ArticlePDF AvailableLiterature Review

AllergoOncology: The role of IgE-mediated allergy in cancer

Authors:

Abstract and Figures

Epidemiological studies have suggested inverse associations between allergic diseases and malignancies. As a proof of concept for the capability of immunoglobulin E (IgE) to destruct tumor cells, several experimental strategies have evolved to specifically target this antibody class towards relevant tumor antigens. It could be demonstrated that IgE antibodies specific to overexpressed tumor antigens have been superior to any other immunoglobulin class with respect to antibody-dependent cellular cytotoxicity (ADCC) and phagocytosis (ADCP) reactions. In an alternative approach, IgE nonspecifically attached to tumor cells proved to be a powerful adjuvant establishing tumor-specific immune memory. Active Th2 immunity could also be achieved by applying an oral immunization regimen using mimotopes, i.e. epitope mimics of tumor antigens. The induced IgE antibodies could be cross-linked by live tumor cells leading to tumoricidic mediator release. Thus, IgE antibodies may not only act in natural tumor surveillance, but could possibly also be exploited for tumor control in active and passive immunotherapy settings. Thereby, eosinophils, mast cells and macrophages can be armed with the cytophilic IgE and become potent anti-tumor effectors, able to trace viable tumor cells in the tissues. It is strongly suggested that the evolving new field AllergoOncology will give new insights into the role of IgE-mediated allergy in malignancies, possibly opening new avenues for tumor therapy.
Content may be subject to copyright.
Review article
AllergoOncology: the role of IgE-mediated allergy in cancer
Before their identification as immunoglobulin E (IgE)
antibodies (1–4), research on ÔreaginsÕwas conducted
during five decades (5–7). Since this era, allergology and
parasitology have been traditionally considered as a
whole and this notion helped in evolving a lively picture
of the IgE-mediated immune response. It is, however,
more or less neglected today that the association of
allergic diseases and tumor occurrence has been discussed
already very early (8, 9). In the 1950s, precise experiments
investigated Ôallergic responsesÕtowards tumor trans-
plants (10). Consequently, the observed immunological
phenomena were even termed tumor allergy (11). The
discussion went on asking about the biological relevance
of tumor allergy for tumor progression (12, 13), until a
negative association between allergy and cancer was
announced for the first time (14). Later, the IgE levels and
atopy reactions in the skin of cancer patients were
examined (15, 16), finally rendering knowledge that the
prevalence of atopy was decreased in cancer patients (17).
Passive anaphylaxis or weekly injections of histamine and
serotonin inhibited tumor growth in a transplant mouse
model, pointing towards a possible role of anaphylactic
Epidemiological studies have suggested inverse associations between allergic
diseases and malignancies. As a proof of concept for the capability of immu-
noglobulin E (IgE) to destruct tumor cells, several experimental strategies have
evolved to specifically target this antibody class towards relevant tumor antigens.
It could be demonstrated that IgE antibodies specific to overexpressed tumor
antigens have been superior to any other immunoglobulin class with respect to
antibody-dependent cellular cytotoxicity (ADCC) and phagocytosis (ADCP)
reactions. In an alternative approach, IgE nonspecifically attached to tumor cells
proved to be a powerful adjuvant establishing tumor-specific immune memory.
Active Th2 immunity could also be achieved by applying an oral immunization
regimen using mimotopes, i.e. epitope mimics of tumor antigens. The induced
IgE antibodies could be cross-linked by live tumor cells leading to tumoricidic
mediator release. Thus, IgE antibodies may not only act in natural tumor sur-
veillance, but could possibly also be exploited for tumor control in active and
passive immunotherapy settings. Thereby, eosinophils, mast cells and macro-
phages can be armed with the cytophilic IgE and become potent anti-tumor
effectors, able to trace viable tumor cells in the tissues. It is strongly suggested
that the evolving new field AllergoOncology will give new insights into the role
of IgE-mediated allergy in malignancies, possibly opening new avenues for
tumor therapy.
E. Jensen-Jarolim
1
, G. Achatz
2
,
M. C. Turner
3
, S. Karagiannis
4
,
F. Legrand
5
, M. Capron
5
,
M. L. Penichet
6
, J. A. Rodrguez
7
,
A. G. Siccardi
8
, L. Vangelista
8
,
A. B. Riemer
1
, H. Gould
4
1
Department of Pathophysiology, Center of
Physiology, Pathophysiology and Immunology,
Medical University Vienna, Austria;
2
Division of
Genetics and General Biology, Hellbrunnerstraße,
Salzburg, Austria;
3
McLaughlin Centre for
Population Health Risk Assessment, Institute of
Population Health, University of Ottawa, Canada;
4
KingÕs College London, UK;
5
Unit Inserm 547,
Universit Lille 2, Institut Pasteur de Lille, France;
6
Division of Surgical Oncology, Department of
Surgery; Department of Microbiology, Immunology,
and Molecular Genetics; and the Jonsson
Comprehensive Cancer Center, David Geffen School
of Medicine, University of California, LA (UCLA),
USA;
7
Division of Surgical Oncology, Department of
Surgery; David Geffen School of Medicine,
University of California, LA (UCLA), USA;
8
Dibit,
San Raffaele Scientific Institute and Department of
Biology and Genetics, University of Milan, Italy
Key words: AllergoOncology; cancer; eosinophils; IgE;
tumoricidic.
Erika Jensen-Jarolim
IPP– Department of Pathophysiology
Center of Physiology, Pathophysiology and
Immunology
Medical University Vienna
Waehringer G. 18
1090 Vienna,
Austria
Accepted for publication 25 March 2008
Allergy 2008: 63: 1255–1266 2008 The Authors
Journal compilation 2008 Blackwell Munksgaard
DOI: 10.1111/j.1398-9995.2008.01768.x
1255
reactions in tumor immunity. Interestingly, an immuno-
histochemical study on the distribution of immunoglob-
ulin classes in head and neck cancer revealed IgE
antibodies to be the most abundant class, fixed in the
cancer tissues on dispersed macrophage-like cells (18).
This work suggests that IgE may have a natural surveil-
lance function in malignancies. Advancements in immu-
nology and molecular biology enable us today to go one
step further and exploit this knowledge for developing
IgE-based targeted cancer therapies.
This review aims at giving a comprehensive overview
on recent epidemiological and experimental evidence for
the occurrence of Th2-type immune mechanisms in tumor
disease. Moreover, also targeted therapies with IgE
antibodies and vaccination strategies will be discussed
as a novel perspective to combat cancer.
IgE antibodies: prime target unknown
IgE is an evolutionary conserved member of the Ig family
with the highest determined affinity to receptors (19–22)
and antigens among all antibody classes (23). The titer of
IgE is very low (nano- to micrograms per milliliter range) in
plasma of normal healthy individuals and of normal
laboratory mouse strains, but IgE is most prominent in
epithelia and mucosa where it is bound to specific receptors
on very potent effector cells like eosinophil or basophil
granulocytes and mast cells. This suggests that IgE plays a
role in local (rather than systemic) immune defence
mechanisms. In these days, IgE is best known for its
strong, unwanted effector functions, in the form of allergic
reactions (19). However, the prime target for IgE is still
unknown. From an evolutionary point of view, IgE is
conserved and can be found in all mammalia, including
monotremata (24). It therefore originated at least 160
million years ago, possibly even more than 300 million
years ago (25), from a gene duplication of IgY, in which the
anaphylactic and opsonic activities of IgY were separated,
giving rise to IgE and IgG, respectively (26). Apparently, in
an evolutionary sense, anaphylactic defence mechanisms
are needed. The division of anaphylactic and opsonic
activities in separate genes allowed principally a tighter and
more specific control of both immune mechanisms.
In the recent past, five B-cell specific control mecha-
nisms have been described that indicate a tight control of
the IgE response, in agreement with the arguments shown,
and that are different from the opsonic type of response.
1. IgE has the shortest free serum half-life (t
½
)ofall
immunoglobulins averaging 12 h in mice (27) and
1–5 days in humans (28, 29), limiting the danger of a
systemic anaphylactic reaction. After production
serum IgE is rapidly bound by the high-affinity
FceRI on the surface of effector cells like mast cells
and basophils where it acquires a long half-life time
(weeks to months).
2. Several studies with mice deficient of, or overex-
pressing the low-affinity IgE receptor CD23, clearly
demonstrated CD23Õs role as a major negative
feedback regulator of IgE production. Yu et al. (30)
showed that disruption of the CD23 gene led to
increased specific IgE levels after immunization with
2,4-dinitrophenyl-ovalbumin (DNP-OVA), while
IgG1 levels were twice as high, specific IgE levels
were 6–12 times higher in these CD23
–/–
mice.
3. Two mIgE knock-out mice (31) underlined the key
function of the IgE receptor in the regulation of IgE
expression in vivo and by analyzing the phenotypes
of the mice strains it could clearly be demonstrated
that the antigen receptor is the only device for an
effective antigen presentation (32, 33). In the first
strain, the intracellular domain of IgE was removed
except for three amino acids (Lys, Val, Lys;
KVKDtail line). The cytoplasmic domain of IgE in
these mice is the same as that of mIgM and mIgD. In
the second line, both the intracellular and trans-
membrane domains of IgE (DM1M2 line) are lack-
ing (Fig. 1). In DM1M2 mice serum IgE is reduced
to less than 10% of normal mice, while KVKDtail
mice show a reduction of 50%, reflecting a serious
impairment of the IgE-mediated immune response.
Upon stimulation of isolated spleen cells of wild
type, DM1M2 and KVKDtail mice with LPS and IL4
in vitro, concentrations of IgE and IgG1 in the
culture supernatants were comparable in wild-type
and mutant mice. These results imply that the
reduced IgE titers found in both mutant lines are
solely a reflection of the loss of biological activities
associated with the transmembrane and cytoplasmic
domains of IgE.
4. The process of alternative polyadenylation restricts
surface IgE expression and thus influences further
serum IgE production. mRNA for the membrane
form of both the murine and human epsilon (e)
heavy chain is poorly expressed, compared with the
mRNA for the secreted form in activated, mIgE-
bearing B cells (32, 34).
5. Finally, also the establishment of humoral memory is
limited for IgE responses in vivo. IgE plasmablasts
have an intrinsic, lower chance to contribute to the
long-lived plasma cell pool and thus to humoral
immunologic memory than IgG1 plasmablasts.
Apparently, an IgE immune response is in all stages
of the response negatively regulated. The IgE anti-
bodies may have strong effector functions, but the
IgE response is slow and limited in developing
memory responses.
Therefore, the IgE antigen receptor itself plays a major
role in the decision of quantity and quality of serum IgE
antibodies and besides, is probably the most powerful
surface receptor, influencing developmental processes of
the cell. It was suggested that the signaling cascade
Jensen-Jarolim et al.
2008 The Authors
1256 Journal compilation 2008 Blackwell Munksgaard Allergy 2008: 63: 1255–1266
underlies a permanent stochastic fluctuation, which
induces no cellular response, but is important for main-
tenance of cell viability [reviewed in (35)]. Taken together,
all these observations point towards the existence of
mechanisms to restrain potentially dangerous, but appar-
ently necessary, serum IgE titers.
Epidemiological association of IgE and malignancies
Hints from clinical observations triggered numerous
epidemiological studies to examine the potential associ-
ation between a history of IgE-mediated allergy and
cancer. These studies, and important methodological
considerations, were summarized in recent review articles
(36, 37). Although the results are not entirely clear, there
is some limited evidence to suggest a possible inverse
relation.
In perhaps the largest study of its kind, over 1.1 mil-
lion US adults, for whom self-reported physician-diag-
nosed asthma or hay fever status at baseline was known,
were followed-up for a period of 18 years (38). Results
suggested a significant inverse association between a
history of both asthma and hay fever, possibly the most
relevant indicator of allergic status examined here, and
all cancer mortality [relative risk (RR) = 0.88, 95%
confidence interval (CI) 0.83–0.93]. The risk of mortality
from several site-specific cancers was also reduced, with
that for colorectal cancer significant (RR = 0.76, 95%
CI 0.64–0.91). In a separate analysis of never smokers,
results were similar, although they attenuated slightly
and were no longer significant. Results from other
smaller prospective studies were mixed (39–42), includ-
ing some who used skin-prick testing to define allergic
status (43–45). Conversely, one study recently followed
70 136 patients for whom data on total serum IgE
levels were known and 57 815 patients for whom data
on allergen-specific IgE levels were known (46). No
association with cancer incidence was reported for either
measure.
Case-control studies using self-reported allergy history
information have also suggested several potential inverse
associations between allergic status and site-specific
cancers including pancreatic (47) and glioma (48) with
the magnitude of the effect ranging from approximately
30–40% reductions in risk. Inverse associations were also
reported in studies of childhood leukaemia (49–51) or
myeloma of adults (52). Case-control studies, measuring
allergen-specific IgE in cases following cancer diagnosis,
have reported mixed findings. Wiemels et al. (53)
reported inverse associations between glioma and total
IgE levels [odds ratio (OR) elevated total IgE = 0.37,
95% CI 0.22–0.64) and allergen-specific IgE levels,
particularly food IgE (OR = 0.12, 95% CI 0.04–0.41).
Melbye et al. (54), in a large multi-center study, reported
a significant inverse association between allergen-specific
IgE and non-HodgkinÕs lymphoma (NHL; OR = 0.68,
95% CI 0.58–0.80). However, upon further analysis,
NHL dissemination in cases was found to be inversely
associated with specific IgE, and in a second, prospective,
study, an inverse association was found only immediately
prior to NHL diagnosis, leading the authors to conclude
that inverse associations reported in previous studies were
likely due to the suppression of the allergic response in
NHL cases. Positive associations were reported between
allergen-specific IgE and both prostate (55, 56) and breast
cancer (57).
The following discussion will focus on the elucidation
of the immunological mechanistic principles potentially
underlying the epidemiological observations.
IgE and its receptors: interaction with intratumoral effector cells
If IgE antibodies (Abs) were directed against tumor-
associated antigens (TAA) they could mediate the cell-to-cell
association between tumor and effector cells, possibly
resulting in antibody-dependent cellular cytotoxicity
(ADCC) and antibody-dependent cellular phagocytosis
(ADCP). For these reactions, the high- and low-affinity
receptors, FceRI and CD23 (19, 20), on effector cells being
on site in the tumor tissue are crucial. The affinity of IgE for
FceRI is by two to five orders of magnitude higher than that
of IgG for their receptors, making IgE the only antibody
signal transduction
AG processing
Figure 1. In the immunoglobulin E (IgE)
KVKDtail
mouse strain,
the intracellular domain of IgE was removed except for the three
conserved amino acids (Lys, Val, Lys). Thus, the cytoplasmic
domain of IgE in these mice is identical to the cytoplasmic tail
expressed by mIgM and mIgD. In the IgE
DM1M2
line, both the
intracellular and transmembrane domains of IgE are lacking.
The cytoplasmic tails of the membrane immunoglobulins have
the capacity to bind proteins, committing a signal transduction
pathway which is independent of the well-known Ig-a/Ig-b
pathways. The green domains represent Ig heavy chains, red
domains indicate Ig light chains, white domain indicates trans-
membrane domain, clewed domain indicates intracellular tail
and pink and brown domains indicate Ig coat proteins Ig-a/
Ig-b.
AllergoOncology: the role of allergy in cancer
2008 The Authors
Journal compilation 2008 Blackwell Munksgaard Allergy 2008: 63: 1255–1266 1257
class being strongly retained by effector cells in the absence
of antigen. Thus, IgE engaged to FceRI in tissues could be
more effective in anti-tumor responses than IgG and its
receptors. FceRI-expressing monocytes exerted primarily
ADCC towards tumor cells. Upon upregulation of FceRI
by preincubation with IgE specific for the ovarian tumor
antigen folate receptor, an increase in ADCC was observed
(58–60).
Recent strategies, aiming at enhancing the anti-tumor
responses of T cells, exploited the high affinty of the IgE–
FceRI complex by transfecting T lymphocytes with a
chimeric molecule comprising the extracellular domain of
FceRI with the cytoplasmic domains of CD28 and T cell
receptor zeta chains (61). FceRI-expressing human T cells
engaged by tumor-specific IgE monoclonal anti-
body (mAb) secreted cytokines, proliferated and medi-
ated cytotoxic functions also in vivo following antigen
ligation.
CD23, the low-affinity IgE receptor, exists in two
forms, CD23a and CD23b, differing at the cytoplasmic
N-terminus which contains different signaling motifs that
determine their functions (62). The expression of CD23b
by interleukin (IL)-4, e.g. derived from natural killer cells
in breast cancer tissues, is induced on various cells,
notably mast cells, basophils and monocytes (63, 64). By
the way, IL-4, like IL-13, is also an important switch
factor for IgE production, the latter being also directly
derived from cancer cells (65). Engagement of CD23b by
IgE–antigen complexes promotes monocyte/macrophage
activation, induces nitric oxide synthase (iNOS) and
generates pro-inflammatory cytokines (66, 67). CD23b
mediates IgE ADCP (63) and may thereby, besides its
control function for IgE production (30), mediate tumor
cell death. This function has been confirmed as a CD23–
IgE complex-driven mechanism of engaging monocytes in
ADCP of ovarian tumor cells upregulated by IL-4 (59).
IgE antibodies can thus engage both cell surface IgE
receptors, FceRI and CD23, and activate several lines of
effector cells against tumor cells in vitro and in vivo.
Indeed, solid tumors are associated with inflammatory
responses involving the infiltration by not only B and
T lymphocytes, neutrophils and natural killer cells, but
also mast cells, macrophages and eosinophils expressing
the IgE receptors (68).
It has been shown that mast cells infiltrate the invasive
fronts of tumor lesions where their degranulation pro-
motes the remodeling of tissue architecture, angiogenesis,
tumor cell growth and metastasis (69). The location of
mast cells away from the core of solid tumors, loss of local
tissue architecture and production of angiogenic factors
[vascular endothelial growth factor (VEGF), basic fibro-
blast growth factor (bFGF), IL-8, tumor necrosis factor
(TNF)a, matrix metalloproteinase (MMP)-9], may play a
role in their inability to target tumor cells. Mast cells have
even been described to be a negative prognostic marker in
Merkel cell carcinoma (70). Moreover, there is genetic
evidence for a role of mast cells for tumor expansion in a
pancreatic islet cell tumor model (71). On the other hand,
the intensity of mast cell activation/response in the
absence of antigenic stimuli may also contribute to tumor
cell death (69, 72), e.g. by releasing preformed proapop-
totic TNFaupon triggering (73, 74), as well as histamine,
which acts to promote or inhibit tumor growth dependent
on the type of histamine receptor expressed (75).
Tumor-associated macrophages (TAM) are found in
virtually all types of tumors and can comprise more than
50% of the total tumor mass (76). Blood monocytes are
recruited to the tumor sites by chemokines and cytokines
released by tumor cells and neighboring endothelial cells.
They can be stimulated to either kill tumor cells and
release angiostatic compounds, or, like mast cells, pro-
mote tumor growth and metastasis by producing angio-
genic factors and MMP (77). Immunohistochemical
studies in breast cancer confirm the presence of CD23
on the surface of TAM and the expression of iNOS (78,
79). Cytotoxic mediators such as NO are produced by
monocytes/macrophages in cancer tissues at levels below
those required for tumor cell cytotoxicity, so that the
balance of NO production has been suggested to be
tipped in favor of tumor growth (80, 81). In contrast,
recent studies report that NO is necessary to mediate the
antiangiogenic effect of TNF via its receptor TNFR2 in
the tumor (74). It has been suggested that these potential
cytotoxic effects could be enhanced to activate TAM
against tumor cells. Activation signals by tumor antigen-
specific IgE may ÔawakeÕTAM to this end.
Like macrophages, also other antigen-presenting cells
such as B lymphocytes, Langerhans cells and dendritic
cells present in tumor infiltrates express CD23 and/or
FceRI and can be activated by locally secreted lympho-
kines as well as tumor antigen-specific IgE. All these
events can initiate IgE antibody-dependent antigen pre-
sentation to autologous T cells and induce active immu-
nity (82–86). Therefore, although the mechanisms by
which IgE antibodies exert their anti-tumor effects against
cancer cells are only starting to emerge, it seems clear that
all the ingredients necessary for an IgE-mediated response
are in place in human malignant disease.
Eosinophils: exquisite effectors in anti-tumor immunity
Eosinophils are today considered as multifunctional
leukocytes, involved in inflammatory processes, tissue
remodeling as well as in modulation of innate and
adaptive immunity (87). Typically, differentiation and
function of eosinophils is regulated by the CD4 Th2
cytokines IL-3, IL-5 and granulocyte macrophage (GM)-
colony-stimulating factor (CSF), and may be regarded as
typical cells in type 2 immunity (88). Eosinophils express
membrane receptors involved in adaptive immune re-
sponses like Fc receptors for IgG, IgA or IgE, class II
major histocompatability complex (MHC) or co-stimu-
latory molecules including CD86 and CD28 that allow
interactions with lymphocytes and antigen-presenting
Jensen-Jarolim et al.
2008 The Authors
1258 Journal compilation 2008 Blackwell Munksgaard Allergy 2008: 63: 1255–1266
cells (89). Triggering eosinophils by engagement of
receptors for cytokines, immunoglobulins or complement
lead to the secretion of numerous cytokines [IL-2, -4, -5, -
10, -12, -13, -16, -18, (transforming growth factor (TGF)-
a/b], chemokines (RANTES, eotaxin) and lipid mediators
(platelet-activating factor, leukotriene C4). Moreover,
release of Th2 or Th1 immunoregulatory cytokines and
cationic proteins can be specific and dependent upon the
nature of stimulus (90, 91). In their specific granules,
eosinophils store great quantities of cytokines and
cationic proteins, known to be highly cytotoxic mediators
[reviewed in (92)]. They can also generate reactive oxygen
species (ROS). This high cytotoxic potential can lead to
deleterious effects in inflammatory and allergic processes,
and eosinophil numbers have also been documented to be
elevated in peripheral blood and/or infiltrate the tissues in
some malignant disorders (93, 94). This eosinophilia is
named TATE for Ôtumor-associated tissue eosinophiliaÕ.
High infiltrating eosinophil counts were associated with a
significantly better prognosis in esophageal squamous cell
carcinoma, gastric cancer, head and neck cancer and
colorectal carcinoma (95–99). Activated, they may more-
over prevent the metastasis potential of prostate cancer
(100, 101). Although TATE may have favorable prog-
nostic value, little is known however on the exact role
played by eosinophils in anti-tumor responses.
Recent in vitro (102) and in vivo (103, 104) studies have
suggested a potential tumoricidal activity of eosinophils.
For the evaluation of the tumoricidal activity of human
eosinophils two tumor cell lines, a T lymphoma and a
colorectal adenocarcinoma, Jurkat and Colo-205 respec-
tively, were used. Several in vitro arguments allow
postulating that human eosinophils purified from the
peripheral blood of different eosinophilic donors are able
to use effectively their cytotoxic potential towards these
human tumor cell lines by inducing their apoptosis in
different ways (Capron et al., unpublished observations).
Indeed, eosinophils from allergic donors are more effi-
cient, and eosinophils from patients with HES (hyper
eosinophilic syndrome) less efficient than eosinophils from
normal donors. The heterogeneity of eosinophil-mediated
tumor cytotoxicity according to eosinophil donors led to
the suggestion that allergic patients are more efficient
towards tumor development with a potential tumor
sensing role of IgE. Experiments in the Capron-lab are
in progress to define the respective cytotoxic properties of
specific eosinophil cationic proteins vs tumoricidal mole-
cules shared with other effector cells, such as granzymes
and perforin. Besides IgE-dependent tumor cell cytotox-
icity of eosinophils (59), there is evidence for IgE
antibody-independent killing mechanisms.
IgE for passive immunotherapy of cancer patients?
In the late 1980s, experiments comparing the capacity of
various mouse/human chimeric antibodies of different
classes and subclasses to elicit ADCC and complement-
dependent cytotoxicity (CDC), found that IgG1 was
more effective than the other IgG subclasses tested
suggesting its superior tumoricidal effect (105). In con-
trast to IgG, human IgE is incapable of directing
complement activation against the targeted tumors
(106). The use of recombinant DNA technology has then
allowed the construction of recombinant IgG antibodies
that can recognize TAA and confer protection against
cancer cells by passive immunotherapy. The success of
these efforts has resulted in FDA-approved therapeutics
such as rituximab (Rituxan) and trastuzumab (Hercep-
tin) both with human IgG1 constant regions and
variable regions targeting CD20 (B-cell lymphoma
TAA) or HER2/neu (breast and ovarian cancer TAA),
respectively (107).
Coupled with the logical concern over IgE-induced
type I hypersensitivity, these results have demotivated
most researchers from further pursuing IgE for passive
immunotherapy. However, the use of IgE for the passive
immunotherapy of cancer would offer several advantages
over conventional IgG-based approaches, including its
high- and low-affinity receptors present on a broad
spectrum of effector cells, its capacity for antigen uptake
and presentation leading to a secondary immune
response. In contrast to IgE where serum concentrations
compose only 0.02% of the total antibody population,
IgG constitutes up to 85%, suggesting that a larger pool
of endogenous IgG competitors for cell surface receptors
could reduce the ADCC efficacy of a therapeutic IgG
dose (108). From this fact and its cytophilicity it may be
expected that lower passive doses of IgE antibody
preparations than necessary for IgG will be sufficient to
achieve therapeutic effects at the targeted tumor.
The first studies using IgE mAbs for the passive
immunotherapy of tumors were performed in the early
1990s (109). Using two IgE-producing hybridomas, Nagi
et al. (109) generated murine IgE antibodies targeting a
glycoprotein (gp36) of the mouse mammary tumor virus
(MMTV). Intraperitoneal injections of the monoclonal
IgE were given at 4-day intervals to CH3/HeJ mice to
treat a syngeneic subcutaneously injected MMTV-secret-
ing mammary carcinoma (H2712). After 6 or 8 weeks
of treatment, the monoclonal IgE therapy was capable of
preventing subcutaneous tumor development in 50% of
the animals treated, but did not protect mice exposed to
MMTV-negative mammary carcinoma cells (MA16/c).
However, this study did not offer side-by-side comparison
of IgE and IgG for the treatments of tumors. It was not
until the late 1990s that studies by Kershaw et al. (110)
investigated both the IgE- and IgG1-mediated growth
inhibition of solid tumors using the human colorectal
carcinoma (COLO 205) cell line. A murine IgE (30.6)
targeting an antigenic determinant on the colorectal
carcinoma cells was shown to transiently inhibit the
growth of COLO 205 cells injected subcutaneously into
SCID mice while both a mouse/human chimeric IgG1
AllergoOncology: the role of allergy in cancer
2008 The Authors
Journal compilation 2008 Blackwell Munksgaard Allergy 2008: 63: 1255–1266 1259
and IgE containing the (30.6) variable region and
corresponding human constant regions showed no effect.
The immune response against the COLO 205 tumors
directed by the murine IgE offered a robust yet transient
reduction of tumor growth using a dose of 1 lg per
mouse of IgE in contrast to the optimum dose of
4·250 lg per mouse required for IgG1. This effect
may be attributed to the superior affinity of IgE to its
receptor FceRI and to the usage of highly tumoricidic
effector cells. In fact, the lack of an anti-tumor effect
observed with the mouse/human chimeric IgE containing
human constant regions is not surprising as human IgE
does not cross-react with murine FceRI.
The potency of IgE interactions with effector cells and
their receptors in tumor cell killing has been demonstrated
using chimeric antibodies (MOv18 IgE and MOv18 IgG1)
against an ovarian tumor-specific antigen, folate binding
protein, expressed in 80% of ovarian cancers. In two
mouse xenograft models of human ovarian carcinoma,
treatment of the mice with IgE, combined with human
peripheral blood mononuclear cells (PBMC), had a
longer-lasting effect in restricting tumor growth than the
same treatment with IgG1 (105). In the second model,
grown orthotopically in nude mice, MOv18 IgE with
human PBMC, gave significantly greater protection than
PBMC alone, while MOv18 IgG
1
with human PBMC
offered no survival advantage (111). Immunohistochem-
ical studies of tumor sections showed the infiltration
of human monocytes/macrophages into tumor lesions,
associated with tumor necrosis and increased survival.
Although a number of studies have demonstrated the
potential of IgE in the passive immunotherapy of tumors,
most attempts have been riddled with limitations includ-
ing several incompatibilities between the mouse and
human immune systems in the domain of IgE-mediated
immunity (112). The complete potential of IgE-mediated
therapy that would be expected in humans has not been
achieved in many studies possibly because the expression
of FceRI in mice is limited to mast cells and basophils,
whereas in humans it is also expressed in monocytes,
macrophages, eosinophils, Langerhans cells and dendritic
cells. Moreover, the experiments described thus far using
mouse/human chimeric IgE for the treatment of tumor
xenografts in immune-suppressed mice suffered from a
limited supply of exogenous human PBMC; better results
may be potentially achieved in patients that have a fully
active immune system and a constant supply of effector
cells.
Some of the incompatibilities challenging IgE-mediated
immunity models have been solved with the development
of FceRI alpha chain transgenic mice. One of these mice
carries both, a human and a murine FceRI alpha chain
(113, 114), while in the other the murine FceRI alpha
chain is replaced by the human FceRI alpha chain (113,
114). The latter transgenic mouse exhibits the cell-specific
pattern of expression of the human alpha chain as well as
the correct structure of the receptor for each cell type.
This unique model can be potentially used in future IgE
immunotherapy studies in which a true IgE-directed
effector cell response could be mounted against a synge-
neic tumor model in vivo. Treatment of tumors using a
mouse/human chimeric IgE should generate a response in
mice closer to that of a human IgE response, but would
still be limited by the risk of a mouse anti-human
hypersensitivity response, preventing sequential adminis-
trations of chimeric IgE. These and other improvements
of transgenic models may soon provide a fully functional
in vivo readout system to test the efficacy of IgE against
various types of cancers in murine models.
IgE as adjuvant in tumor vaccination
Considering that the activation of the antigen–IgE–FceRI
axis profoundly affects the immune system both in its
cellular orchestration and programming, resulting in a
potent inflammatory state, an attempt of using IgE as a
potent adjuvant in anti-tumor vaccination has been
undertaken. In an initial study, the capability of targeting
mouse IgE on tumor cells and the consecutive effect on
tumor vaccination was assessed in C57BL/6 mice using a
T cell lymphoma line and an adenocarcinoma cell line
(115). Targeting of tumor cells was obtained exploiting
biotin–avidin bridges both in vitro, prior to cell admin-
istration, or in vivo by the three-step targeting method
(116). The use of biotin–avidin bridges eliminated the
need for IgE tumor specificity, however biotinylated IgG
anti-tumor antigen monoclonals were required to achieve
selective IgE targeting on tumor cells. Immunization
protocols were implemented in which irradiated tumor
cells (loaded with IgE or IgG) were administered,
followed by tumor cell challenge. Vaccinated mice
exhibited a strong tumor protection in the IgE-targeted
group, indicating that an IgE-specific adjuvant effect was
established. Conversely, nonimmunized mice or mice
immunized with IgG-targeted cells presented comparably
fast tumor growth and death. As IgE was absent during
tumor cell challenge, the protective effect should be
ascribed to the immune system stimulation upon the IgE-
targeted cell vaccination. Indeed, both eosinophils and
T cells (CD4
+
and CD8
+
) proved to be crucial for the
anti-tumor immunity, as their depletion led to abolish-
ment of tumor protection also in the IgE-treated mice. In
addition, dendritic cells loaded with tumor cell fragments
were found exclusively on draining lymph nodes of
IgE-treated animals (Siccardi et al., unpublished). This
represents a strong evidence for tumor antigen-loaded
dendritic cell migration from the tumor site to peripheral
lymph nodes, an important piece in the IgE-driven T cell
immune memory picture.
Based on these evidences, the anti-tumor IgE adjuvan-
ticity approach is being progressed trying to address three
major issues: (i) improvement of the vaccination strategy;
(ii) understanding IgE receptors involvement (FceRI and
Jensen-Jarolim et al.
2008 The Authors
1260 Journal compilation 2008 Blackwell Munksgaard Allergy 2008: 63: 1255–1266
CD23); and (iii) humanization of the reagents to move
towards clinical applications.
In view of the possibility to engineer modified vaccinia
virus Ankara (MVA) to produce recombinant anti-tumor
factors, MVA additive/synergic effect in combination
with IgE tumor cell loading was tested in mice immuni-
zation studies with encouraging results; in addition,
several IgE-targeting protocols have been tested, with
tumor cell haptenization and surface biotinylation as the
most promising ones (Nigro et al., in preparation). In
perspective, MVA being widely recognized as a safe and
promising tool for human therapy (117), its use in IgE-
based anti-tumor vaccination goes in the right direction
for future clinical trials. Use of recombinant MVA to
expressed membrane tumor antigens on the surface of
(IgE-loaded) infected tumor cells should then induce an
enhanced anti-tumor immunity. Most importantly, MVA
could be engineered to express membrane IgE (mIgE) on
the surface of tumor cells, exploiting the reported
capability by mIgE to activate FceRI (118). This strategy
would simplify the vaccination protocol of MVA and IgE
by fusing them into a single mIgE-expressing MVA.
Moreover, knowledge on the IgE–FceRI binding features
(119) allowed the production of a truncated heavy chain
mIgE (cleaved at the joint between the Ce2 and the Ce3
domain) that could represent the ultimate IgE version for
anti-tumor vaccination: guaranteeing FceRI activation,
preventing soluble IgE circulation and excluding antigen
binding. Testing FceRI a-chain knock out BALB/c mice
(120) in the MVA infection/IgE-loading tumor cell
vaccination system provided strong evidence on FceRI
importance for the IgE-orchestrated adjuvanticity
(unpublished observation). However, keeping in mind
the differences in FceRI expression between mouse and
human cell types, a humanized FceRI a-chain mouse
(113) is presently being investigated.
Overall, the mechanism behind IgE-driven tumor
vaccination appears to imply IgE–FceRI recognition, in
a tumor cell–FceRI
+
cell scenario. Following, it is highly
likely that tumor cell killing occurs, either directly by the
FceRI
+
-activated cells (mast cells/basophils) and/or by
the on-site recruitment of specialized killer cells such as
eosinophils. Cell killing releases tumor cell debris con-
taining tumor-specific antigenic determinant available for
dendritic cells that in turn transfer the information to
peripheral immune districts for the establishment and
consolidation of a tumor-specific immune memory
(Fig. 2).
From allergy to oncology: how to make a tumor antigen
an allergen
From the discussion shown, it is clear that IgE antibodies
targeted or fixed to tumor antigens in contrast to overall
elevated IgE levels cause a marked effect on tumor
development and growth. However, all these strategies
used passive applications of IgE. A combination of
knowledge in active cancer immunotherapy on the one
hand, and in basic allergy mechanisms on the other,
prompted recently the development of a vaccine that
would induce tumor-specific IgE in vivo. Two strategies
were combined: first, an epitope-specific vaccination
against the tumor antigen HER-2, rendering antibodies
with similar biological properties as the monoclonal
antibody trastuzumab (121). Second, an oral immuniza-
tion regimen discovered in food allergy research, resulting
in IgE induction (122–125).
Figure 2. Schematics of the mechanisms behind immunoglobulin E (IgE) adjuvanticity in anti-tumor vaccination. The IgE–FceRI
interaction is represented by the crystal structure of the FceRI-achain–IgE Ce3Ce4 complex.
AllergoOncology: the role of allergy in cancer
2008 The Authors
Journal compilation 2008 Blackwell Munksgaard Allergy 2008: 63: 1255–1266 1261
HER-2 is a member of the epidermal growth factor
receptor (EGFR, also know as ErbB) family. This
receptor is overexpressed in approximately 30% of breast
cancer patients and confers a detrimental prognosis in
the course of early as well as advanced breast cancer
(126). The monoclonal antibody trastuzumab (Hercep-
tin) targets this molecule, and beneficially influences
progression of early and advanced HER-2 overexpress-
ing tumors. In active cancer immunotherapy studies,
trastuzumab was used to generate an epitope-specific
vaccine targeting HER-2. Peptide mimics (so-called
mimotopes) were generated from the antibody-binding
site. These mimotopes induced trastuzumab-like IgG
antibodies in mice upon intraperitoneal immunization.
Importantly, actively induced antibodies upon mimotope
vaccination induce antibodies with similar biologic anti-
tumor features as the original antibody itself (121, 127,
128).
Second, food proteins can effectively lead to IgE
formation and sensitization when they persist the gastric
passage undegraded (129). Gastric digestion is impaired
in conditions of hypoacidity, as e.g. under anti-ulcer
treatment. Consequently, an oral immunization regimen
was developed that leads to induction of a Th2 bias in
mice, namely high levels of food-protein-specific IgG1
and IgE antibodies, eosinophilia and hypersensitivity of
skin and mucosa, when antigens are fed under concom-
itant gastric acid suppression (125, 130).
When the oral immunization regimen was adapted for
the mimotope vaccine construct, it could indeed demon-
strate that feedings of trastuzumab mimotopes under
anti-acidic conditions induced HER-2-specific IgE anti-
bodies in BALB/c mice. Apparently, this was the first
documented active induction of tumor-specific IgE
in vivo. The antibodies proved to be functional in an
allergologic mediator release assay, where mimotope-
induced IgE was found to react specifically with HER-2
overexpressing breast cancer cells. In an assay evaluating
the ADCC-mediating potential of the induced anti-HER-
2 IgE antibodies, they were shown to be effective in
mediating lysis of these breast cancer cells. The degree of
IgE–ADCC mediated by the sera correlated with the level
of HER-2-specific IgE detected by the mediator release
assay. These experiments show that it is possible to turn a
tumor antigen into an allergen (Fig. 3).
In the context of eliciting IgE towards self-antigens, it
is self-evident that safety issues are paramount, and the
risk of allergic reactions has to be carefully ruled out.
Similarly as on allergens, the optimal target antigen/
epitope should be present on the tumor cell surface in a
sufficient density to cause cross-linking of IgE antibodies
bound to the effector cells (131). On the other hand, the
target antigen should either not be shed into the blood-
stream, or its soluble form should not form aggregates
that could lead to undue effector cell activation and
systemic anaphylaxis. These issues will have to be
thoroughly monitored in future in vivo trials in the setting
of active as well as passive administrations of anti-tumor
IgE antibodies.
Taken together, it is feasible to induce IgE anti-tumor
antibodies by active immunization and it can be hypoth-
esized that for suitable antigens, this approach could be
developed into an easily applicable oral vaccination.
Conclusions
IgE antibodies favor the recognition of conformationally
intact antigens with a dense epitope display. Intact, viable
tumor cells fulfill these requirements because they over-
express antigens. Via interaction with its receptors on
numerous defence cells, IgE directs potent effector cells
into tumor tissues with proven tumoricidic activity. Thus
it can be hypothesized that IgE antibodies might phys-
iologically survey malignant cells. These principles will
hopefully in the future be exploited for vaccines and
passive antibody therapies.
Acknowledgments
This work was supported by grant PA-18238-B13 of the Austrian
Science Fund (FWF), by the Italian MURST (Cofin 2005), Uni-
versity of California Cancer Research Coordinating Committee
(CRCC), Susan G. Komen Breast Cancer Foundation, Howard
Hughes Medical Institute Gilliam Fellowship for Advanced Studies,
and UCLA MBI Whitcome Fellowship.
Figure 3. The principle of food allergy induction exploited for
oral cancer vaccination. When BALB/c mice were fed with
mimotopes for the important cancer antigen HER-2 under
concomitant anti-ulcer therapy, they developed immunoglobu-
lin E (IgE) specific for HER-2. The induced IgE was able to bind
to FceRI-positive RBL effector cells. Upon consecutive chal-
lenge with HER-2-positive SKBR-3 breast cancer cells, media-
tors were released which acted tumoricidic on the target cells.
The triggering was specific, as non-HER-2-expressing control
cancer cells (A432) were not able to lead to mediator release.
Jensen-Jarolim et al.
2008 The Authors
1262 Journal compilation 2008 Blackwell Munksgaard Allergy 2008: 63: 1255–1266
References
1. Ishizaka K, Ishizaka T. Physicochemi-
cal properties of reaginic antibody. 1.
Association of reaginic activity with an
immunoglobulin other than gammaA-
or gammaG-globulin. J Allergy 1966;
37:169–185.
2. Ishizaka K, Ishizaka T. Identification
of gamma-E-antibodies as a carrier of
reaginic activity. J Immunol 1967;
99:1187–1198.
3. Johansson SG. Raised levels of a
new immunoglobulin class (IgND)
in asthma. Lancet 1967;2:951–953.
4. Johansson SG, Bennich H. Immuno-
logical studies of an atypical (myeloma)
immunoglobulin. Immunology 1967;
13:381–394.
5. Scherrer E. The distribution of reagins
in the blood plasma. J Allergy
1930;2:467.
6. Layton LL, Yamanaka E. Demonstra-
tion of human reagins to foods, cat
dander, an insect, and ragweed and
grass pollens. J Allergy 1962;33:271–
275.
7. Augustin RDemonstration of reagins in
the serum of allergic subjects. In:
WEIR, Handbook of experimental
immunology. Oxford-Edinburgh:
Blackwell, 1967:1076.
8. Martin EG. Predisposing factors and
diagnosis of rectal cancer: a
discussion of allergy. Ann Surg
1935;102:56–61.
9. Bienengraber A. [Tumor metastasis in
the light of allergology.]. Zentralbl Chir
1952;77:1873–1881.
10. Molomut N, Spain DM, Kreisler L,
Warshaw LJ. The effect of an
allergic inflammatory response in the
tumor bed on the fate of transplanted
tumors in mice. Cancer Res
1955;15:181–183.
11. Berdel W, Nass G, Wiedemann G.
[Mechanism of tumor allergy and its
importance in tumor pathogenesis.]. Int
Arch Allergy Appl Immunol
1956;9:200–221.
12. Schlitter HE. [Is there an allergy against
malignant tumor tissue and what can it
signify in regard to the defense of the
body against cancer?]. Strahlentherapie
1961;114:203–224.
13. McCormick DP, Ammann AJ,
Ishizaka K, Miller DG, Hong R. A
study of allergy in patients with
malignant lymphoma and chronic
lymphocytic leukemia. Cancer
1971;27:93–99.
14. Ure DM. Negative assoication between
allergy and cancer. Scott Med J 1969;
14:51–54.
15. Augustin R, Chandradasa KD. IgE
levels and allergic skin reactions in
cancer and non-cancer patients. Int
Arch Allergy Appl Immunol
1971;41:141–143.
16. Jacobs D, Landon J, Houri M, Merrett
TG. Circulating levels of immunoglob-
ulin E in patients with cancer. Lancet
1972;2:1059–1061.
17. Allegra J, Lipton A, Harvey H, Luderer
J, Brenner D, Mortel R et al. De-
creased prevalence of immediate
hypersensitivity (atopy) in a cancer
population. Cancer Res 1976;36:3225–
3226.
18. Neuchrist C, Kornfehl J, Grasl M,
Lassmann H, Kraft D, Ehrenberger K
et al. Distribution of immunoglobulins
in squamous cell carcinoma of the head
and neck. Int Arch Allergy Immunol
1994;104:97–100.
19. Gould HJ, Sutton BJ, Beavil AJ, Beavil
RL, McCloskey N, Coker HA et al.
The biology of IGE and the basis of
allergic disease. Annu Rev Immunol
2003;21:579–628.
20. Zhang M, Murphy RF, Agrawal DK.
Decoding IgE Fc receptors. Immunol
Res 2007;37:1–16.
21. Keegan AD, Conrad DH. The
receptor for the Fc region of IgE.
Springer Semin Immunopathol
1990;12:303–326.
22. Macglashan D Jr. IgE and FceRI reg-
ulation. Ann N Y Acad Sci
2005;1050:73–88.
23. Hantusch B, Scholl I, Harwanegg C,
Krieger S, Becker WM, Spitzauer S
et al. Affinity determinations of
purified IgE and IgG antibodies
against the major pollen allergens
Phl p 5a and Bet v 1a: discrepancy
between IgE and IgG
binding strength. Immunol Lett
2005;97:81–89.
24. Vernersson M, Aveskogh M, Hellman
L. Cloning of IgE from the echidna
(Tachyglossus aculeatus) and a
comparative analysis of epsilon chains
from all three extant mammalian
lineages. Dev Comp Immunol 2004;
28:61–75.
25. Kumar S, Hedges SB. A molecular
timescale for vertebrate evolution.
Nature 1998;392:917–920.
26. Warr GW, Magor KE, Higgins DA.
IgY: clues to the origins of modern
antibodies. Immunol Today 1995;
16:392–398.
27. Vieira P, Rajewsky K. The half-lives of
serum immunoglobulins in adult mice.
Eur J Immunol 1988;18:313–316.
28. Meno-Tetang GM, Lowe PJ. On the
prediction of the human response: a
recycled mechanistic pharmacokinetic/
pharmacodynamic approach. Basic
Clin Pharmacol Toxicol 2005;96:182–
192.
29. Waldmann TA, Iio A, Ogawa M,
McIntyre OR, Strober W. The metab-
olism of IgE. Studies in normal indi-
viduals and in a patient with IgE
myeloma. J Immunol 1976;117:1139–
1144.
30. Yu P, Kosco-Vilbois M, Richards M,
Kohler G, Lamers MC. Negative feed-
back regulation of IgE synthesis by
murine CD23. Nature 1994;369:753–
756.
31. Achatz G, Nitschke L, Lamers MC.
Effect of transmembrane and cytoplas-
mic domains of IgE on the IgE
response. Science 1997;276:409–411.
32. Achatz G, Lamers MC. In vivo analysis
of the cytoplasmic domain of mIgE
antibodies. Int Arch Allergy Immunol
1997;113:142–145.
33. Luger E, Lamers M, Achatz-
Straussberger G, Geisberger R, Infuhr
D, Breitenbach M et al. Somatic
diversity of the immunoglobulin reper-
toire is controlled in an isotype-specific
manner. Eur J Immunol 2001;31:2319–
2330.
34. Karnowski A, Achatz-Straussberger G,
Klockenbusch C, Achatz G, Lamers
MCInefficient processing of mRNA for
the membrane form of IgE is a genetic
mechanism to limit recruitment of IgE-
secreting cells. Eur J Immunol 2006;
36:1917–1925.
35. Geisberger R, Crameri R, Achatz G.
Models of signal transduction through
the B-cell antigen receptor. Immunol-
ogy 2003;110:401–410.
36. Turner MC, Chen Y, Krewski D,
Ghadirian P. An overview of
the association between allergy
and cancer. Int J Cancer
2006;118:3124–3132.
37. Wang H, Diepgen TL. Is atopy a pro-
tective or a risk factor for cancer? A
review of epidemiological studies
Allergy 2005;60:1098–1111.
38. Turner MC, Chen Y, Krewski D,
Ghadirian P, Thun MJ, Calle EE.
Cancer mortality among US men
and women with asthma and hay
fever. Am J Epidemiol 2005;162:212–
221.
39. Kallen B, Gunnarskog J, Conradson
TB. Cancer risk in asthmatic subjects
selected from hospital discharge regis-
try. Eur Respir J 1993;6:694–697.
AllergoOncology: the role of allergy in cancer
2008 The Authors
Journal compilation 2008 Blackwell Munksgaard Allergy 2008: 63: 1255–1266 1263
40. McWhorter WP. Allergy and risk of
cancer. A prospective study using
NHANESI follow-up data. Cancer
1988;62:451–455.
41. Mills PK, Beeson WL, Fraser GE,
Phillips RL. Allergy and cancer: organ
site-specific results from the Adventist
Health Study. Am J Epidemiol
1992;136:287–295.
42. Vesterinen E, Pukkala E, Timonen T,
Aromaa A. Cancer incidence among
78,000 asthmatic patients. Int J Epi-
demiol 1993;22:976–982.
43. Eriksson NE, Holmen A, Hogstedt B,
Mikoczy Z, Hagmar L. A prospective
study of cancer incidence in a cohort
examined for allergy. Allergy
1995;50:718–722.
44. Gergen PJ, Turkeltaub PC, Sempos
CT. Is allergen skin test reactivity a
predictor of mortality? Findings from a
national cohort Clin Exp Allergy
2000;30:1717–1723.
45. Talbot-Smith A, Fritschi L, Divitini
ML, Mallon DF, Knuiman MW. Al-
lergy, atopy, and cancer: a prospective
study of the 1981 Busselton cohort. Am
J Epidemiol 2003;157:606–612.
46. Lindelof B, Granath F, Tengvall-
Linder M, Ekbom A. Allergy and can-
cer. Allergy 2005;60:1116–1120.
47. Gandini S, Lowenfels AB, Jaffee EM,
Armstrong TD, Maisonneuve
PAllergies and the risk of pancreatic
cancer: a meta-analysis with review of
epidemiology and biological mecha-
nisms. Cancer Epidemiol Biomarker
Prev 2005;14:1908–1916.
48. Linos E RT, Alonso A, Michaud D.
Atopy and risk of brain tumors: a
meta-analysis. J Natl Cancer Inst
2007;99:1544–1550.
49. Rosenbaum PF, Buck GM, Brecher
ML. Allergy and infectious disease
histories and the risk of childhood
acute lymphoblastic leukaemia. Paedi-
atr Perinat Epidemiol 2005;19:152–164.
50. Schuz J, Morgan G, Bohler E, Kaatsch
P, Michaelis J. Atopic disease and
childhood acute lymphoblastic leuke-
mia. Int J Cancer 2003;105:255–260.
51. Wen W, Shu XO, Linet MS, Neglia JP,
Potter JD, Trigg ME et al. Allergic
disorders and the risk of childhood
acute lymphoblastic leukemia (United
States). Cancer Causes Control
2000;11:303–307.
52. Matta GM, Battaglio S, Dibello C,
Napoli P, Baldi C, Ciccone G et al.
Polyclonal immunoglobulin E levels are
correlated with hemoglobin values and
overall survival in patients with multi-
ple myeloma. Clin Cancer Res
2007;13:5348–5354.
53. Wiemels JL, Wiencke JK, Patoka J,
Moghadassi M, Chew T, McMillan A
et al. Reduced immunoglobulin E and
allergy among adults with glioma
compared with controls. Cancer Res
2004;64:8468–8473.
54. Melbye M, Smedby KE, Lehtinen T,
Rostgaard K, Glimelius B,
Munksgaard L et al. Atopy and risk of
non-Hodgkin lymphoma. J Natl Can-
cer Inst 2007;99:158–166.
55. Wang H, Diepgen TL. Atopic derma-
titis and cancer risk. Br J Dermatol
2006;154:205–210.
56. Wang H, Rothenbacher D, Low M,
Stegmaier C, Brenner H, Diepgen TL.
Atopic diseases, immunoglobulin E and
risk of cancer of the prostate, breast,
lung and colorectum. Int J Cancer
2006;119:695–701.
57. Petridou ET, Chavelas C, Dikalioti SK,
Dessypris N, Terzidis A, Nikoulis DI
et al. Breast cancer risk in relation to
most prevalent IgE specific antibodies:
a case control study in Greece. Anti-
cancer Res 2007;27:1709–1713.
58. Bracher M, Gould HJ, Sutton BJ,
Dombrowicz D, Karagiannis SN.
Three-colour flow cytometric method
to measure antibody-dependent tumour
cell killing by cytotoxicity and phago-
cytosis. J Immunol Methods
2007;323:160–171.
59. Karagiannis SN, Bracher MG, Hunt J,
McCloskey N, Beavil RL, Beavil AJ
et al. IgE-antibody-dependent immu-
notherapy of solid tumors: cytotoxic
and phagocytic mechanisms of eradi-
cation of ovarian cancer cells. J
Immunol 2007;179:2832–2843.
60. Karagiannis SN, Bracher MG, Beavil
RL, Beavil AJ, Hunt J, McCloskey N
et al. Role of IgE receptors in IgE
antibody-dependent cytotoxicity and
phagocytosis of ovarian tumor cells by
human monocytic cells. Cancer Immu-
nol Immunother 2008;57:247–263.
61. Teng MW, Kershaw MH, Jackson JT,
Smyth MJ, Darcy PK. Adoptive
transfer of chimeric FceRI gene-modi-
fied human T cells for cancer immu-
notherapy. Hum Gene Ther 2006;
17:1134–1143.
62. Yokota A, Kikutani H, Tanaka T, Sato
R, Barsumian EL, Suemura M et al.
Two species of human Fcereceptor II
(FceRII/CD23): tissue-specific and IL-
4-specific regulation of gene expression.
Cell 1988;55:611–618.
63. Yokota A, Yukawa K, Yamamoto A,
Sugiyama K, Suemura M, Tashiro Y
et al. Two forms of the low-affinity
Fc receptor for IgE differentially
mediate endocytosis and phagocytosis:
identification of the critical cytoplasmic
domains. Proc Natl Acad Sci USA
1992;89:5030–5034.
64. Lorenzen J, Lewis CE, McCracken D,
Horak E, Greenall M, McGee JO.
Human tumour-associated NK cells
secrete increased amounts of inter-
feron-gamma and interleukin-4. Br
J Cancer 1991;64:457–462.
65. Aspord C, Pedroza-Gonzalez A,
Gallegos M, Tindle S, Burton EC, Su D
et al. Breast cancer instructs dendritic
cells to prime interleukin 13-secreting
CD4+ T cells that facilitate tumor
development. J Exp Med 2007;
204:1037–1047.
66. Paul-Eugene N, Mossalayi D, Sarfati
M, Yamaoka K, Aubry JP, Bonnefoy
JY et al. Evidence for a role of FceRII/
CD23 in the IL-4-induced nitric oxide
production by normal human mono-
nuclear phagocytes. Cell Immunol
1995; 163:314–318.
67. Mossalayi MD, Paul-Eugene N, Ouaaz
F, Arock M, Kolb JP, Kilchherr E
et al. Involvement of FceRII/CD23
and l-arginine-dependent pathway in
IgE-mediated stimulation of human
monocyte functions. Int Immunol
1994;6:931–934.
68. Brigati C, Noonan DM, Albini A,
Benelli R. Tumors and inflammatory
infiltrates: friends or foes? Clin Exp
Metastasis 2002;19:247–258.
69. Ribatti D, Crivellato E, Roccaro AM,
Ria R, Vacca A. Mast cell contribution
to angiogenesis related to tumour pro-
gression. Clin Exp Allergy
2004;34:1660–1664.
70. Beer T, Ng L, Murray K. Mast cells
have prognostic value in Merkel cell
carcinoma. Am J Dermatopathol
2008;30:27–30.
71. Soucek L, Lawlor E, Soto D, Shchors
K, Swigart L, Evan G. Mast cells are
required for angiogenesis and macro-
scopic expansion of Myc-induced pan-
creatic islet tumors. Nat Med 2007;
13:1211–1218.
72. Crivellato E, Beltrami CA, Mallardi F,
Ribatti D. The mast cell: an active
participant or an innocent bystander?
Histol Histopathol 2004;19:259–270.
73. Gordon JR, Galli SJ. Mast cells as a
source of both preformed and immu-
nologically inducible TNF-alpha/cac-
hectin. Nature 1990;346:274–276.
Jensen-Jarolim et al.
2008 The Authors
1264 Journal compilation 2008 Blackwell Munksgaard Allergy 2008: 63: 1255–1266
74. Zhao X, Mohaupt M, Jiang J, Liu S, Li
B, Qin Z. Tumor necrosis factor
receptor 2-mediated tumor suppression
is nitric oxide dependent and involves
angiostasis. Cancer Res 2007;67:4443–
4450.
75. Medina V, Croci M, Crescenti E,
Mohamad N, Sanchez-Jimenez F,
Massari N et al.The role of histamine
in human mammary carcinogenesis: H3
and H4 receptors as potential thera-
peutic targets for breast cancer treat-
ment. Cancer Biol Ther 2008;7:28–35.
76. Leek RD, Lewis CE, Whitehouse R,
Greenall M, Clarke J, Harris AL.
Association of macrophage infiltration
with angiogenesis and prognosis in
invasive breast carcinoma. Cancer Res
1996;56:4625–4629.
77. Mantovani A, Sozzani S, Locati M,
Allavena P, Sica A. Macrophage
polarization: tumor-associated macro-
phages as a paradigm for polarized M2
mononuclear phagocytes. Trends
Immunol 2002;23:549–555.
78. Schoppmann SF, Birner P, Stockl J,
Kalt R, Ullrich R, Caucig C et al.
Tumor-associated macrophages express
lymphatic endothelial growth factors
and are related to peritumoral lym-
phangiogenesis. Am J Pathol
2002;161:947–956.
79. Thomsen LL, Miles DW. Role of nitric
oxide in tumour progression: lessons
from human tumours. Cancer Metas-
tasis Rev 1998;17:107–118.
80. Xie K, Huang S, Dong Z, Juang SH,
Gutman M, Xie QW et al. Transfection
with the inducible nitric oxide synthase
gene suppresses tumorigenicity and
abrogates metastasis by K-1735 murine
melanoma cells. J Exp Med
1995;181:1333–1343.
81. Keller R, Geiges M, Keist Rl-arginine-
dependent reactive nitrogen intermedi-
ates as mediators of tumor cell killing
by activated macrophages. Cancer Res
1990;50:1421–1425.
82. Luiten RM, Fleuren GJ, Warnaar SO,
Litvinov SV. Target-specific activation
of mast cells by immunoglobulin E
reactive with a renal cell carcinoma-
associated antigen. Lab Invest
1996;74:467–475.
83. Luiten RM, Warnaar SO, Schuurman
J, Pasmans SG, Latour S, Daeron M
et al. Chimeric immunoglobulin E
reactive with tumor-associated antigen
activates human FceRI bearing cells.
Hum Antibodies 1997;8:169–180.
84. Sapino A, Cassoni P, Ferrero E,
Bongiovanni M, Righi L, Fortunati N
et al. Estrogen receptor alpha is a novel
marker expressed by follicular dendritic
cells in lymph nodes and tumor-asso-
ciated lymphoid infiltrates. Am J
Pathol 2003;163:1313–1320.
85. Dadabayev AR, Sandel MH, Menon
AG, Morreau H, Melief CJ, Offringa R
et al. Dendritic cells in colorectal can-
cer correlate with other tumor-infil-
trating immune cells. Cancer Immunol
Immunother 2004;53:978–986.
86. de Visser KE, Korets LV, Coussens
LM. De novo carcinogenesis promoted
by chronic inflammation is B lympho-
cyte dependent. Cancer Cell
2005;7:411–423.
87. Rothenberg ME, Hogan SP. The
eosinophil. Annu Rev Immunol
2006;24:147–174.
88. Asquith KL, Ramshaw HS, Hansbro
PM, Beagley KW, Lopez AF, Foster
PS. The IL-3/IL-5/GM-CSF common
receptor plays a pivotal role in the
regulation of Th2 immunity and aller-
gic airway inflammation. J Immunol
2008;180:1199–1206.
89. Woerly G, Roger N, Loiseau S,
Dombrowicz D, Capron A, Capron M.
Expression of CD28 and CD86 by hu-
man eosinophils and role in the secre-
tion of type 1 cytokines (interleukin 2
and interferon gamma): inhibition by
immunoglobulin a complexes. J Exp
Med 1999;190:487–495.
90. Woerly G, Roger N, Loiseau S, Capron
M. Expression of Th1 and Th2 immu-
noregulatory cytokines by human
eosinophils. Int Arch Allergy Immunol
1999;118:95–97.
91. Woerly G, Lacy P, Younes AB, Roger
N, Loiseau S, Moqbel R et al. Human
eosinophils express and release IL-13
following CD28-dependent activation.
J Leukoc Biol 2002;72:769–779.
92. Decot V, Capron M. [Eosinophils:
structure and functions]. Presse Med
2006;35:113–124.
93. Ionescu MA, Rivet J, Daneshpouy M,
Briere J, Morel P, Janin A. In situ
eosinophil activation in 26 primary
cutaneous T-cell lymphomas with
blood eosinophilia. J Am Acad Der-
matol 2005;52:32–39.
94. Munitz A, Levi-Schaffer F. Eosinoph-
ils: ÔnewÕroles for ÔoldÕcells. Allergy
2004;59:268–275.
95. Ishibashi S, Ohashi Y, Suzuki T,
Miyazaki S, Moriya T, Satomi S et al.
Tumor-associated tissue eosinophilia in
human esophageal squamous cell car-
cinoma. Anticancer Res 2006;26:1419–
1424.
96. Fernandez-Acenero MJ, Galindo-
Gallego M, Sanz J, Aljama A. Prog-
nostic influence of tumor-associated
eosinophilic infiltrate in colorectal car-
cinoma. Cancer 2000;88:1544–1548.
97. Nielsen HJ, Hansen U, Christensen IJ,
Reimert CM, Brunner N, Moesgaard
F. Independent prognostic value of
eosinophil and mast cell infiltration in
colorectal cancer tissue. J Pathol
1999;189:487–495.
98. Samoszuk M. Eosinophils and human
cancer. Histol Histopathol
1997;12:807–812.
99. Iwasaki K, Torisu M, Fujimura T.
Malignant tumor and eosinophils. I.
Prognostic significance in gastric
cancer. Cancer 1986;58:1321–1327.
100. Furbert-Harris PM, Parish-Gause D,
Hunter KA, Vaughn TR, Howland C,
Okomo-Awich J et al. Activated eo-
sinophils upregulate the metastasis
suppressor molecule E-cadherin on
prostate tumor cells. Cell Mol Biol
(Noisy-le-grand) 2003;49:1009–1016.
101. Ellyard JI, Simson L, Parish CR. Th2-
mediated anti-tumour immunity: friend
or foe? Tissue Antigens 2007;70:1–11.
102. Munitz A, Bachelet I, Fraenkel S, Katz
G, Mandelboim O, Simon HU
et al.2B4 (CD244) is expressed and
functional on human eosinophils.
J Immunol 2005;174:110–118.
103. Cormier SA, Taranova AG, Bedient C,
Nguyen T, Protheroe C, Pero R et al.
Pivotal advance: eosinophil infiltration
of solid tumors is an early and persis-
tent inflammatory host response. J
Leukoc Biol 2006;79:1131–1139.
104. Mattes J, Hulett M, Xie W, Hogan S,
Rothenberg ME, Foster P et al.
Immunotherapy of cytotoxic T cell-
resistant tumors by T helper 2 cells: an
eotaxin and STAT6-dependent process.
J Exp Med 2003;197:387–393.
105. Gould HJ, Mackay GA, Karagiannis
SN, OÕToole CM, Marsh PJ, Daniel BE
et al. Comparison of IgE and IgG
antibody-dependent cytotoxicity
in vitro and in a SCID mouse xenograft
model of ovarian carcinoma. Eur J
Immunol 1999;29:3527–3537.
106. Murphy KM, Travers P, Walpost M.
Janeway’s Immunobiology, 7th edn.
Oxford: Garland Science Publishing,
2008.
107. Carter PJ. Potent antibody therapeutics
by design. Nat Rev Immunol
2006;6:343–357.
108. Manz RA, Hauser AE, Hiepe F,
Radbruch A. Maintenance of serum
antibody levels. Annu Rev Immunol
2005;23:367–386.
AllergoOncology: the role of allergy in cancer
2008 The Authors
Journal compilation 2008 Blackwell Munksgaard Allergy 2008: 63: 1255–1266 1265
109. Nagy E, Berczi I, Sehon AH. Growth
inhibition of murine mammary carci-
noma by monoclonal IgE antibodies
specific for the mammary tumor virus.
Cancer Immunol Immunother
1991;34:63–69.
110. Kershaw MH, Darcy PK, Trapani JA,
MacGregor D, Smyth MJ. Tumor-
specific IgE-mediated inhibition of hu-
man colorectal carcinoma xenograft
growth. Oncol Res 1998;10:133–142.
111. Karagiannis SN, Wang Q, East N,
Burke F, Riffard S, Bracher MG et al.
Activity of human monocytes in IgE
antibody-dependent surveillance and
killing of ovarian tumor cells. Eur J
Immunol 2003;33:1030–1040.
112. Kinet JP. The high-affinity IgE receptor
(FceRI): from physiology to pathology.
Annu Rev Immunol 1999;17:931–972.
113. Dombrowicz D, Brini AT, Flamand V,
Hicks E, Snouwaert JN, Kinet JP et al.
Anaphylaxis mediated through a
humanized high affinity IgE receptor. J
Immunol 1996;157:1645–1651.
114. Fung-Leung WP, De Sousa-Hitzler J,
Ishaque A, Zhou L, Pang J, Ngo K
et al. Transgenic mice expressing the
human high-affinity immunoglobulin
(Ig) E receptor alpha chain respond to
human IgE in mast cell degranulation
and in allergic reactions. J Exp Med
1996;183:49–56.
115. Reali E, Greiner JW, Corti A, Gould
HJ, Bottazzoli F, Paganelli G et al.
IgEs targeted on tumor cells: thera-
peutic activity and potential in the de-
sign of tumor vaccines. Cancer Res
2001;61:5517–5522.
116. Paganelli G, Magnani P, Zito F, Villa
E, Sudati F, Lopalco L et al. Three-
step monoclonal antibody tumor tar-
geting in carcinoembryonic antigen-
positive patients. Cancer Res
1991;51:5960–5966.
117. Drexler I, Staib C, Sutter G. Modified
vaccinia virus Ankara as antigen deliv-
ery system: how can we best use its
potential? Curr Opin Biotechnol
2004;15:506–512.
118. Vangelista L, Soprana E, Cesco-
Gaspere M, Mandiola P, Di Lullo G,
Fucci RN et al. Membrane IgE binds
and activates FceRI in an antigen-
independent manner. J Immunol
2005;174:5602–5611.
119. Vangelista L. Current progress in the
understanding of IgE-FceRI interac-
tion. Int Arch Allergy Immunol
2003;131:222–233.
120. Dombrowicz D, Flamand V, Brigman
KK, Koller BH, Kinet JP. Abolition of
anaphylaxis by targeted disruption of
the high affinity immunoglobulin E
receptor alpha chain gene. Cell
1993;75:969–976.
121. Riemer AB, Klinger M, Wagner S,
Bernhaus A, Mazzucchelli L,
Pehamberger H et al. Generation of
peptide mimics of the epitope recog-
nized by trastuzumab on the oncogenic
protein Her-2/neu. J Immunol
2004;173:394–401.
122. Untersmayr E, Jensen-Jarolim E. The
effect of gastric digestion on food al-
lergy. Curr Opin Allergy Clin Immunol
2006;6:214–219.
123. Untersmayr E, Bakos N, Scholl I,
Kundi M, Roth-Walter F, Szalai K
et al. Anti-ulcer drugs promote IgE
formation toward dietary antigens in
adult patients. FASEB J 2005;19:656–
658.
124. Scholl I, Untersmayr E, Bakos N,
Roth-Walter F, Gleiss A, Boltz-
Nitulescu G et al. Antiulcer drugs
promote oral sensitization and hyper-
sensitivity to hazelnut allergens
in BALB/c mice and humans. Am J
Clin Nutr 2005;81:154–160.
125. Untersmayr E, Scholl I, Swoboda I,
Beil WJ, Forster-Waldl E, Walter F
et al. Antacid medication inhibits
digestion of dietary proteins and causes
food allergy: a fish allergy model in
BALB/c mice. J Allergy Clin Immunol
2003;112:616–623.
126. Yarden Y, Sliwkowski MX. Untan-
gling the ErbB signalling network. Nat
Rev Mol Cell Biol 2001;2:127–137.
127. Riemer AB, Kurz H, Klinger M,
Scheiner O, Zielinski CC, Jensen-
Jarolim E. Vaccination with cetuximab
mimotopes and biological properties of
induced anti-epidermal growth factor
receptor antibodies. J Natl Cancer Inst
2005;97:1663–1670.
128. Bramswig KH, Knittelfelder R, Gruber
S, Untersmayr E, Riemer AB, Szalai K
et al. Immunization with mimotopes
prevents growth of carcinoembryonic
antigen positive tumors in BALB/c
mice. Clin Cancer Res 2007;13:6501–
6508.
129. Riemer AB, Untersmayr E,
Knittelfelder R, Duschl A,
Pehamberger H, Zielinski CC et al.
Active induction of tumor-specific IgE
antibodies by oral mimotope vaccina-
tion. Cancer Res 2007;67:3406–3411.
130. Untersmayr E, Ellinger A, Beil WJ,
Jensen-Jarolim E. Eosinophils accu-
mulate in the gastric mucosa of food-
allergic mice. Int Arch Allergy Immu-
nol 2004;135:1–2.
131. Scholl I, Kalkura N, Shedziankova Y,
Bergmann A, Verdino P, Knittelfelder
R et al. Dimerization of the major
birch pollen allergen Bet v 1 is impor-
tant for its in vivo IgE cross-linking
potential in mice. J Immunol
2005;175:6645–6650.
Jensen-Jarolim et al.
2008 The Authors
1266 Journal compilation 2008 Blackwell Munksgaard Allergy 2008: 63: 1255–1266
... This system provides a defense mechanism against infections and toxins [47], but a consensus on whether this is the most relevant function of this Ig isotype has never been reached. Multiple epidemiological studies have described inverse associations between allergic diseases and malignancies [48,49] and led to the hypothesis-supported by experimental studies on animal models [50]-that IgE may be involved in antitumor surveillance, an area of investigation purposedly denominated 'allergo-oncology' [51]. In the largest of these epidemiological studies, among 1,102,247 US men and women who were cancer-free at baseline, with 18 years of follow-up from 1982 to 2000, a total of 81,114 cancer deaths were registered. ...
... Nevertheless, it has been proposed that FcεRI, whose expression is upregulated by IgE, mediate through IgE themselves antibody-dependent cellular phagocytosis (ADCP) and antibody-dependent cellular cytotoxicity (ADCC) against tumor cells which membrane antigens are recognized by IgE antibodies [55]. According to allergo-oncologists, these IgE antibodies recognizing tumor-associated antigens might activate different categories of antitumor effector cells, including macrophages, mast cells, and eosinophils [51]. Tumor-specific IgE have been induced in BALB/c mice with a mimotope vaccine orally administered under anti-acidic conditions, and these antibodies proved to be functional against breast cancer cells [51]. ...
... According to allergo-oncologists, these IgE antibodies recognizing tumor-associated antigens might activate different categories of antitumor effector cells, including macrophages, mast cells, and eosinophils [51]. Tumor-specific IgE have been induced in BALB/c mice with a mimotope vaccine orally administered under anti-acidic conditions, and these antibodies proved to be functional against breast cancer cells [51]. These outcomes led to proposing protocols to transform tumor-associated antigens in allergens, orally induce IgE antibodies, and activate strong antitumor ADCP and ADCC responses [51]. ...
Article
Full-text available
Opposite to other immunoglobulin (Ig) classes and subclasses, there is no consensus on the definition of normal levels of serum total IgE. However, longitudinal studies on birth cohorts produced growth charts of total IgE levels in helminth-free and never atopic children and defining the normal ranges of total serum IgE concentration at the individual, rather than population, level. Accordingly, very ‘low IgE producers’ (i.e., children whose tIgE level belong to the lowest percentiles) became atopic while keeping their total IgE levels in a range considered ‘normal’ if compared to the general age-matched population but ‘abnormally high’ if projected on the tIgE growth chart against the trajectory of that child’s own percentile levels. In ‘low IgE producers’, the IgE-specific activity, i.e., the ratio between allergen-specific and total IgE, is more important than the absolute specific IgE levels to confirm causality between allergen exposure and allergic symptoms. Patients with allergic rhinitis or peanut anaphylaxis but low or undetectable allergen-specific IgE levels must therefore be reconsidered considering their total IgE levels. Low IgE producers have been also associated with common variable immunodeficiency, lung diseases, and malignancies. A few epidemiological studies have shown a higher risk of malignancies in very low IgE producers, leading to a debated hypothesis proposing a novel, evolutionistic-relevant function for IgE antibodies for antitumor immune surveillance.
... Aside from molecular researches on cancer itself, other studies have been conducted to define the correlation of cancers with other chronic diseases like allergies, yielding a new area of research known as "allergooncology" associated with the basic mechanisms of allergic diseases affecting specific cancers (9). ...
... Aside from that, as IgE overproduction is considered to be a hallmark of allergic reactions, cross reactions of this immune component with some special tumor associated antigens may be related with cancer prognosis, as if the tumor antigen was formerly an allergic an antigen. Therefore, it may be intriguing to design an oral vaccine containing tumor antigens to induce IgE antibodies through an indigestible gastric passage (9). ...
Article
Full-text available
Both innate and adaptive arms of immune system play role in tumor development. Moreover, genetic and epigenetic alterations, widely demonstrated in cancer cells, result in disturbances in molecular pathways regulating cell growth, survival, and metastasis. Knowledge of molecular and cellular mechanisms involved in carcinogenesis leads to improvement of targeted treatments for cancers. Similarly, allergies are immune related entities and are treated according to the molecular mechanisms of hypersensitivity reactions. Some studies support the hypothesis of inverse association between cancer and susceptibility for allergies but the correlation is not simple and some demonstrate positive relation. For example, in some studies histamine released in response to allergens plays a role in tumor progression, probably through maintaining survival of myeloid derived suppressor cells (MDSCs). However, in other studies, the protective role of IgE against carcinogenesis have been reported. In this review, the role of immune system and specific molecular mechanisms in cancer and allergy will be discussed. Based on separately mentioned factors, interactions between these two seemingly disparate entities will also be presented. We conducted this review to illustrate potential molecular and cellular mechanisms underlying the association of cancer and allergy and make a basis for future interventions.
... Immunological responses, notably T helper type 2 (Th2)-associated asthma, are crucial, intertwined with hypersensitivity and allergies. Research on the relationship between atopic diseases and malignancies is a burgeoning interdisciplinary eld within allergic tumor biology [14,15] , This research focuses on elucidating the interplay between cancer and IgE-mediated and Th2-biased cellular responses. ...
Preprint
Full-text available
Objective: To investigate the phenotypic and genetic associations between asthma and melanoma. Methods: Initially, we conducted a cross-sectional study utilizing data sourced from the National Health and Nutrition Examination Survey (NHANES) spanning the years 2009 to 2016. In the survey, individuals answering "Yes" to "Ever been told you have asthma?" were categorized as asthma patients, while those answering "yes" to "Ever told you had cancer or malignancy?" and specifying "melanoma" were defined as melanoma patients. We employed a multivariate logistic regression model to evaluate the observed association between asthma and melanoma risk. Subsequently, the Mendelian randomization (MR) approach, a novel method minimizing bias and establishing causal relationships, was utilized asthma dataincluding childhood-onset (347,481 controls and 46,802 cases), adult-onset (347,481 controls and 9,676 cases), and overall onset (347,481 controls and 22,296 cases) were retrieved from GWAS databases, alongside melanoma data (32697 samples) obtained from the FinnGen. Two-sample MR analysis was conducted to evaluate the causal relationship between asthma and melanoma, with the Inverse Variance Weighting (IVW) method as the primary data analysis approach. Results: The observational study encompassed a total of 23,218 participants, including 129 melanoma patients. In the unadjusted multivariate logistic regression model, asthma was associated with a reduced risk of melanoma (odds ratio [OR] = 0.404, 95% confidence interval [CI] = 0.188-0.869, P=0.023). Similar results were observed after adjusting for gender, age, race, and education level (OR = 0.439, 95% CI = 0.209-0.922, P=0.034). However, further adjustments for variables such as BMI, poverty to income ratio, diabetes, alcohol use, smoking, hypertension, hyperlipidemia, and skin reaction to sun after non-exposure, outdoor time did not yield statistically significant results (OR = 0.675, 95% CI = 0.205-2.223, P=0.522). Mendelian analysis also failed to demonstrate a causal relationship between asthma and melanoma risk. Conclusion: While there may be an association between asthma and a reduced risk of melanoma, no causal relationship exists between the two. Further prospective studies with larger sample sizes are warranted to corroborate our findings.
... In addition, the NCCN recommends omalizumab for the treatment of severe ICI-induced pruritus [3]. Concerns of increased malignancy risk with omalizumab have been raised, especially given the role that IgE plays in cancer surveillance [123]. In fact, some investigators have found that low IgE levels may predict a higher risk for cancer development [124,125]. ...
Article
Full-text available
Immune-related cutaneous adverse events (ircAEs) commonly occur in patients on treatment with immune checkpoint inhibitors and can significantly reduce patient quality of life. These are often treated with immunomodulatory agents, including glucocorticoids, immunosuppressants, and biologics. While often effective at managing symptoms, these therapies can cause several adverse events which may limit their use. In addition, immunomodulatory agents should be used with particular caution in patients receiving immunotherapy, as the efficacy of the oncologic regimen may potentially be undermined. In this review, we summarize the safety of systemic therapies that are used in the management of ircAEs, with a particular focus on the resultant risk of secondary tumor progression in patients with active cancer.
... Applications of IgE-mediated ICs or IgE per se in clinical settings are not only limited to the allergy or asthma diseases but rapidly extending to the other diseases, including the treatment of cancer. In this regard, AllergoOncology is emerging as a new field, which suggests the exploitation of IgE antibodies to treat malignant diseases due to the presence of modulatory mechanisms for IgE antibodies [58]. In this study, we exhibited the modulatory mechanisms for IgE when it is in complex with an innocuous antigen with respect to eosinophilic responses to inflammatory stimuli. ...
Article
Full-text available
Immune complexes (ICs) skew immune responses toward either a pro- or anti-inflammatory direction based on the type of stimulation. Immunoglobulin E (IgE) is associated with Th2 immune responses and known to activate innate immune cells. However, roles of antigen (Ag)-specific-IgE ICs in regulating human eosinophil responses remain elusive; therefore, this study builts upon the mechanism of which ovalbumin (Ova)-IgE ICs affects eosinophilic responses utilizing human EoL-1 cell line as a model. Eosinophils are granulocytes functioning through pattern recognition receptors (PRRs) and destructive granule contents in allergic inflammation and parasitic infections. One of the PRRs that eosinophils express is NLRC4, a member of the CARD domain containing nucleotide-binding oligomerization (NOD)-like receptor (NLR) family. Upon recognition of its specific ligand flagellin, NLRC4 inflammasome is formed and leads to the release of interleukin-1β (IL-1β). We exhibited that Ova-IgE ICs induced the NLRC4-inflammasome components, including NLRC4, caspase-1, intracellular IL-1β, and secretion of IL-1β, as well as the granule contents MMP9, TIMP1, and TIMP2 proteins via TLR2 signaling; these responses were suppressed, when NLRC4 inflammasome got actived in the presence of ICs. Furthermore, Ova-IgE ICs induced mRNA expressions of MMP9, TIMP2, and ECP and protein expressions of MMP9 and TIMP2 in EoL-1 through FcɛRII. Interestingly, TLR2 ligand and Ova-IgE ICs costimulation elevated the number of CD63+ cells, a degranulation marker, as compared to the native IgE. Collectively, our findings provide a mechanism for the impacts of Ova-IgE ICs on eosinophilic responses via NLRC4-inflammasome and may help understand eosinophil-associated diseases, including chronic eosinophilic pneumonia, eosinophilic esophagitis, eosinophilic granulomatosis, parasitic infections, allergy, and asthma.
... IgE can bind to cancer cells nonspecifically, promoting the establishment of tumor-specific immune memory and serving as a potent adjuvant. In addition, eosinophils, macrophages, and mast cells attached to cytosolic IgE could serve as potent anti-tumor effectors [24]. Studies have shown that having allergies may be linked to a lower risk of certain types of cancers, including brain, hematological, lymphatic, gastrointestinal, and gynecological cancers. ...
Preprint
Introduction: The association between allergy and cancer is contradictory, whereas some forms of cancer have inverse associations with allergy. Allergic rhinitis (AR) is the most prevalent form of allergy, and lung cancer is one of the most prevalent forms of cancer, with the highest percentage of mortality. Recent studies report a positive association between asthma and lung cancer, although that, too, is not conclusive yet. However, AR has a positive association with asthma; therefore, our research question was to explore whether there is any correlation between AR and lung cancer epidemiologically. Methods and Results: Seven eligible articles were found to be eligible for this systematic review and meta-analysis, including 4663 cases and 9056 controls, five from the USA and one each from Canada and Germany. Pooled analysis [OR: 0.56; 95% CI: 0.45-0.70; p-value<0.00001], and [RR: 0.63; 95% CI: 0.51-0.77; p-value<0.00001] depicted a strong inverse relationship between AR and lung cancer. Among small-cell and non-small-cell lung cancer, the inverse association was stronger with Small-cell lung cancer (SCLC): [OR: 0.58, 95% CI: 0.46-0.73], and [RR: 0.64, 95% CI: 0.53-0.77], with a p-value <0.00001; however, this finding was only based on one study among females. The study in Canada (OR: 0.35 and RR: 0.38) and in Germany (OR: 0.18 and RR: 0.19) had lower OR and RR values compared to the studies in the USA (OR 0.62 and RR 0.69). However, two studies in Canada and Germany were the outlier studies in our study, and sensitivity analyses reduced the heterogeneity from 64% to 27% (for OR) and 72% to 40% (for RR) while analyses were conducted eliminating those studies. No low-quality studies were obtained. Conclusion: New epidemiological studies are required to observe the current scenario more comprehensively.
... In addition, basophils were found to improve the function of humoral immunity, and could release pro-inflammatory cytokines, which augmented the inflammatory antitumor reaction [13][14][15][16]. IgEmediated immune responses against tumors are known as allergo-oncology [17]. ...
Article
Full-text available
Purpose Recent reports have suggested that basophils influence allergic reactions and tumor immunity. In this study, we aimed to elucidate the association between preoperative circulating basophil (CB) counts and the outcomes of patients who underwent esophagectomy for esophageal cancer. Methods A total of 783 consecutive patients who underwent esophagectomy for esophageal cancer were eligible. The clinicopathological factors and prognoses were compared between the groups stratified by the preoperative counts of CB. Results There were more advanced clinical T and N stages in the low CB group than in the high CB group (P = 0.01 and = 0.04, respectively). The incidences of postoperative complications were comparable between the groups. The low CB count was associated with unfavorable overall and recurrence-free survivals (P = 0.04 and 0.01, respectively). In the multivariate analysis, low CB count was one of the independent prognostic factors for poor recurrence-free survival (HR 1.33; 95% CI 1.04–1.70; P = 0.02). In addition, hematogenous recurrence occurred more frequently in the low CB group than in the high CB group (57.6% vs. 41.4%, P = 0.04). Conclusion A preoperative low CB count was an unfavorable prognosticator in patients who underwent esophagectomy for esophageal cancer.
Article
Full-text available
Epidemiological studies have explored the relationship between allergic diseases and cancer risk or prognosis in AllergoOncology. Some studies suggest an inverse association, but uncertainties remain, including in IgE‐mediated diseases and glioma. Allergic disease stems from a Th2‐biased immune response to allergens in predisposed atopic individuals. Allergic disorders vary in phenotype, genotype and endotype, affecting their pathophysiology. Beyond clinical manifestation and commonly used clinical markers, there is ongoing research to identify novel biomarkers for allergy diagnosis, monitoring, severity assessment and treatment. Gliomas, the most common and diverse brain tumours, have in parallel undergone changes in classification over time, with specific molecular biomarkers defining glioma subtypes. Gliomas exhibit a complex tumour‐immune interphase and distinct immune microenvironment features. Immunotherapy and targeted therapy hold promise for primary brain tumour treatment, but require more specific and effective approaches. Animal studies indicate allergic airway inflammation may delay glioma progression. This collaborative European Academy of Allergy and Clinical Immunology (EAACI) and European Association of Neuro‐Oncology (EANO) Position Paper summarizes recent advances and emerging biomarkers for refined allergy and adult‐type diffuse glioma classification to inform future epidemiological and clinical studies. Future research is needed to enhance our understanding of immune–glioma interactions to ultimately improve patient prognosis and survival.
Article
Background: There is increasing evidence that allergic rhinitis (AR) is associated with cancer. However, these results are inconsistent. Because of common risk factors, there may be reverse causality and confounding factors that affect our understanding of the relationship between AR and cancer. We aimed to explore the role of AR in cancer development using Mendelian randomization (MR) analysis. Materials and Methods: We performed a two-sample MR analysis using summary data from genome-wide association studies (GWAS). Single nucleotide polymorphisms (SNPs) strongly associated with AR (or hay fever) were used as instrumental variables, mainly using the inverse variance weighted analysis method, supplemented by MR Egger, maximum likelihood, weighted media, and penalized weighted media for MR analysis. Sensitivity analyses included heterogeneity and horizontal pleiotropy; and leave-one-out analyses were performed to test the robustness of our results. Results: MR analysis revealed no evidence of a causal relationship between AR and any of the examined cancers (all p > 0.05). The results using five different analytical approaches were similar. Sensitivity analyses showed no evidence of heterogeneity nor horizontal pleiotropy. According to the leave-one-out sensitivity analyses, no individual SNP was significantly influencing the causal effect of AR on cancers. Conclusions: These findings do not provide evidence to support that AR has a large impact on the risk of eight common cancers in the European population. However, we cannot rule out a very minor effect of AR on cancer. Further large-scale studies are necessary to validate our findings.
Article
The anti-cancer impact of an allergic reaction is strongly linked to immunity enhancement. Trimethoprim-sulfamethoxazole (TMP-SMX), an antibiotic, has potential immunomodulatory effects, but has side effects such as allergies. Thus far, the effects and underlying mechanisms of TMP-SMX in melanoma have not been clarified. This study examined the potential roles of TMP-SMX in melanoma skin cancer using an immunodeficient mouse model. TMP-SMX significantly improved the survival rate and reduced the tumor weight and growth and vascular endothelial growth factor levels in melanoma skin cancer of immunodeficient mice. In the forced swimming test, TMP-SMX significantly reduced immobility time compared to the melanoma skin cancer of immunodeficient mice, indicating improved immunity. TMP-SMX significantly increased infiltration of mast cells and release of allergy-related mediators (IgE, histamine, interleukin (IL)-4, IL-5, IL-13, and IL-33) and immune-enhancing mediators (tumor necrosis factor-α, IL-2, IL-6, and IL-12). In addition, the administration of TMP-SMX significantly increased the caspase-3, 8, and 9 activities. Furthermore, mice given TMP-SMX showed no adverse reactions according to the blood biochemical parameters. TMP-SMX significantly inhibits the growth of melanoma skin cancer by triggering an allergic reaction and promotingimmunity. Hence, we propose that TMP-SMX may be used as an immune booster in cancer chemotherapy.
Article
Full-text available
Tumor-associated eosinophilia has been observed in numerous human cancers and several tumor models in animals; however, the details sur- rounding this eosinophilia remain largely unde- fined and anecdotal. We used a B16-F10 mela- noma cell injection model to demonstrate that eo- sinophil infiltration of tumors occurred from the earliest palpable stages with significant accumula- tions only in the necrotic and capsule regions. Fur- thermore, the presence of diffuse extracellular ma- trix staining for eosinophil major basic protein was restricted to the necrotic areas of tumors, indicat- ing that eosinophil degranulation was limited to this region. Antibody-mediated depletion of CD4 T cells and adoptive transfer of eosinophils sug- gested, respectively, that the accumulation of eo- sinophils is not associated with T helper cell type 2-dependent immune responses and that recruit- ment is a dynamic, ongoing process, occurring throughout tumor growth. Ex vivo migration stud- ies have identified what appears to be a novel che- motactic factor(s) released by stressed/dying mela- noma cells, suggesting that the accumulation of eosinophils in tumors occurs, in part, through a unique mechanism dependent on a signal(s) re- leased from areas of necrosis. Collectively, these studies demonstrate that the infiltration of tumors by eosinophils is an early and persistent response that is spatial-restricted. It is more important that these data also show that the mechanism(s) that elicit this host response occur, independent of im- mune surveillance, suggesting that eosinophils are part of an early inflammatory reaction at the site of tumorigenesis. J. Leukoc. Biol. 79: 000-000; 2006.
Article
Allergic reactions are mediated by IgE antibodies bound to high‐affinity receptors on mast cells in peripheral tissues and are characterized by their immediacy and hypersensitivity. These properties could also be advantageous in immunotherapy against cancer growth in peripheral tissues. We have constructed chimeric IgE and IgG1 antibodies with murine V regions and human C regions corresponding to the MOv18 monoclonal antibody against the human ovarian tumor‐associated antigen, folate binding protein. The antibodies exhibited the expected binding affinities for antigen and Fc receptors, and effector activities with human basophils and platelets in vitro. The protective activities of MOv18‐IgE and MOv18‐IgG1 were compared in a SCID mouse xenograft model of ovarian carcinoma. The beneficial effects of MOv18‐IgE were greater and of longer duration than those of MOv18‐IgG1. Our results suggest that the allergic reaction could be harnessed for the suppression of ovarian tumors.
Article
In a survey of 140 patients in the gynaecology wards of a large general hospital, the subjects were asked whether they suffered from any of the common hay-fever type allergies, for example, to flowers, bedding dust, etc. An incidence of 20 per cent was established. Surgical and biopsy reports on this ward population established an incidence of malignant conditions at 28 per cent. These two groups of comparable age turned out to be mutually exclusive. Reports of pregnancy nausea were also taken. Previous susceptibility to this varied from none to severe, and correlated positively with reports of common allergic type reactions, and consequently, inversely, with liability to gynaecological cancers in later life. The possible significance of this result in the studies of cancer is discussed.
Article
Infectious disease histories were evaluated in a population-based case-control study of childhood acute lymphoblastic leukaemia (ALL) as it has been hypothesised that delays in early infections are associated with an increased risk of disease. Allergy histories were also assessed as part of a broader evaluation of the role of immune factors in ALL. Cases (n = 255) were diagnosed between 1980 and 1991 at one of four referral centres in a 31-county area of New York State; controls (n = 760) were a random sample of live births from the same region, frequency matched to cases by sex, race and birth year. Data were collected by mailed questionnaire, completed by case and control parents in 1995. Allergy and infectious histories before the age at leukaemia diagnosis for cases and an equivalent age for controls were evaluated. The adjusted odds ratio and 95% confidence interval [CI] associated with a positive history of any allergy was 0.58 [95% Cl 0.38, 0.88] compared with a negative allergy history The occurrence of several common childhood illnesses before 25 months of age and ALL were assessed, with both weak positive and weak inverse associations observed. Overall, these analyses provide little support for the hypothesis that infection delay in early life is associated with an increased risk of ALL. Children with positive allergy histories reported significantly more infections than those with negative histories; however, effect modification of the infection-ALL associations by child allergy history was not observed. Nonetheless, these observations suggest the importance of assessing both allergy and infectious histories and their possible interactions when evaluating the association between these immune factors and childhood ALL.
Article
Background: In 1983, Pretlow et al. reported one classical study of the prognostic influence of eosinophil infiltration in human colon carcinoma. Since then, very few reports have analyzed this supposed prognostic influence in this type of tumor, although there have been several reports of other types of tumors with different results. Eosinophils seem to play a central role in the immunologic defense against tumors; their activity can be induced through immunotherapy with interleukin in cases that are unresponsive to conventional therapies. Methods: To analyze the prognostic influence of eosinophils in colorectal carcinoma, the authors selected 126 patients with this type of tumor treated only with curative surgery and followed for a minimal period of 5 years. They divided these patients into 4 categories according to the number of eosinophils per high-power field (400x): 0 (Grade 1), 1-9 (Grade 2), 10-50 (Grade 3), and more than 50 (Grade 4). Results: The results of this series confirm that high eosinophil counts are associated with a significantly better prognosis. In multivariate analysis, this factor was independent from staging, vascularization, p53 expression, and histologic grade. Conclusions: Eosinophil count served as a significant independent favorable prognosticator in colorectal carcinoma patients.
Article
The associations among certain allergic disorders, atopy upon skin-prick testing, and specific cancers were evaluated in a prospective study. Information regarding history of asthma and hay fever was collected by questionnaire from 3,308 cancer-free participants in the 1981 Busselton Health Survey. A subset of 1,005 participants also underwent skin-prick testing. The cohort was followed for a new diagnosis of cancer or death until the end of 1999. Cox proportional hazards regression analysis was used to estimate adjusted hazard ratios (relative risks) for breast, prostate, colorectal, lung, and hematologic cancers and melanoma. Having a skin reaction to house dust mites nearly tripled the risk of prostate cancer (relative risk = 2.90, 95% confidence interval: 1.26, 6.68). History of asthma and hay fever were associated with a trend toward a reduced risk of colorectal cancer and increased risk of leukemia, but these results were not statistically significant. Hay fever was associated with melanoma risk in men but not in women. No association was found between breast and lung cancers and allergic disorders or atopy.
Article
Long considered to be secondary cells characterized mainly by their ability to be recruited to inflammation sites, these cells are now known to release a wide array of cytotoxic mediators. Moreover they participate in immune response regulation by producing Th1 and Th2 cytokines as well as regulatory cytokines and chemokines. This review describes recent findings about their expression of surface molecules, eosinophil mediators, and the role of both in these novel eosinophil functions.