Hamelmann E, Joetham A, Shultz LD, Larsen GL, Irvin CG, et al. Development of eosinophilic airway inflammation and airway hyperresponsiveness in mast cell-deficient mice. J Exp Med. 1997;186:449–454. doi: 10.1084/jem.186.3.449. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Oettgen HC, Burton OT. IgE and mast cells: the endogenous adjuvant. Adv Immunol. 2015;127:203–256. doi: 10.1016/bs.ai.2015.03.001. [DOI] [PubMed] [Google Scholar]
  • 162.Finkelman F, Morris S, Orekhova T, Sehy D. The in vivo cytokine capture assay for measurement of cytokine production in the mouse. Curr Protoc Immunol. 2003 doi: 10.1002/0471142735.im0628s54. Chapter 6:Unit 6.28. [DOI] [PubMed] [Google Scholar]
  • 163.Kambayashi T, Allenspach EJ, Chang JT, Zou T, Shoag JE, Reiner SL, et al. Inducible MHC class II expression by mast cells supports effector and regulatory T cell activation. J Immunol. 2009;182:4686–4695. doi: 10.4049/jimmunol.0803180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Poncet P, Arock M, David B. MHC class II-dependent activation of CD4+ T cell hybridomas by human mast cells through superantigen presentation. J Leukoc Biol. 1999;66:105–112. doi: 10.1002/jlb.66.1.105. [DOI] [PubMed] [Google Scholar]
  • 165.Kitawaki T, Kadowaki N, Sugimoto N, Kambe N, Hori T, Miyachi Y, et al. IgE-activated mast cells in combination with pro-inflammatory factors induce Th2-promoting dendritic cells. Int Immunol. 2006;18:1789–1799. doi: 10.1093/intimm/dxl113. [DOI] [PubMed] [Google Scholar]
  • 166.Mazzoni A, Siraganian RP, Leifer CA, Segal DM. Dendritic cell modulation by mast cells controls the Th1/Th2 balance in responding T cells. J Immunol. 2006;177:3577–3581. doi: 10.4049/jimmunol.177.6.3577. [DOI] [PubMed] [Google Scholar]
  • 167.Dardalhon V, Awasthi A, Kwon H, Galileos G, Gao W, Sobel RA, et al. IL-4 inhibits TGF-beta-induced Foxp3+ T cells and, together with TGF-beta, generates IL-9+ IL-10+ Foxp3(-) effector T cells. Nat Immunol. 2008;9:1347–1355. doi: 10.1038/ni.1677. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Veldhoen M, Hocking RJ, Atkins CJ, Locksley RM, Stockinger B. TGFbeta in the context of an inflammatory cytokine milieu supports de novo differentiation of IL-17-producing T cells. Immunity. 2006;24:179–189. doi: 10.1016/j.immuni.2006.01.001. [DOI] [PubMed] [Google Scholar]
  • 169.Noval Rivas M, Burton OT, Wise P, Charbonnier L-M, Georgiev P, Oettgen HC, et al. Regulatory T cell reprogramming toward a Th2 cell-like lineage impairs oral tolerance and promotes food allergy. Immunity. 2015;42:512–523. doi: 10.1016/j.immuni.2015.02.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Piconese S, Gri G, Tripodo C, Musio S, Gorzanelli A, Frossi B, et al. Mast cells counteract regulatory T-cell suppression through interleukin-6 and OX40/OX40L axis toward Th17-cell differentiation. Blood. 2009;114:2639–2648. doi: 10.1182/blood-2009-05-220004. [DOI] [PubMed] [Google Scholar]
  • 171.Frossi B, D’Inca F, Crivellato E, Sibilano R, Gri G, Mongillo M, et al. Single-cell dynamics of mast cell-CD4+ CD25+ regulatory T cell interactions. Eur J Immunol. 2011;41:1872–1882. doi: 10.1002/eji.201041300. [DOI] [PubMed] [Google Scholar]
  • 172.Seder RA, Paul WE, Dvorak AM, Sharkis SJ, Kagey-Sobotka A, Niv Y, et al. Mouse splenic and bone marrow cell populations that express high-affinity Fc epsilon receptors and produce interleukin 4 are highly enriched in basophils. Proc Natl Acad Sci USA. 1991;88:2835–2839. doi: 10.1073/pnas.88.7.2835. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Min B, Prout M, Hu-Li J, Zhu J, Jankovic D, Morgan ES, et al. Basophils produce IL-4 and accumulate in tissues after infection with a Th2-inducing parasite. J Exp Med. 2004;200:507–517. doi: 10.1084/jem.20040590. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Voehringer D, Shinkai K, Locksley RM. Type 2 immunity reflects orchestrated recruitment of cells committed to IL-4 production. Immunity. 2004;20:267–277. doi: 10.1016/s1074-7613(04)00026-3. [DOI] [PubMed] [Google Scholar]
  • 175.Mukai K, Matsuoka K, Taya C, Suzuki H, Yokozeki H, Nishioka K, et al. Basophils play a critical role in the development of IgE-mediated chronic allergic inflammation independently of T cells and mast cells. Immunity. 2005;23:191–202. doi: 10.1016/j.immuni.2005.06.011. [DOI] [PubMed] [Google Scholar]
  • 176.Cheng LE, Sullivan BM, Retana LE, Allen CD, Liang HE, Locksley RM. IgE-activated basophils regulate eosinophil tissue entry by modulating endothelial function. J Exp Med. 2015;212:513–524. doi: 10.1084/jem.20141671. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Poddighe D, Mathias CB, Freyschmidt EJ, Kombe D, Caplan B, Marseglia GL, et al. Basophils are rapidly mobilized after initial aeroallergen encounter in naïve mice and provide a priming source of IL-4 in adaptive immune responses. J Biol Regul Homeost Agents. 2014;28:91–103. [PubMed] [Google Scholar]
  • 178.Lambrecht BN, Hammad H. Allergens and the airway epithelium response: gateway to allergic sensitization. J Allergy Clin Immunol. 2014;134:499–507. doi: 10.1016/j.jaci.2014.06.036. [DOI] [PubMed] [Google Scholar]
  • 179.Zhong W, Su W, Zhang Y, Liu Q, Wu J, Di C, et al. Basophils as a primary inducer of the T helper type 2 immunity in ovalbumin-induced allergic airway inflammation. Immunology. 2014;142:202–215. doi: 10.1111/imm.12240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Noti M, Kim BS, Siracusa MC, Rak GD, Kubo M, Moghaddam AE, et al. Exposure to food allergens through inflamed skin promotes intestinal food allergy through the thymic stromal lymphopoietin-basophil axis. J Allergy Clin Immunol. 2014;133:1390–1399. doi: 10.1016/j.jaci.2014.01.021. e1-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Rodewald HR, Feyerabend TB. Widespread immunological functions of mast cells: fact or fiction? Immunity. 2012;37:13–24. doi: 10.1016/j.immuni.2012.07.007. [DOI] [PubMed] [Google Scholar]
  • 182.Lilla JN, Chen CC, Mukai K, BenBarak MJ, Franco CB, Kalesnikoff J, et al. Reduced mast cell and basophil numbers and function in Cpa3-Cre; Mcl-1fl/fl mice. Blood. 2011;118:6930–6938. doi: 10.1182/blood-2011-03-343962. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Reber LL, Marichal T, Galli SJ. New models for analyzing mast cell functions in vivo. Trends Immunol. 2012;33:613–625. doi: 10.1016/j.it.2012.09.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Denzel A, Maus UA, Rodriguez Gomez M, Moll C, Niedermeier M, Winter C, et al. Basophils enhance immunological memory responses. Nat Immunol. 2008;9:733–742. doi: 10.1038/ni.1621. [DOI] [PubMed] [Google Scholar]
  • 185.Hammad H, Plantinga M, Deswarte K, Pouliot P, Willart MA, Kool M, et al. Inflammatory dendritic cells—not basophils—are necessary and sufficient for induction of Th2 immunity to inhaled house dust mite allergen. J Exp Med. 2010;207:2097–2111. doi: 10.1084/jem.20101563. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Simpson A, Tan VY, Winn J, Svensen M, Bishop CM, Heckerman DE, et al. Beyond atopy: multiple patterns of sensitization in relation to asthma in a birth cohort study. Am J Respir Crit Care Med. 2010;181:1200–1206. doi: 10.1164/rccm.200907-1101OC. [DOI] [PubMed] [Google Scholar]
  • 187.Stoltz DJ, Jackson DJ, Evans MD, Gangnon RE, Tisler CJ, Gern JE, et al. Specific patterns of allergic sensitization in early childhood and asthma & rhinitis risk. Clin Exp Allergy. 2013;43:233–241. doi: 10.1111/cea.12050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Simpson A, Soderstrom L, Ahlstedt S, Murray CS, Woodcock A, Custovic A. IgE antibody quantification and the probability of wheeze in preschool children. J Allergy Clin Immunol. 2005;116:744–749. doi: 10.1016/j.jaci.2005.06.032. [DOI] [PubMed] [Google Scholar]
  • 189.Illi S, von Mutius E, Lau S, Niggemann B, Gruber C, Wahn U. Perennial allergen sensitisation early in life and chronic asthma in children: a birth cohort study. Lancet. 2006;368:763–770. doi: 10.1016/S0140-6736(06)69286-6. [DOI] [PubMed] [Google Scholar]

RESOURCES