Next Article in Journal
Neuropsychiatric Disorders Due to Limbic Encephalitis: Immunologic Aspect
Next Article in Special Issue
AhR and Cancer: From Gene Profiling to Targeted Therapy
Previous Article in Journal
System, Method and Software for Calculation of a Cannabis Drug Efficiency Index for the Reduction of Inflammation
Previous Article in Special Issue
The Aryl Hydrocarbon Receptor in Energy Balance: The Road from Dioxin-Induced Wasting Syndrome to Combating Obesity with Ahr Ligands
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

How the AHR Became Important in Cancer: The Role of Chronically Active AHR in Cancer Aggression

1
Department of Environmental Health, Boston University School of Public Health, Boston, MA 02118, USA
2
Graduate Program in Genetics and Genomics, Division of Graduate Medical Sciences, Boston University School of Medicine, Boston, MA 02118, USA
3
Department of Pathology and Laboratory Medicine, Boston University School of Medicine, Boston, MA 02118, USA
4
Department of Environmental Health, Boston University, Boston, MA 02118, USA
5
Boston Medical Center, Boston, MA 02118, USA
6
Division of Computational Biomedicine, Department of Medicine, Boston University School of Medicine, Boston, MA 02118, USA
*
Author to whom correspondence should be addressed.
Equal Contributions.
Current address: 72 East Concord Street (R-408a), Boston, MA 02118, USA.
Int. J. Mol. Sci. 2021, 22(1), 387; https://doi.org/10.3390/ijms22010387
Submission received: 8 December 2020 / Revised: 25 December 2020 / Accepted: 28 December 2020 / Published: 31 December 2020

Abstract

:
For decades, the aryl hydrocarbon receptor (AHR) was studied for its role in environmental chemical toxicity i.e., as a quirk of nature and a mediator of unintended consequences of human pollution. During that period, it was not certain that the AHR had a “normal” physiological function. However, the ongoing accumulation of data from an ever-expanding variety of studies on cancer, cancer immunity, autoimmunity, organ development, and other areas bears witness to a staggering array of AHR-controlled normal and pathological activities. The objective of this review is to discuss how the AHR has gone from a likely contributor to genotoxic environmental carcinogen-induced cancer to a master regulator of malignant cell progression and cancer aggression. Particular focus is placed on the association between AHR activity and poor cancer outcomes, feedback loops that control chronic AHR activity in cancer, and the role of chronically active AHR in driving cancer cell invasion, migration, cancer stem cell characteristics, and survival.

1. The First Hints of a Role for the AHR in Cancer: Carcinogenic Environmental AHR Ligands

Using environmental chemicals as probes, a hydroxylase “inducer” was first discovered by Poland and Glover in 1973 in what may have been one of the most important discoveries in toxicology [1,2,3]. By 1974, it was known that mice of disparate genetic backgrounds exhibited different sensitivities to the environmental chemical 2,3,7,8-tetrachlorodibenzo(p)dioxin (TCDD) and that these differences were likely due to polymorphisms in this unidentified “induction” receptor [4,5,6]. With mounting evidence that TCDD is a type 1 carcinogen in animals and humans [7,8] came the realization that the carcinogenicity of at least some environmental chemicals might not require mutagenesis but might be by this postulated receptor. With the biochemical purification in 1988 of the aryl hydrocarbon receptor (AHR) [9], the sequencing in 1991 of the AHR’s highly conserved N-terminal sequence [10], and the cloning of the AHR gene in 1992 [11,12], came a better understanding of the AHR’s environmental ligand reactivity and its contribution to the induction of hydroxylases that generate mutagenic intermediates. In that vein, a variety of environmental chemicals, including polycyclic aromatic hydrocarbons (PAHs) [13,14,15], aromatic amines [16], and non-ortho-substituted planar polychlorinated biphenyls (e.g., PCBs-118, PCB-156, PCB-126) [17,18,19], were shown to act largely through the AHR, in some cases through ligand-induced, AHR-mediated transcriptional up-regulation of cytochrome P450 Phase 1 hydroxylases (CYP1A1, CYP1A2, CYP1B1) and biotransformation of the parent pro-carcinogen into mutagenic epoxide intermediates [20,21,22,23,24,25,26]. As would be predicted from this understanding of AHR activity, mice lacking these hydroxylases were shown to have a lower incidence of malignant lymphomas and other tumors induced, for example, by PAH [27,28,29]. Notably, it could have been worse. Functional changes in the AHR during human evolution resulted in lower reactivity to PAH, relative to non-human primates, and thereby lower sensitivity to toxic PAHs present in smoke while retaining sensitivity to biologically important nontoxic endogenous AHR ligands [30].
These findings provided a ready explanation for the association between the AHR and cancer. That is, nominally “resting” AHR was seen to be activated by environmental chemicals to induce Phase 1 P450 hydroxylases that generate mutagenic intermediates from the pro-carcinogen parent compounds or from some endogenous substrates (e.g., estradiol or polyunsaturated fats [31,32,33]). These intermediates mutate DNA and initiate cancer. As uncomplicated and attractive as that theory was, it still did not account for the carcinogenicity of TCDD, a non-genotoxic AHR ligand, or for the AHR-driven induction of a growing list of genes unrelated to chemical metabolism but clearly related to cancer (see Section 6.1, Section 6.2, Section 6.3 and Section 6.4). In addition, the seminal demonstration that the AHR is highly conserved throughout evolution [34,35,36] argued for some important function(s) for the AHR in normal cell physiology. Colloquially speaking, if starfish, sea urchins, arthropods, nematodes, mollusks, and fish express AHR homologues (as reviewed in [37,38]), then the AHR must be doing something important in mammals. Since many critical cellular functions in mammals were first identified in the setting of cancer in which malignant cells compromise these cellular functions, often by exaggerating them (e.g., growth, survival, migration), it should come as no surprise that some “nominal” AHR functions were first identified in the cancer context.

2. AHR Transcriptional Signaling

As summarized in a comprehensive review on AHR signaling [39], the AHR is a basic Helix-Loop-Helix-Per/ARNT/Sim (bHLH-PAS) family member and its canonical transcriptional AHR signaling pathway began to be defined in the early 1990s (Figure 1). During that period, “latent” AHR was shown to be confined to the cytoplasm and to exist in a complex with pp60 Src [40], Immunophilin-like Ah Receptor-interacting Protein (AIP) (also known as Hepatitis B virus X-associated Protein 2 (XAP2) [41]), two molecules of HSP90 [42], and the HSP90 co-chaperone, p23 [43]. Both HSP90 and p23 protect the AHR from degradation [44]. Furthermore, HSP90 blocks nuclear translocation and represses DNA binding [45]. Ligand binding to the PAS A and PAS B domains exposes a protein kinase C target site, the phosphorylation of which effects conformational changes and subsequent translocation of the complex to the nucleus. In the nucleus, HSP90, p23, and pp60 Src are released and the ligand-bound AHR complexes are released with the aryl hydrocarbon nuclear translocator (ARNT), first described in 1992 [46]. Domains within the AHR responsible for ARNT binding were identified in 1994 [47]. The C’ terminus of the AHR/ARNT heterodimer then binds to promoters bearing Aryl Hydrocarbon Response Elements (AHREs), also known as Xenobiotic Responses Elements (XREs) [48], and recruits a variety of co-activators including NCoA-2 and p/CIP [49]. Most transcriptionally-mediated AHR activities have been attributed to AHR binding to a consensus AHRE, 5′-(C)GCGTG-3′ [50]. However, alternative AHR complexes and corresponding DNA binding sites have been identified and are likely to be important in physiological AHR functions [51]. In this vein, the AHR can bind to the NF-κB subunits RelA [52] and RelB [53] and to KLF6 [54], enabling the AHR complex to bind to non-canonical “alternative” DNA binding sites. The AHR also signals through its binding to other receptors or transcription factors including, but not limited to, the estrogen receptor [55,56], E2F1 [57], Rb [58], and c-Maf [59]. Furthermore, the AHR contributes to intracellular signaling through non-transcriptional pathways. For example, the AHR associates with tissue factor (TF), preventing its degradation and enhancing thrombosis [60,61]. The AHR may also signal though its associated Src kinase [40,62,63,64] or through E3 ubiquitin ligase activity [65,66]. It is likely that this diversity of AHR-associated proteins, along with differential cofactor recruitment [67], is responsible for the extraordinary variety of AHR responses to a galaxy of endogenous, microbial, dietary, and environmental ligands (reviewed in [68]). Indeed, “normal” physiological AHR activity is involved in oocyte development [69,70], oculomotor development [71], blood vessel development [72,73], cardiomyocyte development [74,75,76], hematopoietic stem cell development [77,78], development of the intestinal immune system and the gut epithelial barrier [79], UVB repair responses in skin [80], and the development and function of a variety of immune cells (reviewed in [56,68,81]).

3. Circumstantial Evidence: High AHR Expression in Many Cancers

Suspicions that the AHR plays some key role(s) in cancer arose from pioneering studies demonstrating dramatically increased AHR expression in numerous cancer subtypes, including Hodgkin’s lymphoma and chronic lymphocytic leukemia [82], adult T-cell leukemia [83], and cancers of the breast [52,56,84,85], head and neck [86,87], brain [88,89,90], kidney [91], lung [92,93], pancreas [94], and GI tract [95,96,97,98]. Increased AHR expression is so consistent in some tumor types that it has been proposed as a prognostic marker [91,99]. The demonstration that both malignant cells and stromal or fibroblast-like cells within the tumor microenvironment (TME) express high AHR levels, as first shown in 2000 [85] and later confirmed [100], suggested that the influence of the AHR in cancer may not be limited to acute induction of mutagenic intermediates and that the contribution of the AHR to cancer progression in the TME, as with most things AHR, may be complex and ongoing.

4. Evidence Builds: An Association between Chronic “Constitutive” AHR Activity and Cancer Patient Outcomes

In addition to being hyper-expressed in some malignant cells, the AHR was shown, as early as 2000 [85], to be “constitutively active” in adult T cell leukemias [83] and cancers of the stomach [101,102], liver [103,104], prostate [105], head and neck [86,87,106], breast [100,107,108,109], brain [88,90], and skin [110,111]. The use of the term “constitutive” in this context reflected the field’s former lack of understanding about endogenous ligands in the tumor, and not the absence of chronic production of endogenous ligands (see Section 5.1). As illustrated in Figure 2, this AHR activity is readily identified by nuclear AHR localization in bladder [112], cervical [113], brain [114], pancreatic [115], head and neck squamous [116], lung adeno [117], lung squamous [117] and skin [118] carcinomas. Increased nuclear AHR localization was positively correlated with a higher tumor grade, more poorly differentiated cells, and/or poor prognosis in prostate, oral, and breast cancers [99,105,106,119,120], suggesting that the AHR may be contributing to increasing cancer aggression.
As would be expected, high AHR expression and activity was usually, although not always [87], correlated with up-regulated CYP1A1 and/or CYP1B1 in cancers of the GI tract, bladder, head and neck, and breast [84,106,119,121,122,123]. Studies in the 1990s suggested that activated AHR, in at least some cancers, predominantly induced CYP1B1 rather than CYP1A1 transcription [84,93,124,125,126], an outcome that could reflect the contribution of distinct AHR-associated proteins such as the AHR interacting protein (AIP) [127] or differential recruitment of co-activators. Indeed, AHR-driven CYP1B1 was proposed as a universal tumor marker that could be immunologically targeted with CYP1B1-derived peptide vaccines [128,129,130,131]. While AHR levels frequently correlated with increased tumor aggression [119], AHR levels per se were understood to not necessarily represent levels of AHR activity. To illustrate this point here, we developed a biomarker set of genes regulated by the AHR in cancer by transcriptional profiling triple negative MDA-MB-231 breast cancer cells and triple negative SUM149 inflammatory breast cancer cells after CRISPR/Cas9-mediated AHR deletion (Figure 3a) and correlated the set of genes down-regulated in both knockouts with patient survival. Using a q value (FDR adjusted p value) of ≤0.05 as a cutoff, we demonstrated that 644 genes were significantly down-regulated in both cell types as compared with matched cells transduced with Cas9 without a guide RNA (“Cas9 controls”) (Supplementary Table S1) (Figure 3b). Using this set of AHR-regulated genes as a biomarker set, Gene Set Variation Analysis (GSVA) [132] was used to test the association between the AHR biomarker set and patient survival using multiple TCGA data sets for the eight cancers shown in Figure 2, all of which express nuclear AHR. Higher expression in human cancers of the genes down-regulated by AHR knockout (i.e., those driven by baseline AHR activity) significantly correlated with poorer survival in all eight cancer types (Figure 3c). These new results add support to the hypothesis that increased AHR activity plays an important role in tumor aggression.

5. Regulators of AHR Activity

5.1. An AHR Amplification Loop: A Possible Driver of “Constitutively Active” AHR

The demonstration of chronic AHR activity in a variety of cancers begged the question of what was persistently driving the AHR in the TME. While a limited number of rare AHR polymorphisms were identified, none appeared to play a key role in human cancer [133,134,135]. Historically, therefore, it has been assumed that the malignant cell itself and/or other cells in the TME produce endogenous AHR ligands that drive chronic AHR activity. Over time, various AHR ligands, including several members of the tryptophan metabolite pathway, were shown to be produced by malignant cells. For example, kynurenine (Kyn), a tryptophan metabolite in the dominant kynurenine pathway, was shown to be produced by breast, head and neck, and brain (glioblastoma) cancers at levels sufficient to activate the AHR (e.g., ~90 µM) [88,90,108] (Figure 4, black font). Note that Kyn should be considered to be an AHR ligand as well as a surrogate for downstream metabolites/AHR ligands, such as kynurenic acid [136], xanthurenic acid [108], cinnabarinic acid [137], and trace kynurenine derivatives [138], any of which could serve as the ultimate effector ligands in a given cancer type.
This production of Kyn and other tryptophan-derived AHR ligands is rate-limited by indoleamine 2,3,-diozygenase (IDO) and tryptophan dioxygenase (TDO) [139], hydroxylases that generate N-formyl-Kyn (Figure 4). N-formyl-Kyn is further reduced to Kyn followed by catabolism to kynurenic, xanthurenic, and cinnabarinic acids. IDO and TDO were first shown to be regulated by the AHR in macrophages and dendritic cells [140,141,142] and later in malignant cells [108]. Therefore, chronic AHR activity may be sustained in malignancies by AHR-regulated induction of IDO or TDO and production of Kyn pathway ligands in a positive amplification loop [108,143] (Figure 4). Since stromal cells in the TME may also express nuclear (i.e., active) AHR [85], non-malignant cells may contribute to this amplification loop. That said, it still is not clear what factor(s) primes the pump. It is possible that environmental, microbial, or dietary [30,144,145,146,147] AHR ligands, many of which can be detected in human sera or urine [148,149,150], initiate AHR activation and start a self-perpetuating AHR circuit. This would suggest that even transient exposure to environmental AHR ligands may initiate a domino effect that results in enhanced AHR activity and ends in malignant transformation. Conversely, it is possible that short-term exposure to AHR inhibitors may reset the amplification loop at a lower steady state. Indeed, bistable or multistable biological circuits, defined as positive or negative feedback pathways that can reset at high low, or intermediate activity states after perturbation with inhibitors or activators, have been described and modeled in various biological contexts [151,152,153,154,155]. For example, the level of the Cdc2-cyclinB/Wee1 signaling pathway or the strength of MAPK activity can be stably reset depending on the magnitude of the response to a perturbation [151]. This could have important implications for the consequences of even brief exposures to environmental AHR ligands and, conversely, for the use of AHR inhibitors as cancer therapeutics.

5.2. Negative Regulators of the AHR Amplification Loop

In counterbalance to the AHR-IDO/TDO amplification circuit is a feedback pathway that limits AHR activity (Figure 4, red font). As early as 1999, it was known that the AHR transcriptionally induces its own inhibitory protein, the AHR Repressor (AHRR) [156,157]. Initially thought to work solely by competing with the AHR for its dimerization partner ARNT, the AHRR was subsequently shown to suppress AHR activity without affecting DNA binding [158,159]. Whatever the mechanism through which the AHRR represses AHR activity, low AHRR expression in breast, lung, stomach, cervical, and ovarian cancers (likely mediated by DNA hyper-methylation and gene silencing [160]) has led some to suggest that the AHRR is a “tumor suppressor” [160], a moniker consistent with the description of the AHR as a tumor promoter, at least in some cancers. Notably, low level AHRR expression in breast cancer is associated with poorer survival [100] and ectopic AHRR expression is associated with decreased invasion [107].
It also has been suggested that a second level of negative feedback within the AHR circuit is likely mediated by AHR induction of prototypic target genes such as CYP1A1 and CYP1B1. In 2004, it was shown that pharmacological inhibition of CYP1A1 increased baseline AHR activity in rat hepatoma cells through the inhibition of CYP1-mediated catabolism of endogenous AHR agonist(s) [161]. To illustrate this point in cancer cells here, we generated CYP1B1 knockout SUM149 inflammatory breast cancer cells (Figure 5a) and quantified AHR activity in the presence or absence of AHR agonists using an AHR-driven (pGudLuc) reporter construct. The AHR knockout cells described in Section 4 were used as a positive control. AHR knockout significantly reduced baseline AHR reporter (pGudLuc) activity in naïve and DMSO groups (Figure 5b, first two green bars). CYP1B1 knockout enhanced baseline AHR activity (Figure 5b, first two red bars). Furthermore, AHR activity induced by several AHR ligands, including environmental (B[a]P) and endogenous AHR ligands 6-formylindolo[3,2-b]carbazole (FICZ), Kyn, or xanthurenic acid, was enhanced by CYP1B1 knockout (Figure 5, remaining red bars). These data illustrate the general conclusion in the field that the classic AHR-inducible hydroxylases likely participate in a negative feedback loop in cancer cells.

6. Consequences of Chronic AHR Activity in Cancer

6.1. AHR-Mediated Cell Migration and Invasion

The ability of malignant cells to migrate from a primary site into adjacent tissue and/or vasculature is a major determinant of metastatic potential [163]. One of the earliest studies implicating the AHR in tumor cell migration was published in 2005 and demonstrated that the migratory potential of immortalized mouse mammary fibroblasts was significantly decreased when the AHR was ablated, an effect likely due to the removal of the AHR’s stimulatory effect on the ERK-FAK-Rac-1 pathway [164]. Shortly thereafter, a second study leveraged the effects of environmental AHR ligands by demonstrating the pro-migratory effects of B[a]P and TCDD on breast cancer cells [165]. In a complementary approach, DiNatale et al. demonstrated that AHR inhibitors slowed the migration of oral squamous carcinoma cells [86]. Similar results were obtained in triple negative breast cancers cells, in which migration and anchorage-independent growth was diminished after AHR knockdown [166]. Our laboratory demonstrated that AHR inhibitors or AHR knockout slows triple negative breast cancer and oral squamous carcinoma cell migration [106,108,162]. Furthermore, ectopic IDO expression, excess Kyn, xanthurenic acid, pyocyanin (a bacterial AHR ligand), B[a]P, TCDD, and FICZ accelerate the migration of breast and/or oral cancer cells in an AHR-dependent fashion [106,108,162]. Similar trends have been seen in other cancers. For example, the AHR transcriptionally regulates Memo-1, a gene implicated in colorectal cancer migration [97]. AHR activation with TCDD induces MMP-9 expression and gastric cancer cell invasiveness, an effect likely mediated through a c-Jun-dependent pathway [96]. TCDD and B[a]P also up-regulate MMP-9 in prostate cancer cells [167], while AHR knockdown decreases invasion of prostate cancer cells in matrigel [168].

6.2. AHR-Mediated Epithelial-to-Mesenchymal Transition (EMT) and Metastasis

Epithelial-to-mesenchymal transition (EMT) is a critical process during which epithelial cells lose apicobasal polarity, connective junctions, and the ability to bind to the basal lamina, which collectively leads to migration and metastasis [169]. E-cadherin is a central contributor to an epithelial morphology, although E-cadherin deficiency in and of itself is insufficient to induce metastasis [170]. E-cadherin expression is repressed by Snail family members Snail, Slug, and Twist, and their upregulation, along with intermediate filament proteins such as Vimentin, are markers of metastatic potential [169,171,172].
Perhaps the earliest and most convincing study implicating the AHR in EMT was provided by Brooks and Eltom in 2011 [173]. These investigators demonstrated that retroviral transduction of an AHR plasmid into non-transformed human mammary epithelial cells was sufficient to induce motility, migration, invasion in Matrigel, anchorage-independent growth, and markers of EMT including increased Vimentin and morphologic changes consistent with EMT. Consistent with these results, AHR hyperactivation with FICZ increased Snai1, Twist1, Twist2, and Vim expression and migration in triple negative breast cancer. These genes have two to five consensus AHR binding site sequences (5′-GCGTG-3′) in their promoter regions, suggesting direct transcriptional regulation [162].
High AHR expression also correlates with lymph node metastases and/or poor prognosis in inflammatory breast and esophageal squamous cell carcinomas (ESCC) [119,174]. In the ESCC context, AHR modulation with the partial AHR agonist 3,3′-diindolylmethane not only down-regulated Vimentin and Slug, but also inhibited the RhoA/ROCK1 pathway, which in turn suppressed COX2/PGE2 signaling, prostaglandin E2 production, migration, metastasis and EMT [174,175,176,177]. This appears to be a generalizable metastasis pathway in that both RhoA/ROCK1 and PGE2 have been implicated in lung and endometrial carcinoma metastasis [178,179,180,181].
Data from hepatocellular carcinomas (HCC) show a similar AHR effect, albeit through different signaling pathways. For example, AHR induction with a prototypic environmental AHR ligand and carcinogen, benzo[a]pyrene (BaP), induced long interspersed nuclear element-1 (Line-1) expression through TGF-α signaling [182], a known inducer of EMT and a facilitator of metastasis [100,183]. Given that Line-1 and other retrotransposon elements mobilize throughout the mammalian genome and damage host DNA via mutational insertions, these results suggest a wide-ranging effect of AHR activation on cancer progression to a highly metastatic state.
While efforts to understand how the AHR affects tumor migration, invasion, EMT and metastasis have focused on classic cancer progression-associated genes discussed above, one significant contributor could be one of the most obvious AHR target genes, CYP1B1, especially in hormone-driven cancers. Kwon et al. demonstrated that ectopic CYP1B1 expression enhances Wnt/β-catenin signaling, a driver of EMT [184,185], and increased invasion in MCF10A breast epithelial cells and/or ER+ MCF7 breast cancer cells at least in part by increasing Snai1, Twist1, and Vimentin and decreasing E-cadherin expression [186]. The effector of this EMT gene profile appears to be the transcription factor, Sp1, induced through CYP1B1-mediated estradiol metabolism. Similarly, our data indicate that CY1B1 knockout in ER- breast cancer cells reduces Wnt5b expression and invasion in matrigel (data not shown).
The translational implications of all of these studies on migration and invasion is exemplified by the ability of non-toxic AHR inhibitors to completely block metastasis of cervical (HeLa), TNBC (MDA-MB-231), and OSCC (HSC3) metastasis in a zebrafish model [107], or by AHR knockout to block melanoma metastasis to the lung [187].

6.3. AHR Role in Cancer Stem Cell (CSC) Development

Perhaps the earliest indication that the AHR could be involved in attainment of stem cell qualities came in 2000 in a toxicology paper by Murante and Gasiewicz, which demonstrated that in vivo treatment with TCDD increased the percentage of bone marrow cells expressing phenotypic markers of hematopoietic stem cells (HSCs) [188]. Subsequent studies from the Gasiewicz group and others extended these results by demonstrating that the AHR plays a central role in HSC growth and differentiation [77,78,188,189,190,191,192] and in lineage commitment of bipotential (erythroid/megakaryocyte) stem cells [78]. Similarly, the AHR has been associated with normal embryonic stem cell function [76,193]. In a classic example of how basic toxicology leads to translational outcomes, follow-up studies demonstrated that an AHR inhibitor, Stemregulin-1, expands HSCs in vitro and that these expanded HSC populations shorten the recovery time in high dose chemotherapy-treated, stem cell-rescued cancer patients [194,195].
These studies are of relevance here given the importance of aberrant organ stem cells in cancer. Cancer stem cells (CSCs) are a relatively small population of chemo- and radio-resistant malignant cells that have the ability to self-renew and to generate progenitor cells that form the bulk of a tumor. CSCs commonly over-express normal organ stem cell-associated genes and have an increased propensity to invade, migrate and metastasize. A rapidly enlarging body of evidence implicates the AHR in these processes. For example, in head and neck carcinoma, lung carcinoma, and choriocarcinoma cell lines, the AHR regulates expression of an ABC transporter, ABCG2, which contributes to chemoresistance by exporting drugs out of the cell against a concentration gradient [86,196,197,198]. AHR expression is elevated and nuclear in choriocarcinoma [197], TNBC [162], and oral squamous cell CSCs [106]. Similarly, in oral cancer and triple negative or ER+ breast cancer cells, the AHR was shown to regulate aldehyde dehydrogenase (ALDH) [106,162,199], which, like ABCG2, is associated with chemotherapy export [200]. ALDH also is associated with increased tumor cell invasion, higher tumor grade, and poor survival [201,202]. Further, AHR hyper-activation with the endogenous ligand FICZ [148,203,204] induced migration and invasion-associated (Snai1, Twist1, Twist2, Tgfb1, Vim) and stem cell-associated (Notch1, Notch2, Bmi1, Nanog, Sox2, Dppa3) genes in triple negative ALDHhigh breast cancer CSCs [162]. Promoters from 11 of these genes contain 3–13 consensus AHR binding site sequences and the AHR interacts directly with the Sox2 promoter [162]. Wnt5a/β-catenin signaling also correlates with CSC phenotype and disease progression in inflammatory breast cancers [119].
Functional studies are consistent with the findings on AHR-regulated stem cell genes. Thus, as would be predicted from the studies on ABCG2 and ALDH1, AHR inhibitors increased the sensitivity of ER+ and ER- breast carcinoma cells to Adriamycin and Paclitaxel [162,199], oral cancer cells to Cisplatin [106,199] and choriocarcinoma cells to Methotrexate [197], directly connecting AHR-driven CSCs to cancer treatment outcomes. With regard to these studies, the ability of some of the chemotherapeutics (e.g., Cisplatin) to reduce AHR signaling in and of themselves, may have contributed to the increased effectiveness of the combination of AHR inhibitor and chemotherapeutics [205]. That said, Paclitaxel and Adriamycin have been shown to enhance AHR expression in MDA-MB-231 breast cancer cells [206] and AHR activity in cardiomyocytes [206], results that one would have expected would lead to a reduction in efficacy of a chemotherapeutic plus AHR inhibitor regimen. The fact that the combination proved more and not less effective demonstrates how elusive it is in some contexts to predict the outcomes of AHR manipulation.
CSC-dependent low adherence spheroid formation of choriocarcinoma or breast carcinoma cells in vitro was suppressed with AHR inhibitors or AHR knockdown and increased with AHR agonists (e.g., TCDD) [162,197,199]. Importantly, AHRhigh/ALDHhigh breast cancer CSCs were significantly more efficient at initiating tumors than AHRlow/ALDHlow cells, and AHR knockdown with siRNA significantly reduced tumor-initiating capacity [162], a sine qua non of CSCs [207].
Finally, it is important to note that, in some instances, AHR down-regulation, not up-regulation, characterizes CSC maintenance. For example, decreased AHR signaling contributes to CSC maintenance in human acute lymphocytic leukemia [208]. This type of paradoxical result, elegantly reviewed in Murray et al. [209] and [210], is not unusual in the AHR field and is likely attributable to the ability of a variety of endogenous and exogenous ligands [211] to differentially recruit AHR co-factors in different tissue contexts effecting different outcomes [67]. Indeed, we have specifically addressed this paradox in invasion, migration, and in vivo metastasis assays in head-to-head comparisons between AHR agonists and inhibitors [107]. It was shown that AHR agonists TCDD and/or 3,3′-diindolylmethane and AHR inhibitors CH223191 [212] and CB7993113 [109] inhibited triple negative breast cancer invasion in vitro and metastasis in vivo. The ability of agonists to reduce these measures of tumor aggression is consistent with previous studies, for example, from Safe et al. [213,214,215] and Kolluri et al. [210,216], showing that AHR hyper-activation with agonists can be anti-tumorigenic. While the molecular signal that results in these apparently contradictory results is still unknown, it has been postulated that agonists such as TCDD, 3,3′-diindolylmethane, or Omeprazole [217] induce differential cofactor recruitment by the AHR then those recruited by endogenous ligands in cancer [107,218]. In essence, some agonists may “divert” the AHR from a pro- to an anti-tumorigenic signaling pathway.

6.4. The AHR’s Role in Malignant Cell Apoptosis

In addition to relatively unfettered growth, an increased ability to invade local tissue, and a propensity to migrate from the primary site, aggressive malignant cells possess the ability to survive what, to a normal cell, would be a lethal signal to initiate apoptosis. Relatively recent studies in cancer cells suggest that the AHR may play a role in that resistance to death. The AHR was first implicated in apoptosis control in 1996, with studies demonstrating that activated T cells were more susceptible to Fas-mediated death signals when exposed to TCDD [219]. In 2005, these studies were extended by Park et al., who showed that the AHR potentiates Fas-mediated apoptosis in hepatocytes [220]. Remarkably, attrition of primordial ovarian follicles, an apoptosis-mediated event responsible for normal ovarian germ cell development, was shown to be driven by the AHR, potentially explaining the fertility issues seen in women who smoke [221]. Although in these cases AHR had a pro-apoptotic effect, they demonstrated that apoptotic and AHR signaling pathways could be linked.
In cancer, the AHR appears to do the opposite, i.e., to suppress apoptosis. As noted in Section 6.3, AHR inhibition increases tumor cell susceptibility to chemotherapeutics [106,162,197,199]. As a corollary, AHR activation with TCDD inhibited apoptosis in lymphoma cells in vitro and in vivo through the induction of COX2 and dysregulation of BCL2 [222]. AHR-dependent COX2 induction was also shown to play a role in blocking apoptosis induced by UVB irradiation, Adriamycin, or the dual tyrosine kinase inhibitor Lapatinib in breast cancer cells [223], or in non-transformed, AHR-transfected breast epithelial cells [224]. AHR-mediated resistance to UVB radiation-induced apoptosis may be most relevant in skin photo-carcinogenesis [14,225], where it was shown, in 2013, that AHR desensitized keratinocytes to UVB-induced apoptosis signaling in consort with increased expression of E2F1 and CHK1 [226]. More recently, it was shown that the AHR also suppresses pyrimidine dimer repair in vitro and in vitro and blocks the formation of double strand breaks that lead to apoptosis [227]. Remarkably, AHR knockout mice exhibited 50% fewer UVB-induced cutaneous squamous cell carcinomas than wildtype mice [227]. Finally, Kyn increased expression of anti-apoptotic proteins cIAP-1, cIAP-2, XIAP and Bcl-2 and decreased pro-apoptotic Bax in a pancreatic cancer cell line [228].

7. Caveats: Interspecies Differences

While the studies summarized above begin to reveal commonalities in how the AHR influences carcinogenesis, some important caveats should be kept in mind, not the least of which is the interspecies differences between murine and human models. The murine AHR has approximately a 10-fold higher affinity for TCDD than the human AHR [229], a difference that suggests that humans may have a higher tolerance for TCDD than mice [229,230]. Furthermore, a relative lack of similarity of the carboxy (DNA-binding) terminus between the mouse and human AHRs may result in different co-factor recruitment, specifically LXXLL binding motifs, leading to distinct transcriptional activity [231,232]. Furthermore, using a transgenic model in which mice expressed the human AHR in liver, it was shown that the murine and human AHRs induce different transcriptional responses to a given ligand, including TCDD and endogenous indole compounds [229]. With regard to the latter class of agonists, the human AHR has a higher affinity for indirubin and generates a different transcriptional profile than the murine AHR, suggesting an evolutionary preference for endogenous ligands [30,233]. This species-specific ligand selectivity may explain many of the differences observed in human and murine models.

8. Conclusions

Here we have summarized a variety of studies, all of which rest on a foundation of toxicological research that was designed to define the basic molecular mechanisms through which environmental AHR ligands generate adverse biological responses. Many of the studies, conducted with diverse cancer types, exploited environmental and, more recently, endogenous AHR ligands to untangle the AHR’s role in cancer. The results clearly indicate a complex association between the AHR and several critical cancer features, including increased malignant cell invasion, migration, metastasis, CSC formation, and survival. For the most part, these studies have been conducted on a basic science level. However, the implications of these studies for our understanding of the genesis of many cancers, their prevention, and their treatment are far-reaching. Thus, these studies have lent credence to the argument that: (1) primary cancer prevention can be effected by minimizing exposure to subsets of environmental AHR ligands, (2) cancer interception can be considered prior to full blown malignancy if early markers of AHR activity can be identified (e.g., Figure 3), and (3) several cancers in which AHR levels correlate with poor survival may be treatable with specific AHR inhibitors.

9. Materials and Methods

9.1. Generation of AHR or CYP1B1 Knockout MDA-MB-231 and SUM149 Cell Lines with CRISPR-Cas9 Gene Editing

Human AHR and CYP1B1 knock-out SUM149 and MDA-MB-231 cell lines were created using lentiCRISPR v2 (Addgene no. 52961, Cambridge, MA, USA), which contains Cas9 and a guide RNA cloning site (BsmBI). Two target sequences (5′-CCTACGCCAGTCGCAAGCGG-3′ and 5′-CCGAGCGCGTCCTCATCGCG-3′) for AHR or the two target sequences (5′-TTAGCGGCCAAGGGTCGTTC-3′ and 5′-CCTGCTACTCCTGTCGGTGC-3′) for CYP1B1 knock-out were used. The target sequences are located in the first exon of the AHR and CYP1B1 genes, respectively. Lentivirus particles were generated in the HEK293NT cells by co-transfecting the lentiCRISPR v2, or AHR-lentiCRISPR v2 plasmids, and the packaging plasmids (pLenti-P2A and pLenti-P2B, Cat. # LV003, Applied Biological Materials Inc. Richmond, BC, Canada), using Lipofectamine 2000 (Invitrogen, Grand Island, NY, USA), according to the manufacturer’s instructions. Virus-containing media were collected 72 h later and filtered through a 0.45 μm filter. SUM149 and MDA-MB-231 cells were transduced with lentiviruses in the presence of 5 µg/mL polybrene as described [234]. Forty-eight hours after transduction, cells were selected with 2 µg/mL Puromycin for 10 days. The efficiency of AHR and CYP1B1 knockout was validated by DNA sequencing and immunoblot analyses.

9.2. Western Blotting

Cells were lysed and protein extracted with RIPA (Radio Immune Precipitation Assay) buffer (Boston BioProducts, Ashland, MA, USA). Protein concentrations were quantified with a Bradford protein assay. Equal amounts of protein (30 µg) were subjected to 10% SDS-PAGE and then transferred to a nitrocellulose membrane. Non-specific binding sites were blocked with blocking buffer containing Tris-buffered saline and 0.1% Tween-20 with 5% nonfat milk powder for 1 h at room temperature, and the blot was incubated with 1:1000 dilution AHR- or beta-actin-specific antibody in blocking buffer. AHR antibody was purchased from Cell Signaling. β–Actin antibodies were from Sigma-Aldrich (St. Louis, MO, USA).

9.3. AHR-Driven Reporter Assay

SUM149 cells were co-transfected with the pGudluc reporter plasmid (0.5 µg) (generously provided by Dr. M. Denison, UC, Davis), and CMV-green (0.1 µg) (for normalization) using TransIT-2020 transfection reagent (Mirus, Madison, WI, USA). Transfection medium was replaced after 24 h. The cells were left untreated or dosed with vehicle (DMSO, 0.1% final concentration), 10 uM B[a]P, 0.5 uM FICZ, 100 uM kynurenine, or 100 uM xanthurenic acid and harvested after 24 h in Glo Lysis Buffer (Promega, San Luis Obispo, CA, USA). Luciferase activity was determined with the Bright-Glo Luciferase System according to the manufacturer’s instructions (Promega, Madison, WI, USA). Luminescence and fluorescence were determined using a Synergy2 multifunction plate reader (Bio-Tek, Winooski, VT, USA).

Supplementary Materials

The following are available online at https://www.mdpi.com/1422-0067/22/1/387/s1.

Author Contributions

Conceptualization, Z.W., D.H.S., J.E.K., Computational analysis, S.M., A.F., Western blotting and gene knockouts, Z.W., Drafting, editing, and review of the manuscript, O.N., Z.W., M.S., J.E.K., K.Y., B.L., E.L., K.B., O.N., A.F., S.M., D.H.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by R01 ES029136, R21 ES029624, R01ES025409 and grants from the Find the Cause Breast Cancer Foundation and from the Hahnemann Foundation.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AHRAryl hydrocarbon receptor
AHRRAHR repressor
AIP/XAP2Immunophilin-like AHR-interacting protein/Hepatitis B virus X-Associated Protein 2
ARNTAryl hydrocarbon receptor nuclear translocator
bHLH-PASBasic helix-loop-helix-Per Arnt Sim
FICZ6-formylindolo[3,2-b]carbazole
KynKynurenine
TCDD2,3,7,8-tetrachlorodibenzo-p-dioxin

References

  1. Poland, A.; Glover, E. Chlorinated dibenzo-p-dioxins: Potent inducers of delta-aminolevulinic acid synthetase and aryl hydrocarbon hydroxylase. II. A study of the structure-activity relationship. Mol. Pharmacol. 1973, 9, 736–747. [Google Scholar] [PubMed]
  2. Poland, A.; Glover, E. Studies on the mechanism of toxicity of the chlorinated dibenzo-p-dioxins. Environ. Health Perspect. 1973, 5, 245–251. [Google Scholar] [CrossRef] [PubMed]
  3. Poland, A.; Glover, E. 2,3,7,8-Tetrachlorodibenzo-p-dioxin: A potent inducer of -aminolevulinic acid synthetase. Science 1973, 179, 476–477. [Google Scholar] [CrossRef] [PubMed]
  4. Nebert, D.W.; Gielen, J.E. Genetic regulation of aryl hydrocarbon hydroxylase induction in the mouse. Fed. Proc. USA 1972, 31, 1315–1325. [Google Scholar]
  5. Poland, A.; Glover, E.; Kende, A.S. Stereospecific, high affinity binding of 2,3,7,8-tetrachlorodibenzo-p-dioxin by hepatic cytosol. Evidence that the binding species is receptor for induction of aryl hydrocarbon hydroxylase. J. Biol. Chem. 1976, 251, 4936–4946. [Google Scholar] [PubMed]
  6. Poland, A.; Glover, E.; Robinson, J.; Nebert, D. Genetic expression of aryl hydrocarbon hydroxylase activity: Induction of monooxygenase activites and cytochrome P1-450 formation by 2,3,7,8-tetrachlorodibenzo-p-dioxin in mice genetically “nonresponsive” to other aromatic hydrocarbons. J. Biol. Chem. 1974, 249, 5599–5606. [Google Scholar]
  7. Ayres, S.M.; Webb, K.B.; Evans, R.G.; Mikes, J. Is 2,3,7,8-TCDD (dioxin) a carcinogen for humans? Environ. Health Perspect. 1985, 62, 329–335. [Google Scholar] [CrossRef]
  8. Catalani, S. IARC revision on dioxin and some dioxin-like compounds. G. Ital. Med. Lav. Ergon. 2010, 32, 79–81. [Google Scholar]
  9. Perdew, G.H.; Poland, A. Purification of the Ah receptor from C57BL/6J mouse liver. J. Biol. Chem. 1988, 263, 9848–9852. [Google Scholar]
  10. Bradfield, C.A.; Glover, E.; Poland, A. Purification and N-terminal amino acid sequence of the Ah receptor from the C57BL/6J mouse. Mol. Pharmacol. 1991, 39, 13–19. [Google Scholar]
  11. Burbach, K.M.; Poland, A.; Bradfield, C.A. Cloning of the Ah-receptor cDNA reveals a distinctive ligand-activated transcription factor. Proc. Natl. Acad. Sci. USA 1992, 89, 8185–8189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Ema, M.; Sogawa, K.; Watanabe, N.; Chujoh, Y.; Matsushita, N.; Gotoh, O.; Funae, Y.; Fujii-Kuriyama, Y. cDNA cloning and structure of mouse putative Ah receptor. Biochem. Biophys. Res. Commun. 1992, 184, 246–253. [Google Scholar] [CrossRef]
  13. Shimizu, Y.; Nakatsuru, Y.; Ichinose, M.; Takahashi, Y.; Kume, H.; Mimura, J.; Fujii-Kuriyama, Y.; Ishikawa, T. Benzo[a]pyrene carcinogenicity is lost in mice lacking the aryl hydrocarbon receptor. Proc. Natl. Acad. Sci. USA 2000, 97, 779–782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Hidaka, T.; Fujimura, T.; Aiba, S. Aryl Hydrocarbon Receptor Modulates Carcinogenesis and Maintenance of Skin Cancers. Front. Med. (Lausanne) 2019, 6, 194. [Google Scholar] [CrossRef] [PubMed]
  15. Matsumoto, Y.; Ide, F.; Kishi, R.; Akutagawa, T.; Sakai, S.; Nakamura, M.; Ishikawa, T.; Fujii-Kuriyama, Y.; Nakatsuru, Y. Aryl hydrocarbon receptor plays a significant role in mediating airborne particulate-induced carcinogenesis in mice. Environ. Sci. Technol. 2007, 41, 3775–3780. [Google Scholar] [CrossRef] [PubMed]
  16. Juricek, L.; Bui, L.C.; Busi, F.; Pierre, S.; Guyot, E.; Lamouri, A.; Dupret, J.M.; Barouki, R.; Coumoul, X.; Rodrigues-Lima, F. Activation of the aryl hydrocarbon receptor by carcinogenic aromatic amines and modulatory effects of their N-acetylated metabolites. Arch. Toxicol. 2015, 89, 2403–2412. [Google Scholar] [CrossRef]
  17. Glazer, L.; Hahn, M.E.; Aluru, N. Delayed effects of developmental exposure to low levels of the aryl hydrocarbon receptor agonist 3,3′,4,4′,5-pentachlorobiphenyl (PCB126) on adult zebrafish behavior. Neurotoxicology 2015, 52, 134–143. [Google Scholar] [CrossRef] [Green Version]
  18. Ovesen, J.L.; Schnekenburger, M.; Puga, A. Aryl hydrocarbon receptor ligands of widely different toxic equivalency factors induce similar histone marks in target gene chromatin. Toxicol. Sci. 2011, 121, 123–131. [Google Scholar] [CrossRef] [Green Version]
  19. Zhang, S.; Rowlands, C.; Safe, S. Ligand-dependent interactions of the Ah receptor with coactivators in a mammalian two-hybrid assay. Toxicol. Appl. Pharmacol. 2007, 227, 196–206. [Google Scholar] [CrossRef] [Green Version]
  20. Shimada, T.; Hayes, C.L.; Yamazaki, H.; Amin, S.; Hecht, S.S.; Guengerich, F.P.; Sutter, T.R. Activation of chemically diverse procarcinogens by human cytochrome P-450 1B1. Cancer Res. 1996, 56, 2979–2984. [Google Scholar]
  21. Tang, Y.M.; Wo, Y.Y.; Stewart, J.; Hawkins, A.L.; Griffin, C.A.; Sutter, T.R.; Greenlee, W.F. Isolation and characterization of the human cytochrome P450 CYP1B1 gene. J. Biol. Chem. 1996, 271, 28324–28330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Walker, N.; Lax, S.; Wyde, M.; Crofts, F.; Lucier, C.; Sutter, T. CYP1B1 and CYP1A1 exhibit differential inducibility by TCDD in a Sprague-Dawley rat tumor promotion model. Fund. Appl. Toxicol. 1997, 36, 215. [Google Scholar]
  23. Sutter, T.R.; Tang, Y.M.; Hayes, C.L.; Wo, Y.Y.; Jabs, E.W.; Li, X.; Yin, H.; Cody, C.W.; Greenlee, W.F. Complete cDNA sequence of a human dioxin-inducible mRNA identifies a new gene subfamily of cytochrome P450 that maps to chromosome 2. J. Biol. Chem. 1994, 269, 13092–13099. [Google Scholar] [PubMed]
  24. Crespi, C.L.; Penman, B.W.; Steimel, D.T.; Smith, T.; Yang, C.S.; Sutter, T.R. Development of a human lymphoblastoid cell line constitutively expressing human CYP1B1 cDNA: Substrate specificity with model substrates and promutagens. Mutagenesis 1997, 12, 83–89. [Google Scholar] [CrossRef] [Green Version]
  25. Shimada, T.; Oda, Y.; Gillam, E.M.; Guengerich, F.P.; Inoue, K. Metabolic activation of polycyclic aromatic hydrocarbons and other procarcinogens by cytochromes P450 1A1 and P450 1B1 allelic variants and other human cytochromes P450 in Salmonella typhimurium NM2009. Drug. Metab. Dispos. 2001, 29, 1176–1182. [Google Scholar]
  26. Thier, R.; Bruning, T.; Roos, P.H.; Bolt, H.M. Cytochrome P450 1B1, a new keystone in gene-environment interactions related to human head and neck cancer? Arch. Toxicol. 2002, 76, 249–256. [Google Scholar] [CrossRef]
  27. Buters, J.; Quintanilla-Martinez, L.; Schober, W.; Soballa, V.J.; Hintermair, J.; Wolff, T.; Gonzalez, F.J.; Greim, H. CYP1B1 determines susceptibility to low doses of 7,12-dimethylbenz[a]anthracene-induced ovarian cancers in mice: Correlation of CYP1B1-mediated DNA adducts with carcinogenicity. Carcinogenesis 2003, 24, 327–334. [Google Scholar] [CrossRef]
  28. Buters, J.T.; Mahadevan, B.; Quintanilla-Martinez, L.; Gonzalez, F.J.; Greim, H.; Baird, W.M.; Luch, A. Cytochrome P450 1B1 determines susceptibility to dibenzo[a,l]pyrene-induced tumor formation. Chem. Res. Toxicol. 2002, 15, 1127–1135. [Google Scholar] [CrossRef]
  29. Heidel, S.M.; Holston, K.; Buters, J.T.; Gonzalez, F.J.; Jefcoate, C.R.; Czupyrynski, C.J. Bone marrow stromal cell cytochrome P4501B1 is required for pre-B cell apoptosis induced by 7,12-dimethylbenz[a]anthracene. Mol. Pharmacol. 1999, 56, 1317–1323. [Google Scholar] [CrossRef]
  30. Hubbard, T.D.; Murray, I.A.; Bisson, W.H.; Sullivan, A.P.; Sebastian, A.; Perry, G.H.; Jablonski, N.G.; Perdew, G.H. Divergent Ah Receptor Ligand Selectivity during Hominin Evolution. Mol. Biol. Evol. 2016, 33, 2648–2658. [Google Scholar] [CrossRef] [Green Version]
  31. Embrechts, J.; Lemiere, F.; Van Dongen, W.; Esmans, E.L.; Buytaert, P.; Van Marck, E.; Kockx, M.; Makar, A. Detection of estrogen DNA-adducts in human breast tumor tissue and healthy tissue by combined nano LC-nano ES tandem mass spectrometry. J. Am. Soc. Mass Spectrom. 2003, 14, 482–491. [Google Scholar] [CrossRef] [Green Version]
  32. Markushin, Y.; Zhong, W.; Cavalieri, E.L.; Rogan, E.G.; Small, G.J.; Yeung, E.S.; Jankowiak, R. Spectral characterization of catechol estrogen quinone (CEQ)-derived DNA adducts and their identification in human breast tissue extract. Chem. Res. Toxicol. 2003, 16, 1107–1117. [Google Scholar] [CrossRef] [PubMed]
  33. Huerta-Yepez, S.; Tirado-Rodriguez, A.; Montecillo-Aguado, M.R.; Yang, J.; Hammock, B.D.; Hankinson, O. Aryl Hydrocarbon Receptor-Dependent inductions of omega-3 and omega-6 polyunsaturated fatty acid metabolism act inversely on tumor progression. Sci. Rep. 2020, 10, 7843. [Google Scholar] [CrossRef] [PubMed]
  34. Hahn, M.E.; Karchner, S.I.; Shapiro, M.A.; Perera, S.A. Molecular evolution of two vertebrate aryl hydrocarbon (dioxin) receptors (AHR1 and AHR2) and the PAS family. Proc. Natl. Acad. Sci. USA 1997, 94, 13743–13748. [Google Scholar] [CrossRef] [Green Version]
  35. Hahn, M.E.; Woodin, B.R.; Stegeman, J.J.; Tillitt, D.E. Aryl hydrocarbon receptor function in early vertebrates: Inducibility of cytochrome P450 1A in agnathan and elasmobranch fish. Comp. Biochem. Physiol. Part C Pharmacol. Toxicol. Endocrinol. 1998, 120, 67–75. [Google Scholar]
  36. Hahn, M.E.; Karchner, S.I.; Evans, B.R.; Franks, D.G.; Merson, R.R.; Lapseritis, J.M. Unexpected diversity of aryl hydrocarbon receptors in non-mammalian vertebrates: Insights from comparative genomics. J. Exp. Zoolog. A Comp. Exp. Biol. 2006, 305, 693–706. [Google Scholar] [CrossRef]
  37. Hahn, M.E. The aryl hydrocarbon receptor: A comparative perspective. Comp. Biochem. Physiol. Part C Pharmacol. Toxicol. Endocrinol. 1998, 121, 23–53. [Google Scholar] [CrossRef]
  38. Hahn, M.E. Aryl hydrocarbon receptors: Diversity and evolution. Chem. Biol. Interact. 2002, 141, 131–160. [Google Scholar] [CrossRef]
  39. Larigot, L.; Juricek, L.; Dairou, J.; Coumoul, X. AhR signaling pathways and regulatory functions. Biochim. Open 2018, 7, 1–9. [Google Scholar] [CrossRef]
  40. Kumar, M.B.; Perdew, G.H. Nuclear receptor coactivator SRC-1 interacts with the Q-rich subdomain of the AhR and modulates its transactivation potential. Gene Expr. 1999, 8, 273–286. [Google Scholar]
  41. Ma, Q.; Whitlock, J.P., Jr. A novel cytoplasmic protein that interacts with the Ah receptor, contains tetratricopeptide repeat motifs, and augments the transcriptional response to 2,3,7,8-tetrachlorodibenzo-p-dioxin. J. Biol. Chem. 1997, 272, 8878–8884. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Denis, M.; Cuthill, S.; Wikstroem, A.-C.; Poellinger, L.; Gustafsson, J.-A. Association of the Dioxin Receptor with Mr 90,000 Heat Shock protein. Biochem. Biophys. Res. Commun. 1988, 155, 801–807. [Google Scholar] [CrossRef]
  43. Kazlauskas, A.; Poellinger, L.; Pongratz, I. Evidence that the co-chaperone p23 regulates ligand responsiveness of the dioxin (Aryl hydrocarbon) receptor. J. Biol. Chem. 1999, 274, 13519–13524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Pappas, B.; Yang, Y.; Wang, Y.; Kim, K.; Chung, H.J.; Cheung, M.; Ngo, K.; Shinn, A.; Chan, W.K. p23 protects the human aryl hydrocarbon receptor from degradation via a heat shock protein 90-independent mechanism. Biochem. Pharmacol. 2018, 152, 34–44. [Google Scholar] [CrossRef]
  45. Pongratz, I.; Grant, G.; Mason, F.; Poellinger, L. Dual roles of the 90-kDa heat shock protein hsp90 in modulating fuctional activites of the dioxin receptor. J. Biol. Chem. 1992, 267, 13728–13733. [Google Scholar]
  46. Reyes, H.; Reisz-Porszasz, S.; Hankinson, O. Identification of the Ah receptor nuclear translocator protein (Arnt) as a component of the DNA binding form of the Ah receptor. Science 1992, 256, 1193–1195. [Google Scholar] [CrossRef]
  47. Reisz-Porszasz, S.; Probst, M.R.; Fukunaga, B.N.; Hankinson, O. Identification of functional domains of the aryl hydrocarbon receptor nuclear translocator protein (ARNT). Mol. Cell. Biol. 1994, 14, 6075–6086. [Google Scholar] [CrossRef] [Green Version]
  48. Harper, P.A.; Giannone, J.V.; Okey, A.B.; Denison, M.S. In vitro transformation of the human Ah receptor and its binding to a dioxin response element. Mol. Pharmacol. 1992, 42, 603–612. [Google Scholar]
  49. Beischlag, T.V.; Wang, S.; Rose, D.W.; Torchia, J.; Reisz-Porszasz, S.; Muhammad, K.; Nelson, W.E.; Probst, M.R.; Rosenfeld, M.G.; Hankinson, O. Recruitment of the NCoA/SRC-1/p160 family of transcriptional coactivators by the aryl hydrocarbon receptor/aryl hydrocarbon receptor nuclear translocator complex. Mol. Cell Biol. 2002, 22, 4319–4333. [Google Scholar] [CrossRef] [Green Version]
  50. Denison, M.S.; Fisher, J.M.; Whitlock, J.P., Jr. Protein-DNA interactions at recognition sites for the dioxin-Ah receptor complex. J. Biol. Chem 1989, 264, 16478–16482. [Google Scholar]
  51. Carlson, D.B.; Perdew, G.H. A dynamic role for the Ah receptor in cell signaling? Insights from a diverse group of Ah receptor interacting proteins. J. Biochem. Mol. Toxicol. 2002, 16, 317–325. [Google Scholar] [CrossRef] [PubMed]
  52. Kim, D.W.; Gazourian, L.; Quadri, S.A.; Romieu-Mourez, R.; Sherr, D.H.; Sonenshein, G.E. The RelA NF-kB subunit and the aryl hydrocarbon receptor (AhR) cooperate to transactivate the c-myc promoter in mammary cells. Oncogene 2000, 19, 5498–5506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Vogel, C.F.; Li, W.; Wu, D.; Miller, J.K.; Sweeney, C.; Lazennec, G.; Fujisawa, Y.; Matsumura, F. Interaction of aryl hydrocarbon receptor and NF-kappaB subunit RelB in breast cancer is associated with interleukin-8 overexpression. Arch. Biochem. Biophys. 2011, 512, 78–86. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Wilson, S.R.; Joshi, A.D.; Elferink, C.J. The tumor suppressor Kruppel-like factor 6 is a novel aryl hydrocarbon receptor DNA binding partner. J. Pharmacol. Exp. Ther. 2013, 345, 419–429. [Google Scholar] [CrossRef] [Green Version]
  55. Wang, F.; Hoivik, D.; Pollenz, R.; Safe, S. Functional and physical interactions between the estrogen receptor Sp1 and nuclear aryl hydrocarbon receptor complexes. Nucleic Acids Res. 1998, 26, 3044–3052. [Google Scholar] [CrossRef] [Green Version]
  56. Baker, J.R.; Sakoff, J.A.; McCluskey, A. The aryl hydrocarbon receptor (AhR) as a breast cancer drug target. Med. Res. Rev. 2020, 40, 972–1001. [Google Scholar] [CrossRef]
  57. Marlowe, J.L.; Fan, Y.; Chang, X.; Peng, L.; Knudsen, E.S.; Xia, Y.; Puga, A. The aryl hydrocarbon receptor binds to E2F1 and inhibits E2F1-induced apoptosis. Mol. Biol. Cell 2008, 19, 3263–3271. [Google Scholar] [CrossRef] [Green Version]
  58. Puga, A.; Xia, Y.; Elferink, C. Role of the aryl hydrocarbon receptor in cell cycle regulation. Chem. Biol. Interact. 2002, 141, 117–130. [Google Scholar] [CrossRef]
  59. Apetoh, L.; Quintana, F.J.; Pot, C.; Joller, N.; Xiao, S.; Kumar, D.; Burns, E.J.; Sherr, D.H.; Weiner, H.L.; Kuchroo, V.K. The aryl hydrocarbon receptor interacts with c-Maf to promote the differentiation of type 1 regulatory T cells induced by IL-27. Nat. Immunol. 2010, 11, 854–861. [Google Scholar] [CrossRef] [Green Version]
  60. Shashar, M.; Belghasem, M.E.; Matsuura, S.; Walker, J.; Richards, S.; Alousi, F.; Rijal, K.; Kolachalama, V.B.; Balcells, M.; Odagi, M.; et al. Targeting STUB1-tissue factor axis normalizes hyperthrombotic uremic phenotype without increasing bleeding risk. Sci. Transl. Med. 2017, 9, eaam8475. [Google Scholar] [CrossRef] [Green Version]
  61. Shivanna, S.; Kolandaivelu, K.; Shashar, M.; Belghasim, M.; Al-Rabadi, L.; Balcells, M.; Zhang, A.; Weinberg, J.; Francis, J.; Pollastri, M.P.; et al. The Aryl Hydrocarbon Receptor is a Critical Regulator of Tissue Factor Stability and an Antithrombotic Target in Uremia. J. Am. Soc. Nephrol. 2016, 27, 189–201. [Google Scholar] [CrossRef] [PubMed]
  62. Peng, Z.; Raufman, J.P.; Xie, G. Src-mediated cross-talk between farnesoid X and epidermal growth factor receptors inhibits human intestinal cell proliferation and tumorigenesis. PLoS ONE 2012, 7, e48461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Xie, G.; Peng, Z.; Raufman, J.P. Src-mediated aryl hydrocarbon and epidermal growth factor receptor cross talk stimulates colon cancer cell proliferation. Am. J. Physiol. Gastrointest. Liver Physiol. 2012, 302, G1006–G1015. [Google Scholar] [CrossRef] [PubMed]
  64. Endler, A.; Chen, L.; Shibasaki, F. Coactivator recruitment of AhR/ARNT1. Int. J. Mol. Sci. 2014, 15, 11100–11110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Ohtake, F.; Baba, A.; Takada, I.; Okada, M.; Iwasaki, K.; Miki, H.; Takahashi, S.; Kouzmenko, A.; Nohara, K.; Chiba, T.; et al. Dioxin receptor is a ligand-dependent E3 ubiquitin ligase. Nature 2007, 446, 562–566. [Google Scholar] [CrossRef]
  66. Dou, H.; Duan, Y.; Zhang, X.; Yu, Q.; Di, Q.; Song, Y.; Li, P.; Gong, Y. Aryl hydrocarbon receptor (AhR) regulates adipocyte differentiation by assembling CRL4B ubiquitin ligase to target PPARgamma for proteasomal degradation. J. Biol. Chem. 2019, 294, 18504–18515. [Google Scholar] [CrossRef]
  67. Hestermann, E.V.; Brown, M. Agonist and chemopreventative ligands induce differential transcriptional cofactor recruitment by aryl hydrocarbon receptor. Mol. Cell Biol. 2003, 23, 7920–7925. [Google Scholar] [CrossRef] [Green Version]
  68. Quintana, F.J.; Sherr, D.H. Aryl hydrocarbon receptor control of adaptive immunity. Pharmacol. Rev. 2013, 65, 1148–1161. [Google Scholar] [CrossRef] [Green Version]
  69. Matikainen, T.; Perez, G.I.; Jurisicova, A.; Pru, J.K.; Schlezinger, J.J.; Ryu, H.Y.; Laine, J.; Sakai, T.; Korsmeyer, S.J.; Casper, R.F.; et al. Aromatic hydrocarbon receptor-driven Bax gene expression is required for premature ovarian failure caused by biohazardous environmental chemicals. Nat. Genet. 2001, 28, 355–360. [Google Scholar] [CrossRef]
  70. Matikainen, T.M.; Moriyama, T.; Morita, Y.; Perez, G.I.; Korsmeyer, S.J.; Sherr, D.H.; Tilly, J.L. Ligand activation of the aromatic hydrocarbon receptor transcription factor drives Bax-dependent apoptosis in developing fetal ovarian germ cells. Endocrinology 2002, 143, 615–620. [Google Scholar] [CrossRef]
  71. Chevallier, A.; Mialot, A.; Petit, J.M.; Fernandez-Salguero, P.; Barouki, R.; Coumoul, X.; Beraneck, M. Oculomotor deficits in aryl hydrocarbon receptor null mouse. PLoS ONE 2013, 8, e53520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Lahvis, G.P.; Lindell, S.L.; Thomas, R.S.; McCuskey, R.S.; Murphy, C.; Glover, E.; Bentz, M.; Southard, J.; Bradfield, C.A. Portosystemic shunting and persistent fetal vascular structures in aryl hydrocarbon receptor-deficient mice. Proc. Natl. Acad. Sci. USA 2000, 97, 10442–10447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Lahvis, G.P.; Pyzalski, R.W.; Glover, E.; Pitot, H.C.; McElwee, M.K.; Bradfield, C.A. The aryl hydrocarbon receptor is required for developmental closure of the ductus venosus in the neonatal mouse. Mol. Pharmacol. 2005, 67, 714–720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Fernandez-Salguero, P.M.; Ward, J.M.; Sundberg, J.P.; Gonzalez, F.J. Lesions of aryl-hydrocarbon receptor-deficient mice. Vet. Pathol. 1997, 34, 605–614. [Google Scholar] [CrossRef] [PubMed]
  75. Kurita, H.; Carreira, V.S.; Fan, Y.; Jiang, M.; Naticchioni, M.; Koch, S.; Rubinstein, J.; Puga, A. Ah receptor expression in cardiomyocytes protects adult female mice from heart dysfunction induced by TCDD exposure. Toxicology 2016, 355–356, 9–20. [Google Scholar] [CrossRef] [PubMed]
  76. Wang, Q.; Chen, J.; Ko, C.I.; Fan, Y.; Carreira, V.; Chen, Y.; Xia, Y.; Medvedovic, M.; Puga, A. Disruption of aryl hydrocarbon receptor homeostatic levels during embryonic stem cell differentiation alters expression of homeobox transcription factors that control cardiomyogenesis. Environ. Health Perspect. 2013, 121, 1334–1343. [Google Scholar] [CrossRef] [Green Version]
  77. Leung, A.; Zulick, E.; Skvir, N.; Vanuytsel, K.; Morrison, T.A.; Naing, Z.H.; Wang, Z.; Dai, Y.; Chui, D.H.K.; Steinberg, M.H.; et al. Notch and Aryl Hydrocarbon Receptor Signaling Impact Definitive Hematopoiesis from Human Pluripotent Stem Cells. Stem Cells 2018. [Google Scholar] [CrossRef] [Green Version]
  78. Smith, B.W.; Rozelle, S.S.; Leung, A.; Ubellacker, J.; Parks, A.; Nah, S.K.; French, D.; Gadue, P.; Monti, S.; Chui, D.H.; et al. The aryl hydrocarbon receptor directs hematopoietic progenitor cell expansion and differentiation. Blood 2013, 122, 376–385. [Google Scholar] [CrossRef] [Green Version]
  79. Kiss, E.A.; Vonarbourg, C.; Kopfmann, S.; Hobeika, E.; Finke, D.; Esser, C.; Diefenbach, A. Natural aryl hydrocarbon receptor ligands control organogenesis of intestinal lymphoid follicles. Science 2011, 334, 1561–1565. [Google Scholar] [CrossRef]
  80. Jux, B.; Kadow, S.; Luecke, S.; Rannug, A.; Krutmann, J.; Esser, C. The aryl hydrocarbon receptor mediates UVB radiation-induced skin tanning. J. Investig. Dermatol. 2011, 131, 203–210. [Google Scholar] [CrossRef] [Green Version]
  81. Trikha, P.; Lee, D.A. The role of AhR in transcriptional regulation of immune cell development and function. Biochim. Biophys. Acta Rev. Cancer 2020, 1873, 188335. [Google Scholar] [CrossRef] [PubMed]
  82. Sherr, D.H.; Monti, S. The role of the aryl hydrocarbon receptor in normal and malignant B cell development. Semin. Immunopathol. 2013, 35, 705–716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Hayashibara, T.; Yamada, Y.; Mori, N.; Harasawa, H.; Sugahara, K.; Miyanishi, T.; Kamihira, S.; Tomonaga, M. Possible involvement of aryl hydrocarbon receptor (AhR) in adult T-cell leukemia (ATL) leukemogenesis: Constitutive activation of AhR in ATL. Biochem. Biophys. Res. Commun. 2003, 300, 128–134. [Google Scholar] [CrossRef]
  84. Yang, X.; Solomon, S.; Fraser, L.R.; Trombino, A.F.; Liu, D.; Sonenshein, G.E.; Hestermann, E.V.; Sherr, D.H. Constitutive regulation of CYP1B1 by the aryl hydrocarbon receptor (AhR) in pre-malignant and malignant mammary tissue. J. Cell. Biochem. 2008, 104, 402–417. [Google Scholar] [CrossRef] [PubMed]
  85. Trombino, A.F.; Near, R.I.; Matulka, R.A.; Yang, S.; Hafer, L.J.; Toselli, P.A.; Kim, D.W.; Rogers, A.E.; Sonenshein, G.E.; Sherr, D.H. Expression of the aryl hydrocarbon receptor/transcription factor (AhR) and AhR-regulated CYP1 gene transcripts in a rat model of mammary tumorigenesis. Breast Cancer Res. Treat. 2000, 63, 117–131. [Google Scholar] [CrossRef] [PubMed]
  86. DiNatale, B.C.; Smith, K.; John, K.; Krishnegowda, G.; Amin, S.G.; Perdew, G.H. Ah receptor antagonism represses head and neck tumor cell aggressive phenotype. Mol. Cancer Res. 2012, 10, 1369–1379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. DiNatale, B.C.; Schroeder, J.C.; Perdew, G.H. Ah receptor antagonism inhibits constitutive and cytokine inducible IL6 production in head and neck tumor cell lines. Mol. Carcinog. 2011, 50, 173–183. [Google Scholar] [CrossRef] [Green Version]
  88. Opitz, C.A.; Litzenburger, U.M.; Sahm, F.; Ott, M.; Tritschler, I.; Trump, S.; Schumacher, T.; Jestaedt, L.; Schrenk, D.; Weller, M.; et al. An endogenous tumour-promoting ligand of the human aryl hydrocarbon receptor. Nature 2011, 478, 197–203. [Google Scholar] [CrossRef]
  89. Litzenburger, U.M.; Opitz, C.A.; Sahm, F.; Rauschenbach, K.J.; Trump, S.; Winter, M.; Ott, M.; Ochs, K.; Lutz, C.; Liu, X.; et al. Constitutive IDO expression in human cancer is sustained by an autocrine signaling loop involving IL-6, STAT3 and the AHR. Oncotarget 2014, 5, 1038–1051. [Google Scholar] [CrossRef] [Green Version]
  90. Takenaka, M.C.; Gabriely, G.; Rothhammer, V.; Mascanfroni, I.D.; Wheeler, M.A.; Chao, C.C.; Gutierrez-Vazquez, C.; Kenison, J.; Tjon, E.C.; Barroso, A.; et al. Control of tumor-associated macrophages and T cells in glioblastoma via AHR and CD39. Nat. Neurosci. 2020, 22, 729–740. [Google Scholar] [CrossRef]
  91. Ishida, M.; Mikami, S.; Kikuchi, E.; Kosaka, T.; Miyajima, A.; Nakagawa, K.; Mukai, M.; Okada, Y.; Oya, M. Activation of the aryl hydrocarbon receptor pathway enhances cancer cell invasion by upregulating the MMP expression and is associated with poor prognosis in upper urinary tract urothelial cancer. Carcinogenesis 2010, 31, 287–295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Lin, P.; Chang, H.; Tsai, W.T.; Wu, M.H.; Liao, Y.S.; Chen, J.T.; Su, J.M. Overexpression of aryl hydrocarbon receptor in human lung carcinomas. Toxicol. Pathol. 2003, 31, 22–30. [Google Scholar] [CrossRef] [PubMed]
  93. Lin, P.; Chang, H.; Ho, W.L.; Wu, M.H.; Su, J.M. Association of aryl hydrocarbon receptor and cytochrome P4501B1 expressions in human non-small cell lung cancers. Lung Cancer 2003, 42, 255–261. [Google Scholar] [CrossRef]
  94. Koliopanos, A.; Kleeff, J.; Xiao, Y.; Safe, S.; Zimmermann, A.; Buchler, M.W.; Friess, H. Increased arylhydrocarbon receptor expression offers a potential therapeutic target for pancreatic cancer. Oncogene 2002, 21, 6059–6070. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Chen, J.; Rocken, C.; Klein-Hitpass, L.; Gotze, T.; Leodolter, A.; Malfertheiner, P.; Ebert, M.P. Microarray analysis of gene expression in metastatic gastric cancer cells after incubation with the methylation inhibitor 5-aza-2′-deoxycytidine. Clin. Exp. Metastasis 2004, 21, 389–397. [Google Scholar] [CrossRef]
  96. Peng, T.L.; Chen, J.; Mao, W.; Song, X.; Chen, M.H. Aryl hydrocarbon receptor pathway activation enhances gastric cancer cell invasiveness likely through a c-Jun-dependent induction of matrix metalloproteinase-9. BMC Cell Biol. 2009, 10, 27. [Google Scholar] [CrossRef] [Green Version]
  97. Bogoevska, V.; Wolters-Eisfeld, G.; Hofmann, B.T.; El Gammal, A.T.; Mercanoglu, B.; Gebauer, F.; Vashist, Y.K.; Bogoevski, D.; Perez, D.; Gagliani, N.; et al. HRG/HER2/HER3 signaling promotes AhR-mediated Memo-1 expression and migration in colorectal cancer. Oncogene 2017, 36, 2394–2404. [Google Scholar] [CrossRef]
  98. Wang, Z.; Monti, S.; Sherr, D.H. The diverse and important contributions of the AHR to cancer and cancer immunity. Curr. Opin. Toxicol. 2017, 2, 102–107. [Google Scholar] [CrossRef]
  99. Su, J.M.; Lin, P.; Chang, H. Prognostic value of nuclear translocation of aryl hydrocarbon receptor for non-small cell lung cancer. Anticancer Res. 2013, 33, 3953–3961. [Google Scholar]
  100. Vacher, S.; Castagnet, P.; Chemlali, W.; Lallemand, F.; Meseure, D.; Pocard, M.; Bieche, I.; Perrot-Applanat, M. High AHR expression in breast tumors correlates with expression of genes from several signaling pathways namely inflammation and endogenous tryptophan metabolism. PLoS ONE 2018, 13, e0190619. [Google Scholar] [CrossRef] [Green Version]
  101. Andersson, P.; McGuire, J.; Rubio, C.; Gradin, K.; Whitelaw, M.L.; Pettersson, S.; Hanberg, A.; Poellinger, L. A constitutively active dioxin/aryl hydrocarbon receptor induces stomach tumors. Proc. Natl. Acad. Sci. USA 2002, 99, 9990–9995. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Kuznetsov, N.V.; Andersson, P.; Gradin, K.; Stein, P.; Dieckmann, A.; Pettersson, S.; Hanberg, A.; Poellinger, L. The dioxin/aryl hydrocarbon receptor mediates downregulation of osteopontin gene expression in a mouse model of gastric tumourigenesis. Oncogene 2005, 24, 3216–3222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Moennikes, O.; Loeppen, S.; Buchmann, A.; Andersson, P.; Ittrich, C.; Poellinger, L.; Schwarz, M. A constitutively active dioxin/aryl hydrocarbon receptor promotes hepatocarcinogenesis in mice. Cancer Res. 2004, 64, 4707–4710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Lu, P.; Cai, X.; Guo, Y.; Xu, M.; Tian, J.; Locker, J.; Xie, W. Constitutive Activation of the Human Aryl Hydrocarbon Receptor in Mice Promotes Hepatocarcinogenesis Independent of Its Coactivator Gadd45b. Toxicol. Sci. 2019, 167, 581–592. [Google Scholar] [CrossRef] [PubMed]
  105. Richmond, O.; Ghotbaddini, M.; Allen, C.; Walker, A.; Zahir, S.; Powell, J.B. The aryl hydrocarbon receptor is constitutively active in advanced prostate cancer cells. PLoS ONE 2014, 9, e95058. [Google Scholar] [CrossRef] [PubMed]
  106. Stanford, E.A.; Ramirez-Cardenas, A.; Wang, Z.; Novikov, O.; Alamoud, K.; Koutrakis, P.; Mizgerd, J.P.; Genco, C.A.; Kukuruzinska, M.; Monti, S.; et al. Role for the aryl hydrocarbon receptor and diverse ligands in oral squamous cell carcinoma migration and tumorigenesis. Mol. Cancer Res. 2016, 14, 696–706. [Google Scholar] [CrossRef] [Green Version]
  107. Narasimhan, S.; Stanford Zulick, E.; Novikov, O.; Parks, A.J.; Schlezinger, J.J.; Wang, Z.; Laroche, F.; Feng, H.; Mulas, F.; Monti, S.; et al. Towards Resolving the Pro- and Anti-Tumor Effects of the Aryl Hydrocarbon Receptor. Int. J. Mol. Sci. 2018, 19, 1388. [Google Scholar] [CrossRef] [Green Version]
  108. Novikov, O.; Wang, Z.; Stanford, E.A.; Parks, A.J.; Ramirez-Cardenas, A.; Landesman, E.; Laklouk, I.; Sarita-Reyes, C.; Gusenleitner, D.; Li, A.; et al. An aryl hydrocarbon receptor-mediated amplification loop that enforces cell migration in ER-/PR-/Her2- human breast cancer cells. Mol. Pharmacol. 2016, 90, 674–688. [Google Scholar] [CrossRef] [Green Version]
  109. Parks, A.J.; Pollastri, M.P.; Hahn, M.E.; Stanford, E.A.; Novikov, O.; Franks, D.G.; Haigh, S.E.; Narasimhan, S.; Ashton, T.D.; Hopper, T.G.; et al. In silico identification of an aryl hydrocarbon receptor antagonist with biological activity in vitro and in vivo. Mol. Pharmacol. 2014, 86, 593–608. [Google Scholar] [CrossRef] [Green Version]
  110. Corre, S.; Tardif, N.; Mouchet, N.; Leclair, H.M.; Boussemart, L.; Gautron, A.; Bachelot, L.; Perrot, A.; Soshilov, A.; Rogiers, A.; et al. Sustained activation of the Aryl hydrocarbon Receptor transcription factor promotes resistance to BRAF-inhibitors in melanoma. Nat. Commun. 2018, 9, 4775. [Google Scholar] [CrossRef]
  111. Leclair, H.M.; Tardif, N.; Paris, A.; Galibert, M.D.; Corre, S. Role of Flavonoids in the Prevention of AhR-Dependent Resistance During Treatment with BRAF Inhibitors. Int. J. Mol. Sci 2020, 21, 5025. [Google Scholar] [CrossRef] [PubMed]
  112. The Human Protein Atlas. HPA029722. Available online: https://www.proteinatlas.org/ENSG00000106546-AHR/pathology/urothelial+cancer#img (accessed on 25 October 2020).
  113. The Human Protein Atlas. HPA029722. Available online: https://www.proteinatlas.org/ENSG00000106546-AHR/pathology/cervical+cancer#img (accessed on 25 October 2020).
  114. The Human Protein Atlas. HPA029723. Available online: https://www.proteinatlas.org/ENSG00000106546-AHR/pathology/glioma#img (accessed on 25 October 2020).
  115. The Human Protein Atlas. HPA029723. Available online: https://www.proteinatlas.org/ENSG00000106546-AHR/pathology/pancreatic+cancer#img (accessed on 25 October 2020).
  116. The Human Protein Atlas. HPA029723. Available online: https://www.proteinatlas.org/ENSG00000106546-AHR/pathology/head+and+neck+cancer#img (accessed on 25 October 2020).
  117. The Human Protein Atlas. HPA029722. Available online: https://www.proteinatlas.org/ENSG00000106546-AHR/pathology/lung+cancer#img (accessed on 25 October 2020).
  118. The Human Protein Atlas. HPA029723. Available online: https://www.proteinatlas.org/ENSG00000106546-AHR/pathology/melanoma#img (accessed on 25 October 2020).
  119. Mohamed, H.T.; Gadalla, R.; El-Husseiny, N.; Hassan, H.; Wang, Z.; Ibrahim, S.A.; El-Shinawi, M.; Sherr, D.H.; Mohamed, M.M. Inflammatory breast cancer: Activation of the aryl hydrocarbon receptor and its target CYP1B1 correlates closely with Wnt5a/b-beta-catenin signalling, the stem cell phenotype and disease progression. J. Adv. Res. 2019, 16, 75–86. [Google Scholar] [CrossRef] [PubMed]
  120. Jeschke, U.; Zhang, X.; Kuhn, C.; Jalaguier, S.; Colinge, J.; Pfender, K.; Mayr, D.; Ditsch, N.; Harbeck, N.; Mahner, S.; et al. The Prognostic Impact of the Aryl Hydrocarbon Receptor (AhR) in Primary Breast Cancer Depends on the Lymph Node Status. Int. J. Mol. Sci. 2019, 20, 1016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. McKay, J.A.; Murray, G.I.; Ah-See, A.K.; Greenlee, W.F.; Marcus, C.B.; Burke, M.D.; Melvin, W.T. Differential expression of CYP1A1 and CYP1B1 in human breast cancer. Biochem. Soc. Trans. 1996, 24, 327S. [Google Scholar] [CrossRef] [PubMed]
  122. Murray, G.I.; Taylor, M.C.; McFadyen, M.C.; McKay, J.A.; Greenlee, W.F.; Burke, M.D.; Melvin, W.T. Tumor-specific expression of cytochrome P450 CYP1B1. Cancer Res. 1997, 57, 3026–3031. [Google Scholar] [PubMed]
  123. Ma, J.X.; Zhang, K.L.; Liu, X.; Ma, Y.L.; Pei, L.N.; Zhu, Y.F.; Zhou, L.; Chen, X.Y.; Kong, Q.Y.; Li, H.; et al. Concurrent expression of aryl hydrocarbon receptor and CYP1A1 but not CYP1A1 MspI polymorphism is correlated with gastric cancers raised in Dalian, China. Cancer Lett. 2006, 240, 253–260. [Google Scholar] [CrossRef]
  124. Spink, D.C.; Spink, B.C.; Cao, J.Q.; DePasquale, J.A.; Pentecost, B.T.; Fasco, M.J.; Li, Y.; Sutter, T.R. Differential expression of CYP1A1 and CYP1B1 in human breast epithelial cells and breast tumor cells. Carcinogenesis 1998, 19, 291–298. [Google Scholar] [CrossRef] [Green Version]
  125. McKay, J.A.; Melvin, W.T.; Ah-See, A.K.; Ewen, S.W.; Greenlee, W.F.; Marcus, C.B.; Burke, M.D.; Murray, G.I. Expression of cytochrome P450 CYP1B1 in breast cancer. FEBS Lett. 1995, 374, 270–272. [Google Scholar] [CrossRef] [Green Version]
  126. Chang, H.; Su, J.M.; Huang, C.C.; Liu, L.C.; Tsai, C.H.; Chou, M.C.; Lin, P. Using a combination of cytochrome P450 1B1 and beta-catenin for early diagnosis and prevention of colorectal cancer. Cancer Detect. Prev. 2005, 29, 562–569. [Google Scholar] [CrossRef]
  127. Nukaya, M.; Lin, B.C.; Glover, E.; Moran, S.M.; Kennedy, G.D.; Bradfield, C.A. The aryl hydrocarbon receptor-interacting protein (AIP) is required for dioxin-induced hepatotoxicity but not for the induction of the Cyp1a1 and Cyp1a2 genes. J. Biol. Chem. 2010, 285, 35599–35605. [Google Scholar] [CrossRef] [Green Version]
  128. Flies, A.; Ahmadi, T.; Parks, A.J.; Prokaeva, T.; Weng, L.; Rolfe, S.S.; Seldin, D.C.; Sherr, D.H. Immunoglobulin light chain, Blimp-1 and cytochrome P4501B1 peptides as potential vaccines for AL amyloidosis. Immunol. Cell Biol. 2012, 90, 528–539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Gribben, J.G.; Ryan, D.P.; Boyajian, R.; Urban, R.G.; Hedley, M.L.; Beach, K.; Nealon, P.; Matulonis, U.; Campos, S.; Gilligan, T.D.; et al. Unexpected association between induction of immunity to the universal tumor antigen CYP1B1 and response to next therapy. Clin. Cancer Res. 2005, 11, 4430–4436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Luby, T.M. Targeting cytochrome P450 CYP1B1 with a therapeutic cancer vaccine. Expert Rev. Vaccines 2008, 7, 995–1003. [Google Scholar] [CrossRef] [PubMed]
  131. McFadyen, M.C.; Murray, G.I. Cytochrome P450 1B1: A novel anticancer therapeutic target. Future Oncol. 2005, 1, 259–263. [Google Scholar] [CrossRef]
  132. Hanzelmann, S.; Castelo, R.; Guinney, J. GSVA: Gene set variation analysis for microarray and RNA-seq data. BMC Bioinform. 2013, 14, 7. [Google Scholar] [CrossRef] [Green Version]
  133. Abnet, C.C.; Fagundes, R.B.; Strickland, P.T.; Kamangar, F.; Roth, M.J.; Taylor, P.R.; Dawsey, S.M. The influence of genetic polymorphisms in Ahr, CYP1A1, CYP1A2, CYP1B1, GST M1, GST T1 and UGT1A1 on urine 1-hydroxypyrene glucuronide concentrations in healthy subjects from Rio Grande do Sul, Brazil. Carcinogenesis 2007, 28, 112–117. [Google Scholar] [CrossRef]
  134. Anttila, S.; Lei, X.D.; Elovaara, E.; Karjalainen, A.; Sun, W.; Vainio, H.; Hankinson, O. An uncommon phenotype of poor inducibility of CYP1A1 in human lung is not ascribable to polymorphisms in the AHR, ARNT, or CYP1A1 genes. Pharmacogenetics 2000, 10, 741–751. [Google Scholar] [CrossRef]
  135. Cauchi, S.; Stucker, I.; Solas, C.; Laurent-Puig, P.; Cenee, S.; Hemon, D.; Jacquet, M.; Kremers, P.; Beaune, P.; Massaad-Massade, L. Polymorphisms of human aryl hydrocarbon receptor (AhR) gene in a French population: Relationship with CYP1A1 inducibility and lung cancer. Carcinogenesis 2001, 22, 1819–1824. [Google Scholar] [CrossRef] [Green Version]
  136. DiNatale, B.C.; Murray, I.A.; Schroeder, J.C.; Flaveny, C.A.; Lahoti, T.S.; Laurenzana, E.M.; Omiecinski, C.J.; Perdew, G.H. Kynurenic acid is a potent endogenous aryl hydrocarbon receptor ligand that synergistically induces interleukin-6 in the presence of inflammatory signaling. Toxicol. Sci 2010, 115, 89–97. [Google Scholar] [CrossRef] [Green Version]
  137. Lowe, M.M.; Mold, J.E.; Kanwar, B.; Huang, Y.; Louie, A.; Pollastri, M.P.; Wang, C.; Patel, G.; Franks, D.G.; Schlezinger, J.; et al. Identification of cinnabarinic acid as a novel endogenous aryl hydrocarbon receptor ligand that drives IL-22 production. PLoS ONE 2014, 9, e87877. [Google Scholar] [CrossRef]
  138. Seok, S.H.; Ma, Z.X.; Feltenberger, J.B.; Chen, H.; Chen, H.; Scarlett, C.; Lin, Z.; Satyshur, K.A.; Cortopassi, M.; Jefcoate, C.R.; et al. Trace derivatives of kynurenine potently activate the aryl hydrocarbon receptor (AHR). J. Biol. Chem. 2018, 293, 1994–2005. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Van Baren, N.; Van den Eynde, B.J. Tryptophan-degrading enzymes in tumoral immune resistance. Front. Immunol. 2015, 6, 34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Nguyen, N.T.; Kimura, A.; Nakahama, T.; Chinen, I.; Masuda, K.; Nohara, K.; Fujii-Kuriyama, Y.; Kishimoto, T. Aryl hydrocarbon receptor negatively regulates dendritic cell immunogenicity via a kynurenine-dependent mechanism. Proc. Natl. Acad. Sci. USA 2010, 107, 19961–19966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Vogel, C.F.; Goth, S.R.; Dong, B.; Pessah, I.N.; Matsumura, F. Aryl hydrocarbon receptor signaling mediates expression of indoleamine 2,3-dioxygenase. Biochem. Biophys. Res. Commun. 2008, 375, 331–335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Simones, T.; Shepherd, D.M. Consequences of AhR activation in steady-state dendritic cells. Toxicol. Sci. 2011, 119, 293–307. [Google Scholar] [CrossRef]
  143. Li, Q.; Harden, J.L.; Anderson, C.D.; Egilmez, N.K. Tolerogenic Phenotype of IFN-gamma-Induced IDO+ Dendritic Cells Is Maintained via an Autocrine IDO-Kynurenine/AhR-IDO Loop. J. Immunol. 2016, 197, 962–970. [Google Scholar] [CrossRef] [Green Version]
  144. Hwang, S.J.; Hwang, Y.J.; Yun, M.O.; Kim, J.H.; Oh, G.S.; Park, J.H. Indoxyl 3-sulfate stimulates Th17 differentiation enhancing phosphorylation of c-Src and STAT3 to worsen experimental autoimmune encephalomyelitis. Toxicol. Lett. 2013, 220, 109–117. [Google Scholar] [CrossRef]
  145. Rothhammer, V.; Mascanfroni, I.D.; Bunse, L.; Takenaka, M.C.; Kenison, J.E.; Mayo, L.; Chao, C.C.; Patel, B.; Yan, R.; Blain, M.; et al. Type I interferons and microbial metabolites of tryptophan modulate astrocyte activity and central nervous system inflammation via the aryl hydrocarbon receptor. Nat. Med. 2016, 22, 586–597. [Google Scholar] [CrossRef]
  146. Schroeder, J.C.; Dinatale, B.C.; Murray, I.A.; Flaveny, C.A.; Liu, Q.; Laurenzana, E.M.; Lin, J.M.; Strom, S.C.; Omiecinski, C.J.; Amin, S.; et al. The uremic toxin 3-indoxyl sulfate is a potent endogenous agonist for the human aryl hydrocarbon receptor. Biochemistry 2010, 49, 393–400. [Google Scholar] [CrossRef] [Green Version]
  147. Benson, J.M.; Shepherd, D.M. Dietary ligands of the aryl hydrocarbon receptor induce anti-inflammatory and immunoregulatory effects on murine dendritic cells. Toxicol. Sci. 2011, 124, 327–338. [Google Scholar] [CrossRef] [Green Version]
  148. Wincent, E.; Amini, N.; Luecke, S.; Glatt, H.; Bergman, J.; Crescenzi, C.; Rannug, A.; Rannug, U. The suggested physiologic aryl hydrocarbon receptor activator and cytochrome P4501 substrate 6-formylindolo[3,2-b]carbazole is present in humans. J. Biol. Chem. 2009, 284, 2690–2696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Adachi, J.; Mori, Y.; Matsui, S.; Takigami, H.; Fujino, J.; Kitagawa, H.; Miller, C.A., 3rd; Kato, T.; Saeki, K.; Matsuda, T. Indirubin and indigo are potent aryl hydrocarbon receptor ligands present in human urine. J. Biol. Chem. 2001, 276, 31475–31478. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Schlezinger, J.J.; Bernard, P.L.; Haas, A.; Grandjean, P.; Weihe, P.; Sherr, D.H. Direct assessment of cumulative aryl hydrocarbon receptor agonist activity in sera from experimentally exposed mice and environmentally exposed humans. Environ. Health Perspect. 2010, 118, 693–698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Angeli, D.; Ferrell, J.E., Jr.; Sontag, E.D. Detection of multistability, bifurcations, and hysteresis in a large class of biological positive-feedback systems. Proc. Natl. Acad. Sci. USA 2004, 101, 1822–1827. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Avendano, M.S.; Leidy, C.; Pedraza, J.M. Tuning the range and stability of multiple phenotypic states with coupled positive-negative feedback loops. Nat. Commun. 2013, 4, 2605. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Kato, A.; Mitrophanov, A.Y.; Groisman, E.A. A connector of two-component regulatory systems promotes signal amplification and persistence of expression. Proc. Natl. Acad. Sci. USA 2007, 104, 12063–12068. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Mitrophanov, A.Y.; Groisman, E.A. Positive feedback in cellular control systems. Bioessays 2008, 30, 542–555. [Google Scholar] [CrossRef] [Green Version]
  155. Mitrophanov, A.Y.; Hadley, T.J.; Groisman, E.A. Positive autoregulation shapes response timing and intensity in two-component signal transduction systems. J. Mol. Biol. 2010, 401, 671–680. [Google Scholar] [CrossRef] [Green Version]
  156. Baba, T.; Mimura, J.; Gradin, K.; Kuroiwa, A.; Watanabe, T.; Matsuda, Y.; Inazawa, J.; Sogawa, K.; Fujii-Kuriyama, Y. Structure and expression of the Ah receptor repressor gene. J. Biol. Chem. 2001, 276, 33101–33110. [Google Scholar] [CrossRef] [Green Version]
  157. Mimura, J.; Ema, M.; Sogawa, K.; Fujii-Kuriyama, Y. Identification of a novel mechanism of regulation of Ah (dioxin) receptor function. Genes Dev. 1999, 13, 20–25. [Google Scholar] [CrossRef] [Green Version]
  158. Evans, B.R.; Karchner, S.I.; Allan, L.L.; Pollenz, R.S.; Tanguay, R.L.; Jenny, M.J.; Sherr, D.H.; Hahn, M.E. Repression of aryl hydrocarbon receptor (AHR) signaling by AHR repressor: Role of DNA binding and competition for AHR nuclear translocator. Mol. Pharmacol. 2008, 73, 387–398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Hahn, M.E.; Allan, L.L.; Sherr, D.H. Regulation of constitutive and inducible AHR signaling: Complex interactions involving the AHR repressor. Biochem. Pharmacol. 2009, 77, 485–497. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Zudaire, E.; Cuesta, N.; Murty, V.; Woodson, K.; Adams, L.; Gonzalez, N.; Martinez, A.; Narayan, G.; Kirsch, I.; Franklin, W.; et al. The aryl hydrocarbon receptor repressor is a putative tumor suppressor gene in multiple human cancers. J. Clin. Investig. 2008, 118, 640–650. [Google Scholar] [CrossRef] [PubMed]
  161. Levine-Fridman, A.; Chen, L.; Elferink, C.J. Cytochrome P4501A1 promotes G1 phase cell cycle progression by controlling aryl hydrocarbon receptor activity. Mol. Pharmacol. 2004, 65, 461–469. [Google Scholar] [CrossRef] [Green Version]
  162. Stanford, E.A.; Wang, Z.; Novikov, O.; Mulas, F.; Landesman-Bollag, E.; Monti, S.; Smith, B.W.; Seldin, D.C.; Murphy, G.J.; Sherr, D.H. The role of the aryl hydrocarbon receptor in the development of cells with the molecular and functional characteristics of cancer stem-like cells. BMC Biol. 2016, 14, 20. [Google Scholar] [CrossRef] [Green Version]
  163. Friedl, P.; Wolf, K. Tumour-cell invasion and migration: Diversity and escape mechanisms. Nat. Rev. Cancer 2003, 3, 362–374. [Google Scholar] [CrossRef]
  164. Mulero-Navarro, S.; Pozo-Guisado, E.; Perez-Mancera, P.A.; Alvarez-Barrientos, A.; Catalina-Fernandez, I.; Hernandez-Nieto, E.; Saenz-Santamaria, J.; Martinez, N.; Rojas, J.M.; Sanchez-Garcia, I.; et al. Immortalized mouse mammary fibroblasts lacking dioxin receptor have impaired tumorigenicity in a subcutaneous mouse xenograph model. J. Biol. Chem. 2005, 280, 28731–28741. [Google Scholar] [CrossRef] [Green Version]
  165. Diry, M.; Tomkiewicz, C.; Koehle, C.; Coumoul, X.; Bock, K.W.; Barouki, R.; Transy, C. Activation of the dioxin/aryl hydrocarbon receptor (AhR) modulates cell plasticity through a JNK-dependent mechanism. Oncogene 2006, 25, 5570–5574. [Google Scholar] [CrossRef] [Green Version]
  166. Goode, G.D.; Ballard, B.R.; Manning, H.C.; Freeman, M.L.; Kang, Y.; Eltom, S.E. Knockdown of aberrantly upregulated aryl hydrocarbon receptor reduces tumor growth and metastasis of MDA-MB-231 human breast cancer cell line. Int. J. Cancer 2013, 133, 2769–2780. [Google Scholar] [CrossRef] [Green Version]
  167. Haque, M.; Francis, J.; Sehgal, I. Aryl hydrocarbon exposure induces expression of MMP-9 in human prostate cancer cell lines. Cancer Lett. 2005, 225, 159–166. [Google Scholar] [CrossRef]
  168. Ide, H.; Lu, Y.; Yu, J.; Noguchi, T.; Kanayama, M.; Muto, S.; Yamaguchi, R.; Kawato, S.; Horie, S. Aryl hydrocarbon receptor signaling involved in the invasiveness of LNCaP cells. Hum. Cell 2017, 30, 133–139. [Google Scholar] [CrossRef] [PubMed]
  169. Thiery, J.P.; Acloque, H.; Huang, R.Y.; Nieto, M.A. Epithelial-mesenchymal transitions in development and disease. Cell 2009, 139, 871–890. [Google Scholar] [CrossRef] [PubMed]
  170. Chen, A.; Beetham, H.; Black, M.A.; Priya, R.; Telford, B.J.; Guest, J.; Wiggins, G.A.; Godwin, T.D.; Yap, A.S.; Guilford, P.J. E-cadherin loss alters cytoskeletal organization and adhesion in non-malignant breast cells but is insufficient to induce an epithelial-mesenchymal transition. BMC Cancer 2014, 14, 552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Vincent-Salomon, A.; Thiery, J.P. Host microenvironment in breast cancer development: Epithelial-mesenchymal transition in breast cancer development. Breast Cancer Res. 2003, 5, 101–106. [Google Scholar] [CrossRef] [Green Version]
  172. Richardson, A.M.; Havel, L.S.; Koyen, A.E.; Konen, J.M.; Shupe, J.; Wiles, W.G.T.; Martin, W.D.; Grossniklaus, H.E.; Sica, G.; Gilbert-Ross, M.; et al. Vimentin Is Required for Lung Adenocarcinoma Metastasis via Heterotypic Tumor Cell-Cancer-Associated Fibroblast Interactions during Collective Invasion. Clin. Cancer Res. 2018, 24, 420–432. [Google Scholar] [CrossRef] [Green Version]
  173. Brooks, J.; Eltom, S.E. Malignant transformation of mammary epithelial cells by ectopic overexpression of the aryl hydrocarbon receptor. Curr. Cancer Drug Targets 2011, 11, 654–669. [Google Scholar] [CrossRef]
  174. Zhu, P.; Yu, H.; Zhou, K.; Bai, Y.; Qi, R.; Zhang, S. 3,3′-Diindolylmethane modulates aryl hydrocarbon receptor of esophageal squamous cell carcinoma to reverse epithelial-mesenchymal transition through repressing RhoA/ROCK1-mediated COX2/PGE2 pathway. J. Exp. Clin. Cancer Res. 2020, 39, 113. [Google Scholar] [CrossRef]
  175. Zhu, P.; Zhou, K.; Lu, S.; Bai, Y.; Qi, R.; Zhang, S. Modulation of aryl hydrocarbon receptor inhibits esophageal squamous cell carcinoma progression by repressing COX2/PGE2/STAT3 axis. J. Cell Commun. Signal. 2020, 14, 175–192. [Google Scholar] [CrossRef]
  176. Gilkes, D.M.; Xiang, L.; Lee, S.J.; Chaturvedi, P.; Hubbi, M.E.; Wirtz, D.; Semenza, G.L. Hypoxia-inducible factors mediate coordinated RhoA-ROCK1 expression and signaling in breast cancer cells. Proc. Natl. Acad. Sci. USA 2014, 111, E384–E393. [Google Scholar] [CrossRef] [Green Version]
  177. Yuan, J.; Chen, L.; Xiao, J.; Qi, X.K.; Zhang, J.; Li, X.; Wang, Z.; Lian, Y.F.; Xiang, T.; Zhang, Y.; et al. SHROOM2 inhibits tumor metastasis through RhoA-ROCK pathway-dependent and -independent mechanisms in nasopharyngeal carcinoma. Cell Death Dis. 2019, 10, 58. [Google Scholar] [CrossRef]
  178. Liu, J.; Seibold, S.A.; Rieke, C.J.; Song, I.; Cukier, R.I.; Smith, W.L. Prostaglandin endoperoxide H synthases: Peroxidase hydroperoxide specificity and cyclooxygenase activation. J. Biol. Chem. 2007, 282, 18233–18244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Che, D.; Zhang, S.; Jing, Z.; Shang, L.; Jin, S.; Liu, F.; Shen, J.; Li, Y.; Hu, J.; Meng, Q.; et al. Macrophages induce EMT to promote invasion of lung cancer cells through the IL-6-mediated COX-2/PGE2/beta-catenin signalling pathway. Mol. Immunol. 2017, 90, 197–210. [Google Scholar] [CrossRef] [PubMed]
  180. Li, S.; Xu, X.; Jiang, M.; Bi, Y.; Xu, J.; Han, M. Lipopolysaccharide induces inflammation and facilitates lung metastasis in a breast cancer model via the prostaglandin E2-EP2 pathway. Mol. Med. Rep. 2015, 11, 4454–4462. [Google Scholar] [CrossRef] [PubMed]
  181. Ke, J.; Yang, Y.; Che, Q.; Jiang, F.; Wang, H.; Chen, Z.; Zhu, M.; Tong, H.; Zhang, H.; Yan, X.; et al. Prostaglandin E2 (PGE2) promotes proliferation and invasion by enhancing SUMO-1 activity via EP4 receptor in endometrial cancer. Tumour Biol. 2016, 37, 12203–12211. [Google Scholar] [CrossRef] [Green Version]
  182. Reyes-Reyes, E.M.; Ramos, I.N.; Tavera-Garcia, M.A.; Ramos, K.S. The aryl hydrocarbon receptor agonist benzo(a)pyrene reactivates LINE-1 in HepG2 cells through canonical TGF-beta1 signaling: Implications in hepatocellular carcinogenesis. Am. J. Cancer Res. 2016, 6, 1066–1077. [Google Scholar]
  183. Warusavitarne, J.; McDougall, F.; de Silva, K.; Barnetson, R.; Messina, M.; Robinson, B.G.; Schnitzler, M. Restoring TGFbeta function in microsatellite unstable (MSI-H) colorectal cancer reduces tumourigenicity but increases metastasis formation. Int. J. Colorectal Dis. 2009, 24, 139–144. [Google Scholar] [CrossRef]
  184. Singh, A.; Settleman, J. EMT, cancer stem cells and drug resistance: An emerging axis of evil in the war on cancer. Oncogene 2010, 29, 4741–4751. [Google Scholar] [CrossRef] [Green Version]
  185. Jang, G.B.; Kim, J.Y.; Cho, S.D.; Park, K.S.; Jung, J.Y.; Lee, H.Y.; Hong, I.S.; Nam, J.S. Blockade of Wnt/beta-catenin signaling suppresses breast cancer metastasis by inhibiting CSC-like phenotype. Sci. Rep. 2015, 5, 12465. [Google Scholar] [CrossRef] [Green Version]
  186. Kwon, Y.J.; Baek, H.S.; Ye, D.J.; Shin, S.; Kim, D.; Chun, Y.J. CYP1B1 Enhances Cell Proliferation and Metastasis through Induction of EMT and Activation of Wnt/beta-Catenin Signaling via Sp1 Upregulation. PLoS ONE 2016, 11, e0151598. [Google Scholar] [CrossRef] [Green Version]
  187. Mengoni, M.; Braun, A.D.; Gaffal, E.; Tuting, T. The aryl hydrocarbon receptor promotes inflammation-induced dedifferentiation and systemic metastatic spread of melanoma cells. Int. J. Cancer 2020, 147, 2902–2913. [Google Scholar] [CrossRef]
  188. Murante, F.G.; Gasiewicz, T.A. Hemopoietic progenitor cells are sensitive targets of 2,3,7,8-tetrachlorodibenzo-p-dioxin in C57BL/6J mice. Toxicol. Sci. 2000, 54, 374–383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Singh, K.P.; Casado, F.L.; Opanashuk, L.A.; Gasiewicz, T.A. The aryl hydrocarbon receptor has a normal function in the regulation of hematopoietic and other stem/progenitor cell populations. Biochem. Pharmacol. 2009, 77, 577–587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Singh, K.P.; Garrett, R.W.; Casado, F.L.; Gasiewicz, T.A. Aryl hydrocarbon receptor-null allele mice have hematopoietic stem/progenitor cells with abnormal characteristics and functions. Stem Cells Dev. 2011, 20, 769–784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Thurmond, T.S.; Silverstone, A.E.; Baggs, R.B.; Quimby, F.W.; Staples, J.E.; Gasiewicz, T.A. A chimeric aryl hydrocarbon receptor knockout mouse model indicates that aryl hydrocarbon receptor activation in hematopoietic cells contributes to the hepatic lesions induced by 2,3,7, 8- tetrachlorodibenzo-p-dioxin. Toxicol. Appl. Pharmacol. 1999, 158, 33–40. [Google Scholar] [CrossRef]
  192. Singh, K.P.; Bennett, J.A.; Casado, F.L.; Walrath, J.L.; Welle, S.L.; Gasiewicz, T.A. Loss of aryl hydrocarbon receptor promotes gene changes associated with premature hematopoietic stem cell exhaustion and development of a myeloproliferative disorder in aging mice. Stem Cells Dev. 2014, 23, 95–106. [Google Scholar] [CrossRef] [Green Version]
  193. Ko, C.I.; Wang, Q.; Fan, Y.; Xia, Y.; Puga, A. Pluripotency factors and Polycomb Group proteins repress aryl hydrocarbon receptor expression in murine embryonic stem cells. Stem Cell Res. 2014, 12, 296–308. [Google Scholar] [CrossRef] [Green Version]
  194. Boitano, A.E.; Wang, J.; Romeo, R.; Bouchez, L.C.; Parker, A.E.; Sutton, S.E.; Walker, J.R.; Flaveny, C.A.; Perdew, G.H.; Denison, M.S.; et al. Aryl hydrocarbon receptor antagonists promote the expansion of human hematopoietic stem cells. Science 2010, 329, 1345–1348. [Google Scholar] [CrossRef] [Green Version]
  195. Wagner, J.E., Jr.; Brunstein, C.G.; Boitano, A.E.; DeFor, T.E.; McKenna, D.; Sumstad, D.; Blazar, B.R.; Tolar, J.; Le, C.; Jones, J.; et al. Phase I/II Trial of StemRegenin-1 Expanded Umbilical Cord Blood Hematopoietic Stem Cells Supports Testing as a Stand-Alone Graft. Cell Stem Cell 2016, 18, 144–155. [Google Scholar] [CrossRef] [Green Version]
  196. Tan, K.P.; Wang, B.; Yang, M.; Boutros, P.C.; Macaulay, J.; Xu, H.; Chuang, A.I.; Kosuge, K.; Yamamoto, M.; Takahashi, S.; et al. Aryl hydrocarbon receptor is a transcriptional activator of the human breast cancer resistance protein (BCRP/ABCG2). Mol. Pharmacol. 2010, 78, 175–185. [Google Scholar] [CrossRef] [Green Version]
  197. Wu, C.; Yu, S.; Tan, Q.; Guo, P.; Liu, H. Role of AhR in regulating cancer stem cell-like characteristics in choriocarcinoma. Cell Cycle 2018, 17, 2309–2320. [Google Scholar] [CrossRef] [Green Version]
  198. Yan, B.; Liu, S.; Shi, Y.; Liu, N.; Chen, L.; Wang, X.; Xiao, D.; Liu, X.; Mao, C.; Jiang, Y.; et al. Activation of AhR with nuclear IKKalpha regulates cancer stem-like properties in the occurrence of radioresistance. Cell Death Dis. 2018, 9, 490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Al-Dhfyan, A.; Alhoshani, A.; Korashy, H.M. Aryl hydrocarbon receptor/cytochrome P450 1A1 pathway mediates breast cancer stem cells expansion through PTEN inhibition and beta-Catenin and Akt activation. Mol. Cancer 2017, 16, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Clark, D.W.; Palle, K. Aldehyde dehydrogenases in cancer stem cells: Potential as therapeutic targets. Ann. Transl. Med. 2016, 4, 518. [Google Scholar] [CrossRef] [PubMed]
  201. Charafe-Jauffret, E.; Ginestier, C.; Iovino, F.; Tarpin, C.; Diebel, M.; Esterni, B.; Houvenaeghel, G.; Extra, J.M.; Bertucci, F.; Jacquemier, J.; et al. Aldehyde dehydrogenase 1-positive cancer stem cells mediate metastasis and poor clinical outcome in inflammatory breast cancer. Clin. Cancer Res. 2010, 16, 45–55. [Google Scholar] [CrossRef] [Green Version]
  202. Ginestier, C.; Hur, M.H.; Charafe-Jauffret, E.; Monville, F.; Dutcher, J.; Brown, M.; Jacquemier, J.; Viens, P.; Kleer, C.G.; Liu, S.; et al. ALDH1 is a marker of normal and malignant human mammary stem cells and a predictor of poor clinical outcome. Cell Stem Cell 2007, 1, 555–567. [Google Scholar] [CrossRef] [Green Version]
  203. Oberg, M.; Bergander, L.; Hakansson, H.; Rannug, U.; Rannug, A. Identification of the tryptophan photoproduct 6-formylindolo[3,2-b]carbazole, in cell culture medium, as a factor that controls the background aryl hydrocarbon receptor activity. Toxicol. Sci. 2005, 85, 935–943. [Google Scholar] [CrossRef] [Green Version]
  204. Wincent, E.; Bengtsson, J.; Mohammadi Bardbori, A.; Alsberg, T.; Luecke, S.; Rannug, U.; Rannug, A. Inhibition of cytochrome P4501-dependent clearance of the endogenous agonist FICZ as a mechanism for activation of the aryl hydrocarbon receptor. Proc. Natl. Acad. Sci. USA 2012, 109, 4479–4484. [Google Scholar] [CrossRef] [Green Version]
  205. Sasaki-Kudoh, E.; Kudo, I.; Kakizaki, Y.; Hosaka, M.; Ikeda, S.; Uemura, S.; Grave, E.; Togashi, S.; Sugawara, T.; Shimizu, H.; et al. Cisplatin Inhibits AhR Activation. Am. J. Mol. Biol. 2018, 8, 69–82. [Google Scholar] [CrossRef] [Green Version]
  206. Gao, W.Q.; Ma, J.; Sun, L.L.; Li, Q.; Zhu, R.Y.; Jin, J. Paclitaxel-mediated human aryl hydrocarbon receptor mRNA translation by an internal ribosomal entry site-dependent mechanism. Oncol. Rep. 2017, 38, 3211–3219. [Google Scholar] [CrossRef]
  207. Neuzil, J.; Stantic, M.; Zobalova, R.; Chladova, J.; Wang, X.; Prochazka, L.; Dong, L.; Andera, L.; Ralph, S.J. Tumour-initiating cells vs. cancer ’stem’ cells and CD133: What’s in the name? Biochem. Biophys. Res. Commun. 2007, 355, 855–859. [Google Scholar] [CrossRef]
  208. Ly, M.; Rentas, S.; Vujovic, A.; Wong, N.; Moreira, S.; Xu, J.; Holzapfel, N.; Bhatia, S.; Tran, D.; Minden, M.D.; et al. Diminished AHR Signaling Drives Human Acute Myeloid Leukemia Stem Cell Maintenance. Cancer Res. 2019, 79, 5799–5811. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Murray, I.A.; Patterson, A.D.; Perdew, G.H. Aryl hydrocarbon receptor ligands in cancer: Friend and foe. Nat. Rev. Cancer 2014, 14, 801–814. [Google Scholar] [CrossRef] [PubMed]
  210. Kolluri, S.K.; Jin, U.H.; Safe, S. Role of the aryl hydrocarbon receptor in carcinogenesis and potential as an anti-cancer drug target. Arch. Toxicol. 2017, 91, 2497–2513. [Google Scholar] [CrossRef] [PubMed]
  211. Murray, I.A.; Morales, J.L.; Flaveny, C.A.; Dinatale, B.C.; Chiaro, C.; Gowdahalli, K.; Amin, S.; Perdew, G.H. Evidence for ligand-mediated selective modulation of aryl hydrocarbon receptor activity. Mol. Pharmacol. 2010, 77, 247–254. [Google Scholar] [CrossRef] [Green Version]
  212. Zhao, B.; Degroot, D.E.; Hayashi, A.; He, G.; Denison, M.S. CH223191 is a ligand-selective antagonist of the Ah (Dioxin) receptor. Toxicol. Sci. 2010, 117, 393–403. [Google Scholar] [CrossRef] [Green Version]
  213. Safe, S. Molecular biology of the Ah receptor and its role in carcinogenesis. Toxicol. Lett. 2001, 120, 1–7. [Google Scholar] [CrossRef]
  214. Safe, S.; Cheng, Y.; Jin, U.H. The Aryl Hydrocarbon Receptor (AhR) as a Drug Target for Cancer Chemotherapy. Curr. Opin. Toxicol. 2017, 2, 24–29. [Google Scholar] [CrossRef]
  215. Safe, S.; Qin, C.; McDougal, A. Development of selective aryl hydrocarbon receptor modulators for treatment of breast cancer. Expert Opin. Investig. Drugs 1999, 8, 1385–1396. [Google Scholar] [CrossRef]
  216. O’Donnell, E.F.; Jang, H.S.; Pearce, M.; Kerkvliet, N.I.; Kolluri, S.K. The aryl hydrocarbon receptor is required for induction of p21cip1/waf1 expression and growth inhibition by SU5416 in hepatoma cells. Oncotarget 2017, 8, 25211–25225. [Google Scholar] [CrossRef] [Green Version]
  217. Jin, U.H.; Lee, S.O.; Pfent, C.; Safe, S. The aryl hydrocarbon receptor ligand omeprazole inhibits breast cancer cell invasion and metastasis. BMC Cancer 2014, 14, 498. [Google Scholar] [CrossRef] [Green Version]
  218. Schlezinger, J.J.; Liu, D.; Farago, M.; Seldin, D.C.; Belguise, K.; Sonenshein, G.E.; Sherr, D.H. A role for the aryl hydrocarbon receptor in mammary gland tumorigenesis. Biol. Chem. 2006, 387, 1175–1187. [Google Scholar] [CrossRef]
  219. Rhile, M.J.; Nagarkatti, M.; Nagarkatti, P.S. Role of Fas apoptosis and MHC genes in 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD)-induced immunotoxicity of T cells. Toxicology 1996, 110, 153–167. [Google Scholar] [CrossRef]
  220. Park, K.T.; Mitchell, K.A.; Huang, G.; Elferink, C.J. The aryl hydrocarbon receptor predisposes hepatocytes to Fas-mediated apoptosis. Mol. Pharmacol. 2005, 67, 612–622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. Robles, R.; Morita, Y.; Mann, K.K.; Perez, G.I.; Yang, S.; Matikainen, T.; Sherr, D.H.; Tilly, J.L. The aryl hydrocarbon receptor, a basic helix-loop-helix transcription factor of the PAS gene family, is required for normal ovarian germ cell dynamics in the mouse. Endocrinology 2000, 141, 450–453. [Google Scholar] [CrossRef] [PubMed]
  222. Vogel, C.F.; Li, W.; Sciullo, E.; Newman, J.; Hammock, B.; Reader, J.R.; Tuscano, J.; Matsumura, F. Pathogenesis of aryl hydrocarbon receptor-mediated development of lymphoma is associated with increased cyclooxygenase-2 expression. Am. J. Pathol. 2007, 171, 1538–1548. [Google Scholar] [CrossRef] [Green Version]
  223. Bekki, K.; Vogel, H.; Li, W.; Ito, T.; Sweeney, C.; Haarmann-Stemmann, T.; Matsumura, F.; Vogel, C.F. The aryl hydrocarbon receptor (AhR) mediates resistance to apoptosis induced in breast cancer cells. Pestic. Biochem. Physiol. 2015, 120, 5–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Fujisawa, Y.; Li, W.; Wu, D.; Wong, P.; Vogel, C.; Dong, B.; Kung, H.J.; Matsumura, F. Ligand-independent activation of the arylhydrocarbon receptor by ETK (Bmx) tyrosine kinase helps MCF10AT1 breast cancer cells to survive in an apoptosis-inducing environment. Biol. Chem. 2011, 392, 897–908. [Google Scholar] [CrossRef] [PubMed]
  225. Vogeley, C.; Esser, C.; Tuting, T.; Krutmann, J.; Haarmann-Stemmann, T. Role of the Aryl Hydrocarbon Receptor in Environmentally Induced Skin Aging and Skin Carcinogenesis. Int. J. Mol. Sci. 2019, 20, 6005. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Frauenstein, K.; Sydlik, U.; Tigges, J.; Majora, M.; Wiek, C.; Hanenberg, H.; Abel, J.; Esser, C.; Fritsche, E.; Krutmann, J.; et al. Evidence for a novel anti-apoptotic pathway in human keratinocytes involving the aryl hydrocarbon receptor, E2F1, and checkpoint kinase 1. Cell Death Differ. 2013, 20, 1425–1434. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Pollet, M.; Shaik, S.; Mescher, M.; Frauenstein, K.; Tigges, J.; Braun, S.A.; Sondenheimer, K.; Kaveh, M.; Bruhs, A.; Meller, S.; et al. The AHR represses nucleotide excision repair and apoptosis and contributes to UV-induced skin carcinogenesis. Cell Death Differ. 2018, 25, 1823–1836. [Google Scholar] [CrossRef] [Green Version]
  228. Leja-Szpak, A.; Goralska, M.; Link-Lenczowski, P.; Czech, U.; Nawrot-Porabka, K.; Bonior, J.; Jaworek, J. The Opposite Effect of L-kynurenine and Ahr Inhibitor Ch223191 on Apoptotic Protein Expression in Pancreatic Carcinoma Cells (Panc-1). Anticancer Agents Med. Chem. 2019, 19, 2079–2090. [Google Scholar] [CrossRef] [PubMed]
  229. Flaveny, C.A.; Perdew, G.H. Transgenic Humanized AHR Mouse Reveals Differences between Human and Mouse AHR Ligand Selectivity. Mol. Cell Pharmacol. 2009, 1, 119–123. [Google Scholar] [CrossRef] [PubMed]
  230. Moriguchi, T.; Motohashi, H.; Hosoya, T.; Nakajima, O.; Takahashi, S.; Ohsako, S.; Aoki, Y.; Nishimura, N.; Tohyama, C.; Fujii-Kuriyama, Y.; et al. Distinct response to dioxin in an arylhydrocarbon receptor (AHR)-humanized mouse. Proc. Natl. Acad. Sci. USA 2003, 100, 5652–5657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Flaveny, C.; Reen, R.K.; Kusnadi, A.; Perdew, G.H. The mouse and human Ah receptor differ in recognition of LXXLL motifs. Arch. Biochem. Biophys. 2008, 471, 215–223. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Flaveny, C.A.; Murray, I.A.; Perdew, G.H. Differential Gene Regulation by the Human and Mouse Aryl Hydrocarbon Receptor. Toxicol. Sci. 2009, 114, 217–225. [Google Scholar] [CrossRef] [Green Version]
  233. Hubbard, T.D.; Murray, I.A.; Bisson, W.H.; Lahoti, T.S.; Gowda, K.; Amin, S.G.; Patterson, A.D.; Perdew, G.H. Adaptation of the human aryl hydrocarbon receptor to sense microbiota-derived indoles. Sci. Rep. 2015, 5, 12689. [Google Scholar] [CrossRef]
  234. Sanjana, N.E.; Shalem, O.; Zhang, F. Improved vectors and genome-wide libraries for CRISPR screening. Nat. Methods 2014, 11, 783–784. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The Transcriptional AHR Signaling Pathway. Cytoplasmic aryl hydrocarbon receptor (AHR) exists in a complex with HSP90, Immunophilin-like Ah Receptor-interacting Protein (AIP), p23, and Src 40. Once engaged with exogenous or endogenous ligands, the AHR sheds AIP and Src and translocates to the nucleus. In the nucleus the AHR dimerizes with the Aryl Hydrocarbon Nuclear Translocator (ARNT), binds to consensus Aryl Hydrocarbon Response Elements (AHREs), recruits coactivators, and transactivates a battery of genes including the hydroxylases CYP1B1 and CYP1A1, which metabolize some environmental AHR ligands into mutagenic epoxide intermediates. (Alternative AHR complexes containing NF-κB subunits, KLF6, or potentially other proteins bind to hybrid (AHRE/NFκB sites) or alternative DNA sequences to activate different sets of AHR responsive genes). CYP1B1 degrades at least some endogenous and exogenous AHR ligands in a negative feedback loop. The AHR complex can also induce IDO1/2 and/or TDO dioxygenases, which metabolize tryptophan into endogenous AHR ligands including, but not limited to, Kyn (kynurenine), in a positive feedback loop (see Section 5). Distinct sets of genes are activated by different AHR ligands, likely a result of differential co-factor recruitment. The AHR also functions through non-transcriptional pathways not represented here.
Figure 1. The Transcriptional AHR Signaling Pathway. Cytoplasmic aryl hydrocarbon receptor (AHR) exists in a complex with HSP90, Immunophilin-like Ah Receptor-interacting Protein (AIP), p23, and Src 40. Once engaged with exogenous or endogenous ligands, the AHR sheds AIP and Src and translocates to the nucleus. In the nucleus the AHR dimerizes with the Aryl Hydrocarbon Nuclear Translocator (ARNT), binds to consensus Aryl Hydrocarbon Response Elements (AHREs), recruits coactivators, and transactivates a battery of genes including the hydroxylases CYP1B1 and CYP1A1, which metabolize some environmental AHR ligands into mutagenic epoxide intermediates. (Alternative AHR complexes containing NF-κB subunits, KLF6, or potentially other proteins bind to hybrid (AHRE/NFκB sites) or alternative DNA sequences to activate different sets of AHR responsive genes). CYP1B1 degrades at least some endogenous and exogenous AHR ligands in a negative feedback loop. The AHR complex can also induce IDO1/2 and/or TDO dioxygenases, which metabolize tryptophan into endogenous AHR ligands including, but not limited to, Kyn (kynurenine), in a positive feedback loop (see Section 5). Distinct sets of genes are activated by different AHR ligands, likely a result of differential co-factor recruitment. The AHR also functions through non-transcriptional pathways not represented here.
Ijms 22 00387 g001
Figure 2. Nuclear AHR expression in eight human cancers. Immunohistochemistry images were obtained from the Human Protein Atlas. Nuclear staining for the AHR is evident in all eight cancer types.
Figure 2. Nuclear AHR expression in eight human cancers. Immunohistochemistry images were obtained from the Human Protein Atlas. Nuclear staining for the AHR is evident in all eight cancer types.
Ijms 22 00387 g002
Figure 3. A 644 AHR-driven gene signature correlates with survival in eight human cancers. (a) Top: Western blotting for AHR protein in AHR knockout (KO), wildtype (WT) or Cas9 control SUM149 triple negative inflammatory human breast cancer cells or MDA-MB-231 triple negative human breast cancer cells. Bottom: Quantification of β-actin normalized band densities. Data are presented as means + SE from three independent experiments. No significant differences (ns) were found between WT and Cas9 (transfected with Cas9 but not guide RNA) control cells. **** p <0.00001 compared to WT or Cas9 controls. (b) Affymetrix whole human genome microarrays were used to determine transcriptional profiles of Cas9 control or AHR knockout SUM149 or MDA-MB-231 cells. Genes differentially expressed upon AHR knockout were identified across cell types. A significant decrease after AHR knockout was defined as an adjusted false discovery rate (q value) of ≤0.05. (c) The 644 AHR-driven gene signature in “b” was correlated with survival for the eight cancers in Figure 1 using TCGA data and Gene Set Variation Analysis (129).
Figure 3. A 644 AHR-driven gene signature correlates with survival in eight human cancers. (a) Top: Western blotting for AHR protein in AHR knockout (KO), wildtype (WT) or Cas9 control SUM149 triple negative inflammatory human breast cancer cells or MDA-MB-231 triple negative human breast cancer cells. Bottom: Quantification of β-actin normalized band densities. Data are presented as means + SE from three independent experiments. No significant differences (ns) were found between WT and Cas9 (transfected with Cas9 but not guide RNA) control cells. **** p <0.00001 compared to WT or Cas9 controls. (b) Affymetrix whole human genome microarrays were used to determine transcriptional profiles of Cas9 control or AHR knockout SUM149 or MDA-MB-231 cells. Genes differentially expressed upon AHR knockout were identified across cell types. A significant decrease after AHR knockout was defined as an adjusted false discovery rate (q value) of ≤0.05. (c) The 644 AHR-driven gene signature in “b” was correlated with survival for the eight cancers in Figure 1 using TCGA data and Gene Set Variation Analysis (129).
Ijms 22 00387 g003
Figure 4. AHR Regulatory Circuit. Tryptophan is metabolized to Kynurenine (Kyn) and downstream AHR ligands that drive AHR activity and cancer aggression (migration, invasion, metastasis, “stem-ness”). Activated, nuclear AHR upregulates Ido1, Ido2, and/or Tdo2, producing more AHR ligands in an amplification loop that sustains AHR activity. AHR induces CYP1A1 and CYP1B1 (hydroxylases that can metabolize some AHR ligands) and AHR repressor (which limits AHR activity). Environmental, microbial or dietary AHR ligands may “prime the pump” and/or exacerbate the endogenous AHR signaling circuit.
Figure 4. AHR Regulatory Circuit. Tryptophan is metabolized to Kynurenine (Kyn) and downstream AHR ligands that drive AHR activity and cancer aggression (migration, invasion, metastasis, “stem-ness”). Activated, nuclear AHR upregulates Ido1, Ido2, and/or Tdo2, producing more AHR ligands in an amplification loop that sustains AHR activity. AHR induces CYP1A1 and CYP1B1 (hydroxylases that can metabolize some AHR ligands) and AHR repressor (which limits AHR activity). Environmental, microbial or dietary AHR ligands may “prime the pump” and/or exacerbate the endogenous AHR signaling circuit.
Ijms 22 00387 g004
Figure 5. AHR Knockout Reduces and CYP1B1 Knockout Enhances Baseline and Ligand-induced AHR Activity. The AHR or CYP1B1 was deleted from triple negative, inflammatory human breast cancer cells [162]. (a) Western blotting showing the absence of detectable CYP1B1. (See Figure 3 for AHR knockout western blots). (b) CMV-driven GFP-normalized, AHR-driven pGudLuc reporter activity was assayed in wildtype, Cas9 control, AHR knockout, or CYP1B1 knockout SUM149 cells treated for 24 hours with 0.1% DMSO (vehicle), 10 uM B[a]P, 0.5 uM FICZ, 100 uM kynurenine (Kyn), or 100 uM xanthurenic acid (XA). Data are presented as normalized means + SE. from a minimum of three experiments. ++ p < 0.01 relative to similarly treated Cas9 control. *** p < 0.001 relative to similarly treated Cas9 control. **** p < 0.0001 relative to similarly treated Cas9 control.
Figure 5. AHR Knockout Reduces and CYP1B1 Knockout Enhances Baseline and Ligand-induced AHR Activity. The AHR or CYP1B1 was deleted from triple negative, inflammatory human breast cancer cells [162]. (a) Western blotting showing the absence of detectable CYP1B1. (See Figure 3 for AHR knockout western blots). (b) CMV-driven GFP-normalized, AHR-driven pGudLuc reporter activity was assayed in wildtype, Cas9 control, AHR knockout, or CYP1B1 knockout SUM149 cells treated for 24 hours with 0.1% DMSO (vehicle), 10 uM B[a]P, 0.5 uM FICZ, 100 uM kynurenine (Kyn), or 100 uM xanthurenic acid (XA). Data are presented as normalized means + SE. from a minimum of three experiments. ++ p < 0.01 relative to similarly treated Cas9 control. *** p < 0.001 relative to similarly treated Cas9 control. **** p < 0.0001 relative to similarly treated Cas9 control.
Ijms 22 00387 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, Z.; Snyder, M.; Kenison, J.E.; Yang, K.; Lara, B.; Lydell, E.; Bennani, K.; Novikov, O.; Federico, A.; Monti, S.; et al. How the AHR Became Important in Cancer: The Role of Chronically Active AHR in Cancer Aggression. Int. J. Mol. Sci. 2021, 22, 387. https://doi.org/10.3390/ijms22010387

AMA Style

Wang Z, Snyder M, Kenison JE, Yang K, Lara B, Lydell E, Bennani K, Novikov O, Federico A, Monti S, et al. How the AHR Became Important in Cancer: The Role of Chronically Active AHR in Cancer Aggression. International Journal of Molecular Sciences. 2021; 22(1):387. https://doi.org/10.3390/ijms22010387

Chicago/Turabian Style

Wang, Zhongyan, Megan Snyder, Jessica E. Kenison, Kangkang Yang, Brian Lara, Emily Lydell, Kawtar Bennani, Olga Novikov, Anthony Federico, Stefano Monti, and et al. 2021. "How the AHR Became Important in Cancer: The Role of Chronically Active AHR in Cancer Aggression" International Journal of Molecular Sciences 22, no. 1: 387. https://doi.org/10.3390/ijms22010387

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop