IMR Press / FBL / Volume 26 / Issue 3 / DOI: 10.2741/4899
Review
Environmental pollutants: an immunoendocrine perspective on phthalates
Show Less
1 Laboratorio de Especiacion Quimica de Aerosoles Organicos Atmosfericos y Desarrollo de Tecnologias Verdes, Departamento de Ciencias Ambientales, Centro de Ciencias de la Atmosfera, Universidad Nacional Autonoma de Mexico, CP 04510, Ciudad de Mexico, Mexico
2 Departamento de Inmunologia, Instituto de Investigaciones Biomedicas, Universidad Nacional Autonoma de Mexico, AP 70228, Ciudad de Mexico, CP, 04510, Mexico
3 Laboratorio de Genotoxicologia y Mutagenesis Ambientales, Departamento de Ciencias Ambientales. Centro de Ciencias de la Atmosfera, Universidad Nacional Autonoma de Mexico, CP 04510, Ciudad de Mexico, Mexico
Send correspondence to: Karen Elizabeth Nava-Castro, Laboratorio de Genotoxicologia y Mutagenesis Ambientales, Departamento de Ciencias Ambientales. Centro de Ciencias de la Atmosfera, Universidad Nacional Autonoma de Mexico, CP 04510, Ciudad de Mexico, Mexico, Tel: 52-1-55-56224077, Fax: 52-1-55-56223369, E-mail: karlenc@atmosfera.unam.mx
Front. Biosci. (Landmark Ed) 2021, 26(3), 401–430; https://doi.org/10.2741/4899
Published: 1 October 2020
Abstract

Phthalates are endocrine disrupting compounds (EDCs) used as plasticizers in a wide array of daily-use products, from flooring and automotive parts to medical devices and are even present in the children´s toys. Since these compounds are not covalently bound other molecules, they leach from these synthetic products, causing a high level of human exposure to them. EDCs exert several endocrine effects, most typically, reduced biosynthesis of the male hormone, testosterone and disturbances in estrogen, androgen, PPAR-gamma and AhR that control complex immunoendocrine regulatory networks. Besides impacting the developmental processes and long-term adverse effects, since cells of the immune system express endocrine receptors, and synthetize and respond to several hormones and other endocrine ligands, phthalates also cause dysregulation of immune system.

Keywords
Phthalates
Immunomodulation
Immune
Endocrine
Disruptor
EDC
Review
2. INTRODUCTION

Phthalates are considered endocrine disrupting compounds (EDCs), which are molecules that interfere with hormonal homeostasis. Particularly, phthalates have been shown to interact with several receptors involved in estrogenic and androgenic pathways (1). It is important to note that even though the main target of EDCs is the endocrine system, they may affect other tissues. One of such tissues are cells from the immune system, which are greatly susceptible to endocrine modulation. This relationship was first recognized with epidemiological studies that found a higher incidence of some autoimmune diseases -such as rheumatoid arthritis and systemic erythematous lupus- in the female population (2, 3). Later, it was demonstrated that sex steroids regulate multiple immunological functions like lymphocyte maturation, cellular migration, expression of MHC molecules, and cytokine production and that this regulation is mediated by hormonal receptors expressed by immune cell populations (4).

Though some immunological effects of EDCs have been revised previously (5), this review aims to summarize the current evidence of immune dysregulation caused specifically by phthalates; and linking them with the acknowledged molecular mechanisms under an immunoendocrine perspective.

2.1. Phthalates

Phthalates are alkyl diesters of phthalic acid and are used as plasticizers in PVC products, as solvents and fixatives in personal care products and as additives in the enteric coating of some drug tablets. Thus, phthalates are present in a wide array of daily-use products, from flooring and automotive parts to medical devices and even children´s toys. Since these compounds are not covalently bound to the products, they easily migrate from them (6), leading to high human exposure.

Phthalate exposure has been estimated by biomonitoring studies and ranges between 1 and 2 ug/kg bw/day of a single compound for adults (7, 8) and up to 4 ug/kg bw/day for children (9). While oral consumption of contaminated food and dermal absorption from personal care products are considered the main exposure routes (10–12), phthalates are ubiquitously found in environmental matrices such as in the atmosphere, soil and water bodies (13). Of particular interest is the presence of phthalates in the atmospheric aerosol, forming part of the organic content of fine particulate matter (PM2.5). Phthalate concentrations can range from 1 to 100 ng/m3 outdoors (14, 15) and up to 1,000 ng/m3 indoors (16, 17), being Bis-(2-ethylhexyl) phthalate (DEHP) and Di-n-Butyl phthalate (DBP) the most common.

2.1.1. Phthalates as endocrine disrupting compounds (EDCs)

An Endocrine disruptor compounds is defined by the United States Environmental Protection Agency (EPA) as an “exogenous agent that interferes with synthesis, secretion, transport, metabolism, binding action, or elimination of natural blood-borne hormones that are present in the body and are responsible for homeostasis, reproduction, and developmental process” (18). EDCs mimic endogenous hormones and bind to their receptors inducing functional and regulatory activities of gene expression.

Phthalates exert sexually dimorphic effects, affecting in different ways to male and female individuals. Bases on epidemiological data, elevated levels of DEHP urinary metabolites were linked to female patients experiencing premature thelarche (19). On the other hand, phthalates effects on males are grouped into the so called “phthalate syndrome”, which is a type of testicular dysgenesis featuring cryptorquidism hypospadias, incomplete testicular descend, reduced ano-genital distance, low sperm count and/or quality, elevated risk of infertility and testicular cancer (6, 20–22), caused by a reduced testosterone synthesis.

2.1.2. Phthalate metabolism

Phthalates are readily metabolized once in the organism, leading to the urinary and fecal excretion of their metabolites. The first metabolic phase is an enzymatic hydrolysis of one of the side chains, which turns the original diester into a monoester (e.g. DEHP into MEHP). Phase I is catalyzed by esterases, which are present in blood plasma and most tissues (23). It is important to note that phthalate monoesters can also exert biological effects and maintain binding potential to some receptors (24–27). After this first phase, the monoesters can be further transformed by hydroxylation and/or carboxylation of the remaining side chain, before being further metabolized to glucuronide conjugates in phase II (23).

2.2. Molecular signaling induced by phthalates

The biological action induced by phthalates differ from the target cell in which they have an effect. The best described mechanisms involve signaling through the hormone receptors: estrogen (ER) and androgen (AR) receptors; and by the peroxisome proliferator-activated receptor gamma (PPAR-gamma), although other non-transcriptional mechanisms have been described, such as the activation of G protein-coupled receptors (Figure 1).

Figure 1

Molecular signaling induced by phthalates. Phthalates and some phthalate metabolites may exert fast non-transcriptional effects by engaging and activating G protein-coupled receptors (GPCR), which activate downstream pathways that lead to intracellular calcium increases and activation of diverse protein kinases. Phthalates and their metabolites readily cross cell membranes, entering the cytoplasm. Once in the cytoplasm, they may bind to estrogen receptors (ER) and induce their dimerization and therefore their activation, leading to transcriptional activity, which may vary depending on the existing coregulators. Phthalates may also bind androgen receptor (AR), triggering transcriptional effects and/or regulating the activation of kinases and transcription factors. Aryl hydrocarbon receptor (AhR) may also bind phthalates, enter the nucleus and regulate transcription of responsive genes. Upon nucleus localization, phthalates may also act as Peroxisome proliferation-activated receptor gamma (PPARγ) ligands, leading to transcriptional activity.

2.2.1. Estrogen receptor (ER) pathways

Classical nuclear estrogen receptors (ER-alpha and ER-beta) belong to the ligand-activated transcription factor family. These receptors are widely expressed in mammal tissues, some of them expressing ER-alpha (uterus, liver, kidney and heart) or ER-beta (ovary, prostate, hematopoietic and central nervous systems) and some others expressing both (mammary gland, thyroid and adrenal glands and some brain regions). Even though both receptors are highly homologous and display a similar affinity for their natural ligand 17-beta-estradiol and for estrogen response elements (EREs) in the DNA, (28, 29), ligand binding to one receptor or another triggers different responses. The key to such differential responses lays in the signaling pathways of ERs.

Estrogen receptors reside in the cytoplasm and form homo- or heterodimers upon ligand binding, translocate to the nucleus and bind to the EREs in the DNA, typically located near the promoter regions of target genes. This activation leads to the recruitment of coregulators, which associate to the ER dimers, forming the initial transcriptional complex (30). It is important to note that coregulators can either increase a gene’s transcription (coactivator) or inhibit it (corepressor) and that recruitment of one or another depends on the type of ER dimer (ER-alpha or ER-beta homodimers or heterodimer) and, of course, on which coregulators are present in a particular moment in a particular cell type.

The activity of estrogen receptors is not limited to target genes containing EREs. Ligand-activated ERs can modulate gene expression by attaching to other transcription factors which in turn act on their respective response elements (30). In addition to the former, estrogen receptors may also be embedded into the plasma membrane, where they can bind extracellular ligands. Moreover, some non-classical membrane estrogen receptors have been described, such as GPER (or Gpr30), which is a G protein-coupled receptor. These membrane-bound receptors trigger rapid effects such as calcium flux, second messenger production (cAMP, DAG, IP3) and kinase activation (31, 32).

Considering the former, estrogens effects depend not only on the dose, but on the homo- or heterodimers formation, their target (either nucleus or cytoplasmic proteins), the coregulators and active transcription factors existing and all of it varies according to the context and specific cell type.

2.2.2. Androgen receptor (AR) pathway

Similar to estrogen receptors, androgen receptor (AR) has a widespread distribution, from reproductive tissues to muscle, bone, adipose tissue, and cardiovascular, immune and nervous system (33). Endogenous ligands for AR are testosterone and dihydrotestosterone (DHT), a metabolite of the former with a greater agonistic activity.

AR resides in cytoplasm, attached to the cytoskeleton by associating with heat shock proteins or other chaperones. Upon androgen binding, AR suffers conformational changes that dissociate it from the chaperones, which allows it to interact with coregulators and translocate to the nucleus, where it can dimerize and bind to DNA regions denominated androgen response elements (AREs). Once there, AR together with coregulators forms an initial transcription factor that recruits the whole transcription machinery (33, 34).

Androgens can also indirectly modulate genetic expression without binding to AREs. The ligand-AR complex can activate other signaling pathways such as ERK and MAPK or even sequestrate other transcription factors (e.g. AP-1), repressing in this way the expression of their target genes (33).

2.2.3. The peroxisome proliferator-activated receptor gamma (PPAR-gamma) pathway

Phthalates have been also described as peroxisome proliferators, inducing hepatocyte peroxisome and cellular proliferation, which in some studies in mice and rats, have an impact in the incidence of hepatocellular adenomas. This effect is dependent on the activation of the nuclear receptor peroxisome proliferator activated receptor gamma (PPAR-gamma). This receptor belongs to the nuclear receptor subfamily of ligand-inducible transcription factors, which were originally described as key transcriptional regulator in the metabolism of glucose and fatty acids, lipid storage and insulin sensitivity (35–37). In fact, it is one of the main targets in the treatment of metabolic syndrome, insulin resistance and type 2 diabetes (38), since it is highly expressed in white and brown adipose tissues.

Constitutively, PPAR-gamma forms a heterodimer with the retinoid X receptor (RXR) and they bind to specific DNA sequences termed PPAR response elements (PPRE). In basal conditions, PPAR-gamma is bound to corepressor molecules such as the silencing mediator of retinoid and thyroid hormone receptors (SMRT), the nuclear receptor corepressor (NCoR) and some histone deacetylases (HDACs). Upon ligand binding, there is a conformational change in the PPAR-gamma protein that allows the dissociation of its repressors and the recruitment of its coactivators. Among these are the CREB-binding protein (CBP), the histone acetyl transferase (p300) and the PPAR-binding protein 4 (PBP). During adipocyte differentiation, changes in the recruitment of coactivators allow the formation of different transcription complexes that modify the responses to metabolic stimuli and ensure the activation of specific signaling pathways (39).

2.2.4. The Aryl hydrocarbon receptor (AhR) pathway

The AhR was initially described when analyzing the effects of the herbicide 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) induced in exposed chemical workers. This receptor belongs to the family of the bHLH Pas domain transcription factors that are involved in sensing environmental changes. The canonical signaling pathway is induced when its ligand reaches the cytoplasm and bind to the AhR –which in the absence of ligand, is attached to actin filaments by the interaction with the chaperon proteins HSP90, AIP and p23. The ligand-AhR/HSP90/p23 complex then translocate to the nucleus, where it is released from this complex and heterodimerizes with the ARNT protein. This dimer binds to genomic regions containing the dioxin response element (DRE) and induce the transcription of genes such as CYP1A1, CYP1A2, CYP1B1 and the AhR repressor (AhRR). The expression and function of these proteins regulate negatively this signaling pathway by three mechanisms: 1) AhR proteosomal degradation – mediated by the covalently union of ubiquitin-; 2) Ligand degradation by CYP1A1 and 3) AhR/ARNT complex disruption by the repressor AhRR. This repressor is a protein similar to AhR but differs in two characteristics: it lacks its ligand-binding domain and contains a transactivation domain in the C-terminal domain that recruit other co-repressors (40, 41).

In addition to this canonical pathway, the AhR/ARNT complex can recruit other proteins and regulate either transcription or chromatin remodeling. Among these proteins, the SW1/SNF remodeling complex and the steroid receptor coactivator-1 (SRC-1), CBP/p300, p160/bHLH-PAS, NCoA2/GRIP1/TIF2, pCIP (p300/CPB/CoIntegrator associated protein) and RIP140 or the ATP-dependent chromatin remodeling components such as BRG-1 (40, 41).

There are other signaling pathways related to AhR activation. For example, the ligand TCDD increase Ca2+ concentration by activating the Src tyrosine kinase and the Focal adhesion kinase (FAK). These two pathways converge in MAPK activation and therefore, there is also an induced migration, adhesion and inflammation-related functions such as prostaglandin and arachidonic production.

2.3. The role of endocrine receptor pathways on the immune system

The immune and endocrine systems are part of the neuroimmunoendocrine network, a recently recognized multisystem circuitry that regulates homeostasis. Regarding the immune system, its cellular components not only express endocrine receptors, but synthetize and respond to several hormones and other endocrine ligands. Some molecular pathways typically associated to the endocrine system play a developmental and regulatory role on the immune system, such as the thymus involution through age, the sexual dimorphism in immune response and the tolerogenic character of immune response during pregnancy.

Estrogen receptors are expressed in most immune cell populations, from hematopoietic precursors to mature B and T lymphocytes, NK cells, monocytes and dendritic cells (42). However, expression patterns are different, with some cells expressing preferentially ER-alpha (e.g. CD4+ T cells) and other expressing higher levels of ER-beta (e.g. B cells) (43, 44). In homeostatic conditions, ER-alpha acts on myeloid and lymphoid precursors, inducing developmental pathways (42).

The developmental importance of ER-alpha is depicted by the thymic and splenic hypoplasia observed in ER-alpha-deficient mice (45). ER has a profound effect on T lymphocytes which is sensitive to the dose or physiological levels of estradiol, the natural ligand. A good example is the shift from Th1 response towards Th2 and Treg expansion due to high levels of estrogen, as occurs during pregnancy. B cells are deeply modulated by estrogen, both ER-alpha and ER-beta pathways are involved in maturation, function and survival of this population (42), so that females tend to display a greater humoral response than males (46). Differentiation and maturation of dendritic cells, as well as macrophage effector functions are also modulated by estrogen receptor pathways (44).

AR is expressed in several cell populations from both myeloid and lymphoid origin and has been involved in sexual immunological dimorphism and immunosuppression (47). The majority of the studies regard androgens and and androgen receptor as suppressive or regulatory. AR pathways downregulate T cell and B cell proliferation, as well as the inflammatory response (48). Androgens also modulate the functions of dendritic cells, downregulating expression levels of MHC, HLA and costimulatory molecules, which further modulates T cell response (47). One exception to the predominantly suppressive role of androgens is the neutrophil subpopulation, since androgens (via AR) promotes their differentiation and maturation, probably by enhancing Granulocyte colony-stimulating factor (G-CSF) signaling (48). Though histamine release by mast cells appears to be downregulated by androgens, testosterone has been shown to induce IL-33 expression, which drives the generation of innate lymphoid cells and basophils, which in turn produce Th2 cytokines and drive IgE antibody switch (47). A very interesting observation is that IgE is found to be higher in young males over females affected by allergic rhinitis (49).

PPAR-gamma is expressed in lower amounts in some immune cells such as macrophages, eosinophils, dendritic cells, T and B lymphocytes; however, it has a critical role in both differentiation and activation of these cells in diseases that involve chronic inflammation such as atherosclerosis, inflammatory bowel disease and rheumatoid arthritis (50).

Expression of PPAR-gamma is differentially expressed in different subtypes of macrophages. Resting bone-marrow-derived macrophages express low levels of PPAR-gamma mRNA, whereas activated peritoneal macrophages express high levels of PPAR-gamma (51). However, peritoneal macrophages incubated with 15-dPGJ2 –a PPAR-gamma agonist- showed morphological features typical of resting cells, even in the presence of IFN-gamma. In addition, it also downregulates iNOS expression and inhibit IFN-gamma-dependent nitrite production. This suggests that might inhibit the expression of genes that become upregulated during macrophage differentiation and activation (51). In contrast, other studies report that activation of PPAR-gama primes primary human monocytes into M2 differentiation, but it is not able to modify marker expression of differentiated M1 or M2 macrophages (52).

Eosinophils, an immune cell population related to allergic responses and asthma, express PPAR-alpha, PPAR-gamma and PPAR-delta. In fact, the administration of PPAR-gamma agonists (rosiglitazone or pioglitazone) in a murine model of allergic asthma, induce a reduction in the number and activation state of eosinophils in the airways, determined by the expression of IL-4, IL-5 and ovalbumin-specific IgE, which suggest that PPAR-gamma may be protective factor in the pathogenesis of the asthma (53). This correlate with in vitro experiments in which the presence of high concentrations of PPAR-gamma agonists reduce the eotaxin-dependent eosinophil migration; while low concentrations of this compounds increase migration (54). Interestingly, this effect is mediated by changes in calcium influx, but not by the expression of CCR3 or phosphorylation of p38 or ERK (55).

As mentioned for eosinophils, PPAR-gamma expression in dendritic cells have been related to allergic responses. Particularly, PPAR-gamma deficient CD11b+ dendritic cells fail to migrate to the lung draining lymph nodes and therefore a reduced capacity to polarize naïve CD4+ T cells toward a Th2 phenotype (56, 57). Interestingly, Tuna and collaborators mention that the effect of PPAR-gamma activation depend on the tissue microenvironment. While in the lung it may induce a mucosal phenotype in mDCs and that loss of PPAR-gamma promotes an inflammatory phenotype (determined by the ability of BM-DC to polarize CD4 T cells toward iTregs and to induce CCR9 – a mucosal homing receptor- on T cells); deficiency of this receptor showed no change in the frequency or phenotype of mDC in the colon. This suggest that the intestinal microenvironment can maintain the mucosal DC phenotype of via PPAR-gamma-independent mechanisms (58).

T cells express PPAR-gamma 1, but not PPAR-gamma 2 and it is upregulated on activated T cells. Interestingly, its activation through agonists (such as 15-deoxy-Δ12, 14-prostaglandin J2 or troglitazone) inhibits PMA- mediated proliferation and induce a decrement in cell viability (59). This effect was also described in human and B cell lines. Incubation of B lineage cells with PPAR-gamma against but not with PPAR-alpha agonists, induce apoptosis as demonstrated by the increase of AnnexinV staining and the evaluation of DNA fragmentation in TUNEL assays (60).

Because of its function as an environmental sensor, AhR is widely expressed in the body and the immune system. However, it is highly expressed at the barrier sites such as the skin, lung and gut. In the immune system, the expression of AhR has been described in dendritic cells, B cells and some subpopulations of T cells such as Th17, and in lesser amount, in regulatory T cells (61).

In the T cell subpopulations, AhR is highly expressed in Th17 and in a lesser amount in regulatory T cells (Tregs), but it has not been found in naïve, Th1 or Th2 cells (41). The role of this receptor in Th17 cells remains mostly elusive, but some studies suggest that AhR interacts with NF-κB and STAT1, inhibiting those proinflammatory pathways (62) and promoting the Th17 phenotype by enabling cells to produce high levels of IL-17 and IL-22 (63). Regarding B cells, it has been reported that AhR is expressed in several subpopulations, although at different levels (41). AhR role in B cell function is still not well understood, but an in vitro study reported B cell maturation impairment and antibody production suppression when LPS stimulation occurred in presence of a potent AhR ligand (64). AhR is also expressed by antigen presenting cells, such as dendritic cells, whose differentiation and maturation can be affected by AhR activation. Furthermore, activation of AhR decreases the expression of class II MHC, as well as costimulatory molecules, promoting regulatory T cells differentiation (65).

3. PHTHALATES AS IMMUNOMODULATORS
3.1. In vitro studies

As summarized in Table 1, most studies about the impact of phthalates in the immune system are developed in murine macrophages either cell lines or in primary cultures of peritoneal exudates. In these cells, it was found that some phthalates such as DEHP, DEP and MEHP promoted a pro-inflammatory state characterized by an increase in the expression and secretion of IL-1, IL-6, TNF-alpha, and the chemokine CXCL1 and ROS (66). Other studies found that the secretion of IL-6, IL-10 and the chemokine CXCL8 was enhanced, while TNF-alpha was impaired by DEP and DnBP presence during stimulation (67). They are also capable of increase the phagocytic index of rabbit lung macrophages, which is accompanied by an increment in the release of lysosomal hydrolases (68). However, Shertzer found that even when the phagocytic index after the infection with S. aureous was increased, there was a reduction in the rate of pathogen destruction; which may impact as a deficient response in subsequent exposures (69). On the contrary, other phthalates such as the DBP, BzBP, DEP and DPrP inhibit the production of IL-1beta, IL-6 and IFN-beta (70, 71), apparently by the inactivation of the IFN-beta promoter (70).

Table 1 In vitro studies in animal cells
Country Cellular type Immune system component Phthalate Findings Reference
USA Primary culture of rat testicle germ cell Macrophages, IL-6, TNF-alpha and CXCL1 DEHP and DEP (100 uM up to 7 days) The expression of proinflammatory cytokines (IL-6, TNF-alpha and CXCL1) are increased (66)
USA Rabbit alveolar macrophages Macrophages DEHP (2.5 uL per ml of serum) Significant stimulation of phagocytic activity in macrophages pre-exposed to DEHP as well as release of lysosomal hydrolases. (68)
USA Rabbit lung macrophages Macrophages DEHP (2.5 uL per ml of serum) The cultures previously exposed to DEHP, after infection with S. Aureus, show an increase in the phagocytic index and a decrease between 2 and 14 times of the rate of destruction of the pathogen. (69)
Japan RAW 264 IFN-beta BBP, DBP, DOP, DEP, DHP, DPP and DPrP (50-200 uM) Inhibition of IL-1beta and IL-6. Only DEP and DPrP inhibit the LPS-induced activation of IFN-beta promoter. (70)
Japan Mice (abdominal macrophages) and RAW 264 cell line TNF-alpha BBP (10 ng/mL) Inhibits the production of TNF-alpha in RAW 264. In abdominal macrophages no response was seen. (71)
Japan Bone marrow-derived dendritic cells (BMDC) and splenocytes from mouse Dendritic cells, splenocytes, IL-4 and IFN-gamma DEHP (0.1-10 uM) Increases BMDC differentiation but not the activation. Higher IL-4 and IFN-gamma expression on T lymphocytes in response to BMDC differentiation. Increased Th2 response in splenocytes. (72)
China Primary cultures of mouse splenocytes IL-4, IL-6 and IFN-gamma DEHP (1-50 uM) and MEHP (5-80 uM) DEHP: Increases IL-4 and IFN-γ, decreases the Th1/Th2 ratio (IFN-gamma / IL-4). MEHP: decreases the Th1/Th2 ratio (IFN-gamma / IL-6. (73)
Korea Primary cultures of mouse splenocytes B lymphocytes DEHP (100 uM) Increases B cell proliferation by inducing Ca2+ flux, PKC, ERK1/2 and NF-kappa B activation. (74)
Japan Mouse thymocytes and splenocytes Thymocytes (IL-3, IL-4 and IFN-gamma) and splenocytes DEHP (1 nM-1uM) Stimulates the proliferative response in thymocytes and splenocytes. Increases the production of IL-3 in thymocytes, IL-4 and IFN-gamma in Con A-treated thymocytes (75)
USA Mouse bone marrow-derived B cells and BU-11 cell line B lymphocytes MEHP (10-200 uM) Induces apoptosis and suppresses B cell proliferation in developing bone marrow (76)
USA PD31 cell line (murine pre-B cells) and primary cultures of rainbow trout kidney B cells, Plasmatic cells and IgM DEHP (1-256 uM) DEHP: Inhibits B cells proliferation and reduces the number of plasmatic cells secretory of IgM. Accelerates the terminal differentiation of B cells (77)
Japan Rat basophils RBL-2H3 Mast cells DBP, DIBP and DEHP (50-500 uM) Without Ag stimulation, the phthalates don´t cause a significant increase in degranulation. In cultures sensitized with Anti IgE, the three induce degranulation: DIBP > DBP > DEHP. (78)
Korea RAW 264.7 cell line (murine macrophages) iNOS, TNF-α and IL-1beta DCHP, DEP and DBP (2-200 uM) DCHP, DEP, DBP: changes in iNOS mRNA expression. DBP: Suppression of TNF-alpha and IL-1beta transcription (80)
Korea RAW 264.7 Macrophages, iNOS BBP (2-200 uM) Reduces the production of NO via regulation of iNOS mRNA expression (81)
China Macrophages of murine peritoneal exudate IL-1β, IL-6, IL-10, IL-12 and TNF-alpha DBP (1-100 uM) Inhibition of IL-1beta, IL-6 and TNF-alpha production. No change in IL-10 and IL-12 synthesis (82)
Norway (RAW 264.7 TNF-alpha MEHP (0.3-1 mM) Increases TNF-alpha and changes the cellular morphology of macrophages to a more elongated form, reminiscent of differentiated M2 macrophages (83)
Japan Peripheral blood mononuclear cells (PBMC) of mice Mature and immature DCs DEHP and MEHP (0.1-10 uM) DEHP inhibits the expression of DC differentiation and maturation markers. MEHP has no effect. (84)
Norway Primary cultures of rat alveolar macrophages TNF-alpha, ROS, CXCL2 and leukotriene B4 (LTB4) MEHP (0.1 – 1 mM) MEHP increases the production TNF-alpha, ROS, CXCL2 and (LTB4) (85)
Finland RAW 264.7 ROS production and cell death DEHP, DBP, BBP and DIBP (1 nM – 1 mM) Increased ROS production. Only DBP and DBIP caused cell death due to necrosis at their highest doses (effect>with DIBP). (86)
USA primary cultures of murine cells IL-4, IFN-gamma and IgE DEHP and DINP (0.1–10 uM) By themselves they don´t induce the production of IL-4 and IFN-gamma in mouse lymph nodes. In stimulated cells with PMA or metalloproteins an increase in IL-4 is observed. (87)
UK Human and rat neutrophils CD11b (lymphocyte activation marker) DEHP (1-3 mg/L) DEHP initiates inflammatory response in human and rat neutrophils (by expression of CD11b) (88)
Japan Carp leukocytes Leukocytes and kidney macrophages DEHP and DBP (1–1000nM) Both regulate the function of carp phagocytic cells (89)
UK Rat peritoneal leukocytes Leukocytes DEHP, MEHP and DBP (1–100 ug/mL) Inhibition of the formation of prostaglandins and leukotrienes (PGE-2 y LTB4) in leukocytes. (90)

In T cells, phthalates seem to promote the polarization to the Th2 phenotype. Hansen et al. showed that 100 uM of DEP and DnBP were capable to suppress the secretion of IL-2 and IL-4, TNF-alpha and IFN-gamma, with no impact in cell viability. It was found that DEHP stimulation might also increase the expression of TCR and CD3 molecules and antigen-induced proliferation (72). In total splenocyte cultures, DEHP significantly increased IL-4 and IFN-gamma concentrations, while MEHP increased IL-6 in supernatants. Interestingly, the effect of DEHP was mediated by the activation of the Ca2+/CaN/NF-AT signaling pathway, which was not observed with MEHP. These data suggest that both phthalates may induce the same outcome but by altering different cellular mechanisms (73). DEHP also induces B cell proliferation; activate Ca2+ flux, membrane PKC translocation, ERK1/2 phosphorylation and NF-kappa B activation after 2 hrs (74).

In murine thymocytes, DEHP also increases B cell proliferation as demonstrated by Yamashita et al. (75). In contrast, MEHP was found to suppress B cell ((3)H) thymidine incorporation in both bone marrow-derived B cells and the B cell line BU-11 at doses of 25 or 100 uM. This was explained by an increase in B-cell apoptosis at high doses and an apparent arrest without death at low doses (76).

Interestingly, these alterations are also seen in other organisms such as the rainbow trout, in which, in vitro DEHP exposition of B cells inhibited in a dose-dependent way the proliferation of these cells. Also, authors found that there were changes in the differentiation that generates deficient plasmablast expansion and a reduced number of IgM-secreting plasma cells (77).

Regarding to other immune cell types, there is also evidence in which the effect of some phthalates modifies the function of bone marrow-derived dendritic cells (BM-DC) and basophils. In DC’s, the presence of 10 uM of DEHP during the differentiation of BM-derived macrophages increased the expression of maturation markers such as MHC-II, CD80/86, CD11c and DEC205. It also increased their capacity to induce secretion of IFN-gamma, IL-4 and IL-10 by T cells. This effect was also observed when DC’s were differentiated in normal conditions and DEHP was added during the activation. Interestingly, at 100 uM of DEHP during activation, the percentage of maturation/activations markers were reduced and their antigen presenting activity is not affected as observed by the reduced proliferation of co-cultured T cells. Authors suggest that the enhancement on these markers and function might contribute to the aggravating effect of DEHP on allergic disorders, which also may be useful for the treatment in this kind of pathologies (72).

Finally, since air pollutants have been related to airway diseases, the study of immune cell populations related to allergies has gained relevance. The RBL-2H3 cells, a basophil-derived cell line, which has been described also as a mast cell model, was exposed to different phthalates -DPB, DiPB and DEHP- at concentrations of 50-500 uM. Authors described that in the absence of antigen stimulation, none of those compounds induce β-hexosaminidase release. However, when cells were pre-sensitized and then stimulated with an antigen, all of the phthalates increased the degranulation of these cells (78). Another report found an interest correlation with the number of carbon atoms present in the phthalates and the histamine-release of human basophils. In this work, PMBC’s were incubated in the presence different phthalates and the stimulated with any of the following stimuli: anti-IgE, the bacterial derived peptide fMLP, a calcium Ionophore or an allergen (cat hair extract). As described before, there was no effect on the histamine release in the absence of activation. Interestingly, only the phthalates and monophthalates with an 8-carbon atoms alkyl chain length were the stronger histamine-release potentiators, when there were no differences for the 4-, 9- or 10-carbon chain (79).

Some studies have also assessed the effect of phthalates on human cell lines and primary cultures (Table 2). As in the former studies, heterogeneity in exposure times and concentrations is observed, but some general features can be distinguished. It is interesting to note that, in general terms, a more reactive or proinflammatory response is observed in human cells in vitro exposed to different phthalates, regardless of the heterogeneity of models and exposure schemes.

Table 2 In vitro studies with human cells
Country Cellular type Immune system component Phthalate Findings Reference
Denmark Monocytes/Macrophages and T cell (peripheral blood mononuclear cells) Monocytes/Macrophages DEP, DnBP and MnBP (Multiple concentrations) DEP and DnBP increase IL-6, IL-10 and CXCL8. They decrease the synthesis of TNF-α. T cells: decrease in IL-2, IL-4, TNF-alpha and IFN-gamma expression. (67)
Denmark THP-1 and peripheral blood mononuclear cells (PBMC) of allergic and healthy individuals THP-1: IL-1β, IL-6, IL-12α PBMC: IL-4, IL-5 and IFN-alpha MBUP, MBEP, MEHP, MOP, MINP and MIDP. (0.2-200 ug/mL) Mono-phthalates don´t increase the cytokine production in TPH-1 alone but increase it upon LPS stimulation. There seems to be a relationship between the length of the side chains and the toxicity. MBUP tends to increase the levels of IL-4 in PBMC of allergic patients (79)
UK Human and rat neutrophils CD11b (lymphocyte activation marker) DEHP (1-3 mg/mL) DEHP initiates inflammatory response in neutrophils (by expression of CD11b) of humans and rats (88)
Slovak Republic THP-1 TNF-alpha, IL-1beta, IL-8 DiP (0.2-10 uM) At high doses (10 uM), increases IL-8 secretion (91)
France THP-1cell line (differentiated macrophages) IL-1β, TNF-α and IL-8 (secretion of cytokines induced by LPS) DEHP and DBP (0.001-10 uM) DBP: increased expression of TNF-alpha (92)
USA THP-1 and placental macrophages (PM) COX-2 and prostaglandin E2 MEHP (10-180 uM) MEHP: Increases the release of PGE2 via increased expression of COX2 (93)
Portugal Peripheral blood-derived M1 and M2 macrophages IL-1beta, IL-6 and IL-10 DEHP and DBP (1 uM) M1 macrophages: DEHP—>Increase the expression of IL-1beta, IL-10 and decrease IL-6. DBP—> Increase the expression of IL-1β is not affected. IL-10 and IL-6 the same DEHP. M2 macrophages: DEHP—>IL-1beta is not affected. IL-10 and IL-6 decrease. DBP—>Increase the expression of IL-1beta. IL-10 and IL-6 decrease. (94)
Japan Peripheral blood leukocytes Granulocytes DEHP (43.4 ug/cm2 in blood storage bags) Dose-dependent reduction of chemotaxis and bactericidal. Cell count and phagocytic activity aren´t modified. (95)
Italy Primary cultures of granulocytes isolated from blood Chemotherapy and ROS production DEHP (0.0002-0.0024 %) Stimulates chemotaxis and ROS production (96)
Puerto Rico TK6 cell line (Lymphoblasts) Mitochondrial membrane permeability, ROS, Caspases 3 y 7 DEHP and MEHP (10-500 uM) Both compounds affect the viability of lymphoblasts by increasing the permeability of the mitochondrial membrane, generating ROS and activating Caspases 3 and 7. (97)
Norway Peripheral blood mononuclear leukocytes Leukocytes TBEP and DEHP (0.1 uM-0.1M) Affects the binding of beta-adrenergic ligand in mononuclear leukocytes (interacts with specific and nonspecific binding sites). (98)
Germany Peripheral blood leukocytes Leukocytes DBP and DiBP (354 mM) Genotoxicity: DiBP >DEHP (99)
Germany Peripheral blood lymphocytes Lymphocytes MEHP (0.1-2.5 mM) Genotoxicity (100)
USA Primary cultures of umbilical cord blood IL-1beta, IL-8, IL-6 and VEGF MEHP (100–500 uM) Inhibits neutrophil apoptosis (neonatal cells are more sensitive than adults). In adults and neonates increases the production of VEGF and IL-1beta, only increases IL-8 in adult neutrophils. IL-6 is not affected. (101)
Korea HMC-1 cell line (mast cells) TNF-alpha and IL-4 DEHP (100 uM) Increases the expression of TNF-alpha and IL-4 (102)
Japan THP-1 TNF-alpha, IL-1beta, IL-8 and IL-6 DEHP (200uM) Increases TNF-alpha, IL-1b, IL-8 and IL-6 secretion. Increases the mRNA expression levels of IL-8, CXCL1, CXCL2, CXCL3, CXCL6, CCL3, MMP10, CSF2, TNF-a, IL-1beta and IL-6. Induces p65 NF-kappa B translocation to the nucleus. (103)
Taiwan pDCs (Plasmacytoid dendritic cells), serum IFN-alpha, IFN-beta, IFN-gamma and IL-13 DEHP and BBP (1–100 nM) DEHP and BBP: suppress the expression of IFN-alpha, and IFN-beta. BBP: Suppresses the expression of IFN-gamma but increases the IL-13 (for T cell CD4+) (104)
Taiwan MDA-MB-231 cell line (adenocarcinoma) mdDCs and TADCs (CXCL1) BBP (1 and 10 uM) BBP stimulates the production of CXCL1 by the TADCs (105)

Studies with THP, a macrophage cell line seem to concur in that phthalates and their metabolites increase the cytokine and chemokine production (79, 91, 92). Tetz and cols. Also observed a proinflammatory effect of phthalate exposure, with activation of COX pathway and increased prostaglandin production (93). An interesting study by Teixeira noted that the effect of phthalate exposure showed differential features, depending on the macrophage phenotype, either M1 or M2 (94).

One of the first studies that documented the effect of phthalates on immune cells was derived from the concern of donated blood being storage in plastic bags; this study found decreased chemotaxis and bactericidal capacity of neutrophils (95). However, more recent studies using human granulocytes have found rather an increase in chemotaxis, IL-8 (96, 97), and promotion of inflammatory response (88).

Mast cells are of particular interest, given their role in allergic disease. Lee and cols. Reported an increase in IFN-gamma and IL-4 production upon DEHP exposure, which is linked to a more reactive phenotype. However, more studies regarding mast cells differentiation, migration and function are desirable.

3.2. In vivo studies

Given the complex multidirectional regulation that occurs between immune and endocrine system, it is clear that, although they offer valuable information, in vitro studies don´t fully represent what occurs in a whole organism, and it is why in vivo studies are of such importance. Data from the full literature review is detailed in Table 3.

Table 3 In vivo studies in animal models
Country Organism Immune system component Phthalate Findings Reference
China Mice IgE, IgG1, Tfh (follicular T helper), Plasmatic cells, IL-21 and IL-4 DEHP (30-300 ug/Kg, oral) DEHP acts on Thf cells by increasing its production of IL-21 and IL-4, which leads an increase of IgE and IgG1 secretion. (106)
Denmark WT and PPAR-α-deficient mice IgG1, IgE and IgG2a DEHP (100 ug together with OVA, i.p.) It induces increased levels of IgG and IgG2a in both strains of mice (107)
Denmark Mice IgG1, IgE and IgG2a DEHP (0.022-13 mg/m3, aerosol, with OVA, 5 days/week, 2 weeks) DEHP is capable of increasing serum IgG1 levels and inflammatory cell levels in the lungs but only at very high concentrations (108)
Denmark Mice IgG1, IgE, IgG2a, eosinophils and lymphocytes (serum, bronchoalveolar fluid and draining lymph nodes) MEHP (0.03-0.4 mg/m3, aerosol, with OVA, 5 days/week, 2 weeks) Increase in IgG1. > number of eosinophils and lymphocytes in bronchoalveolar fluid (109)
Korea Mouse Cytokines, immunoglubulines, lung inflammation DINP (50 mg/Kg, i.p. on days 0,3 and 10) DINP suppresses Th1 polarization and promotes Th2. In asthmatic mice, DINP increases IL-4, IL-5, IgE, IgG1, decreases IgG2a and IFN-gamma (in serum) and induces lung inflammation s (Caspases 1 and 3 increased) (110)
Sweden Mice IgE, IL-4, IFN-gamma DEHP (30-3000 ug/kg/day, oral, 52 days. In sensitized mice, serum IgE increases in a dose-dependent manner. IL-4 in lung tissue increases in sensitized mice exposed to the highest DEHP dose (111)
USA Salmon B lymphocytes DEHP, DMP, DBP, BBP, DEP, and DnOP (from plastic-polluted lake) DEHP accelerates the differentiation towards the plasmatic cells (secretory of Ab) (112)
China Mice IgE, IFN-gamma (Th1), IL-4 (Th2), IL-17A (Th17), TNF-alpha, IL-5 and IL-13 DBP (0.4-40 mg/Kg/day, 40 days) In lesions due to atopic dermatitis, the exposure to DBP increases Th2 and Th17 cytokines (113)
Japan Mice with induced allergic rhinitis IL-5, IL-6, IL-12 and IL-13 DEHP (0.0004-0.16 ug, nasal instillation) Increased levels of nasal IL-13 in mice treated with the allergen +DEHP (114)
Japan Mice CCL17, CCL22, IL-4 DINP (0.15-150 mg/Kg/day, i.p., days -5, 2, 9 and 16) DINP increases the production of CCL17, CCL22 and IL-4 in induced atopic dermatitis (115)
Japan Mice IL-4 DBP and DPP (80 ul, together with acetone, as dermal vehicle) Increased FITC+ DC in lymph nodes. Increased IL-4 in draining lymph nodes after skin sensitization. (116)
Japan Mice with induced contact sensitivity TSLP (Th2 response cytokine) and IL-4 DBP (Mixed 1:1 with acetone, used as vehicle, dermally administered) Increases the expression of TSLP in sensitized skin. DBP is essential for sensitization with FITC. DBP induces migration of Ag presenting cells (117)
USA Mice TSLP (cytokine related to IL-7) produced by epithelial cells DBP (Mixed 1:1 with acetone, used as vehicle, dermally administered) The epicutaneous application of DBP induced the expression of TSLP in the skin (118)
Japan Mice Macrophages and DC DBP and DPP (Mixed 1:1 with acetone, used as vehicle, dermally administered) In the presence of these phthalates, a number of antigen-presenting DCs are observed in the lymph nodes (they migrate from the skin, presenting FITC) (119)
Taiwan Mice Macrophages, TNF-alpha, IL-1β and IL-6 DEHP (500 mg/Kg/day, 4 weeks) Promotes inflammation by increasing serum of TNF-alpha, IL-1β and IL-6. It also promotes the infiltration of macrophages in white adipose tissue. (120)
USA Rats Macrophages MEHP (1g/Kg, oral) Induces the infiltration of CD11b+ macrophages in testis of immature peripubertal rats, triggering germ cell apoptosis. No infiltration observed in adult rats. (121)
Japan Mice Lymphocytes, macrophages, IL-10 and IFN-gamma DEHP (0.01-0.1 % in food) Greater presence of lymphocytes and macrophages in testis, along with increased expression of IL-10 and IFN-gamma mRNA. (122)
China Rats IL-1beta (of testicular macrophages) DBP (50, 250 mg/Kg/day, 90 days) Increases the expression of IL-1beta mRNA (and suppresses the production of testosterone) (123)
Japan Mice IFN-gamma DEHP (0.01% in food) Increases the expression of IFN-gamma and the infiltration of leukocytes in autoimmune testicular inflammation. (124)
New Zealand Mice Dendritic cells DBP (Mixed 1:1 with acetone, used as vehicle, dermally administered) Stimulation of Th2 response (125)
China Zebrafish Macrophages DBP (0.02 – 2.0 uM, embryo culture) Decrease in macrophage formation. (126)
Japan Mice with induced peritonitis Macrophages, Neutrophils and IL-5 DEHP 1, 100 ug/mouse, i.p., 2/week, 4 times) Exaggerates inflammation by increasing macrophages and neutrophils and increases IL-5 and eotaxin production. (127)
Netherlands Rats TNF-alpha DEHP (1, 1000 mg/Kg/day, oral, PND 10-50 or 50-90) The exposure to DEHP increases TNF-alpha levels in splenocytes of adult rats (stimulated with LPS) (128)
Japan Epicuta-neously sensitized mice IL-4, IFN-gamma (from draining lymph nodes) DEHP and DBP (Mixed 1:1 with acetone, used as vehicle, dermally administered) DBP: Reduction of IL-4. DEHP: Reduction of IL-4 and increase of IFN-gamma (129)
Japan OVA-immunized mice Splenocytes and thymocytes DEHP (10 uM, drinking water, 4 weeks) It has no effect on the production of Ab. Stimulates the proliferation of splenocytes treated with Con A (mitogen for T cells) and of the thymus cells treated with Con A. (130)
USA Rats Kupffer cells DEHP (1.2 g/Kg, oral) An increase in free radicals in the bile of rats is observed after intragastric administration of DEHP. The increase in free radicals is due to the activation of Kupffer cells. (131)

In general terms, phthalate exposure leads to higher immunoglobulin levels. In some models, an increase of IgG -mainly IgG, has been observed (106–110), while in sensitized animals is the IgE class which is augmented (111). For one of the most studied phthalates – DEHP, the increase in IgE seems to be mediated by the promotion of differentiation of B cells into plasma cells (112) and by the stimulation of follicular T helper cells (Thf), producing higher IL-21 and IL-4 (106).

In diverse pathologies that are characterized by hypersensitivity, such as asthma, dermatitis and rhinitis, phthalate exposure has been shown to increase Th2 (DBP, DPP, DEHP and DINP) and Th17 cytokines (DBP) (110, 113, 114). In animal asthma models, the most studied phthalates -DEHP and DINP- have been shown to increase pulmonary inflammation, antibody production (specially IgE), and IL-4 production. This Th2 polarization has also been observed in an allergic rhinitis model, where DEHP exposure leads to higher IL-13 levels in the nasal epithelium (114). In cutaneous hypersensitivity models, DBP, DPP and DINP have been shown to promote Th2 response, as denoted by a high IL-4 production (115, 116).

Moreover, phthalate exposure also stimulates the production of chemoattractant molecules, such as TSLP, CCL17 and CCL22 (115, 117, 118), which coincides with another common finding: a higher recruitment of antigen presenting cells like dendritic cells and macrophages. This phenomenon, known as the adjuvant effect, has been observed in skin and other tissues, including adipose tissue, and for various phthalates (119, 120).

One of the most typical target organs regarding phthalate toxicity is the male gonad. The effect of phthalates on this tissue has been mainly attributed to pure endocrine mechanisms, but it seems that immunomodulatory mechanisms may as well play an important role. It has been observed that exposure to some phthalates, fundamentally DEHP, causes an increased infiltration of macrophages and other leukocytes in the testis, together with a higher expression of proinflammatory cytokines, such as IL-1β and IFN-gamma, which leads to germinal cell apoptosis and lower testosterone levels (121–124).

3.3. Epidemiological associations

Epidemiological studies, albeit heterogeneous, support the immunomodulatory effects of phthalates, particularly due to prenatal exposure. Data from the full literature review is detailed in Table 4. A 70% higher risk of asthma diagnosis at age 5-11 was observed in children born to women with higher urinary levels of BBzP and DnBP metabolites during their pregnancies (132). Another study, carried out in a Spaniard cohort, also found an association between asthma diagnosis at age 7 and maternal urinary levels of MBzP and the sum of DEHP metabolites (133). In a similar fashion, a study in Taiwan reported an association between high maternal DEHP exposition and asthma incidence, particularly in boys (134). In a French cohort, a study found a correlation between maternal urinary concentrations of DiBP and DiNP metabolites and eczema occurrence; interestingly, this association was stronger for boys (135).

Table 4 Epidemiological studies on phthalate exposure
Country Biological sample Study cohort Immune system component Phthalate Findings Reference
USA Urine of pregnant inner-city women CCCEH (1998-2006) Asthma diagnosis in children between 5 and 11 years old BBzP, DnBP, DEHP, DEP Risk of asthma was > 70% higher among children with maternal prenatal BBzP and DnBP metabolite concentrations in the third versus the first tertile. (132)
Spain Urine of pregnant women INMA (2004-2008) Atopy and asthma in progeny Phthalate metabolites Sum of DEHP metabolites, as well as MBzP during pregnancy increase the risk of asthma at age 7 (133)
Taiwan Urine of pregnant women and children Taiwan Maternal and Infant Cohort Study (2000-2009) Asthma occurrence in progeny Phthalate metabolites Prenatal and postnatal exposure to phthalate was associated with the occurrence of asthma in children, particularly for boys (134)
France Urine of pregnant women EDEN (2003-2006) Development of Eczema in progeny DiBP and DiNP metabolites Occurrence of eczema in childhood is probably influenced by prenatal exposure to phthalates, in boys (135)
Canada Umbilical cord serum MIREC (2008-2011) IgE and IL-33 11 phthalate metabolites Inverse non-linear associations between maternal urinary MCPP levels and elevated of IgE and IL-33 levels (136)
Taiwan Serum for children at 2 and 5 years Ad hoc: from medical centers, regional hospitals, local hospitals, and clinics in Taiwan in 2004 IgE MEP, MEHP, MBP and MBzP Phthalate levels positively correlated with serum IgE levels only in boys. (137)
USA Urine NHANES (2005-2006) Allergic symptoms and sensitization Phthalate metabolites MBzP positively associated with allergic symptoms in adults. Mono-(3-carboxypropyl) phthalate and the sum of DEHP metabolites positively associated with allergic sensitization in adults. Phthalate metabolites inversely associated with asthma and hay fever in children. (138)
Austria Serum of healthy and allergic to home dust individuals 16 healthy and 16 HDM-allergic subjects G-CSF, IL-5 and IL-6 DEHP On allergic subjects, low concentrations of DEHP in the dust increased G-CSF, IL-5 and IL-6, while high concentrations of DEHP decreased G-CSF and IL-6. (139)

At the molecular level, Ashley-Martin didn’t find an association between maternal MCPP (a nonspecific secondary metabolite) and levels of IgE, nor IL-33 in neonates (136). However, another study where MEP, MEHP, MBP and MBzP were measured, found a positive association between maternal urinary levels of those metabolites and IgE levels of children at age 2 and 5 (137). Interestingly, this association was only significative in boys.

Regarding current exposure in adults, the NHANES 2005-2006 study found an association between urinary levels of high molecular weight phthalate metabolites and allergic sensitization (defined by the response to at least 19 specific IgE antigens) (138). Furthermore, MBzP concentration correlated with the presence of allergic symptoms at the moment of the sampling. However, this same study found that in children there was an inverse relationship between urinary concentrations of phthalate metabolites and the incidence of asthma and hay fever. In another study, in adults allergic to dust mite, inhalation of DEHP-containing dust modified the expression of G-CSF, IL-5 and IL-6 in a differential fashion, depending on the concentration (139). At low levels, those cytokines increased, while at high concentrations both G-CSF and IL-6 decreased. It is interesting to note that, similarly to what is observed with some hormones, the dose-response relationship seems to be non-monotonic.

4. CONCLUSIONS

The immune system is regulated by a complex crosstalk between its own soluble mediators and several endocrine pathways. Immune cells not only express endocrine receptors but synthetize and respond to several hormones and other endocrine ligands. In this context, perturbations in the endocrine homeostasis may lead to immune dysregulation. Regarded as EDCs, phthalates and their metabolites interact with estrogen, androgen, PPAR-gamma and AhR pathways, and there is vast in vitro evidence of the effects these compounds exert on immune cells. Furthermore, reports on animal models confirm these phenomena occur in a whole organism setting. However, epidemiological evidence is the one that raises the more concerns, since it points out two relevant aspects of immune dysregulation associated to phthalates: developmental and long-term effects.

Taken together, the evidence suggests more consciousness should arise regarding the use of phthalates, especially in daily use personal care items intended for its use by vulnerable populations. We hope the recognition of the immunoendocrine modulation network and the role it plays from early developmental stages will lead a more integrative perspective in toxicological studies. Regarding phthalates, more studies on immunological imprinting and potential neurological effects are desirable, since this system is also profoundly modulated by endocrine signaling.

5. ACKNOWLEDGMENTS

This work was supported by Grant No. IA202919 from Programa de Apoyo a Proyectos de Investigación e Innovación Tecnológica (PAPIIT) and Grant No. FC-2016-2125 from Consejo Nacional de Ciencia y Tecnología. Margarita Isabel Palacios-Arreola is a postdoctoral fellow from Dirección General de Asuntos del Personal Académico, UNAM.

Abbreviations
Abbreviation Expansion
EDCs

Endocrine disrupting compounds

PPAR-gamma

Peroxisome proliferator-activated receptor gamma

AhR

Aryl hidrocarbon receptor

MHC

Major histocompatibility complex

PVC

Polyvinyl chloride

ug

micrograms

Bw

body weight

PM2.5

Fine particulate matter (of less than 2.5 micrometers)

DEHP

Bis-(2-ethylhexyl) phthalate

DBP

Di-n-butyl phthalate

EPA

Environmental Protection Agency

MEHP

Mono(2-ethylhexyl) phthalate

ER

Estrogen receptor

AR

Androgen receptor

EREs

Estrogen response elements

DNA

Deoxyribonucleic acid

eRs

Estrogen receptors

GPER

G protein receptor-coupled estrogen receptor

cAMP

Cyclic adenosine monophosphate

DAG

Diacylglycerol

IP3

Inositol triphosphate

DHT

Dihydrotestosterone

AREs

Androgen response elements

ERK

Extracellular signal-regulated kinase

MAPK

Mitogen-activated protein kinase

RXR

Retinoid X receptor

PPRE

PPAR response element

SMRT

silencing mediator of retinoid and thyroid hormone receptors

nCoR

nuclear receptor corepressor

HDACs

histone deacetylases

CBP

CREB-binding protein

PBP

PPAR-binding protein 4

TCDD

2,3,7,8-tetrachlorodibenzo-p-dioxin

bHLH

basic helix-loop-helix

HSP90

heat-shock protein 90

AIP

Aryl hydrocarbon receptor-interacting protein

ARNT

Aryl hydrocarbon nuclear translocator

DRE

Dioxin response element

CYP1A1

Cytochrome P450, family 1, subfamily A, polypeptide 1

AhRR

AhR receptor

SW1/SNF

SWItch/Sucrose Non-Fermentable remodeling complex

SRC-1

steroid receptor coactivator-1

FAK

focal adhesion kinase

NK

natural killer cells

Th

T helper cells

Treg

Regulatory T cells

HLA

human leukocyte antigen

G-CSF

Granulocyte colony stimulating factor

IL

interleukin

IgE

immunoglobulin E

15-dPGJ2

15-deoxy-Δ12,14-prostaglandin J2

IFN-

Interferon

iNOS

Inducible Nitric oxide synthase

M1/M2

macrophage phenotype 1 or 2

CD

cluster of differentiation

BM-DC

Bone marrow-derived dendritic cells

iTregs

Induced Tregs

CCR9

C-C chemokine receptor type 9

mDC

mature dendritic cells

PMA

phorbol myristate acetate

TUNEL

Terminal deoxynucleotidyl transferase dUTP nick end labeling assay

DEP

Diethyl phthalate

TNF

tumor necrosis factor

CXCL8

C-X-C chemokine ligand 8

DnBP

Di-n-butyl phthalate

BzBP

Benzylbutyl phthalate

DPrP

Dipropyl phthalate

TCR

T cell receptor

Ca

calcium

CaN

calmodulin

NF-AT

Nuclear factor of activated T-cells

PKC

protein kinase C

NF-kappa B

nuclear factor kappa B

DiBP

Diisobutyl phthalate

PBMC

peripheral blood mononuclear cells

fMLP

N-formyl-methionyl-leucyl-phenylalanine

COX

cyclooxygenase

Thf

follicular T helper cells

DPP

D-n-propyl phthalate

DINP

Diisononyl phthalate

TSLP

Thymic stromal lymphopoietin

BBzP

Benzylbutyl phthalate

MBzP

Monobenzyl phthalate

DiNP

Diisononyl phthalate

MCPP

Mono (3-carboxypropyl) phthalate

MEP

Monoetyl phthalate

MBP

Monobutyl phthalate

ConA

concanavalin A

DIBP

Diisobutyl phthalate

cx-MiDP

Carboxi-monoisodecyl phthalate

DAP

Diallyl phthalate

DCHP

Dicyclohexyl phthalate

DHP

Di-n-hexyl phthalate

DiHepP

Diisoheptyl phthalate

DiHP

Diisohexyl phthalate

DiP

Diisononyl phthalate

DiPP

Diisopentyl phthalate

DnOP

Di-n-octyl phthalate

DnPP

Di-n-pentyl phthalate

DPeP

Di-n-pentyl phthalate

DTDP

Ditridecyl phthalate

MBEP

Monobenzyl phthalate

MBUP

Mono-n-butyl phthalte

MIDP

Monoisodecyl phthalate

MiDP

Monoisodecyl phthalate

MINP

Monoisononyl phthalate

MnBP

Mono-n-butyl phthalate

MOP

Mono-n-octyl phthalate

MTDP

Monotridecyl phthalate

ROS

Reactive oxygen species

LTB4

leukotriene B4

PGE-2

prostaglandin 2

STAT

Signal transducer and activator of transcription

TBEP

Tris-(2-butoxyethyl)-phosphate

References
[1]
EngelA BuhrkeT ImberF JesselS SeidelA VölkelW LampenA Agonistic and antagonistic effects of phthalates and their urinary metabolites on the steroid hormone receptors Eralpha, Erbeta, and AR. Toxicol Lett20172775463DOI: 10.1016/j.toxlet.2017.05.028
[2]
CooperGS StroehlaBC The epidemiology of autoimmune diseases. Autoimmun Rev20032119125DOI: 10.1016/S1568-9972(03)00006-5
[3]
BoumanA HeinemanMJ FaasMM Sex hormones and the immune response in humans. Hum Reprod Update200511411423DOI: 10.1093/humupd/dmi008
[4]
Muñoz-CruzS Togno-PierceC Morales-MontorJ Non-Reproductive Effects of Sex Steroids: Their Immunoregulatory Role. Curr Top Med Chem20111117141727DOI: 10.2174/156802611796117630
[5]
NowakK JabłońskaE Ratajczak-WronaW Immunomodulatory effects of synthetic endocrine disrupting chemicals on the development and functions of human immune cells. Environ Int2019125350364DOI: 10.1016/j.envint.2019.01.078
[6]
DutySM CalafatAM SilvaMJ RyanL HauserR Phthalate exposure and reproductive hormones in adult men. Hum Reprod200520604610DOI: 10.1093/humrep/deh656
[7]
WittassekM WiesmüllerGA KochHM EckardR DoblerL MüllerJ AngererJ SchlüterC Internal phthalate exposure over the last two decades - A retrospective human biomonitoring study. Int J Hyg Environ Health2007210319333DOI: 10.1016/j.ijheh.2007.01.037
[8]
KochHM DrexlerH AngererJ An estimation of the daily intake of di(2-ethylhexyl) phthalate (DEHP) and other phthalates in the general population. Int J Hyg Environ Health20032067783DOI: 10.1078/1438-4639-00205
[9]
KochHM WittassekM BrüningT AngererJ HeudorfU Exposure to phthalates in 5-6 years old primary school starters in Germany-A human biomonitoring study and a cumulative risk assessment. Int J Hyg Environ Health2011214188195DOI: 10.1016/j.ijheh.2011.01.009
[10]
MeekerJD SathyanarayanaS SwanSH Phthalates and other additives in plastics: human exposure and associated health outcomes. Philos Trans R Soc B Biol Sci200936420972113DOI: 10.1098/rstb.2008.0268
[11]
VarshavskyJR Morello-FroschR WoodruffTJ ZotaAR Dietary sources of cumulative phthalates exposure among the U. S. general population in NHANES 2005-2014. Environ Int2018115417429DOI: 10.1016/j.envint.2018.02.029
[12]
WittassekM KochHM AngererJ BrüningT Assessing exposure to phthalates - The human biomonitoring approach. Mol Nutr Food Res201155731DOI: 10.1002/mnfr.201000121
[13]
BergéA CladièreM GasperiJ CoursimaultA TassinB MoilleronR Meta-analysis of environmental contamination by phthalates, http://www.ncbi.nlm.nih.gov/pubmed/23917738,2013DOI: 10.1007/s11356-013-1982-5
[14]
JiY WangF ZhangL ShanC BaiZ SunZ LiuL ShenB A comprehensive assessment of human exposure to phthalates from environmental media and food in Tianjin, China. J Hazard Mater2014279133140DOI: 10.1016/j.jhazmat.2014.06.055
[15]
Quintana-BelmaresRO KraisAM EsfahaniBK Rosas-PérezI MucsD López-MarureR BergmanÅ Alfaro-MorenoE Phthalate esters on urban airborne particles: Levels in PM10and PM2.5from Mexico City and theoretical assessment of lung exposure. Environ Res2018161439445DOI: 10.1016/j.envres.2017.11.039
[16]
BerghC TorgripR EmeniusG ÖstmanC Organophosphate and phthalate esters in air and settled dust - a multi-location indoor study. Indoor Air2011216776DOI: 10.1111/j.1600-0668.2010.00684.x
[17]
FrommeH LahrzT PilotyM GebhartH OddoyA RudenH Occurrence of phthalates and musk fragrances in indoor air and dust from apartments and kindergartens in Berlin (Germany). Indoor Air200414188195DOI: 10.1111/j.1600-0668.2004.00223.x
[18]
Diamanti-KandarakisE BourguignonJP GiudiceLC HauserR PrinsGS SotoAM ZoellerRT GoreAC Endocrine-disrupting chemicals: An Endocrine Society scientific statement. Endocr Rev200930293342DOI: 10.1210/er.2009-0002
[19]
ColónI CaroD BourdonyCJ RosarioO Identification of phthalate esters in the serum of young Puerto Rican girls with premature breast development. Environ Health Perspect2000108895900DOI: 10.1289/ehp.00108895
[20]
SwanSH MainKM LiuF StewartSL KruseRL CalafatAM MaoCS RedmonJB TernandCL SullivanS TeagueJL TeamStudy for Future Families Research Decrease in anogenital distance among male infants with prenatal phthalate exposure. Environ Health Perspect200511310561061DOI: 10.1289/ehp.8100
[21]
SwanSH Environmental phthalate exposure in relation to reproductive outcomes and other health endpoints in humans. Environ Res2008108177184DOI: 10.1016/j.envres.2008.08.007
[22]
JoensenUN FrederiksenH JensenMB LauritsenMP OlesenIA LassenTH AnderssonA-M JørgensenN Phthalate Excretion Pattern and Testicular Function: A Study of 881 Healthy Danish Men. Environ Health Perspect201212013971403DOI: 10.1289/ehp.1205113
[23]
FrederiksenH SkakkebækNE AnderssonAM Metabolism of phthalates in humans. Mol Nutr Food Res200751899911DOI: 10.1002/mnfr.200600243
[24]
KwackSJ HanEY ParkJS BaeJY AhnIY LimSK KimDH JangDE ChoiL LimHJ KimTH PatraN ParkKL KimHS LeeBM Comparison of the short-term toxicity of phthalate diesters and monoesters in sprague-dawley male rats. Toxicol Res2010267582DOI: 10.5487/TR.2010.26.1.075
[25]
OkuboT SuzukiT YokoyamaY KanoK KanoI Estimation of estrogenic and anti-estrogenic activities of some phthalate diesters and monoesters by MCF-7 cell proliferation assay in vitro. Biol Pharm Bull20032612191224DOI: 10.1248/bpb.26.1219
[26]
ZhangH ZhangZ NakanishiT WanY HiromoriY NagaseH HuJ Structure-Dependent Activity of Phthalate Esters and Phthalate Monoesters Binding to Human Constitutive Androstane Receptor. Chem Res Toxicol20152811961204DOI: 10.1021/acs.chemrestox.5b00028
[27]
BilityMT ThompsonJT McKeeRH DavidRM ButalaJH HeuvelJP Vanden PetersJM Activation of mouse and human peroxisome proliferator-activated receptors (PPARs) by phthalate monoesters. Toxicol Sci200482170182DOI: 10.1093/toxsci/kfh253
[28]
HeldringN PikeA AnderssonS MatthewsJ ChengG HartmanJ TujagueM StrömA TreuterE WarnerM GustafssonJ-Å Estrogen Receptors: How Do They Signal and What Are Their Targets. Physiol Rev200787905931DOI: 10.1152/physrev.00026.2006
[29]
MatthewsJ GustafssonJA Estrogen signaling: a subtle balance between ER alpha and ER beta. Mol Interv20033281292DOI: 10.1124/mi.3.5.281
[30]
VrtačnikP OstanekB Mencej-BedračS MarcJ The many faces of estrogen signaling. Biochem medica201424329342DOI: 10.11613/BM.2014.035
[31]
LevinER Membrane estrogen receptors signal to determine transcription factor function. Steroids201813214DOI: 10.1016/j.steroids.2017.10.014
[32]
VailG RoepkeTA Membrane-initiated estrogen signaling via Gq-coupled GPCR in the central nervous system. Steroids20191427783DOI: 10.1016/j.steroids.2018.01.010
[33]
Davey RA Grossmann M Androgen Receptor Structure, Function and Biology: From Bench to Bedside. Clin Biochem Rev 2016 37 3 15
[34]
BennettNC GardinerRA HooperJD JohnsonDW GobeGC Molecular cell biology of androgen receptor signalling. Int J Biochem Cell Biol201042813827DOI: 10.1016/j.biocel.2009.11.013
[35]
TontonozP SpiegelmanBM Fat and Beyond: The Diverse Biology of PPARγ. Annu Rev Biochem200877289312DOI: 10.1146/annurev.biochem.77.061307.091829
[36]
PoulsenL la C SiersbækM MandrupS PPARs: Fatty acid sensors controlling metabolism. Semin Cell Dev Biol201223631639DOI: 10.1016/j.semcdb.2012.01.003
[37]
AhmadianM SuhJM HahN LiddleC AtkinsAR DownesM EvansRM Pparγ signaling and metabolism: The good, the bad and the future. Nat Med201319557566DOI: 10.1038/nm.3159
[38]
AmmazzalorsoA MaccalliniC AmoiaP AmorosoR Multitarget PPARγ agonists as innovative modulators of the metabolic syndrome. Eur J Med Chem2019173261273DOI: 10.1016/j.ejmech.2019.04.030
[39]
MaX WangD ZhaoW XuL Deciphering the roles of PPARγ in adipocytes via dynamic change of transcription complex. Front Endocrinol20189473DOI: 10.3389/fendo.2018.00473
[40]
LarigotL JuricekL DairouJ CoumoulX AhR signaling pathways and regulatory functions. Biochim Open2018719DOI: 10.1016/j.biopen.2018.05.001
[41]
StockingerB MeglioP Di GialitakisM DuarteJH The Aryl Hydrocarbon Receptor: Multitasking in the Immune System. Annu Rev Immunol201432403432DOI: 10.1146/annurev-immunol-032713-120245
[42]
KovatsS Estrogen receptors regulate innate immune cells and signaling pathways. Cell Immunol20152946369DOI: 10.1016/j.cellimm.2015.01.018
[43]
PhielKL HendersonRA AdelmanSJ EllosoMM Differential estrogen receptor gene expression in human peripheral blood mononuclear cell populations. Immunol Lett200597107113DOI: 10.1016/j.imlet.2004.10.007
[44]
KhanD AhmedS Ansar The Immune System Is a Natural Target for Estrogen Action: Opposing Effects of Estrogen in Two Prototypical Autoimmune Diseases. Front Immunol20156635DOI: 10.3389/fimmu.2015.00635
[45]
ErlandssonMC OhlssonC Gustafsson CarlstenH Role of oestrogen receptors α and β in immune organ development and in oestrogen-mediated effects on thymus. Immunology20011031725DOI: 10.1046/j.1365-2567.2001.01212.x
[46]
EnglerRJM NelsonMR KloteMM VanRadenMJ HuangCY CoxNJ KlimovA KeitelWA NicholKL CarrWW TreanorJJ Half- vs full-dose trivalent inactivated influenza vaccine (2004-2005): Age, dose, and sex effects on immune responses. Arch Intern Med200816824052414DOI: 10.1001/archinternmed.2008.513
[47]
BuppMR Gubbels JorgensenTN Androgen-Induced Immunosuppression. Front Immunol20189794DOI: 10.3389/fimmu.2018.00794
[48]
LaiJJ LaiKP ZengW ChuangKH AltuwaijriS ChangC Androgen receptor influences on body defense system via modulation of innate and adaptive immune systems: Lessons from conditional AR knockout mice. Am J Pathol201218115041512DOI: 10.1016/j.ajpath.2012.07.008
[49]
CoutoTAP De Paula FalsarellaN MattosCC De De Brandão MattosLC De Total ige plasma levels vary according to gender and age in brazilian patients with allergic rhinitis. Clinics201469740744DOI: 10.6061/clinics/2014(11)06
[50]
SzélesL TöröcsikD NagyL PPARγ in immunity and inflammation: cell types and diseases, Biochim. Biophys. Acta, Mol. Basis Dis2007177110141030DOI: 10.1016/j.bbalip.2007.02.005
[51]
RicoteM LiAC WillsonTM KellyCJ GlassCK The peroxisome proliferator-activated receptor-γ is a negative regulator of macrophage activation. Nature19983917982DOI: 10.1038/34178
[52]
BouhlelMA DerudasB RigamontiE DièvartR BrozekJ HaulonS ZawadzkiC JudeB TorpierG MarxN StaelsB Chinetti-GbaguidiG PPARγ Activation Primes Human Monocytes into Alternative M2 Macrophages with Anti-inflammatory Properties. Cell Metab20076137143DOI: 10.1016/j.cmet.2007.06.010
[53]
KimSR LeeKS ParkHS ParkSJ MinKH JinSM LeeYC Involvement of IL-10 in Peroxisome Proliferator-Activated Receptor γ-Mediated Anti-Inflammatory Response in Asthma. Mol Pharmacol20056815681575DOI: 10.1124/mol.105.017160
[54]
KobayashiY UekiS MahemutiG ChibaT OyamadaH SaitoN KandaA KayabaH ChiharaJ Physiological Levels of 15-Deoxy-Δ 12,14 -Prostaglandin J 2 Prime Eotaxin-Induced Chemotaxis on Human Eosinophils through Peroxisome Proliferator-Activated Receptor-γ Ligation. J Immunol200517557445750DOI: 10.4049/jimmunol.175.9.5744
[55]
SmithSG ImaokaH PuniaN IrshadA JanssenLL SehmiR GauvreauGM The Effect of PPAR Agonists on the Migration of Mature and Immature Eosinophils. PPAR Res2012201218DOI: 10.1155/2012/235231
[56]
NobsSP NataliS PohlmeierL OkreglickaK SchneiderC KurrerM SallustoF KopfM PPARγ in dendritic cells and T cells drives pathogenic type-2 effector responses in lung inflammation. J Exp Med201721430153035DOI: 10.1084/jem.20162069
[57]
HammadH HeerHJ de SoulliéT AngeliV TrotteinF HoogstedenHC LambrechtBN Activation of Peroxisome Proliferator-Activated Receptor-γ in Dendritic Cells Inhibits the Development of Eosinophilic Airway Inflammation in a Mouse Model of Asthma. Am J Pathol2004164263271DOI: 10.1016/S0002-9440(10)63116-1
[58]
TunaH AvdiushkoRG SindhavaVJ WedlundL KaetzelCS KaplanAM BondadaS CohenDA Regulation of the mucosal phenotype in dendritic cells by PPARγ: role of tissue microenvironment. J Leukoc Biol201495471485DOI: 10.1189/jlb.0713408
[59]
ClarkRB Bishop-BaileyD Estrada-HernandezT HlaT PuddingtonL PadulaSJ The Nuclear Receptor PPARγ and Immunoregulation: PPARγ Mediates Inhibition of Helper T Cell Responses. J Immunol200016413641371DOI: 10.4049/jimmunol.164.3.1364
[60]
PadillaJ LeungE PhippsRP Human B lymphocytes and B lymphomas express PPAR-γ and are killed by PPAR-γ agonists. Clin Immunol20021032233DOI: 10.1006/clim.2001.5181
[61]
Gutiérrez-VázquezC QuintanaFJ Regulation of the Immune Response by the Aryl Hydrocarbon Receptor. Immunity2018481933DOI: 10.1016/j.immuni.2017.12.012
[62]
KimuraA NakaT NoharaK Fujii-KuriyamaY KishimotoT Aryl hydrocarbon receptor regulates Stat1 activation and participates in the development of Th17 cells. Proc Natl Acad Sci U S A200810597219726DOI: 10.1073/pnas.0804231105
[63]
VeldhoenM HirotaK WestendorfAM BuerJ DumoutierL RenauldJC StockingerB The aryl hydrocarbon receptor links TH17-cell-mediated autoimmunity to environmental toxins. Nature2008453106109DOI: 10.1038/nature06881
[64]
AbrewKN De KaminskiNE ThomasRS An integrated genomic analysis of aryl hydrocarbon receptor-mediated inhibition of B-cell differentiation. Toxicol Sci2010118454469DOI: 10.1093/toxsci/kfq265
[65]
QuintanaFJ SherrDH Aryl hydrocarbon receptor control of adaptive immunity Pharmacol Rev20136511481161DOI: 10.1124/pr.113.007823
[66]
HarrisS ShubinSP WegnerS NessK Van GreenF HongSW FaustmanEM The presence of macrophages and inflammatory responses in an in vitro testicular co-culture model of male reproductive development enhance relevance to in vivo conditions. Toxicol Vitr201636210215DOI: 10.1016/j.tiv.2016.08.003
[67]
HansenJF BendtzenK BoasM FrederiksenH NielsenCH RasmussenÅK Feldt-RasmussenU Influence of phthalates on cytokine production in monocytes and macrophages: a systematic review of experimental trials. PLoS One201510e0120083DOI: 10.1371/journal.pone.0120083
[68]
BallyMB OpheimDJ ShertzerHG Di-(2-ethylhexyl) phthalate enhances the release of lysosomal enzymes from alveolar macrophages during phagocytosis. Toxicology1980184960DOI: 10.1016/0300-483X(80)90037-2
[69]
ShertzerHG BallyMB OpheimDJ Inhibition of alveolar macrophage killing by di(2-ethylhexyl) phthalate. Arch Environ Contam Toxicol198514605608DOI: 10.1007/BF01055391
[70]
OhnishiT YoshidaT IgarashiA MuroiM TanamotoKI Effects of possible endocrine disruptors on MyD88-independent TLR4 signaling. FEMS Immunol Med Microbiol200852293295DOI: 10.1111/j.1574-695X.2007.00355.x
[71]
HongC-C Shimomura-ShimizuM MuroiM TanamotoK Effect of endocrine disrupting chemicals on lipopolysaccharide-induced tumor necrosis factor-alpha and nitric oxide production by mouse macrophages. Biol Pharm Bull20042711361139DOI: 10.1248/bpb.27.1136
[72]
KoikeE InoueK Ichiro YanagisawaR TakanoH Di-(2-ethylhexyl) phthalate affects immune cells from atopic prone mice in vitro. Toxicology20092595460DOI: 10.1016/j.tox.2009.02.002
[73]
PeiX DuanZ MaM ZhangY GuoL Role of Ca/CaN/NFAT signaling in IL-4 expression by splenic lymphocytes exposed to phthalate (2-ethylhexyl) ester in spleen lymphocytes. Mol Biol Rep20144121292142DOI: 10.1007/s11033-014-3062-4
[74]
OhP-S LimK-T Cudrania tricupidata Bureau (CTB) Glycoprotein Inhibits Proliferation by Di(2-ethylhexyl) phthalate in Primary Splenocytes: Responses in Cell Proliferation Signaling. Immunol Invest201140339355DOI: 10.3109/08820139.2010.546468
[75]
YamashitaU SugiuraT YoshidaY KurodaE Effect of endocrine disrupters on thymocytes in vitro. J UOEH200325161170DOI: 10.7888/juoeh.25.161
[76]
SchlezingerJJ HowardGJ HurstCH EmberleyJK WaxmanDJ WebsterT SherrDH Environmental and Endogenous Peroxisome Proliferator-Activated Receptor γ Agonists Induce Bone Marrow B Cell Growth Arrest and Apoptosis: Interactions between Mono(2-ethylhexyl) phthalate, 9- cis -Retinoic Acid, and 15-Deoxy-Δ 12,14 -prostaglandin J 2 . J Immunol200417331653177DOI: 10.4049/jimmunol.173.5.3165
[77]
MartinsK ApplegateB HagedornB KennishJ ZwolloP Di(2-ethylhexyl) phthalate inhibits B cell proliferation and reduces the abundance of IgM-secreting cells in cultured immune tissues of the rainbow trout. Fish Shellfish Immunol201544332341DOI: 10.1016/j.fsi.2015.02.037
[78]
NakamuraR TeshimaR SawadaJ ichi Effect of dialkyl phthalates on the degranulation and Ca2+ response of RBL-2H3 mast cells. Immunol Lett200280119124DOI: 10.1016/S0165-2478(01)00318-2
[79]
GlueC MillnerA BodtgerU JinquanT PoulsenLK In vitro effects of monophthalates on cytokine expression in the monocytic cell line THP-1 and in peripheral blood mononuclear cells from allergic and non-allergic donors. Toxicol In vitro200216657662DOI: 10.1016/S0887-2333(02)00082-6
[80]
KimHG YeonS min KimKH KimH ParkJ Il KangHJ ChaEJ ParkHD KangHJ ParkTW JeonYH ParkY In ChangKT JungYW Estrogenic Endocrine-Disrupting Chemicals Modulate the Production of Inflammatory Mediators and Cell Viability of Lipopolysaccharide-Stimulated Macrophages. Inflammation201538595605DOI: 10.1007/s10753-014-9966-2
[81]
KimKH YeonSM KimHG ChoiHS KangH ParkHD ParkTW PackSP LeeEH ByunY ChoiSE LeeKS HaUH JungYW Diverse influences of androgen-disrupting chemicals on immune responses mounted by macrophages. Inflammation201437649656DOI: 10.1007/s10753-013-9781-1
[82]
LiL LiHS SongNN ChenHM The immunotoxicity of dibutyl phthalate on the macrophages in mice. Immunopharmacol Immunotoxicol201335272281DOI: 10.3109/08923973.2013.768267
[83]
BøllingA Kocbach OvrevikJ SamuelsenJT HolmeJA RakkestadKE MathisenGH PaulsenRE KorsnesM Suárez BecherR Mono-2-ethylhexylphthalate (MEHP) induces TNF-α release and macrophage differentiation through different signalling pathways in RAW264.7 cells. Toxicol Lett20122094350DOI: 10.1016/j.toxlet.2011.11.016
[84]
ItoT InoueKI NishimuraN TakanoH Phthalate esters modulate the differentiation and maturation of mouse peripheral blood mononuclear cell-derived dendritic cells. J Appl Toxicol201232142148DOI: 10.1002/jat.1652
[85]
RakkestadKE HolmeJA PaulsenRE SchwarzePE BecherR Mono(2-ethylhexyl) phthalate induces both pro- and anti-inflammatory responses in rat alveolar macrophages through crosstalk between p38, the lipoxygenase pathway and PPARα. Inhal Toxicol201022140150DOI: 10.3109/08958370903019885
[86]
NaaralaJ KorpiA Cell death and production of reactive oxygen species by murine macrophages after short term exposure to phthalates. Toxicol Lett2009188157160DOI: 10.1016/j.toxlet.2009.04.001
[87]
LeeMH ParkJ ChungSW KangBY KimSH KimTS Enhancement of interleukin-4 production in activated CD4+ T cells by diphthalate plasticizers via increased NF-AT binding activity. Int Arch Allergy Immunol2004134213222DOI: 10.1159/000078768
[88]
GourlayT SamartzisI StefanouD TaylorK Inflammatory response of rat and human neutrophils exposed to Di-(2-ethyl-hexyl)-phthalate-plasticized polyvinyl chloride. Artificial Organs200327256260DOI: 10.1046/j.1525-1594.2003.07107.x
[89]
WatanukiH GushikenY SakaiM In vitro modulation of common carp (Cyprinus carpio L.) phagocytic cells by Di-n-butyl phthalate and Di-2-ethylhexyl phthalate. Aquat Toxicol200363119126DOI: 10.1016/S0166-445X(02)00172-8
[90]
TavaresIA VineND Phthalic acid esters inhibit arachidonate metabolism by rat peritoneal leucocytes. J Pharm Pharmacol1985376768DOI: 10.1111/j.2042-7158.1985.tb04936.x
[91]
BennasrouneA RojasL FoucaudL GoulaouicS Laval-GillyP FickovaM CouleauN DurandetC HenryS FallaJ Effects of 4-nonylphenol and/or diisononylphthalate on THP-1 cells: Impact of endocrine disruptors on human immune system parameters. Int J Immunopathol Pharmacol201225365376DOI: 10.1177/039463201202500206
[92]
CouleauN FallaJ BeillerotA BattagliaE D'InnocenzoM PlançonS Laval-GillyP BennasrouneA Effects of Endocrine Disruptor Compounds, Alone or in Combination, on Human Macrophage-Like THP-1 Cell Response. PLoS One201510e0131428DOI: 10.1371/journal.pone.0131428
[93]
TetzLM AronoffDM Loch-CarusoR Mono-ethylhexyl phthalate stimulates prostaglandin secretion in human placental macrophages and THP-1 cells. Reprod Biol Endocrinol20151356DOI: 10.1186/s12958-015-0046-8
[94]
TeixeiraD MarquesC PestanaD FariaA NorbertoS CalhauC MonteiroR Effects of xenoestrogens in human M1 and M2 macrophage migration, cytokine release, and estrogen-related signaling pathways. Environ Toxicol20163114961509DOI: 10.1002/tox.22154
[95]
MiyamotoM SasakawaS Effects of Plasticizers and Plastic Bags on Granulocyte Function during Storage. Vox Sang1987531922DOI: 10.1111/j.1423-0410.1987.tb04907.x
[96]
PalleschiS RossiB DianaL SilvestroniL Di(2-ethylhexyl) phthalate stimulates Ca2+ entry, chemotaxis and ROS production in human granulocytes. Toxicol Lett20091875257DOI: 10.1016/j.toxlet.2009.01.031
[97]
Rosado-BerriosCA VélezC ZayasB Mitochondrial permeability and toxicity of diethylhexyl and monoethylhexyl phthalates on TK6 human lymphoblasts cells. Toxicol Vitr20112520102016DOI: 10.1016/j.tiv.2011.08.001
[98]
SagerG LittleC The effect of the plasticizers TBEP (tris-(2-butoxyethyl)-phosphate) and DHEP (DI-(2-ethylhexyl)-phthalate) on β-adrenergic ligand binding to α1-acid glycoprotein and mononuclear leukocytes. Biochem Pharmacol19893825512557DOI: 10.1016/0006-2952(89)90101-9
[99]
KleinsasserNH WallnerBC KastenbauerER WeissacherH HarreusUA Genotoxicity of di-butyl-phthalate and di-iso-butyl-phthalate in human lymphocytes and mucosal cells. Teratog Carcinog Mutagen200121189196DOI: 10.1002/tcm.1007
[100]
KleinsasserNH HarréusUA KastenbauerER WallnerBC SassenAW StaudenmaierR RettenmeierAW Mono(2-ethylhexyl)phthalate exhibits genotoxic effects in human lymphocytes and mucosal cells of the upper aerodigestive tract in the comet assay. Toxicol Lett20041488390DOI: 10.1016/j.toxlet.2003.12.013
[101]
VetranoAM LaskinDL ArcherF SyedK GrayJP LaskinJD NwebubeN WeinbergerB Inflammatory effects of phthalates in neonatal neutrophils. Pediatr Res201068134139DOI: 10.1203/PDR.0b013e3181e5c1f7
[102]
LeeJ OhPS LimKT Allergy-related cytokines (IL-4 and TNF-α) are induced by Di(2-ethylhexyl) phthalate and attenuated by plant-originated glycoprotein (75 kDa) in HMC-1 cells. Environ Toxicol201126364372DOI: 10.1002/tox.20563
[103]
NishiokaJ IwaharaC KawasakiM YoshizakiF NakayamaH TakamoriK OgawaH IwabuchiK Di-(2-ethylhexyl) phthalate induces production of inflammatory molecules in human macrophages. Inflamm Res2012616978DOI: 10.1007/s00011-011-0390-x
[104]
KuoCH HsiehCC KuoHF HuangMY YangSN ChenLC HuangSK HungCH Phthalates suppress type I interferon in human plasmacytoid dendritic cells via epigenetic regulation. Allergy Eur J Allergy Clin Immunol201368870879DOI: 10.1111/all.12162
[105]
HsuYL HungJY TsaiEM WuCY HoYW JianSF YenMC ChangWA HouMF KuoPL Benzyl butyl phthalate increases the chemoresistance to doxorubicin/cyclophosphamide by increasing breast cancer-associated dendritic cell-derived CXCL1/GROaα and S100A8/A9. Oncol Rep20153428892900DOI: 10.3892/or.2015.4307
[106]
HanY WangX ChenG XuG LiuX ZhuW HuP ZhangY ZhuC MiaoJ Di-(2-ethylhexyl) phthalate adjuvantly induces imbalanced humoral immunity in ovalbumin-sensitized BALB/c mice ascribing to T follicular helper cells hyperfunction. Toxicology20143248897DOI: 10.1016/j.tox.2014.07.011
[107]
LarsenST NielsenGD The adjuvant effect of di-(2-ethylhexyl) phthalate is mediated through a PPARα-independent mechanism. Toxicol Lett2007170223228DOI: 10.1016/j.toxlet.2007.03.009
[108]
LarsenST HansenJS HansenEW ClausenPA NielsenGD Airway inflammation and adjuvant effect after repeated airborne exposures to di-(2-ethylhexyl) phthalate and ovalbumin in BALB/c mice. Toxicology2007235119129DOI: 10.1016/j.tox.2007.03.010
[109]
HansenJS LarsenST PoulsenLK NielsenGD Adjuvant effects of inhaled mono-2-ethylhexyl phthalate in BALB/cJ mice. Toxicology20072327988DOI: 10.1016/j.tox.2006.12.011
[110]
HwangYH PaikMJ YeeST Diisononyl phthalate induces asthma via modulation of Th1/Th2 equilibrium. Toxicol Lett20172724959DOI: 10.1016/j.toxlet.2017.03.014
[111]
GuoJ HanB QinL LiB YouH YangJ LiuD WeiC NanbergE BornehagCG YangX Pulmonary toxicity and adjuvant effect of Di-(2-exylhexyl) phthalate in Ovalbumin-immunized BALB/c mice. PLoS One20127e39008DOI: 10.1371/journal.pone.0039008
[112]
MartinsK HagedornB AliS KennishJ ApplegateB LeuM EppL PallisterC ZwolloP Tissue Phthalate Levels Correlate With Changes in Immune Gene Expression in a Population of Juvenile Wild Salmon. Arch Environ Contam Toxicol2016713547DOI: 10.1007/s00244-016-0283-7
[113]
LiJ LiL ZuoH KeC YanB WenH ZhangY YangX T-helper type-2 contact hypersensitivity of balb/c mice aggravated by dibutyl phthalate via long-term dermal exposure. PLoS One20149e87887DOI: 10.1371/journal.pone.0087887
[114]
HeM InoueKI YoshidaS TanakaM TakanoH SunG IchinoseT Effects of airway exposure to di-(2-ethylhexyl) phthalate on allergic rhinitis. Immunopharmacol Immunotoxicol201335390395DOI: 10.3109/08923973.2013.787432
[115]
KoikeE YanagisawaR SadakaneK InoueKI IchinoseT TakanoH Effects of diisononyl phthalate on atopic dermatitis in vivo and immunologic responses in vitro. Environ Health Perspect2010118472478DOI: 10.1289/ehp.0901255
[116]
MaruyamaT ShibaT IizukaH MatsudaT KurohaneK ImaiY Effects of phthalate esters on dendritic cell subsets and interleukin-4 production in fluorescein isothiocyanate-induced contact hypersensitivity. Microbiol Immunol200751321326DOI: 10.1111/j.1348-0421.2007.tb03914.x
[117]
ShigenoT KatakuseM FujitaT MukoyamaY WatanabeH Phthalate ester-induced thymic stromal lymphopoietin mediates allergic dermatitis in mice. Immunology 128 (1 Suppl),2009e849e857DOI: 10.1111/j.1365-2567.2009.03094.x
[118]
LarsonRP ZimmerliSC ComeauMR ItanoA OmoriM IsekiM HauserC ZieglerSF Dibutyl Phthalate-Induced Thymic Stromal Lymphopoietin Is Required for Th2 Contact Hypersensitivity Responses. J Immunol201018429742984DOI: 10.4049/jimmunol.0803478
[119]
ImaiY KondoA IizukaH MaruyamaT KurohaneKK Effects of phthalate esters on the sensitization phase of contact hypersensitivity induced by fluorescein isothiocyanate. Clin Exp Allergy20063614621468DOI: 10.1111/j.1365-2222.2006.02574.x
[120]
ZhaoJF HsiaoSH HsuMH PaoKC KouYR ShyueSK LeeTS Di-(2-ethylhexyl) phthalate accelerates atherosclerosis in apolipoprotein E-deficient mice. Arch Toxicol201690181190DOI: 10.1007/s00204-014-1377-5
[121]
MurphyCJ StermerAR RichburgJH Age- and Species-Dependent Infiltration of Macrophages into the Testis of Rats and Mice Exposed to Mono-(2-Ethylhexyl) Phthalate (MEHP)1. Biol Reprod20149118DOI: 10.1095/biolreprod.113.115527
[122]
KitaokaM HiraiS TerayamaH NaitoM QuN HatayamaN MiyasoH MatsunoY KomiyamaM ItohM MoriC Effects on the local immunity in the testis by exposure to di-(2-ethylhexyl) phthalate (DEHP) in mice. J Reprod Dev201359485490DOI: 10.1262/jrd.2012-180
[123]
ZhengSJ TianHJ CaoJ GaoYQ Exposure to di(n-butyl)phthalate and benzo(a)pyrene alters IL-1β secretion and subset expression of testicular macrophages, resulting in decreased testosterone production in rats. Toxicol Appl Pharmacol20102482837DOI: 10.1016/j.taap.2010.07.008
[124]
HiraiS NaitoM KuramasuM OgawaY TerayamaH QuN HatayamaN HayashiS ItohM Low-dose exposure to di-(2-ethylhexyl) phthalate (DEHP) increases susceptibility to testicular autoimmunity in mice. Reprod Biol201515163171DOI: 10.1016/j.repbio.2015.06.004
[125]
ConnorLM TangSC CognardE OchiaiS HilliganKL OldSI PellefiguesC WhiteRF PatelD SmithAAT EcclesDA LamiableO McConnellMJ RoncheseF Th2 responses are primed by skin dendritic cells with distinct transcriptional profiles. J Exp Med2017214125142DOI: 10.1084/jem.20160470
[126]
XuH DongX ZhangZ YangM WuX LiuH LaoQ LiC Assessment of immunotoxicity of dibutyl phthalate using live zebrafish embryos. Fish Shellfish Immunol201545286292DOI: 10.1016/j.fsi.2015.04.033
[127]
TanakaM InoueKI MomoiT TakanoH In vivo immunoamplifying effects of di-(2-ethylhexyl) phthalate on cytokine response. Immunopharmacol Immunotoxicol201335147150DOI: 10.3109/08923973.2012.733705
[128]
TonkECM VerhoefA GremmerER LoverenH van PiersmaAH Relative sensitivity of developmental and immune parameters in juvenile versus adult male rats after exposure to di(2-ethylhexyl) phthalate. Toxicol Appl Pharmacol20122604857DOI: 10.1016/j.taap.2012.01.018
[129]
MatsudaT KurohaneK ImaiY Di-(2-ethylhexyl) phthalate enhances skin sensitization to isocyanate haptens in mice. Toxicol Lett201019297100DOI: 10.1016/j.toxlet.2009.10.009
[130]
YamashitaU KurodaE YoshidaY SugiuraT Effect of Endocrine Disrupters on Immune Responses In vivo. J UOEH200325365374DOI: 10.7888/juoeh.25.365
[131]
RusynI KadiiskaMB DikalovaA KonoH YinM TsuchiyaK MasonRP PetersJM GonzalezFJ SegalBH HollandSM ThurmanRG Phthalates rapidly increase production of reactive oxygen species in vivo: Role of Kupffer cells. Mol Pharmacol200159744750DOI: 10.1124/mol.59.4.744
[132]
WhyattRM PerzanowskiMS JustAC RundleAG DonohueKM CalafatAM HoepnerLA PereraFP MillerRL Asthma in Inner-City Children at 5-11 Years of Age and Prenatal Exposure to Phthalates: The Columbia Center for Children's Environmental Health Cohort. Environ Health Perspect201412211411146DOI: 10.1289/ehp.1307670
[133]
GasconM CasasM MoralesE ValviD Ballesteros-GómezA LuqueN RubioS MonfortN VenturaR MartínezD SunyerJ VrijheidM Prenatal exposure to bisphenol A and phthalates and childhood respiratory tract infections and allergy. J Allergy Clin Immunol2015135370378DOI: 10.1016/j.jaci.2014.09.030
[134]
KuHY SuPH WenHJ SunHL WangCJ ChenHY JaakkolaJJK WangS-L GroupT Prenatal and Postnatal Exposure to Phthalate Esters and Asthma: A 9-Year Follow-Up Study of a Taiwanese Birth Cohort. PLoS One201510e0123309DOI: 10.1371/journal.pone.0123309
[135]
SoomroMH BaizN PhilippatC VernetC SirouxV MaesanoC Nichole SanyalS SlamaR BornehagC-G Annesi-MaesanoI Prenatal Exposure to Phthalates and the Development of Eczema Phenotypes in Male Children: Results from the EDEN Mother-Child Cohort Study. Environ Health Perspect2018126027002DOI: 10.1289/EHP1829
[136]
Ashley-MartinJ DoddsL LevyAR PlattRW MarshallJS ArbuckleTE Prenatal exposure to phthalates, bisphenol A and perfluoroalkyl substances and cord blood levels of IgE, TSLP and IL-33. Environ Res2015140360368DOI: 10.1016/j.envres.2015.04.010
[137]
WangIJ LinCC LinYJ HsiehWS ChenPC Early life phthalate exposure and atopic disorders in children: A prospective birth cohort study. Environ Int2014624854DOI: 10.1016/j.envint.2013.09.002
[138]
HoppinJA JaramilloR LondonSJ BertelsenRJ SaloPM SandlerDP ZeldinDC Phthalate exposure and allergy in the U. S. population: Results from NHANES 2005-2006. Environ Health Perspect201312111291134DOI: 10.1289/ehp.1206211
[139]
DeutschieT ReiterR ButteW HeinzowB KeckT RiechelmannH A controlled challenge study on Di(2-ethylhexyl) phthalate (DEHP) in house dust and the immune response in human nasal mucosa of allergic subjects. Environ Health Perspect200811614871493DOI: 10.1289/ehp.11474
Share
Back to top