Next Article in Journal
Characterization of GLOD4 in Leydig Cells of Tibetan Sheep during Different Stages of Maturity
Previous Article in Journal
An SNP-Based Genetic Map and QTL Mapping for Growth Traits in the Red-Spotted Grouper (Epinephelus akaara)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Transcription Factors That Govern Development and Disease: An Achilles Heel in Cancer

by
Dhananjay Huilgol
,
Prabhadevi Venkataramani
,
Saikat Nandi
* and
Sonali Bhattacharjee
*
Bungtown Road, Cold Spring Harbor Laboratory, Cold Spring Harbor, New York, NY 11724, USA
*
Authors to whom correspondence should be addressed.
Genes 2019, 10(10), 794; https://doi.org/10.3390/genes10100794
Submission received: 15 September 2019 / Revised: 5 October 2019 / Accepted: 8 October 2019 / Published: 12 October 2019
(This article belongs to the Section Human Genomics and Genetic Diseases)

Abstract

:
Development requires the careful orchestration of several biological events in order to create any structure and, eventually, to build an entire organism. On the other hand, the fate transformation of terminally differentiated cells is a consequence of erroneous development, and ultimately leads to cancer. In this review, we elaborate how development and cancer share several biological processes, including molecular controls. Transcription factors (TF) are at the helm of both these processes, among many others, and are evolutionarily conserved, ranging from yeast to humans. Here, we discuss four families of TFs that play a pivotal role and have been studied extensively in both embryonic development and cancer—high mobility group box (HMG), GATA, paired box (PAX) and basic helix-loop-helix (bHLH) in the context of their role in development, cancer, and their conservation across several species. Finally, we review TFs as possible therapeutic targets for cancer and reflect on the importance of natural resistance against cancer in certain organisms, yielding knowledge regarding TF function and cancer biology.

1. Introduction

1.1. Embryonic Development and Cancer: Two Sides of the Same Coin

Embryonic development involves a mass of cells achieving specific cell identities depending on morphogen gradients and the activation of transcription factors (TFs). These genetic changes propel ‘stem cells’ to form terminally differentiated cell types. In cancer, a terminally differentiated cell undergoes dedifferentiation to a stem cell, following which it assumes a new differentiated identity [1]. Interestingly, the cellular and molecular mechanisms are also quite conserved since tumorigenesis is caused by the reactivation of repressed genes [1]. If development is orderliness and regulation, cancer is deregulation. While in cancer, the accumulation of mutations in the genome leads to uncontrolled proliferation and the misdirected establishment of cell identity, embryonic development involves stem cell proliferation, fate specification, and migration must be streamlined in order to ‘assemble’ an organism. Studying embryonic tumors provides a case study as to how embryonic development can progress to cancer. Some examples include retinoblastoma (Rb), neuroblastoma, and nephroblastoma. An error in stem cell differentiation leads to each of these cancers. RB1, a tumor suppressor and well-known cell cycle regulator, has mutations in Rb [2]. However, a significant number of carcinomas are difficult to study since they are embedded amid a large population of differentiated cells [3].
Interestingly, a number of regulators of gene expression are also used as markers for cancer detection. Octamer binding transcription factor 4 (Oct4) is expressed in pluripotent stem cells and is necessary for controlling their pluripotency and self-renewal from gastrulation [4]. Intriguingly, it is also an established marker of tumor initiating cells (TIC) and embryonic carcinoma cells [5,6,7,8,9]. Oct4 plays an important role in the initiation and progression of cancer. The overexpression of Oct4, along with Sox2 and Nanog, has been shown to cause the dedifferentiation of glioblastoma multiforme cell lines into induced glioma stem cells [10,11]. In addition, OCT4 overexpression in human metastatic melanomas causes a loss of melanoma markers and increases cell migration [12,13]. RNAi mediated knockdown of Oct4 leads to reduced TIC phenotypes [13]. In lung adenocarcinoma, Oct4 along with Nanog knockdown reversed the epithelial-to-mesenchymal transition (EMT) and inhibited tumorigenesis and metastasis [14].
Similarly, Myc is also known as a tumor initiation and maintenance factor. In vitro, Myc knockdown in cancer cell lines decreased cell proliferation and, in some cases, induced apoptosis [15,16,17]. Myc overexpression in an embryonic liver leads to proliferation, whereas in an adult liver, it leads to polyploidy and cell growth [18]. These results have been reported in a large number of tumors including epithelial tumors (hepatocellular, breast, squamous carcinoma), hematopoietic tumors (T- and B-cell lymphoma and leukemia) and mesenchymal tumors (osteogenic sarcoma) [19,20,21,22] (Figure 1).

1.2. Epithelial-to-Mesenchymal Transition: in Development and Cancer

Epithelial-to-Mesenchymal Transition (EMT) refers to the fate transformation of a cell from a stable, stationary epithelial cell to a more migratory mesenchymal cell that is resistant to apoptosis [23,24]. This process is as important in implantation, organ development, and embryo formation as in neoplastic transformation [25]. Placental formation, the initiation of the primitive streak, and gastrulation leading to the separation of three germinal layers all involve EMT [26]. Wnt-signaling is particularly important for EMT during these processes—Wnt3 for EMT during gastrulation and Wnt8c for the formation of the primitive streak [27]. Wnt molecules interact with other pathways such as TGF- and FGF receptors in order to regulate EMT during such developmental processes. The development of one of the best studied migratory cells in the embryo, neural crest cells, also requires EMT from the neuroectoderm [28]. These cells travel to different parts of the embryo and contribute to facial musculature and melanocytes for skin pigmentation, among others. Along with signaling through Wnt, Bone Morphogenetic Proteins (BMPs), Fibroblast Growth Factor (FGF), and c-Myb pathways, neural crest cells need to downregulate E-cadherin and N-cadherin expression for migration [29,30,31,32,33,34].
Multiple cancers are marked by the overproliferation of epithelial cells and angiogenesis followed by invasion through the basement membrane [35]. Malignancy is the final stage of cancer cell migration to distant sites. EMT has been shown to be a critical mechanism for the spread of epithelial malignancies. The expression of mesenchymal markers, as is the case with neural crest cells, is a hallmark of EMT in cancer [35]. A spectrum of signaling mechanisms involved in development are implicated in carcinomas. TFs such as Snail, Slug, Twist, and FoxD3 are essential for EMT both in development and cancer progression. Wnt, bone morphogenetic proteins (BMPs), and fibroblast growth factor (FGF) signaling, along with the loss of E-cadherin by epithelial cells, makes physiological processes in embryogenesis and carcinoma metastasis nearly identical [36,37] (Figure 1).
While during embryogenesis, the remodeling and diversification of tissues proceeds to generate a fully functional organism, mutations in the DNA facilitate EMT in cancer and lead to invasion and metastasis [38].

1.3. Cell Migration: Essential for Development and Cancer Progression

Placental or blastocyst cells invading the uterine endometrium and cancer cells invading the juxtaposed epithelial or endothelial cells use similar cellular mechanisms. It is a multistep process including apposition, adherence or attachment, and eventually differentiation following invasion. Angiogenesis is a key process established after invasion and inflammation that ultimately provides nutrition to invading cells [39,40]. As expected, non-classical HLA class I antigens are recruited for both embryonic development and cancer leading to the recognition of both these tissues as self [39,41]. In tumors, and during the second trimester of pregnancy, a TH-2 type anti-inflammatory immune response is initiated in order to fuel tumor growth and provide sustenance for pregnancy, respectively [42,43]. FoxP3, a TF necessary and sufficient for suppressing the immune functions in regulatory T-cells [44], is also known to regulate the differentiation of uterine T-cells into regulatory T-cells [45,46,47,48,49]. Infertility is a consequence of the absence, or reduced expression of FoxP3 [50]. Incidentally, FoxP3 also happens to be a tumor suppressor gene in breast and prostate cancers [51,52]. The misexpression of FoxP3, owing to its location on the X-chromosome can lead to carcinogenesis [52]. Morphogens, TFs, epigenetic factors, and their downstream signaling cascades interact within a cell. We will mention how each of these contribute individually to development and cancer progression in the next section.
A wide range of morphogens such as Wnt, Hedgehog (HH), BMPs, and FGFs are essential for the patterning and development of a range of structures including limbs, the heart, and central nervous system. A reactivation of these pathways has been observed in tumorigenesis and metastasis. For example, the Wnt pathway is necessary for patterning, fate specification and progenitor maintenance [53]. Mutations in the tumor suppressor gene, adenomatous polyposis coli, can lead to colorectal cancer [54]. Ovarian cancers and hepatoblastoma also show -catenin overexpression, which is downstream of Wnt signaling [55,56]. Downstream players at the time of attachment include cell adhesion molecules such as integrins, immunoglobulin-CAMs, selectins, and cadherins. E-cadherins are important for both mammary gland development as well as mammary tumors [57,58].
Epigenetic changes are not just markers of embryonic development and cancer, they also play a major role in both hematopoiesis and the progression of hematological cancer. DNA methyltransferases are particularly high in embryos [59] and tumors [60,61]. Ten eleven translocation (TET) enzymes are multi-domain enzymes, important for regulating DNA methylation [62]. They are highly expressed in blastocysts at the time of attachment and invasion, and are essential for the survival of an organism [63,64]. TET enzymes are essential for DNA repair and chromosomal translocations in normal physiological conditions and therefore prevent carcinogenesis. Mutations in these enzymes have been implicated in myeloid leukemia [65] and late onset B-cell lymphoma [66,67]. TET2 was identified as a tumor suppressor and mutations across the gene have been implicated in acute myeloid leukemia (AML), myeloproliferative neoplasms (MPN), and myelodysplastic syndrome (MDS) [68].
TFs are at the heart of influencing any cancer initiation and progression. The LIM-homeodomain TFs, reversion-induced Lim (RIL) or PDZ and LIM domain 4 (PDLIM4), promote apoptosis in cancer cells and, hence, are silenced epigenetically in AML and MDS [69], as well as breast cancer [70]. Earlier reviews have drawn comparisons between development and cancer with a focus on individual TF families such as GATA or Pax in the context of specific organs [71]. In this review, we introduce four of the many TF families that are essential for the regulation of different aspects of embryonic development as well as cancer; basic Helix-loop-Helix (bHLH), GATA, High Mobility Group box (HMG) and Paired box (Pax) TFs (Table 1). We have chosen these families since they have been extensively studied in both development and disease. Parallels between the role of TFs in development and cancer across multiple TF families, and spanning different organ systems, reveal a broad biological phenomenon, and therefore fundamental to understanding and eventually targeting cancer. We discuss the domain structure of the TFs and their role as regulators of development and cancer. We also highlight their role in development across different species to emphasize their evolutionary conservation. Transcriptional regulation is essential for almost every process in the existence of an organism. We will therefore discuss the role of this regulation in the prevention of cancer progression and conclude our review by discussing some of these molecules as candidates for therapeutics and lessons we can learn from species that are cancer resistant (Figure 1).

2. High Mobility Group Box (HMG)

The HMG-box domain was originally identified as a duplicated 80-amino acid L-shaped domain which binds the DNA minor groove and nucleosomes and, thereby, can induce structural changes in the chromatin fiber [72,73]. HMGs are the most abundant non-histone ubiquitous chromatin proteins in a cell and can be divided into three structurally distinct classes, namely HMG-nucleosome binding family (HMGN), HMG-AT-hook family (HMGA), and HMG-box family (HMGB) [72,74,75,76,77] (Table 1).

2.1. HMG Proteins: A Superfamily of Chromatin Remodelers

HMGN proteins are characterized by a bipartite nuclear localization signal (NLS), a nucleosome-binding domain (NBD) and an acidic C-terminal [78]. HMGN proteins bind specifically to nucleosome core particles to alter and regulate chromatin structure and function [78]. HMGA proteins contain three copies of a conserved DNA-binding peptide motif called the ‘AT-hook’ and an acidic C-terminal tail [79,80]. The AT-hook motif is positively charged and preferentially binds to the AT-rich sequence in the minor groove of DNA [81]. By binding the DNA, HMGA proteins can induce structural and/or conformational changes in the DNA, as well as promote the recruitment of additional components, most of which are TFs [82]. HMGB proteins are the most abundant in the family, with each mammalian nucleus containing approximately 105 to 106 molecules [83]. They interact with proteins implicated in a diverse range of DNA-dependent cellular processes, including DNA replication, recombination, the maintenance of genome integrity, and transposition, among others [84] (Table 1).

2.2. Role in Development

HMGN proteins are expressed in all embryonic tissues and are linked to differentiation [85,86]. HMGN1−/− mice are subfertile, hypersensitive to various stress conditions, such as exposure to UV light or ionizing irradiation (IR), and have slight defects in corneal epithelium development and maintenance [87,88,89]. HMGA proteins have a role in differentiation and spermatogenesis [90]. HMGA1−/− mice suffer from cardiac hypertrophy, hematological malignancies, and type 2 diabetes [91,92]. HMGA2−/− mice are pygmies characterized by reduced fat tissue and craniofacial defects [93,94]. HMGB genes are expressed in both embryonic and adult tissue [95]. HMGB1 and B2 proteins are important for neural stem cell proliferation, differentiation, and maintenance [96]. HMGB1−/− mice show defects in the activation of the glucocorticoid receptor and die within a day of birth due to hypoglycemia. HMGB2−/− mice show defects in spermatogenesis and the maintenance of articular cartilage homeostasis in adults, while HMGB3−/− mice exhibit erythrocythemia [97,98,99,100,101,102] (Table 1).

2.3. Evolutionary Conservation

The HMG boxes of these proteins are well conserved through evolution with homologs in plants, yeast, flies, worms, mammalian cell lines, and animals [192,193,194]. Despite the conservation of the primary sequence among members of this superfamily, several genetic mechanisms have resulted in structural and functional diversity within members [192,193,194]. Phylogenetic and sequence analyses have revealed three possible mechanisms for this divergence, namely (i) gene duplication from an ancient box, (ii) exon shuffling/intragenic duplications to explain why some members of the family carry several HMG boxes, and (iii) the slow accumulation of mutations in newly duplicated genes [192].

2.4. Role in Cancer

Tumor markers are detectable in body fluids such as blood serum and urine, and are powerful tools for cancer detection and prognosis. A change in the transcriptional profile of HMGs has been reported in several cancer types. HMGA1 is overexpressed in colon carcinoma, breast carcinomas, and invasive ovarian carcinomas, whereas it was not detectable in normal colon, breast, or ovarian tissue [106]. HMGA1 and HMGA2 are overexpressed in pancreatic adenocarcinomas and non-small cell lung carcinomas (NSCLC), in both squamous and adenocarcinoma histotypes [106]. HMGN1 regulates the transcription of proto-oncogenes and pro-metastatic genes like c-fos, BCL3, N-cadherin, JunB and c-Jun involved in tumor progression, in a way which may suppress the development of cancer [103]. The mRNA and protein expression levels of HMGB1 are increased in the lungs of patients with NSCLC, pancreatic ductal adenocarcinoma (PDAC), gastric cancer, colorectal cancer, hepatocellular carcinoma (HCC), and correlate with disease development, tumor progression, invasion, poor prognosis, and metastasis [107] (Table 1).

3. GATA Transcription Factors

The discovery of the GATA TF family has transformed the field of hematology. GATA1, the founding member of the GATA family, was initially described as a TF binding to DNA sites within the regulatory regions of several members of α and β-globin families in chickens. Also known as Eryf1, GATA1 was subsequently cloned, purified, and characterized as a ‘switch factor’ in erythroid development [195,196]. This led to the cloning of other members of the GATA family, GATA2 to GATA6. The GATA family shares two highly conserved C2H2-type zinc-finger motifs (Cys-X2-C-X17-Cys-X2-Cys (ZNI and ZNII)) that are involved in DNA-binding by recognizing the GATA element (A/TGATAA/G) [108]. GATA1, along with GATA2 and GATA3 are collectively grouped as a hematopoietic GATA subfamily, while GATA4, GATA5, and GATA6 are classified as an endodermal GATA subfamily [110,197] (Table 1).

3.1. Role in Development

GATA1 functions by promoting the development of erythrocytes, megakaryocytes, mast cells, and eosinophils [109,198,199]. The loss of GATA1 leads to a substantial increase in GATA2 expression, indicating that GATA1 not only suppresses GATA2 transcription during erythropoiesis, but is also partly compensated by GATA2. This phenomenon, also known as the ‘GATA switch’, is facilitated by the displacement of GATA2 from its enhancer by overexpressing GATA1 [111]. GATA3 plays a pivotal role in T-cell lymphopoiesis—from the generation of T-cell progenitors to CD4+ specification. GATA3 has also been shown to regulate the self-renewal and differentiation of long-term hematopoietic stem cells (HSCs) in the bone marrow [115,200,201]. A deficiency of GATA3 during embryogenesis drastically reduces HSC production in the aorta-gonads-mesonephros region [202].
GATA4, 5, and 6 are highly expressed in the mesoderm and endoderm-derived tissues such as the stomach, liver, heart, lung, and gonads. GATA4 induces angiogenic factors such as vascular endothelial growth factor (VEGF) to regulate cardiac angiogenesis by promoting compensation after injury [117]. Cyclin D2 and GATA4 have been shown to interact and form a positive feedback loop that enhances the cardiogenic activity of GATA4 [203]. Furthermore, GATA4 promotes bile absorption in the proximal ileum to restore bile homoeostasis [118]. In the developing heart, GATA5 is expressed in both the myocardium and endocardium of mouse embryos. The deletion of both the isoforms of GATA5 has been shown to result in hypoplastic hearts and partially penetrant bicuspid valve [121,122]. GATA6 has been demonstrated to play a role in the proper patterning of the aortic arch arteries, liver bud growth, and commitment of the endoderm to a hepatic cell fate [124,204]. GATA6, along with its target gene, Wnt2, forms a forward transcriptional loop to control posterior cardiac development [125] (Table 1).

3.2. Evolutionary Conservation

GATA transcriptional regulators are widely distributed in fungi, plants, and metazoans and their DNA-binding domain is characterized by the presence of one or more class IV zinc finger motif(s). Fungal GATA factors have been shown to be involved in diverse functions such as nitrogen control, siderophore biosynthesis, light-regulated photomorphogenesis, circadian regulation, and mating-type switching [205]. In vertebrates, the zinc-finger domains are more than 70% conserved among all the six GATA binding proteins, although lower homology exists among their amino- and carboxy-terminal sequences [121]. In non-vertebrates such as Drosophila melanogaster and Caenorhabditis elegans, GATA TFs contain only a single zinc-finger motif that has undergone modular evolution [206]. Since vertebrates and invertebrates share only one C-terminal zinc finger (ZNII), it is possible that a single tandem duplication event might have occurred before the fungal and metazoan lineages diverged, resulting in two zinc finger motifs in vertebrates [207].
Evolutionary analysis reveals that the plant GATA family is much larger, more varied, and complex. In contrast to one or two-zinc finger motifs in vertebrates and invertebrates, phylogenetic analysis reveals the presence of four different classes of zinc fingers in plants. For example, Arabidopsis thaliana and rice (Oryza sativa) genomes contain 29 and 28 loci respectively that encode putative GATA factors that can be grouped into seven different subfamilies [208]. The GATA subfamily VI in plants consists of a tri-zinc finger protein, which has not been previously reported in eukaryotes. Plant GATA factors, unlike animals and fungi, have also been found to be associated with additional domains, such as CONSTANS, CO-like, and TOC1 (CCT) domain, an acidic domain or a transposase-like domain, involved in light signaling or nitrate-dependent transcriptional pathways [209]. Although it is unclear, multiple models of evolution including gene duplication and exon shuffling may explain the underlying basis of the GATA family expansion in plants.

3.3. Role in Cancer

The loss of expression, overexpression, or mutation of GATA factors have been associated with a multitude of cancers including leukemia, colorectal, lung, and breast cancers. Acute megakaryoblastic leukemia (DS-AMKL) seen in Down Syndrome patients is mostly associated with point mutations within the N-terminal zinc finger motif of GATA1. This results in a truncated form of GATA1 (GATA1s) that lacks N-terminal amino acids [110]. The presence of this mutation inhibits GATA1’s ability to bind the hematopoietic transcription co-factor FOG1 (friend of GATA) and affects platelet production [210,211]. Although less is known about the direct role of GATA2 in cancer, a subset of human chronic myelogenous leukemia (CML) patients harbor two mutations in the zinc finger domain of GATA2 [112]. Furthermore, GATA2 is required for Kras-driven NSCLC tumorigenesis [212]. Nearly 10% of human breast cancers are associated with GATA3 mutations in the C-terminal zinc finger of ZNII. The downregulation of GATA3 is a strong prognostic marker, especially in the cases of estrogen receptor (ER)-negative breast cancers, and is linked with aggressiveness and poor survival [116,213]. GATA3 restoration in breast cancer cell-lines induces miR-29b expression, leading to repressed metastasis and reduced tumor outgrowth [214,215].
The downregulation of GATA4 and GATA5 expression due to epigenetic silencing, such as CpG island hypermethylation and histone hypermethylation is often observed in cases of gastric, lung, ovarian, colorectal, oesophageal cancers, glioblastoma, and diffuse large B-cell lymphoma [119]. GATA6 acts as a double-edged sword in different cancer types. For example, it acts a tumor suppressor in astrocytoma while it is overexpressed in human colon cancer and pancreatic carcinoma [126,216,217]. Although not much is known about GATA factors, improved insights into GATA regulation at transcriptional, translational and post-translational levels can be exploited as novel biomarkers in cancer (Table 1).

4. Pax Transcription Factors

Paired box (Pax) genes encode TFs that orchestrate complex processes such as embryogenesis and are crucial for maintaining stem-cell pluripotency and stem cell-lineage specificity during development [218,219]. Pax proteins are characterized by the presence of three conserved elements: two DNA-binding domains—the paired domain (PD) and homeodomain (HD)—and the short octapeptide sequence (OP) located between the PD and HD domains [220]. Deletion of the OP motif in some Pax proteins is indicative of a transcriptional inhibitory activity [221]. PD is composed of 128 amino acids and makes sequence-specific contacts with DNA. A second paired-type HD domain found in several Pax members consists of 60 highly conserved amino acid residues. It shares strong homology with other homeobox gene products. PD can either bind DNA independently or as a cooperative interaction with HD domain. However, isolated HD domains have not been demonstrated to bind DNA [222,223]. Additionally, a transactivation domain (TD) at the carboxy terminus of Pax is a proline, serine-, and threonine rich region mediating transcriptional regulation [146,224].
The Pax family is composed of nine TFs (PAX1-PAX9) in humans as well as in mice (Pax1-Pax9). They are subdivided into subgroups I–IV based on the presence, absence, or truncation of a homeodomain—subgroup I (PAX1, PAX9), subgroup II (PAX2, PAX5, PAX8), subgroup III (PAX3, PAX7) and subgroup IV (PAX4, PAX6) [225]. The important roles of Pax genes in development underscore their functions in adult tissue regeneration and the repercussions of their aberrant loss, overexpression, or re-expression are associated with pathology (Table 1).

4.1. Role in Development

During development, the temporal and spatial expressions of Pax genes are tightly regulated. Pax expression is observed during proliferation but is switched off during terminal differentiation [223]. Pax gene expression in adult tissues has often been associated with tissue homeostasis. A small fraction of cortical cells in the adult thymus express PAX1, where it promotes the maturation of thymocytes [129]. PAX2 expression has been documented in medullary regions of the adult kidneys, mammary gland, transitional urothelium of the ureter and bladder as well in the epithelial lining of fallopian tubes of females [226,227,228]. Upon kidney injury, Pax2 expression re-emerges and prevents tubular cells from apoptosis in the initial stage of regeneration [131,132]. Pax3, expressed during early neurogenesis, regulates the generation of sensory neurons from precursors that originate from the neural crest [135]. PAX3 is also expressed in muscle stem cells in adults and melanoblasts (melanocyte stem cells) located in the bulge region of hair follicles, where it maintains their undifferentiated state [136]. Although not much is known about Pax4 function, its expression was shown to confer a protective function in pancreatic β-cells, increasing its replicative potential by transcriptionally activating Myc expression. It also protects β-cells from apoptosis by the activation of the anti-apoptotic Bcl-xL [140].
PAX5 is involved in B lymphopoiesis, specifically in the pathway regulating V- to -DJ recombination [142]. Intriguingly, re-programming of mature B-cells to pluripotent stem cells was shown to require Pax5, in addition to Sox2, Oct4, Klf4 and Myc [229]. During development, Pax6 is expressed in multiple brain regions and pancreatic islets, and is essential for eye organogenesis [146,230,231]. Pax6 is crucial for neuroectoderm cell fate determination [232]. Furthermore, the delicate balance between neural stem cell self-renewal and neurogenesis is regulated by Pax6 [233]. It was also reported that Pax6 is re-expressed during corneal wound repair. PAX6 deficiency was correlated with increased stromal cell apoptosis and cell-proliferation [234,235,236]. On the other hand, Pax7 maintains the proliferation and survival of postnatal satellite cells [237]. It is also found in muscle satellite cells, which are needed for tissue repair and regeneration following muscle injury [150].
In the adult thyroid, Pax8 plays a role in regulation of thyroglobulin (Tg), thyroid peroxidase (Tpo), and sodium/iodide symporter (NIS) that are essential for thyroid hormone synthesis [151,152,238]. Pax8 is also important for the maintenance of adult thyroid stem/progenitor cells [239]. Additionally, PAX8 expression has been documented in adult kidneys, specifically in the Bowman’s capsule and medullary regions, which are sites of renal stem/progenitor cells [152,240,241]. PAX9, like PAX8, is also expressed in the adult thymus and eosphagus [129]. Furthermore, PAX9 is also important in development of permanent teeth [154]. Thus, although Pax expression is relatively rare in adult tissues, this expression may be crucial for the survival of stem cell populations and maintenance of pluripotency (Table 1).

4.2. Evolutionary Conservation

Pax genes are specific to the animal lineage and have not been found in unicellular organisms, fungi or plants so far [220]. Four Pax genes (Pax1/9, Pax2/5/8, Pax3/7, and Pax4/6) have been found in the basal chordates, amphioxus (e.g., Brachiostoma floridae) and tunicates (e.g., Ciona intestinalis) [220,242]. Phylogenetic analyses indicated that, in the ancestral chordate, a single Pax gene of each subfamily was present, which gave rise to the amphioxus Pax. Subsequently, two major rounds of whole genome duplications occurred that gave rise to multiple vertebrate Pax subfamily copies [220,242,243]. Another partial duplication occurred subsequently, resulting in the nine Pax genes in mammals [244,245]. An alternative scenario would be that more Pax genes would have arisen after two whole genome duplications and were then lost during the vertebrate evolution [246,247].
Overall, purifying the selection appears to be the main factor responsible for the molecular evolution of the Pax family in chordate species. However, there are some indications of potential group-specific changes that are beyond this general pattern [248]. Phylogenetic analysis revealed that Pax2 and Pax5 ancestors were most likely involved in a round of complete vertebrate duplication while Pax8 was the most recent gene to appear by local gene duplication in this family. Lizards and birds have lost Pax4 and Pax8 [248]. Accelerated evolutionary rates were suggested for the Pax4, Pax8, and Pax7 genes. Thus, the asymmetric evolution of the Pax family genes can be associated with the emergence of adaptive novelties in the chordate evolutionary trajectory [248].
Moreover, two other alternative scenarios have been proposed to explain the evolution of Pax genes. One scenario assumes that the first Pax gene comprised of the PD domain alone (represented by PaxA/neuro) while the second Pax gene appeared as a result of the fusion of PD with an HD-containing gene [249]. Such capturing events could have happened several times and given rise to diverse primary Pax types [250]. The other scenario considers only one capturing event followed by gene duplications giving rise to the distinct Pax forms [249,251,252]. In this model, the PaxA gene is not assumed to denote the progenitor type, but instead is a remnant form lacking the HD domain.

4.3. Role in Cancer

Pax genes belonging to subgroups II and III that contain OP and a partial HD are involved in cell motility, cell survival, and self-sufficiency in growth signals, thus favoring tumor progression [253]. Conversely, Pax genes in subgroups I and IV that only contain one of these domains are rarely involved in cancer, or are indicators of favorable prognosis in cancers [225].
PAX1 was found to be hypermethylated in cervical cancer tissues [130]. On the other hand, PAX2 is expressed in ovarian cancer, renal-cell carcinomas (RCC) and in some bladder carcinomas, where it is crucial for tumor survival since PAX2 regulates the surface protein metallopeptidase, A Disintegrin and metalloproteinase-domain containing protein 10 (ADAM10) [133]. Over 70% of RCC cell-lines bear deletions/mutations in the Von Hippel Lindau (VHL) tumor suppressor gene that, in turn, promotes PAX2 expression in renal tumors [254,255]. In breast cancer, PAX2 was reported to form a complex with the ER and regulate Erythroblastic Oncogene B2 (ERBB2), thus determining the response to tamoxifen [134]. In addition, resistance to apoptosis in Kaposi’s sarcoma is associated with PAX2 expression [256]. In a majority of alveolar rhabdomyosarcomas (ARMS), PAX3 has been shown to undergo chromosome rearrangement with FOXO1/FKHR [137,138]. The PAX3-FKHR fusion in ARMS is a strong transcriptional regulator and is thought to be a dominant-acting oncoprotein [257]. PAX3 is also expressed in primary melanomas and its expression in sentinel lymph nodes has been considered as a prognostic marker for aggressive tumors with a poor outcome [139]. PAX4 is upregulated in human insulinomas [258] and functions as a survival factor in rat insulinoma cells via Bcl-xL upregulation [141].
Most B-cell neoplasms, including B-cell lymphoma demonstrate PAX5 overexpression [143]. However, in HCC, PAX5 acts as a tumor suppressor by interacting with the p53 signaling pathway [144]. In breast cancer, PAX5 expression enhances epithelial behavior and is associated with better prognosis in patients [145]. In PDAC, Pax6 promotes cancer progression by the activation of the receptor tyrosine kinase, c-met [147]. Conversely, PAX6 expression was observed to suppress the invasiveness of glioblastoma cells by regulating the expression of matrix-metalloproteinase 2. In addition, PAX6 also reduced angiogenesis and increased glioma cell susceptibility to detachment and oxidative stress [148,149,259].
Similar to PAX3, albeit less frequently, PAX7 also undergoes rearrangements with FOXO1/FKHR in ARMS [137]. PAX8 undergoes rearrangements with peroxisome proliferator-activated receptor γ (PPARγ) in thyroid adenocarcinomas [260]. PAX8 was also shown to be essential for basal E2F1 transcription and maintaining the stability of its TF c-factor, Rb, in renal, ovarian, and thyroid cancers [241]. In addition, PAX8 also regulates telomerase in certain glioblastoma cell lines [153]. PAX9 is amplified and has been implicated in promoting the proliferation of lung cancer cells [155]. Oncogene-induced cell-survival in oral squamous cell carcinomas is mediated by PAX9 [156]. Thus, Pax genes play a major role in conferring growth and survival advantages to cancer cells by regulating cell plasticity [261] (Table 1).

5. bHLH Transcription Factors

Basic helix loop helix (bHLH) TFs are named on the basis of their structure, and have two evolutionarily conserved domains, namely the basic domain that binds to the E-box DNA sequences (CANNTG) to regulate transcription and the helix-loop-helix (HLH) domain, important for protein homo- or hetero-dimerization. Post dimerization, they bind to the E-box. The dimerization happens via two alpha-helices connected by a non-conserved loop region [262]. Class I bHLH molecules are expressed quite ubiquitously, whereas Class II molecules are tissue specific [263,264]. One such tissue specific bHLH factor is Twist, which regulates EMT in both development and cancer [159,160,265]. The bHLH TF superfamily is imperative for proper development, including the fate specification and cell differentiation of almost all the tissues of any organism from flies to humans [263]. One example of the same TF playing important roles in development and cancer is Myc. Elevated levels of MYC are seen in 60%–70% of all cancers [266]. These are also bHLH TFs and play an important role normally in cell cycle, differentiation, and angiogenesis (Table 1).

5.1. Role in Development

Proneural bHLH TFs were first identified in Drosophila for their ability to confer neural identity to ectodermal tissue. In contrast, vertebrate bHLH genes act after neural identity has been determined. The Achete-Schute complex and Atonal are two neural-specific bHLH gene families in vertebrates, based on their homology in flies, that play a wide range of roles in development [262,267]. Proneural genes Neurogenin 1, 2, and Ascl1 are required for neural differentiation in both the peripheral and central nervous systems (CNS) [268,269,270,271,272,273,274,275]. Neurogenin 2 and Ascl1 have in fact been used for neuronal reprogramming due to their ability to specify cell fates based on their target genes [157,276,277,278,279,280]. Atoh1 of the Atonal family is also important for the differentiation of granule cells of the cerebellum and of inner ear hair cells [165]. The bHLH family of neural specific genes also include NeuroD1, D2 and D6, and the Olig family. These are important factors for differentiation to neurons and oligodendrocytes within the CNS. NeuroD1 is necessary for the differentiation of inner ear sensory neurons, granule cells of the cerebellum and the hippocampus [166,281]. NeuroD2 and D6 are necessary for the formation of the corpus callosum, needed to communicate between the two cerebral hemispheres [168]. Olig1, 2 and 3 are necessary and sufficient for oligodendrocyte differentiation in the neocortex, spinal cord and the cerebellum, respectively [174,282,283].
Besides their extensive role in neural development, bHLH TFs have also been well-studied in the development of other structures. Math1, Neurogenin 3, and NeuroD1 play a sequential role in the development of gastrointestinal entero-endocrine cells—specification, segregation to the secretory lineage and differentiation [284,285,286,287]. Hand1 and Hand2 play critical roles in the proliferation, differentiation, and the morphogenesis of embryonic ventricle cardiomyocytes [170,171,288]. Twist1 and Twist2 play a major role in bone formation or osteogenesis. They are important for osteoprogenitor proliferation and differentiation via FGF signaling [157]. Twist1 is expressed in the skeletal mesenchyme and also important for craniofacial development, also via FGF signaling [158,289,290] (Table 1).

5.2. Evolutionary Conservation

bHLH is a large family of TFs that control the developmental and physiological processes of eukaryotes, and exist in fungi, plants, and animals [267]. Several TFs of this family are evolutionarily conserved across different species and play a crucial role during development. Orthologs of Nephew of atonal 3 (Nato 3), a proneural gene, are conserved across Drosophila, C.elegans, mice, and humans. They are highly similar in their bHLH domain [291]. The Hand gene family is also highly conserved across Drosophila and mammals, and is essential for heart and vascular development [292].
In yeast, bHLH TFs promote cell cycle control and transcriptional enhancement [293,294]. The bHLH members are the second largest class of plant TFs and play a pivotal role in plant growth and maintenance. Phytohormone signaling cascades impinge on to bHLH TFs for Arabidopsis development and defense [295]. SlPRE2, an atypical bHLH member, controls the pigmentation of tomato fruit and the morphology of the plant [296].
The bHLH family has expanded in plants and animals following evolutionarily independent events [267]. It is unclear whether bHLH TFs evolved from a single common ancestor or via domain shuffling from an ancestral protein [297]. Genome segment and tandem duplications are thought to have led to bHLH gene family expansion in plants [298,299], whereas studies in animals suggest single-gene duplication [300]. The field is still debating whether bHLH TFs expanded in parallel with the evolution of multicellularity, or with the colonization of land [301,302]. Evolutionary analyses of several land plants, chlorophytes, and red algae suggest that the first plants had minimal bHLH genes, and that all modern plant bHLH proteins descended and evolved via a large number of gene duplications [302,303].

5.3. Role in Cancer

An innumerable number of bHLH TFs are important for cellular differentiation, cell cycle arrest, and apoptosis. Therefore, it isn’t surprising that they play a major role in tumor growth and progression. Myc is a proto-oncogene that is dysregulated in several types of cancer. Copy number variations in MYC occur very frequently among other genetic events leading to human cancers, for example, in PDAC [163]. Myc is downstream of multiple important signaling pathways such as PI3K [304], Notch [305], Wnt-APC [306], and KRAS-ERK [307] that are implicated in different types of cancers. More importantly, Myc is responsible for both initiating, as well as maintain the tumor [164,308]. Hes1 and Hey1 positively regulate p53 levels, a tumor suppressor gene [188]. Both these TFs are dysregulated in several different cancers [309]. Twist induces EMT and is activated during tumor progression [159,160,161]. BHLHE40 (DEC1) and DEC2 are important for the regulation of the cell cycle via cyclin D1 and cell death in oral and breast cancer cells [179,181]. Dec1 also leads to EMT in pancreatic cancer cells [310]. The expression of TCF3 (E2A) is enhanced in prostate cancer, thereby promoting tumor progression—it provides resistance to apoptosis in prostate cancer [311]. Hypoxia-inducible factor 2 alpha (HIF-2) aids the progression of neuroblastoma and other cancers in non-hypoxic conditions by recruiting Argonaut 2 [312].
The bHLH TFs have been shown to be downregulated in pancreatic cancer and, in fact, a high-throughput screen has identified small molecules as bHLH activators, which may be used as therapeutic targets [313] (Table 1).

6. Discussion

6.1. Transcription Factors—Crucial Proteins for Development and Homeostasis

In this review we have discussed four families of TFs that have been well studied in both, development and cancer. However, there are a multitude of TFs that have important roles in multiple physiological processes and derangements. In fact, 294 cancer-related TFs have been listed in different resources [314,315]. The LIM family of TFs has been exhaustively studied in development [316] and cancer [317]. A few other TFs that have been studied considerably well include the specificity proteins (Sp) family [318], forkhead box (FOX) family [319,320,321,322], HOX genes [323,324], ETS-domain TFs [325,326,327], steroid reproductive hormone receptors [328,329] and zinc finger ZBTB proteins, with N-terminal BTB/POZ domains [330].
While we have limited our review to a subset of TFs, development and cancer are regulated by a number of epigenetic factors and noncoding RNA molecules. They have been discussed at length elsewhere [331]. Recently, these molecules have also been targeted for cancer therapy [314,332].

6.2. Therapeutic Targeting of Transcription Factor: Need of the Hour

TFs regulate a wide range of biological processes and therefore are essential for maintaining homeostasis. They account for nearly 20% of the identified oncogenes and although promising candidates for targeting cancer [314,315], they were considered undruggable up until this decade [333]. A better understanding of their mechanisms of action and structural interactions with the cognate DNA sequence and protein regulators have led to the discovery of useful drug candidates. Despite this progress, the immense repertoire of downstream targets, threshold of expression in normal versus cancerous tissue, redundancy, and compensation by other TFs, epigenetic modulation [334], and vastly different mutations in the same gene across individuals [335] makes it arduous for TFs to be effectively targeted [314].
Targeted genome editing technology mediated by CRISPR shows great promise in both fundamental and clinical research. It has been employed for the increment or attenuation of gene expression more reliably than any other genetic engineering technology [336,337]. Targeting TFs using this approach could be a reasonable therapeutic route since they control the fate of a cell, in normal physiology and in cancer. Catalytically inactive dCas9 can be recruited to specific sites on the DNA, which is particularly useful when fused to TFs. This would allow the activation or repression of certain downstream genes [338,339]. Direct targeting of cancer markers such as MYC has been explored to reduce genetic alterations leading to uncontrolled proliferation and metastasis [340]. CRISPR may prove useful in such targeting. In addition, the CRISPR system has been tested for light-induced spatio-temporal control of gene expression [341]. DNA break caused by CRISPR/Cas9 triggers two mechanisms of DNA repair: non homologous end-joining (NHEJ) and homology-directed repair (HDR). Of these, HDR is high fidelity and therefore allows precise DNA editing [342,343]. A novel CRISPR-barcoding tool utilizing HDR enables identification of mutation such as p53 mutation in breast cancer cells (MCF7) and even correcting a mutation, for example, ALK-F1174L in Kelly neuroblastoma cells [344].
As discussed above GATA, HMG, PAX, and bHLH have been implicated in cancer and the characterization of these molecular targets in vitro and in vivo studies have led to the development of several preclinical and clinical studies. The targeted modulation of these TFs can be used for the development of new cancer treatment [337,345,346,347,348]. We have summarized a list of the ongoing preclinical and clinical trials studies for various TF targets (Figure 2, Table 2).

6.3. Natural Resistance Against Cancer: Learning from Life

The task of suppressing somatic mutations in larger organisms and those with a longer lifespan is more challenging. According to Peto’s paradox, there is no correlation between the body size, longevity and increased risk of developing cancer. Therefore, in evolution, larger animals have mechanisms to suppress cancer by either eliminating certain proto-oncogenes or duplicating tumor suppressor genes [358,359,360]. Elephants appear to have low cancer occurrence rates since they have re-functionalized the leukemia inhibitory factor pseudogene 6 (LIF6) with pro-apoptotic functions [361]. In addition, the duplication/multiplication of tumor-suppressor protein TP53 seems to provide another explanation, even though most are processed pseudogenes [362]. DNA damage leads to TP53 upregulation which, in turn, transcriptionally upregulates LIF6. A TP53 response element perhaps evolved co-incident with large body sizes [363]. The analysis of cancer prevention in elephants suggests a lack of understanding of the full extent of the tumor-suppressive capacity of p53 in humans [364].
Cetacean species, another order of large mammals could also be effective models for studying cancer [365]. The beluga whales of the St. Lawrence estuary have a high occurrence of cancer, sometimes even surpassing humans, but are an exception among other cetaceans. In pilot whales, bottlenose dolphins, and other toothed whales, cancer is a rare event [366]. The bowhead whales have an extraordinarily long lifespan [367]. Comparative genomics and transcriptomics have revealed the duplication of proliferating cell nuclear antigen (PCNA) and other genes involved in DNA repair in these animals [368]. Cross-species comparisons allow us to understand cancer resistance in other mammals as well—for example, naked mole rats and blind mole rats are remarkably resistant to cancer [369].
Fundamentally, plants are different from animals owing to their cell walls. Even though plants develop tumors, the cell wall exerts control on cancer metastasis. Plant tumors are mainly caused by pathogens such as Agrobacterium (crown gall), geminivirus, and Ustinaginales among other fungal infections [172,370]. In the absence of infections, they are remarkably resistant to neoplastic transformation and hence, cancer. However, spontaneous tumors arise in interspecific hybrids of certain plant species, such as Nicotiana (tobacco) [371]. Most of these tumors are caused by phyto-hormonal imbalance [372]. Interestingly, homologs of an extensively studied tumor suppressor and cell cycle regulator, Rb, plays an important role in tumorigenesis in divergent multicellular species [373]. Rb-related (RBR) in plants are implicated in tumor-like growth upon infection with Agrobacterium and geminiviruses [374]. Although exhaustive research exists on mammalian Rb and its role in cancer compared to its plant homolog, they have similar roles in cell cycle progression, regulation of TFs via chromatin modifying proteins and role in cell fate decisions [373]. Comparing the molecular aspects of tumor-initiation and progression with plants may provide insights into cancer prevention and the understanding of its biology.

Author Contributions

D.H., P.V., S.N. and S.B. were involved in the conception and drafting of the manuscript. D.H., P.V. and S.B. were involved in designing the figures and tables. All authors read and approved the final manuscript.

Acknowledgments

We apologize to those colleagues whose work has not been cited due to space limitation.

Conflicts of Interest

S.N. is a Portfolio Manager at AIG Investments. AIG did not have any input into the work presented. No other authors have a conflict of interest.

Abbreviations

Acute megakaryoblastic leukemiaAMKL
Alveolar rhabdomyosarcomas ARMS
basic Helix-loop-HelixbHLH
Bone Morphogenetic ProteinBMP
Central nervous system CNS
Chronic myelogenous leukemiaCML
Cystein-rich Polycomb-like Proteins CPP
Epithelial-to-mesenchymal transitionEMT
Fibroblast Growth Factor FGF
Forkhead box FOX
HedgehogHH
Hepatocellular carcinoma HCC
Hematopoietic stem cells HSCs
Hematopoietic stem/progenitor cells HSPC
High Mobility Group boxHMG
HMG-AT-hook family HMGA
HMG-box familyHMGB
HMG-nucleosome binding familyHMGBN
HomeodomainHD
Homology-directed repair HDR
Leukemia Inhibitory Factor pseudogene 6 LIF6
Myelodysplastic syndrome MDS
Myeloproliferative neoplasmsMPN
Nephew of atonal 3Nato 3
Neurogenin2Ngn2
Non homologous end-joining NHEJ
Non-small cell lung carcinomas NSCLC
Nuclear localization signal NLS
Nucleosome-binding domain NBD
Octamer binding transcription factor 4Oct-04
OctopeptideOP
Paired box genes PAX
Paired domain PD
Pancreatic ductal adenocarcinoma PDAC
Proliferating cell nuclear antigen PCNA
Renal-cell carcinomas RCC
Specificity proteins Sp
RetinoblastomaRb
Ten Eleven Translocation TET
Transactivation domain TD
Tumor initiating cells TIC
ThyroglobulinTy
Thyroid peroxidaseTpo
Transcription FactorsTFs

References

  1. Roy, N.; Hebrok, M. Regulation of Cellular Identity in Cancer. Dev. Cell 2015, 35, 674–684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Knudson, A.G. Mutation and Cancer: Statistical Study of Retinoblastoma. Proc. Natl. Acad. Sci. USA 1971, 68, 820–823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Knudson, A.G. Cancer genetics through a personal retrospectroscope. Genes Chromosom. Cancer 2003, 38, 288–291. [Google Scholar] [CrossRef] [PubMed]
  4. Nichols, J.; Zevnik, B.; Anastassiadis, K.; Niwa, H.; Klewe-Nebenius, D.; Chambers, I.; Scholer, H.; Smith, A. Formation of Pluripotent Stem Cells in the Mammalian Embryo Depends on the POU Transcription Factor Oct4. Cell 1998, 95, 379–391. [Google Scholar] [CrossRef] [Green Version]
  5. Atlasi, Y.; Mowla, S.J.; Ziaee, S.A.; Gokhale, P.J.; Andrews, P.W. OCT4 Spliced Variants Are Differentially Expressed in Human Pluripotent and Nonpluripotent Cells. Stem Cells 2008, 26, 3068–3074. [Google Scholar] [CrossRef]
  6. Ku, J.-L.; Shin, Y.-K.; Kim, D.-W.; Choi, J.-S.; Hong, S.-H.; Jeon, Y.-K.; Park, J.-H. Establishment and characterization of 13 human colorectal carcinoma cell lines: Mutations of genes and expressions of drug-sensitivity genes and cancer stem cell markers. Carcinogenesis 2010, 31, 1003–1009. [Google Scholar] [CrossRef]
  7. Liu, A.; Yu, X.; Liu, S. Pluripotency transcription factors and cancer stem cells: Small genes make a big difference. Chin. J. Cancer 2013, 32, 483–487. [Google Scholar] [CrossRef]
  8. Amini, S.; Fathi, F.; Mobalegi, J.; Sofimajidpour, H.; Ghadimi, T. The expressions of stem cell markers: Oct4, Nanog, Sox2, nucleostemin, Bmi, Zfx, Tcl1, Tbx3, Dppa4, and Esrrb in bladder, colon, and prostate cancer, and certain cancer cell lines. Anat. Cell Boil. 2014, 47, 1–11. [Google Scholar] [CrossRef] [Green Version]
  9. Wang, Y.-J.; Herlyn, M. The emerging roles of Oct4 in tumor-initiating cells. Am. J. Physiol. Physiol. 2015, 309, C709–C718. [Google Scholar] [CrossRef]
  10. Olmez, I.; Shen, W.; McDonald, H.; Ozpolat, B. Dedifferentiation of patient-derived glioblastoma multiforme cell lines results in a cancer stem cell-like state with mitogen-independent growth. J. Cell. Mol. Med. 2015, 19, 1262–1272. [Google Scholar] [CrossRef]
  11. Ghosh, D.; Nandi, S.; Bhattacharjee, S. Combination therapy to checkmate Glioblastoma: Clinical challenges and advances. Clin. Transl. Med. 2018, 7, 33. [Google Scholar] [CrossRef] [PubMed]
  12. Borrull, A.; Ghislin, S.; Deshayes, F.; Lauriol, J.; Alcaide-Loridan, C.; Middendorp, S. Nanog and Oct4 overexpression increases motility and transmigration of melanoma cells. J. Cancer Res. Clin. Oncol. 2012, 138, 1145–1154. [Google Scholar] [CrossRef] [PubMed]
  13. Kumar, S.M.; Liu, S.; Lu, H.; Zhang, H.; Zhang, P.J.; Gimotty, P.A.; Guerra, M.; Guo, W.; Xu, X. Acquired cancer stem cell phenotypes through Oct4-mediated dedifferentiation. Oncogene 2012, 31, 4898–4911. [Google Scholar] [CrossRef] [Green Version]
  14. Chiou, S.-H.; Wang, M.-L.; Chou, Y.-T.; Chen, C.-J.; Hong, C.-F.; Hsieh, W.-J.; Chang, H.-T.; Chen, Y.-S.; Lin, T.-W.; Hsu, H.-S.; et al. Coexpression of Oct4 and Nanog Enhances Malignancy in Lung Adenocarcinoma by Inducing Cancer Stem Cell-Like Properties and Epithelial-Mesenchymal Transdifferentiation. Cancer Res. 2010, 70, 10433–10444. [Google Scholar] [CrossRef] [PubMed]
  15. Cappellen, D.; Schlange, T.; Bauer, M.; Maurer, F.; Hynes, N.E. Novel c-MYC target genes mediate differential effects on cell proliferation and migration. EMBO Rep. 2007, 8, 70–76. [Google Scholar] [CrossRef] [PubMed]
  16. Koh, C.M.; Khattar, E.; Leow, S.C.; Liu, C.Y.; Müller, J.; Ang, W.X.; Li, Y.; Franzoso, G.; Li, S.; Guccione, E.; et al. Telomerase regulates MYC-driven oncogenesis independent of its reverse transcriptase activity. J. Clin. Investig. 2015, 125, 2109–2122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Wang, R.; Dillon, C.P.; Shi, L.Z.; Milasta, S.; Carter, R.; Finkelstein, D.; McCormick, L.L.; Fitzgerald, P.; Chi, H.; Munger, J.; et al. The transcription factor Myc controls metabolic reprogramming upon T lymphocyte activation. Immunity 2011, 35, 871–882. [Google Scholar] [CrossRef]
  18. Beer, S.; Zetterberg, A.; Ihrie, R.A.; McTaggart, R.A.; Yang, Q.; Bradon, N.; Arvanitis, C.; Attardi, L.D.; Feng, S.; Ruebner, B.; et al. Developmental context determines latency of MYC-induced tumorigenesis. PLoS Boil. 2004, 2, e332. [Google Scholar] [CrossRef] [PubMed]
  19. Felsher, D.W.; Bishop, J. Reversible Tumorigenesis by MYC in Hematopoietic Lineages. Mol. Cell 1999, 4, 199–207. [Google Scholar] [CrossRef]
  20. Jain, M.; Arvanitis, C.; Chu, K.; Dewey, W.; Leonhardt, E.; Trinh, M.; Sundberg, C.D.; Bishop, J.M.; Felsher, D.W. Sustained Loss of a Neoplastic Phenotype by Brief Inactivation of MYC. Science 2002, 297, 102–104. [Google Scholar] [CrossRef] [PubMed]
  21. Pelengaris, S.; Khan, M.; I Evan, G. Suppression of Myc-induced apoptosis in beta cells exposes multiple oncogenic properties of Myc and triggers carcinogenic progression. Cell 2002, 109, 321–334. [Google Scholar] [CrossRef]
  22. Marinkovic, D.; Marinkovic, T.; Kokai, E.; Barth, T.; Möller, P.; Wirth, T. Identification of novel Myc target genes with a potential role in lymphomagenesis. Nucleic Acids Res. 2004, 32, 5368–5378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Hay, E. An Overview of Epithelio-Mesenchymal Transformation. Cells Tissues Organs 1995, 154, 8–20. [Google Scholar] [CrossRef] [PubMed]
  24. Kalluri, R.; Neilson, E.G. Epithelial-mesenchymal transition and its implications for fibrosis. J. Clin. Investig. 2003, 112, 1776–1784. [Google Scholar] [CrossRef] [PubMed]
  25. Roche, J. The Epithelial-to-Mesenchymal Transition in Cancer. Cancers 2018, 10, 52. [Google Scholar] [CrossRef] [PubMed]
  26. Ferrer-Vaquer, A.; Viotti, M.; Hadjantonakis, A.-K. Transitions between epithelial and mesenchymal states and the morphogenesis of the early mouse embryo. Cell Adhes. Migr. 2014, 4, 447–457. [Google Scholar] [CrossRef]
  27. Wang, X.; Kopinke, D.; Lin, J.; McPherson, A.D.; Duncan, R.N.; Otsuna, H.; Moro, E.; Hoshijima, K.; Grunwald, D.J.; Argenton, F.; et al. Wnt signaling regulates postembryonic hypothalamic progenitor differentiation. Dev. Cell 2012, 23, 624–636. [Google Scholar] [CrossRef]
  28. Duband, J.L.; Thiery, J.P. Appearance and distribution of fibronectin during chick embryo gastrulation and neurulation. Dev. Boil. 1982, 94, 337–350. [Google Scholar] [CrossRef]
  29. Liem, K.F.; Jessell, T.M.; Briscoe, J. Regulation of the neural patterning activity of sonic hedgehog by secreted BMP inhibitors expressed by notochord and somites. Development 2000, 127, 4855–4866. [Google Scholar]
  30. Sela-Donenfeld, D.; Kalcheim, C. Localized BMP4–Noggin Interactions Generate the Dynamic Patterning of Noggin Expression in Somites. Dev. Boil. 2002, 246, 311–328. [Google Scholar] [CrossRef]
  31. Thiery, J.P. Epithelial–mesenchymal transitions in development and pathologies. Curr. Opin. Cell Boil. 2003, 15, 740–746. [Google Scholar] [CrossRef] [PubMed]
  32. Burstyn-Cohen, T.; Stanleigh, J.; Sela-Donenfeld, D.; Kalcheim, C. Canonical Wnt activity regulates trunk neural crest delamination linking BMP/noggin signaling with G1/S transition. Development 2004, 131, 5327–5339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Karafiat, V.; Dvorakova, M.; Pajer, P.; Cermak, V.; Dvorak, M. Melanocyte fate in neural crest is triggered by Myb proteins through activation of c-kit. Cell. Mol. Life Sci. 2007, 64, 2975–2984. [Google Scholar] [CrossRef] [PubMed]
  34. Villanueva, S.; Glavic, A.; Ruiz, P.; Mayor, R.; Ruiz-Rudolph, P. Posteriorization by FGF, Wnt, and Retinoic Acid Is Required for Neural Crest Induction. Dev. Boil. 2002, 241, 289–301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Kelley, L.C.; Lohmer, L.L.; Hagedorn, E.J.; Sherwood, D.R. Traversing the basement membrane in vivo: A diversity of strategies. J. Cell Boil. 2014, 204, 291–302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Tepass, U.; Theres, C.; Knust, E. crumbs encodes an EGF-like protein expressed on apical membranes of Drosophila epithelial cells and required for organization of epithelia. Cell 1990, 61, 787–799. [Google Scholar] [CrossRef]
  37. Edelman, G.M.; Crossin, K.L. Cell Adhesion Molecules: Implications for a Molecular Histology. Annu. Rev. Biochem. 1991, 60, 155–190. [Google Scholar] [CrossRef]
  38. Lamouille, S.; Xu, J.; Derynck, R. Molecular mechanisms of epithelial–mesenchymal transition. Nat. Rev. Mol. Cell Boil. 2014, 15, 178–196. [Google Scholar] [CrossRef]
  39. Beaman, J.E.; White, C.R.; Seebacher, F. Evolution of Plasticity: Mechanistic Link between Development and Reversible Acclimation. Trends Ecol. Evol. 2016, 31, 237–249. [Google Scholar] [CrossRef]
  40. Holtan, S.G.; Creedon, D.J.; Haluska, P.; Markovic, S.N. Cancer and Pregnancy: Parallels in Growth, Invasion, and Immune Modulation and Implications for Cancer Therapeutic Agents. Mayo Clin. Proc. 2009, 84, 985–1000. [Google Scholar] [CrossRef] [Green Version]
  41. Burger, O.; Baudisch, A.; Vaupel, J.W. Human mortality improvement in evolutionary context. Proc. Natl. Acad. Sci. USA 2012, 109, 18210–18214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Saito, S.; Nakashima, A.; Shima, T.; Ito, M. Th1/Th2/Th17 and Regulatory T-Cell Paradigm in Pregnancy. Am. J. Reprod. Immunol. 2010, 63, 601–610. [Google Scholar] [CrossRef] [PubMed]
  43. Mor, G. The Unique immunologic and microbial aspects of pregnancy. Placenta 2017, 57, 226. [Google Scholar] [CrossRef]
  44. Sakaguchi, S. Regulatory T cells: Key controllers of immunologic self-tolerance. Cell 2000, 101, 455–458. [Google Scholar] [CrossRef]
  45. Fontenot, J.D.; Gavin, M.A.; Rudensky, A.Y. Foxp3 programs the development and function of CD4+CD25+ regulatory T cells. Nat. Immunol. 2003, 4, 330–336. [Google Scholar] [CrossRef]
  46. Hori, S.; Nomura, T.; Sakaguchi, S. Control of Regulatory T Cell Development by the Transcription Factor Foxp3. Science 2003, 299, 1057–1061. [Google Scholar] [CrossRef]
  47. Khattri, R.; Cox, T.; Yasayko, S.A.; Ramsdell, F. An essential role for Scurfin in CD4+CD25+ T regulatory cells. Nat. Immunol. 2003, 4, 337–342. [Google Scholar] [CrossRef]
  48. Fontenot, J.D.; Rasmussen, J.P.; Williams, L.M.; Dooley, J.L.; Farr, A.G.; Rudensky, A.Y. Regulatory T Cell Lineage Specification by the Forkhead Transcription Factor Foxp3. Immunity 2005, 22, 329–341. [Google Scholar] [CrossRef]
  49. Wan, Y.Y.; Flavell, R.A. Identifying Foxp3-expressing suppressor T cells with a bicistronic reporter. Proc. Natl. Acad. Sci. USA 2005, 102, 5126–5131. [Google Scholar] [CrossRef] [Green Version]
  50. Tremellen, K.P.; Jasper, M.J.; Robertson, S.A. Primary unexplained infertility is associated with reduced expression of the T-regulatory cell transcription factor Foxp3 in endometrial tissue. Mol. Hum. Reprod. 2006, 12, 301–308. [Google Scholar] [Green Version]
  51. Wang, L.; Liu, R.; Li, W.; Chen, C.; Katoh, H.; Chen, G.-Y.; McNally, B.; Lin, L.; Zhou, P.; Zuo, T.; et al. Somatic single hits inactivate the X-linked tumor suppressor FOXP3 in the prostate. Cancer Cell 2009, 16, 336–346. [Google Scholar] [CrossRef] [PubMed]
  52. Martin, F.; Ladoire, S.; Mignot, G.; Apetoh, L.; Ghiringhelli, F. Human FOXP3 and cancer. Oncogene 2010, 29, 4121–4129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. De Calisto, J.; Araya, C.; Marchant, L.; Riaz, C.F.; Mayor, R. Essential role of non-canonical Wnt signalling in neural crest migration. Development 2005, 132, 2587–2597. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Giles, R.H.; Van Es, J.H.; Clevers, H. Caught up in a Wnt storm: Wnt signaling in cancer. Biochim. Biophys. Acta (BBA) Bioenergy 2003, 1653, 1–24. [Google Scholar] [CrossRef]
  55. Polakis, P. The many ways of Wnt in cancer. Curr. Opin. Genet. Dev. 2007, 17, 45–51. [Google Scholar] [CrossRef]
  56. Bell, S.M.; Schreiner, C.M.; Goetz, J.A.; Robbins, D.J.; Scott, W.J. Shh signaling in limb bud ectoderm: Potential role in teratogen-induced postaxial ectrodactyly. Dev. Dyn. 2005, 233, 313–325. [Google Scholar] [CrossRef]
  57. Bruner, H.C.; Derksen, P.W.B. Loss of E-Cadherin-Dependent Cell-Cell Adhesion and the Development and Progression of Cancer. Cold Spring Harb. Perspect. Biol. 2018, 10, a029330. [Google Scholar] [CrossRef]
  58. Lee, Y.C.; Baath, J.A.; Bastle, R.M.; Bhattacharjee, S.; Cantoria, M.J.; Dornan, M.; Gamero-Estevez, E.; Ford, L.; Halova, L.; Kernan, J.; et al. Impact of Detergents on Membrane Protein Complex Isolation. J. Proteome Res. 2018, 17, 348–358. [Google Scholar] [CrossRef]
  59. Monks, A.; Scudiero, D.; Skehan, P.; Shoemaker, R.; Paull, K.; Vistica, D.; Hose, C.; Langley, J.; Cronise, P.; Vaigro-Wolff, A.; et al. Feasibility of a High-Flux Anticancer Drug Screen Using a Diverse Panel of Cultured Human Tumor Cell Lines. J. Natl. Cancer Inst. 1991, 83, 757–766. [Google Scholar] [CrossRef]
  60. El-Deiry, W.S.; Kern, S.E.; Pietenpol, J.A.; Kinzler, K.W.; Vogelstein, B. Definition of a consensus binding site for p53. Nat. Genet. 1992, 1, 45–49. [Google Scholar] [CrossRef]
  61. Kautiainen, T.L.; A Jones, P. DNA methyltransferase levels in tumorigenic and nontumorigenic cells in culture. J. Boil. Chem. 1986, 261, 1594–1598. [Google Scholar]
  62. Rasmussen, K.D.; Helin, K. Role of TET enzymes in DNA methylation, development, and cancer. Genes Dev. 2016, 30, 733–750. [Google Scholar] [CrossRef] [PubMed]
  63. Dawlaty, M.M.; Breiling, A.; Le, T.; Raddatz, G.; Barrasa, M.I.; Cheng, A.W.; Gao, Q.; Powell, B.E.; Li, Z.; Xu, M.; et al. Combined deficiency of Tet1 and Tet2 causes epigenetic abnormalities but is compatible with postnatal development. Dev. Cell 2013, 24, 310–323. [Google Scholar] [CrossRef] [PubMed]
  64. Dawlaty, M.M.; Breiling, A.; Le, T.; Barrasa, M.I.; Raddatz, G.; Gao, Q.; Powell, B.E.; Cheng, A.W.; Faull, K.F.; Lyko, F.; et al. Loss of Tet enzymes compromises proper differentiation of embryonic stem cells. Dev. Cell 2014, 29, 102–111. [Google Scholar] [CrossRef]
  65. An, J.; González-Avalos, E.; Chawla, A.; Jeong, M.; López-Moyado, I.F.; Li, W.; Goodell, M.A.; Chavez, L.; Ko, M.; Rao, A. Acute loss of TET function results in aggressive myeloid cancer in mice. Nat. Commun. 2015, 6, 10071. [Google Scholar] [CrossRef] [Green Version]
  66. Cimmino, L.; Dawlaty, M.M.; Ndiaye-Lobry, D.; Yap, Y.S.; Bakogianni, S.; Yu, Y.T.; Bhattacharyya, S.; Shaknovich, R.; Geng, H.M.; Lobry, C.; et al. TET1 is a tumor suppressor of hematopoietic malignancy. Nat. Immunol. 2015, 16, 653. [Google Scholar] [CrossRef]
  67. Zhang, Q.; Zhao, K.; Shen, Q.; Han, Y.; Gu, Y.; Li, X.; Zhao, D.; Liu, Y.; Wang, C.; Zhang, X.; et al. Tet2 is required to resolve inflammation by recruiting Hdac2 to specifically repress IL-6. Nature 2015, 525, 389–393. [Google Scholar] [CrossRef] [Green Version]
  68. Gaidzik, V.I.; Paschka, P.; Späth, D.; Habdank, M.; Köhne, C.-H.; Germing, U.; Von Lilienfeld-Toal, M.; Held, G.; Horst, H.-A.; Haase, D.; et al. TET2 Mutations in Acute Myeloid Leukemia (AML): Results from a Comprehensive Genetic and Clinical Analysis of the AML Study Group. J. Clin. Oncol. 2012, 30, 1350–1357. [Google Scholar] [CrossRef]
  69. Boumber, Y.A.; Kondo, Y.; Chen, X.; Shen, L.; Gharibyan, V.; Konishi, K.; Estey, E.; Kantarjian, H.; Garcia-Manero, G.; Issa, J.J. RIL, a LIM Gene on 5q31, Is Silenced by Methylation in Cancer and Sensitizes Cancer Cells to Apoptos.pdf. Cancer Res. 2007, 67, 1997–2005. [Google Scholar] [CrossRef]
  70. Liu, Z.; Zhan, Y.; Tu, Y.; Chen, K.; Liu, Z.; Wu, C. PDZ and LIM domain protein 1(PDLIM1)/CLP36 promotes breast cancer cell migration, invasion and metastasis through interaction with alpha-actinin. Oncogene 2015, 34, 1300–1311. [Google Scholar] [CrossRef]
  71. Tremblay, M.; Sanchez-Ferras, O.; Bouchard, M. GATA transcription factors in development and disease. Development 2018, 145, dev164384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Thomas, J.; A Travers, A. HMG1 and 2, and related ‘architectural’ DNA-binding proteins. Trends Biochem. Sci. 2001, 26, 167–174. [Google Scholar] [CrossRef]
  73. Goodwin, G.H.; Sanders, C.; Johns, E.W. A New Group of Chromatin-Associated Proteins with a High Content of Acidic and Basic Amino Acids. JBIC J. Boil. Inorg. Chem. 1973, 38, 14–19. [Google Scholar] [CrossRef] [PubMed]
  74. Bianchi, M.E.; Agresti, A. HMG proteins: Dynamic players in gene regulation and differentiation. Curr. Opin. Genet. Dev. 2005, 15, 496–506. [Google Scholar] [CrossRef] [PubMed]
  75. Asin-Cayuela, J.; Gustafsson, C.M. Mitochondrial transcription and its regulation in mammalian cells. Trends Biochem. Sci. 2007, 32, 111–117. [Google Scholar] [CrossRef] [PubMed]
  76. Bonawitz, N.D.; Clayton, D.A.; Shadel, G.S. Initiation and Beyond: Multiple Functions of the Human Mitochondrial Transcription Machinery. Mol. Cell 2006, 24, 813–825. [Google Scholar] [CrossRef] [PubMed]
  77. Štros, M. HMGB proteins: Interactions with DNA and chromatin. Biochim. Biophys. Acta (BBA) Bioenergy 2010, 1799, 101–113. [Google Scholar] [CrossRef]
  78. Furusawa, T.; Cherukuri, S. Developmental function of HMGN proteins. Biochim. Biophys. Acta (BBA) Bioenergy 2010, 1799, 69–73. [Google Scholar] [CrossRef] [Green Version]
  79. Reeves, R. Structure and Function of the HMGI(Y) Family of Architectural Transcription Factors. Environ. Health Perspect. 2000, 108, 803. [Google Scholar] [CrossRef]
  80. Reeves, R. Molecular biology of HMGA proteins: Hubs of nuclear function. Gene 2001, 277, 63–81. [Google Scholar] [CrossRef]
  81. Tkachuk, D.C.; Kohler, S.; Cleary, M.L. Involvement of a homolog of Drosophila trithorax by 11q23 chromosomal translocations in acute leukemias. Cell 1992, 71, 691–700. [Google Scholar] [CrossRef]
  82. Maher, J.F.; Nathans, D. Multivalent DNA-binding properties of the HMG-1 proteins. Proc. Natl. Acad. Sci. USA 1996, 93, 6716–6720. [Google Scholar] [CrossRef] [PubMed]
  83. Duguet, M.; de Recondo, A.M. A deoxyribonucleic acid unwinding protein isolated from regenerating rat liver. Physical and functional properties. J. Biol. Chem. 1978, 253, 1660–1666. [Google Scholar] [PubMed]
  84. Bustin, M. Revised nomenclature for high mobility group (HMG) chromosomal proteins. Trends Biochem. Sci. 2001, 26, 152–153. [Google Scholar] [CrossRef]
  85. Körner, U.; Bustin, M.; Scheer, U.; Hock, R. Developmental role of HMGN proteins in Xenopus laevis. Mech. Dev. 2003, 120, 1177–1192. [Google Scholar] [CrossRef]
  86. Furusawa, T.; Lim, J.H.; Catez, F.; Birger, Y.; Mackem, S.; Bustin, M. Down-regulation of nucleosomal binding protein HMGN1 expression during embryogenesis modulates Sox9 expression in chondrocytes. Mol. Cell. Biol. 2006, 26, 592–604. [Google Scholar] [CrossRef]
  87. Birger, Y.; West, K.L.; Postnikov, Y.V.; Lim, J.; Furusawa, T.; Wagner, J.P.; Laufer, C.S.; Kraemer, K.H.; Bustin, M. Chromosomal protein HMGN1 enhances the rate of DNA repair in chromatin. EMBO J. 2003, 22, 1665–1675. [Google Scholar] [CrossRef] [Green Version]
  88. Birger, Y.; Davis, J.; Furusawa, T.; Rand, E.; Piatigorsky, J.; Bustin, M. A role for chromosomal protein HMGN1 in corneal maturation. Differ. Res. Biol. Divers. 2006, 74, 19–29. [Google Scholar] [CrossRef] [Green Version]
  89. Birger, Y.; Catez, F.; Furusawa, T.; Lim, J.-H.; Prymakowska-Bosak, M.; West, K.L.; Postnikov, Y.V.; Haines, D.C.; Bustin, M. Increased tumorigenicity and sensitivity to ionizing radiation upon loss of chromosomal protein HMGN1. Cancer Res. 2005, 65, 6711–6718. [Google Scholar] [CrossRef]
  90. Chieffi, P.; Battista, S.; Barchi, M.; Di Agostino, S.; Pierantoni, G.M.; Fedele, M.; Chiariotti, L.; Tramontano, D.; Fusco, A. HMGA1 and HMGA2 protein expression in mouse spermatogenesis. Oncogene 2002, 21, 3644–3650. [Google Scholar] [CrossRef]
  91. Fedele, M.; Fidanza, V.; Battista, S.; Pentimalli, F.; Klein-Szanto, A.J.; Visone, R.; De Martino, I.; Curcio, A.; Morisco, C.; Del Vecchio, L.; et al. Haploinsufficiency of the Hmga1 Gene Causes Cardiac Hypertrophy and Myelo-Lymphoproliferative Disorders in Mice. Cancer Res. 2006, 66, 2536–2543. [Google Scholar] [CrossRef] [PubMed]
  92. Foti, D.; Chiefari, E.; Fedele, M.; Iuliano, R.; Brunetti, L.; Paonessa, F.; Manfioletti, G.; Barbetti, F.; Brunetti, A.; Croce, C.M.; et al. Lack of the architectural factor HMGA1 causes insulin resistance and diabetes in humans and mice. Nat. Med. 2005, 11, 765–773. [Google Scholar] [CrossRef] [PubMed]
  93. Zhou, X.; Benson, K.F.; Ashar, H.R.; Chada, K. Mutation responsible for the mouse pygmy phenotype in the developmentally regulated factor HMGI-C. Nature 1995, 376, 771–774. [Google Scholar] [CrossRef] [PubMed]
  94. Anand, A.; Chada, K. In vivo modulation of Hmgic reduces obesity. Nat. Genet. 2000, 24, 377–380. [Google Scholar] [CrossRef]
  95. Itou, J.; Taniguchi, N.; Oishi, I.; Kawakami, H.; Lotz, M.; Kawakami, Y. HMGB factors are required for posterior digit development through integrating signaling pathway activities. Dev. Dyn. 2011, 240, 1151–1162. [Google Scholar] [CrossRef] [Green Version]
  96. Abraham, A.B.; Bronstein, R.; I Chen, E.; Koller, A.; Ronfani, L.; Maletic-Savatic, M.; E Tsirka, S. Members of the high mobility group B protein family are dynamically expressed in embryonic neural stem cells. Proteome Sci. 2013, 11, 18. [Google Scholar] [CrossRef]
  97. Nemeth, M.J.; Kirby, M.R.; Bodine, D.M. Hmgb3 regulates the balance between hematopoietic stem cell self-renewal and differentiation. Proc. Natl. Acad. Sci. USA 2006, 103, 13783–13788. [Google Scholar] [CrossRef] [Green Version]
  98. Nemeth, M.J.; Cline, A.P.; Anderson, S.M.; Garrett-Beal, L.J.; Bodine, D.M. Hmgb3 deficiency deregulates proliferation and differentiation of common lymphoid and myeloid progenitors. Blood 2005, 105, 627–634. [Google Scholar] [CrossRef] [Green Version]
  99. Ronfani, L.; Ferraguti, M.; Croci, L.; Ovitt, C.E.; Scholer, H.R.; Consalez, G.G.; Bianchi, M.E. Reduced fertility and spermatogenesis defects in mice lacking chromosomal protein Hmgb2. Development 2001, 128, 1265–1273. [Google Scholar]
  100. Taniguchi, N.; Carames, B.; Lotz, M. 217 Chromatin Protein Hmgb2 Regulates Articular Cartilage Surface Maintenance Via Beta-Catenin Pathways. Osteoarthr. Cartil. 2009, 17, S123. [Google Scholar] [CrossRef]
  101. Taniguchi, N.; Yoshida, K.; Ito, T.; Tsuda, M.; Mishima, Y.; Furumatsu, T.; Ronfani, L.; Abeyama, K.; Kawahara, K.-I.; Komiya, S.; et al. Stage-Specific Secretion of HMGB1 in Cartilage Regulates Endochondral Ossification. Mol. Cell. Boil. 2007, 27, 5650–5663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Taniguchi, N.; Caramés, B.; Ronfani, L.; Ulmer, U.; Komiya, S.; Bianchi, M.E.; Lotz, M. Aging-related loss of the chromatin protein HMGB2 in articular cartilage is linked to reduced cellularity and osteoarthritis. Proc. Natl. Acad. Sci. USA 2009, 106, 1181–1186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Lim, J.-H.; Catez, F.; Birger, Y.; West, K.L.; Prymakowska-Bosak, M.; Postnikov, Y.V.; Bustin, M. Chromosomal Protein HMGN1 Modulates Histone H3 Phosphorylation. Mol. Cell 2004, 15, 573–584. [Google Scholar] [CrossRef] [PubMed]
  104. Arce-Cerezo, A.; Garcia, M.; Rodríguez-Nuevo, A.; Crosa-Bonell, M.; Enguix, N.; Pero, A.; Muñoz, S.; Roca, C.; Ramos, D.; Franckhauser, S.; et al. HMGA1 overexpression in adipose tissue impairs adipogenesis and prevents diet-induced obesity and insulin resistance. Sci. Rep. 2015, 5, 14487. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Shah, S.N.; Cope, L.; Poh, W.; Belton, A.; Roy, S.; Talbot, C.C.; Sukumar, S.; Huso, D.L.; Resar, L.M.S. HMGA1: A Master Regulator of Tumor Progression in Triple-Negative Breast Cancer Cells. PLoS ONE 2013, 8, e63419. [Google Scholar] [CrossRef] [PubMed]
  106. Pallante, P.; Sepe, R.; Puca, F.; Fusco, A. High Mobility Group A Proteins as Tumor Markers. Front. Med. 2015, 2, 15. [Google Scholar] [CrossRef] [PubMed]
  107. Kang, R.; Chen, R.; Zhang, Q.; Hou, W.; Wu, S.; Cao, L.; Huang, J.; Yu, Y.; Fan, X.G.; Yan, Z.; et al. HMGB1 in health and disease. Mol. Asp. Med. 2014, 40, 1–116. [Google Scholar] [CrossRef] [Green Version]
  108. Simon, M.C. Gotta have GATA. Nat Genet. 1995, 11, 9–11. [Google Scholar] [CrossRef]
  109. Weiss, M.J.; Orkin, S.H. GATA transcription factors: Key regulators of hematopoiesis. Exp. Hematol. 1995, 23, 99–107. [Google Scholar]
  110. Zheng, R.; Blobel, G.A. GATA Transcription Factors and Cancer. Genes Cancer 2010, 1, 1178–1188. [Google Scholar] [CrossRef] [Green Version]
  111. Grass, J.A.; Boyer, M.E.; Pal, S.; Wu, J.; Weiss, M.J.; Bresnick, E.H. GATA-1-dependent transcriptional repression of GATA-2 via disruption of positive autoregulation and domain-wide chromatin remodeling. Proc. Natl. Acad. Sci. USA 2003, 100, 8811–8816. [Google Scholar] [CrossRef]
  112. Zhang, S.-J.; Ma, L.-Y.; Huang, Q.-H.; Li, G.; Gu, B.-W.; Gao, X.-D.; Shi, J.-Y.; Wang, Y.-Y.; Gao, L.; Cai, X.; et al. Gain-of-function mutation of GATA-2 in acute myeloid transformation of chronic myeloid leukemia. Proc. Natl. Acad. Sci. USA 2008, 105, 2076–2081. [Google Scholar] [CrossRef] [PubMed]
  113. Rodriguez-Bravo, V.; Carceles-Cordon, M.; Hoshida, Y.; Cordon-Cardo, C.; Galsky, M.D.; Domingo-Domenech, J. The role of GATA2 in lethal prostate cancer aggressiveness. Nat. Rev. Urol. 2017, 14, 38–48. [Google Scholar] [CrossRef] [PubMed]
  114. Gao, L.; Hu, Y.; Tian, Y.; Fan, Z.; Wang, K.; Li, H.; Zhou, Q.; Zeng, G.; Hu, X.; Yu, L.; et al. Lung cancer deficient in the tumor suppressor GATA4 is sensitive to TGFBR1 inhibition. Nat. Commun. 2019, 10, 1665. [Google Scholar] [CrossRef] [PubMed]
  115. Ho, I.C.; Vorhees, P.; Marin, N.; Oakley, B.K.; Tsai, S.F.; Orkin, S.H.; Leiden, J.M. Human GATA-3: A lineage-restricted transcription factor that regulates the expression of the T cell receptor alpha gene. EMBO J. 1991, 10, 1187–1192. [Google Scholar] [CrossRef] [PubMed]
  116. Usary, J.; Llaca, V.; Karaca, G.; Presswala, S.; Karaca, M.; He, X.; Langerød, A.; Kåresen, R.; Oh, D.S.; Dressler, L.G.; et al. Mutation of GATA3 in human breast tumors. Oncogene 2004, 23, 7669–7678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Heineke, J.; Auger-Messier, M.; Xu, J.; Oka, T.; Sargent, M.A.; York, A.; Klevitsky, R.; Vaikunth, S.; Duncan, S.A.; Aronow, B.J.; et al. Cardiomyocyte GATA4 functions as a stress-responsive regulator of angiogenesis in the murine heart. J. Clin. Investig. 2007, 117, 3198–3210. [Google Scholar] [CrossRef] [Green Version]
  118. Beuling, E.; Baffour–Awuah, N.Y.A.; Stapleton, K.A.; Aronson, B.E.; Noah, T.K.; Shroyer, N.F.; Duncan, S.A.; Fleet, J.C.; Krasinski, S.D. GATA factors regulate proliferation, differentiation, and gene expression in small intestine of mature mice. Gastroenterol. 2011, 140, 1219–1229. [Google Scholar] [CrossRef]
  119. Akiyama, Y.; Watkins, N.; Suzuki, H.; Jair, K.-W.; Van Engeland, M.; Esteller, M.; Sakai, H.; Ren, C.-Y.; Yuasa, Y.; Herman, J.G.; et al. GATA-4 and GATA-5 Transcription Factor Genes and Potential Downstream Antitumor Target Genes Are Epigenetically Silenced in Colorectal and Gastric Cancer. Mol. Cell. Boil. 2003, 23, 8429–8439. [Google Scholar] [CrossRef]
  120. Gong, Y.; Zhang, L.; Zhang, A.; Chen, X.; Gao, P.; Zeng, Q. GATA4 inhibits cell differentiation and proliferation in pancreatic cancer. PLoS ONE 2018, 13, e0202449. [Google Scholar] [CrossRef]
  121. Ip, H.S.; Morrisey, E.E.; Tang, Z.; Parmacek, M.S. GATA-4 Activates Transcription Via Two Novel Domains That Are Conserved within the GATA-4/5/6 Subfamily. J. Boil. Chem. 1997, 272, 8515–8524. [Google Scholar] [Green Version]
  122. Laforest, B.; Nemer, M. GATA5 interacts with GATA4 and GATA6 in outflow tract development. Dev. Boil. 2011, 358, 368–378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Feng, H.; Zhu, M.; Zhang, R.; Wang, Q.; Li, W.; Dong, X.; Chen, Y.; Lu, Y.; Liu, K.; Lin, B.; et al. GATA5 inhibits hepatocellular carcinoma cells malignant behaviours by blocking expression of reprogramming genes. J. Cell. Mol. Med. 2019, 23, 2536–2548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Lepore, J.J.; Mericko, P.A.; Cheng, L.; Lu, M.M.; Morrisey, E.E.; Parmacek, M.S. GATA-6 regulates semaphorin 3C and is required in cardiac neural crest for cardiovascular morphogenesis. J. Clin. Investig. 2006, 116, 929–939. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Tian, Y.; Yuan, L.; Goss, A.M.; Wang, T.; Yang, J.; Lepore, J.J.; Zhou, D.; Schwartz, R.J.; Patel, V.; Cohen, E.D.; et al. Characterization and in vivo pharmacological rescue of a Wnt2-Gata6 pathway required for cardiac inflow tract development. Dev. Cell 2010, 18, 275–287. [Google Scholar] [CrossRef] [PubMed]
  126. Kamnasaran, D.; Qian, B.; Hawkins, C.; Stanford, W.L.; Guha, A. GATA6 is an astrocytoma tumor suppressor gene identified by gene trapping of mouse glioma model. Proc. Natl. Acad. Sci. USA 2007, 104, 8053–8058. [Google Scholar] [CrossRef] [PubMed]
  127. Martinelli, P.; Carrillo-de Santa Pau, E.; Cox, T.; Sainz, B., Jr.; Dusetti, N.; Greenhalf, W.; Rinaldi, L.; Costello, E.; Ghaneh, P.; Malats, N.; et al. GATA6 regulates EMT and tumour dissemination, and is a marker of response to adjuvant chemotherapy in pancreatic cancer. Gut 2017, 66, 1665–1676. [Google Scholar] [CrossRef]
  128. Shen, W.; Niu, N.; Lawson, B.; Qi, L.; Zhang, J.; Li, T.; Zhang, H.; Liu, J. GATA6: A new predictor for prognosis in ovarian cancer. Hum. Pathol. 2019, 86, 163–169. [Google Scholar] [CrossRef]
  129. Peters, H.; Doll, U.; Niessing, J. Differential expression of the chicken Pax-1 and Pax-9 Gene: In situ hybridization and immunohistochemical analysis. Dev. Dyn. 1995, 203, 1–16. [Google Scholar] [CrossRef]
  130. Lai, H.-C.; Lin, Y.-W.; Huang, T.H.; Yan, P.; Huang, R.-L.; Wang, H.-C.; Liu, J.; Chan, M.W.; Chu, T.-Y.; Sun, C.-A.; et al. Identification of novel DNA methylation markers in cervical cancer. Int. J. Cancer 2008, 123, 161–167. [Google Scholar] [CrossRef]
  131. Imgrund, M.; Gröne, E.; Gröne, H.-J.; Kretzler, M.; Holzman, L.; Schlöndorff, D.; Rothenpieler, U.W. Re-expression of the developmental gene Pax-2 during experimental acute tubular necrosis in mice1. Kidney Int. 1999, 56, 1423–1431. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Maeshima, A.; Nojima, Y.; Kojima, I. Involvement of Pax-2 in the Action of Activin A on Tubular Cell Regeneration. J. Am. Soc. Nephrol. 2002, 13, 2850–2859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Doberstein, K.; Pfeilschifter, J.; Gutwein, P. The transcription factor PAX2 regulates ADAM10 expression in renal cell carcinoma. Carcinogenesis 2011, 32, 1713–1723. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Wu, H.; Chen, Y.; Liang, J.; Shi, B.; Wu, G.; Zhang, Y.; Wang, D.; Li, R.; Yi, X.; Zhang, H.; et al. Hypomethylation-linked activation of PAX2 mediates tamoxifen-stimulated endometrial carcinogenesis. Nature 2005, 438, 981–987. [Google Scholar] [CrossRef] [PubMed]
  135. Koblar, S.A.; Murphy, M.; Barrett, G.L.; Underhill, A.; Gros, P.; Bartlett, P.F. Pax3 regulates neurogenesis in neural crest-derived precursor cells. J. Neurosci. Res. 1999, 56, 518–530. [Google Scholar] [CrossRef]
  136. Lang, D.; Lu, M.M.; Huang, L.; Engleka, K.A.; Zhang, M.; Chu, E.Y.; Lipner, S.; Skoultchi, A.; Millar, S.E.; Epstein, J.A. Pax3 functions at a nodal point in melanocyte stem cell differentiation. Nature 2005, 433, 884–887. [Google Scholar] [CrossRef]
  137. Galili, N.; Davis, R.J.; Fredericks, W.J.; Mukhopadhyay, S.; Rauscher, F.J.; Emanuel, B.S.; Rovera, G.; Barr, F.G. Fusion of a fork head domain gene to PAX3 in the solid tumour alveolar rhabdomyosarcoma. Nat. Genet. 1993, 5, 230–235. [Google Scholar] [CrossRef]
  138. Bennicelli, J.L.; Edwards, R.H.; Barr, F.G. Mechanism for transcriptional gain of function resulting from chromosomal translocation in alveolar rhabdomyosarcoma. Proc. Natl. Acad. Sci. USA 1996, 93, 5455–5459. [Google Scholar] [CrossRef]
  139. Takeuchi, H.; Morton, D.L.; Kuo, C.; Turner, R.R.; Elashoff, D.; Elashoff, R.; Taback, B.; Fujimoto, A.; Hoon, D.S. Prognostic significance of molecular upstaging of paraffin-embedded sentinel lymph nodes in melanoma patients. J. Clin. Oncol. 2004, 22, 2671–2680. [Google Scholar] [CrossRef]
  140. Brun, T.; Franklin, I.; St-Onge, L.; Biason-Lauber, A.; Schoenle, E.J.; Wollheim, C.B.; Gauthier, B.R. The diabetes-linked transcription factor PAX4 promotes {beta}-cell proliferation and survival in rat and human islets. J. Cell Biol. 2004, 167, 1123–1135. [Google Scholar] [CrossRef]
  141. Brun, T.; Duhamel, D.L.; Hu He, K.H.; Wollheim, C.B.; Gauthier, B.R. The transcription factor PAX4 acts as a survival gene in INS-1E insulinoma cells. Oncogene 2007, 26, 4261–4271. [Google Scholar] [CrossRef] [PubMed]
  142. Sanz, E.; Alvarez-Mon, M.; Martinez, A.C.; de la Hera, A. Human cord blood CD34+Pax-5+ B-cell progenitors: Single-cell analyses of their gene expression profiles. Blood 2003, 101, 3424–3430. [Google Scholar] [CrossRef] [PubMed]
  143. Krenacs, L.; Himmelmann, A.W.; Quintanilla-Martinez, L.; Fest, T.; Riva, A.; Wellmann, A.; Bagdi, E.; Kehrl, J.H.; Jaffe, E.S.; Raffeld, M. Transcription Factor B-Cell–Specific Activator Protein (BSAP) Is Differentially Expressed in B Cells. Blood 1998, 1308–1316. [Google Scholar]
  144. Liu, W.; Li, X.; Chu, E.S.; Go, M.Y.; Xu, L.; Zhao, G.; Li, L.; Dai, N.; Si, J.; Tao, Q.; et al. Paired box gene 5 is a novel tumor suppressor in hepatocellular carcinoma through interaction with p53 signaling pathway. Hepatology 2011, 53, 843–853. [Google Scholar] [CrossRef] [PubMed]
  145. Vidal, L.J.; Perry, J.K.; Vouyovitch, C.M.; Pandey, V.; Brunet-Dunand, S.E.; Mertani, H.C.; Liu, D.X.; Lobie, P.E. PAX5alpha enhances the epithelial behavior of human mammary carcinoma cells. Mol. Cancer Res. MCR 2010, 8, 444–456. [Google Scholar] [CrossRef] [PubMed]
  146. Chi, N.; Epstein, J.A. Getting your Pax straight: Pax proteins in development and disease. Trends Genet. 2002, 18, 41–47. [Google Scholar] [CrossRef]
  147. Mascarenhas, J.B.; Young, K.P.; Littlejohn, E.L.; Yoo, B.K.; Salgia, R.; Lang, D. PAX6 Is Expressed in Pancreatic Cancer and Actively Participates in Cancer Progression through Activation of the MET Tyrosine Kinase Receptor Gene. J. Boil. Chem. 2009, 284, 27524–27532. [Google Scholar] [CrossRef] [Green Version]
  148. Mayes, D.A.; Hu, Y.; Teng, Y.; Siegel, E.; Wu, X.; Panda, K.; Tan, F.; Yung, W.A.; Zhou, Y.-H. PAX6 Suppresses the Invasiveness of Glioblastoma Cells and the Expression of the Matrix Metalloproteinase-2 Gene. Cancer Res. 2006, 66, 9809–9817. [Google Scholar] [CrossRef] [Green Version]
  149. Zhou, Y.H.; Hu, Y.; Mayes, D.; Siegel, E.; Kim, J.G.; Mathews, M.S.; Hsu, N.; Eskander, D.; Yu, O.; Tromberg, B.J.; et al. PAX6 suppression of glioma angiogenesis and the expression of vascular endothelial growth factor A. J. Neuro-Oncol. 2010, 96, 191–200. [Google Scholar] [CrossRef]
  150. Buckingham, M.; Relaix, F. The Role ofPaxGenes in the Development of Tissues and Organs: Pax3 and Pax7 Regulate Muscle Progenitor Cell Functions. Annu. Rev. Cell Dev. Boil. 2007, 23, 645–673. [Google Scholar] [CrossRef]
  151. Zannini, M.; Francis-Lang, H.; Plachov, D.; Di Lauro, R. Pax-8, a paired domain-containing protein, binds to a sequence overlapping the recognition site of a homeodomain and activates transcription from two thyroid-specific promoters. Mol. Cell. Boil. 1992, 12, 4230–4241. [Google Scholar] [CrossRef] [PubMed]
  152. Tong, G.-X.; Yu, W.M.; Beaubier, N.T.; Weeden, E.M.; Hamele-Bena, D.; Mansukhani, M.M.; O’Toole, K.M. Expression of PAX8 in normal and neoplastic renal tissues: An immunohistochemical study. Mod. Pathol. 2009, 22, 1218–1227. [Google Scholar] [CrossRef] [PubMed]
  153. Chen, Y.-J.; Campbell, H.G.; Wiles, A.K.; Eccles, M.R.; Reddel, R.R.; Braithwaite, A.W.; Royds, J.A. PAX8 Regulates Telomerase Reverse Transcriptase and Telomerase RNA Component in Glioma. Cancer Res. 2008, 68, 5724–5732. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Suda, N.; Ogawa, T.; Kojima, T.; Saito, C.; Moriyama, K. Non-syndromic oligodontia with a novel mutation of PAX9. J. Dent. Res. 2011, 90, 382–386. [Google Scholar] [CrossRef]
  155. Kendall, J.; Liu, Q.; Bakleh, A.; Krasnitz, A.; Nguyen, K.C.Q.; Lakshmi, B.; Gerald, W.L.; Powers, S.; Mu, D. Oncogenic cooperation and coamplification of developmental transcription factor genes in lung cancer. Proc. Natl. Acad. Sci. USA 2007, 104, 16663–16668. [Google Scholar] [CrossRef] [Green Version]
  156. Lee, J.C.; Sharma, M.; Lee, Y.H.; Lee, N.H.; Kim, S.Y.; Yun, J.S.; Nam, S.Y.; Hwang, P.H.; Jhee, E.C.; Yi, H.K. Pax9 mediated cell survival in oral squamous carcinoma cell enhanced by c-myb. Cell Biochem. Funct. 2008, 26, 892–899. [Google Scholar] [CrossRef]
  157. Huang, C.; Chan, J.A.; Schuurmans, C. Proneural bHLH Genes in Development and Disease. Curr. Top Dev. Biol. 2014, 110, 75–127. [Google Scholar] [CrossRef]
  158. Bialek, P.; Kern, B.; Yang, X.; Schrock, M.; Sosic, D.; Hong, N.; Wu, H.; Yu, K.; Ornitz, D.M.; Olson, E.N.; et al. A twist code determines the onset of osteoblast differentiation. Dev. Cell 2004, 6, 423–435. [Google Scholar] [CrossRef]
  159. Yang, J.; A Mani, S.; Donaher, J.L.; Ramaswamy, S.; Itzykson, R.A.; Come, C.; Savagner, P.; Gitelman, I.; Richardson, A.; A Weinberg, R. Twist, a Master Regulator of Morphogenesis, Plays an Essential Role in Tumor Metastasis. Cell 2004, 117, 927–939. [Google Scholar] [CrossRef] [Green Version]
  160. Chang, A.T.; Liu, Y.; Ayyanathan, K.; Benner, C.; Jiang, Y.; Prokop, J.W.; Paz, H.; Wang, D.; Li, H.-R.; Fu, X.-D.; et al. An evolutionarily conserved DNA architecture determines target specificity of the TWIST family bHLH transcription factors. Genes Dev. 2015, 29, 603–616. [Google Scholar] [CrossRef] [Green Version]
  161. Tsai, J.H.; Yang, J. Epithelial–mesenchymal plasticity in carcinoma metastasis. Genes Dev. 2013, 27, 2192–2206. [Google Scholar] [CrossRef] [PubMed]
  162. Hurlin, P.J. Control of Vertebrate Development by MYC. Cold Spring Harb. Perspect. Med. 2013, 3, a014332. [Google Scholar] [CrossRef] [PubMed]
  163. Witkiewicz, A.K.; McMillan, E.A.; Balaji, U.; Baek, G.; Lin, W.-C.; Mansour, J.; Mollaee, M.; Wagner, K.-U.; Koduru, P.; Yopp, A.; et al. Whole-exome sequencing of pancreatic cancer defines genetic diversity and therapeutic targets. Nat. Commun. 2015, 6, 6744. [Google Scholar] [CrossRef] [PubMed]
  164. Chesi, M.; Robbiani, D.F.; Sebag, M.; Chng, W.J.; Affer, M.; Tiedemann, R.; Valdez, R.; Palmer, S.E.; Haas, S.S.; Stewart, A.K.; et al. AID-dependent activation of a MYC transgene induces multiple myeloma in a conditional mouse model of post-germinal center malignancies. Cancer Cell 2008, 13, 167–180. [Google Scholar] [CrossRef]
  165. Bermingham, N.A. Math1: An Essential Gene for the Generation of Inner Ear Hair Cells. Science 1999, 284, 1837–1841. [Google Scholar] [CrossRef]
  166. Price, S.D.; Shope, C.; Himes, D.; Liu, M.; Pereira, F.A.; Chu, M.-J.; Eatock, R.A.; Brownell, W.E.; Lysakowski, A.; Tsai, M.-J. Essential role of BETA2/NeuroD1 in development of the vestibular and auditory systems. Genes Dev. 2000, 14, 2839–2854. [Google Scholar]
  167. Boulay, G.; Sandoval, G.J.; Riggi, N.; Iyer, S.; Buisson, R.; Naigles, B.; Awad, M.E.; Rengarajan, S.; Volorio, A.; McBride, M.J.; et al. Cancer-Specific Retargeting of BAF Complexes by a Prion-like Domain. Cell 2017, 171, 163–178. [Google Scholar] [CrossRef]
  168. Khan, S.; Stott, S.R.; Chabrat, A.; Truckenbrodt, A.M.; Spencer-Dene, B.; Nave, K.-A.; Guillemot, F.; Levesque, M.; Ang, S.-L. Survival of a Novel Subset of Midbrain Dopaminergic Neurons Projecting to the Lateral Septum Is Dependent on NeuroD Proteins. J. Neurosci. 2017, 37, 2305–2316. [Google Scholar] [CrossRef] [Green Version]
  169. Laclé, M.M.; Van Diest, P.J.; Goldschmeding, R.; Van Der Wall, E.; Nguyen, T.Q. Expression of Connective Tissue Growth Factor in Male Breast Cancer: Clinicopathologic Correlations and Prognostic Value. PLoS ONE 2015, 10, e0118957. [Google Scholar] [CrossRef]
  170. Barnes, R.M.; Firulli, B.A.; Conway, S.J.; Vincentz, J.W.; Firulli, A.B. Analysis of the Hand1 cell lineage reveals novel contributions to cardiovascular, neural crest, extra-embryonic, and lateral mesoderm derivatives. Dev. Dyn. 2010, 239, 3086–3097. [Google Scholar] [CrossRef] [Green Version]
  171. Barnes, R.M.; Firulli, B.A.; VanDusen, N.J.; Morikawa, Y.; Conway, S.J.; Cserjesi, P.; Vincentz, J.W.; Firulli, A.B. Hand2 loss-of-function in Hand1-expressing cells reveals distinct roles in epicardial and coronary vessel development. Circ. Res. 2011, 108, 940–949. [Google Scholar] [CrossRef] [PubMed]
  172. Martı́nez-Espinoza, A.D.; Garcı́a-Pedrajas, M.D.; Gold, S.E. The Ustilaginales as Plant Pests and Model Systems. Fungal Genet. Boil. 2002, 35, 1–20. [Google Scholar]
  173. Kato, N.; Iwase, A.; Ishida, C.; Nagai, T.; Mori, M.; Bayasula; Nakamura, T.; Osuka, S.; Ganiyeva, U.; Qin, Y.; et al. Upregulation of Fibroblast Growth Factors Caused by Heart and Neural Crest Derivatives Expressed 2 Suppression in Endometriotic Cells: A Possible Therapeutic Target in Endometriosis. Reprod. Sci. 2019, 26, 979–987. [Google Scholar] [CrossRef] [PubMed]
  174. Zhou, Q.; Anderson, D.J. The bHLH transcription factors OLIG2 and OLIG1 couple neuronal and glial subtype specification. Cell 2002, 109, 61–73. [Google Scholar] [CrossRef]
  175. Brena, R.M.; Morrison, C.; Liyanarachchi, S.; Jarjoura, D.; Davuluri, R.V.; Otterson, G.A.; Reisman, D.; Glaros, S.; Rush, L.J.; Plass, C. Aberrant DNA Methylation of OLIG1, a Novel Prognostic Factor in Non-Small Cell Lung Cancer. PLoS Med. 2007, 4, e108. [Google Scholar] [CrossRef]
  176. Kosty, J.; Lu, F.; Kupp, R.; Mehta, S.; Lu, Q.R. Harnessing OLIG2 function in tumorigenicity and plasticity to target malignant gliomas. Cell Cycle 2017, 16, 1654–1660. [Google Scholar] [CrossRef] [Green Version]
  177. MacLean, H.E.; Kronenberg, H.M. Expression of Stra13 during mouse endochondral bone development. Gene Expr. Patterns 2004, 4, 633–636. [Google Scholar] [CrossRef]
  178. Huang, K.-L.; Mashl, R.J.; Wu, Y.; Ritter, D.I.; Wang, J.; Oh, C.; Paczkowska, M.; Wyczalkowski, M.A.; Oak, N.; Scott, A.D.; et al. Pathogenic Germline Variants in 10,389 Adult Cancers. Cell 2018, 173, 355–370. [Google Scholar] [CrossRef]
  179. Bhawal, U.K.; Sato, F.; Arakawa, Y.; Fujimoto, K.; Kawamoto, T.; Tanimoto, K.; Ito, Y.; Sasahira, T.; Sakurai, T.; Kobayashi, M.; et al. Basic helix-loop-helix transcription factor DEC1 negatively regulates cyclin D1. J. Pathol. 2011, 224, 420–429. [Google Scholar] [CrossRef]
  180. Sasamoto, T.; Fujimoto, K.; Kanawa, M.; Kimura, J.; Takeuchi, J.; Harada, N.; Goto, N.; Kawamoto, T.; Noshiro, M.; Suardita, K.; et al. DEC2 is a negative regulator for the proliferation and differentiation of chondrocyte lineage-committed mesenchymal stem cells. Int. J. Mol. Med. 2016, 38, 876–884. [Google Scholar] [CrossRef]
  181. Liu, J.; Uygur, B.; Zhang, Z.; Shao, L.; Romero, D.; Vary, C.; Ding, Q.; Wu, W.-S. Slug inhibits proliferation of human prostate cancer cells via downregulation of cyclin D1 expression. Prostate 2010, 70, 1768–1777. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Blanpain, C.; Lowry, W.E.; Pasolli, H.A.; Fuchs, E. Canonical notch signaling functions as a commitment switch in the epidermal lineage. Genome Res. 2006, 20, 3022–3035. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Ichijo, R.; Iizuka, Y.; Kubo, H.; Toyoshima, F. Essential roles of Tbx3 in embryonic skin development during epidermal stratification. Genes Cells 2017, 22, 284–292. [Google Scholar] [CrossRef] [PubMed]
  184. Kobayashi, T.; Kageyama, R. Hes1 regulates embryonic stem cell differentiation by suppressing Notch signaling. Genes Cells 2010, 15, 689–698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Imayoshi, I.; Ishidate, F.; Kageyama, R. Real-time imaging of bHLH transcription factors reveals their dynamic control in the multipotency and fate choice of neural stem cells. Front. Cell. Neurosci. 2015, 9, 288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Baek, J.H.; Hatakeyama, J.; Sakamoto, S.; Ohtsuka, T.; Kageyama, R. Persistent and high levels of Hes1 expression regulate boundary formation in the developing central nervous system. Development 2006, 133, 2467–2476. [Google Scholar] [CrossRef] [Green Version]
  187. Carré, A.; Rachdi, L.; Tron, E.; Richard, B.; Castanet, M.; Schlumberger, M.; Bidart, J.-M.; Szinnai, G.; Polak, M. Hes1 Is Required for Appropriate Morphogenesis and Differentiation during Mouse Thyroid Gland Development. PLoS ONE 2011, 6, e16752. [Google Scholar] [CrossRef]
  188. Huang, Q.; Raya, A.; DeJesus, P.; Chao, S.-H.; Quon, K.C.; Caldwell, J.S.; Chanda, S.K.; Izpisua-Belmonte, J.C.; Schultz, P.G. Identification of p53 regulators by genome-wide functional analysis. Proc. Natl. Acad. Sci. USA 2004, 101, 3456–3461. [Google Scholar] [CrossRef] [Green Version]
  189. Fischer, A.; Schumacher, N.; Maier, M.; Sendtner, M.; Gessler, M. The Notch target genes Hey1 and Hey2 are required for embryonic vascular development. Genome Res. 2004, 18, 901–911. [Google Scholar] [CrossRef]
  190. Benito-Gonzalez, A.; Doetzlhofer, A. Hey1 and Hey2 control the spatial and temporal pattern of mammalian auditory hair cell differentiation downstream of Hedgehog signaling. J. Neurosci. 2014, 34, 12865–12876. [Google Scholar] [CrossRef]
  191. Yin, X.; Zeng, Z.; Xing, J.; Zhang, A.; Jiang, W.; Wang, W.; Sun, H.; Ni, L. Hey1 functions as a positive regulator of odontogenic differentiation in odontoblastlineage cells. Int. J. Mol. Med. 2018, 41, 331–339. [Google Scholar] [PubMed]
  192. Laudet, V.; Stéhelin, D.; Clevers, H. Ancestry and diversity of the HMG box superfamily. Nucleic Acids Res. 1993, 21, 2493–2501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Štros, M.; Launholt, D.; Grasser, K.D. The HMG-box: A versatile protein domain occurring in a wide variety of DNA-binding proteins. Cell. Mol. Life Sci. 2007, 64, 2590–2606. [Google Scholar] [CrossRef] [PubMed]
  194. Sessa, L.; Bianchi, M.E. The evolution of High Mobility Group Box (HMGB) chromatin proteins in multicellular animals. Gene 2007, 387, 133–140. [Google Scholar] [CrossRef] [PubMed]
  195. Evans, T.; Felsenfeld, G. The erythroid-specific transcription factor eryf1: A new finger protein. Cell 1989, 58, 877–885. [Google Scholar] [CrossRef]
  196. Katsumura, K.R.; Bresnick, E.H. GATA Factor Mechanisms Group; the GATA Factor Mechanisms Group the GATA factor revolution in hematology. Blood 2017, 129, 2092–2102. [Google Scholar] [CrossRef]
  197. Lentjes, M.H.; Niessen, H.E.; Akiyama, Y.; De Bruïne, A.P.; Melotte, V.; Van Engeland, M. The emerging role of GATA transcription factors in development and disease. Expert Rev. Mol. Med. 2016, 18, e3. [Google Scholar] [CrossRef]
  198. Simon, M.C.; Pevny, L.; Wiles, M.V.; Keller, G.; Costantini, F.; Orkin, S.H. Rescue of erythroid development in gene targeted GATA–1−mouse embryonic stem cells. Nat. Genet. 1992, 1, 92–98. [Google Scholar] [CrossRef]
  199. Fujiwara, Y.; Browne, C.P.; Cunniff, K.; Goff, S.C.; Orkin, S.H. Arrested development of embryonic red cell precursors in mouse embryos lacking transcription factor GATA-1. Proc. Natl. Acad. Sci. USA 1996, 93, 12355–12358. [Google Scholar] [CrossRef]
  200. Hosoya, T.; Maillard, I.; Engel, J.D. From the cradle to the grave: Activities of GATA-3 throughout T-cell development and differentiation. Immunol. Rev. 2010, 238, 110–125. [Google Scholar] [CrossRef]
  201. Frelin, C.; Herrington, R.; Janmohamed, S.; Barbara, M.; Tran, G.; Paige, C.J.; Benveniste, P.; Zúñiga-Pflücker, J.-C.; Souabni, A.; Busslinger, M.; et al. GATA-3 regulates the self-renewal of long-term hematopoietic stem cells. Nat. Immunol. 2013, 14, 1037–1044. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Fitch, S.R.; Kimber, G.M.; Wilson, N.K.; Parker, A.; Mirshekar-Syahkal, B.; Göttgens, B.; Medvinsky, A.; Dzierzak, E.; Ottersbach, K. Signaling from the sympathetic nervous system regulates hematopoietic stem cell emergence during embryogenesis. Cell Stem Cell 2012, 11, 554–566. [Google Scholar] [CrossRef] [PubMed]
  203. Yamak, A.; Latinkić, B.V.; Dali, R.; Temsah, R.; Nemer, M. Cyclin D2 is a GATA4 cofactor in cardiogenesis. Proc. Natl. Acad. Sci. USA 2014, 111, 1415–1420. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Zhao, R.; Watt, A.J.; Li, J.; Luebke-Wheeler, J.; Morrisey, E.E.; Duncan, S.A. GATA6 Is Essential for Embryonic Development of the Liver but Dispensable for Early Heart Formation. Mol. Cell. Boil. 2005, 25, 2622–2631. [Google Scholar] [CrossRef] [Green Version]
  205. Scazzocchio, C. The fungal GATA factors. Curr. Opin. Microbiol. 2000, 3, 126–131. [Google Scholar] [CrossRef]
  206. Lowry, J.A.; Atchley, W.R. Molecular Evolution of the GATA Family of Transcription Factors: Conservation Within the DNA-Binding Domain. J. Mol. Evol. 2000, 50, 103–115. [Google Scholar] [CrossRef]
  207. He, C.; Cheng, H.; Zhou, R. GATA family of transcription factors of vertebrates: Phylogenetics and chromosomal synteny. J. Biosci. 2007, 32, 1273–1280. [Google Scholar] [CrossRef]
  208. Reyes, J.C.; Muro-Pastor, M.I.; Florencio, F.J. The GATA Family of Transcription Factors in Arabidopsis and Rice1. Plant Physiol. 2004, 134, 1718–1732. [Google Scholar] [CrossRef]
  209. Arguello-Astorga, G.; Herrera-Estrella, L. Evolution of Light-Regulated Plant Promoters. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1998, 49, 525–555. [Google Scholar] [CrossRef]
  210. Shimizu, R.; Ohneda, K.; Engel, J.D.; Trainor, C.D.; Yamamoto, M. Transgenic rescue of GATA-1-deficient mice with GATA-1 lacking a FOG-1 association site phenocopies patients with X-linked thrombocytopenia. Blood 2004, 103, 2560–2567. [Google Scholar] [CrossRef]
  211. Chang, A.N.; Cantor, A.B.; Fujiwara, Y.; Lodish, M.B.; Droho, S.; Crispino, J.D.; Orkin, S.H. GATA-factor dependence of the multitype zinc-finger protein FOG-1 for its essential role in megakaryopoiesis. Proc. Natl. Acad. Sci. USA 2002, 99, 9237–9242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Kumar, M.S.; Hancock, D.C.; Molina-Arcas, M.; Steckel, M.; East, P.; Diefenbacher, M.; Armenteros-Monterroso, E.; Lassailly, F.; Matthews, N.; Nye, E.; et al. The GATA2 Transcriptional Network Is Requisite for RAS Oncogene-Driven Non-Small Cell Lung Cancer. Cell 2012, 149, 642–655. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Kouros-Mehr, H.; Bechis, S.K.; Slorach, E.M.; Littlepage, L.E.; Egeblad, M.; Ewald, A.J.; Pai, S.-Y.; Ho, I.-C.; Werb, Z. GATA-3 links tumor differentiation and dissemination in a luminal breast cancer model. Cancer Cell 2008, 13, 141–152. [Google Scholar] [CrossRef] [PubMed]
  214. Dydensborg, A.B.; Rose, A.A.N.; Wilson, B.J.; Grote, D.; Paquet, M.; Giguère, V.; Siegel, P.M.; Bouchard, M. GATA3 inhibits breast cancer growth and pulmonary breast cancer metastasis. Oncogene 2009, 28, 2634–2642. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Chou, J.; Lin, J.H.; Brenot, A.; Kim, J.-W.; Provot, S.; Werb, Z. GATA3 suppresses metastasis and modulates the tumour microenvironment by regulating microRNA-29b expression. Nat. Cell Biol. 2013, 15, 201–213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Fu, B.; Luo, M.; Lakkur, S.; Lucito, R.; Iacobuzio-Donahue, C.A. Frequent genomic copy number gain and overexpression of GATA-6 in pancreatic carcinoma. Cancer Boil. Ther. 2008, 7, 1593–1601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Shureiqi, I.; Zuo, X.; Broaddus, R.; Wu, Y.; Guan, B.; Morris, J.S.; Lippman, S.M. The transcription factor GATA-6 is overexpressed in vivo and contributes to silencing 15-LOX-1 in vitro in human colon cancer. FASEB J. 2007, 21, 743–753. [Google Scholar] [CrossRef]
  218. Lang, D.; Powell, S.K.; Plummer, R.S.; Young, K.P.; Ruggeri, B.A. PAX genes: Roles in development, pathophysiology, and cancer. Biochem. Pharmacol. 2007, 73, 1–14. [Google Scholar] [CrossRef]
  219. Buckingham, M. Skeletal muscle progenitor cells and the role of Pax genes. C. R. Boil. 2007, 330, 530–533. [Google Scholar] [CrossRef]
  220. Vorobyov, E.; Horst, J. Getting the proto-Pax by the tail. J. Mol. Evol. 2006, 63, 153–164. [Google Scholar] [CrossRef]
  221. Eberhard, D.; Jiménez, G.; Heavey, B.; Busslinger, M. Transcriptional repression by Pax5 (BSAP) through interaction with corepressors of the Groucho family. EMBO J. 2000, 19, 2292–2303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Apuzzo, S.; Gros, P. Cooperative Interactions between the Two DNA Binding Domains of Pax3: Helix 2 of the Paired Domain Is in the Proximity of the Amino Terminus of the Homeodomain. Biochemistry 2007, 46, 2984–2993. [Google Scholar] [CrossRef] [PubMed]
  223. Wang, Q.; Fang, W.-H.; Krupinski, J.; Kumar, S.; Slevin, M.; Kumar, P. Pax genes in embryogenesis and oncogenesis. J. Cell. Mol. Med. 2008, 12, 2281–2294. [Google Scholar] [CrossRef] [PubMed]
  224. Robichaud, G.A.; Nardini, M.; Laflamme, M.; Čuperlović-Culf, M.; Ouellette, R.J. Human Pax-5 C-terminal Isoforms Possess Distinct Transactivation Properties and Are Differentially Modulated in Normal and Malignant B Cells. J. Boil. Chem. 2004, 279, 49956–49963. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Robson, E.J.D.; He, S.-J.; Eccles, M.R. A PANorama of PAX genes in cancer and development. Nat. Rev. Cancer 2006, 6, 52–62. [Google Scholar] [CrossRef] [PubMed]
  226. Tong, G.-X.; Melamed, J.; Mansukhani, M.; Memeo, L.; Hernandez, O.; Deng, F.-M.; Chiriboga, L.; Waisman, J. PAX2: A reliable marker for nephrogenic adenoma. Mod. Pathol. 2006, 19, 356–363. [Google Scholar] [CrossRef]
  227. Tong, G.X.; Chiriboga, L.; Hamele-Bena, D.; Borczuk, A.C. Expression of PAX2 in papillary serous carcinoma of the ovary: Immunohistochemical evidence of fallopian tube or secondary Mullerian system origin? Mod. Pathol. 2007, 20, 856–863. [Google Scholar] [CrossRef]
  228. Silberstein, G.B.; Dressler, G.R.; Van Horn, K. Expression of the PAX2 oncogene in human breast cancer and its role in progesterone-dependent mammary growth. Oncogene 2002, 21, 1009–1016. [Google Scholar] [CrossRef] [Green Version]
  229. Hanna, J.; Markoulaki, S.; Schorderet, P.; Carey, B.W.; Beard, C.; Wernig, M.; Creyghton, M.P.; Steine, E.J.; Cassady, J.P.; Foreman, R.; et al. Direct reprogramming of terminally differentiated mature B lymphocytes to pluripotency. Cell 2008, 133, 250–264. [Google Scholar] [CrossRef]
  230. Simpson, T.I.; Price, D.J. Pax6; A pleiotropic player in development. BioEssays 2002, 24, 1041–1051. [Google Scholar] [CrossRef]
  231. Pichaud, F.; Desplan, C. Pax genes and eye organogenesis. Curr. Opin. Genet. Dev. 2002, 12, 430–434. [Google Scholar] [CrossRef]
  232. Zhang, X.; Huang, C.T.; Chen, J.; Pankratz, M.T.; Xi, J.; Li, J.; Yang, Y.; LaVaute, T.M.; Li, X.-J.; Ayala, M.; et al. Pax6 is a human neuroectoderm cell fate determinant. Cell Stem Cell 2010, 7, 90–100. [Google Scholar] [CrossRef] [PubMed]
  233. Sansom, S.N.; Griffiths, D.S.; Faedo, A.; Kleinjan, D.-J.; Ruan, Y.; Smith, J.; Van Heyningen, V.; Rubenstein, J.L.; Livesey, F.J. The Level of the Transcription Factor Pax6 Is Essential for Controlling the Balance between Neural Stem Cell Self-Renewal and Neurogenesis. PLoS Genet. 2009, 5, e1000511. [Google Scholar] [CrossRef] [PubMed]
  234. Sivak, J.M.; Mohan, R.; Rinehart, W.B.; Xu, P.-X.; Maas, R.L.; Fini, M.E. Pax-6 expression and activity are induced in the reepithelializing cornea and control activity of the transcriptional promoter for matrix metalloproteinase gelatinase B. Dev. Boil. 2000, 222, 41–54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Ou, J.; Lowes, C.; Collinson, J.M. Cytoskeletal and cell adhesion defects in wounded and Pax6+/- corneal epithelia. Investig. Ophthalmol. Vis. Sci. 2010, 51, 1415–1423. [Google Scholar] [CrossRef] [PubMed]
  236. Ramaesh, T.; Ramaesh, K.; Martin Collinson, J.; Chanas, S.A.; Dhillon, B.; West, J.D. Developmental and cellular factors underlying corneal epithelial dysgenesis in the Pax6+/− mouse model of aniridia. Exp. Eye Res. 2005, 81, 224–235. [Google Scholar] [CrossRef]
  237. Relaix, F.; Montarras, D.; Zaffran, S.; Gayraud-Morel, B.; Rocancourt, D.; Tajbakhsh, S.; Mansouri, A.; Cumano, A.; Buckingham, M. Pax3 and Pax7 have distinct and overlapping functions in adult muscle progenitor cells. J. Cell Biol. 2006, 172, 91–102. [Google Scholar] [CrossRef]
  238. De Felice, M.; Di Lauro, R. Minireview: Intrinsic and Extrinsic Factors in Thyroid Gland Development: An Update. Endocrinology 2011, 152, 2948–2956. [Google Scholar] [CrossRef] [Green Version]
  239. Thomas, D.; Friedman, S.; Lin, R.-Y. Thyroid stem cells: Lessons from normal development and thyroid cancer. Endocr.-Relat. Cancer 2008, 15, 51–58. [Google Scholar] [CrossRef]
  240. Little, M.H.; Bertram, J.F. Is there such a thing as a renal stem cell? J. Am. Soc. Nephrol. JASN 2009, 20, 2112–2117. [Google Scholar] [CrossRef]
  241. Li, C.G.; E Nyman, J.; Braithwaite, A.W.; Eccles, M.R. PAX8 promotes tumor cell growth by transcriptionally regulating E2F1 and stabilizing RB protein. Oncogene 2011, 30, 4824–4834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  242. Mazet, F.; Hutt, J.A.; Millard, J.; Shimeld, S.M. Pax gene expression in the developing central nervous system of Ciona intestinalis. Gene Expr. Patterns 2003, 3, 743–745. [Google Scholar] [CrossRef]
  243. Sagasser, S.; Hadrys, T.; DeSalle, R.; Fischer, N.; Schierwater, B. The Trichoplax PaxB Gene: A Putative Proto-PaxA/B/C Gene Predating the Origin of Nerve and Sensory Cells. Mol. Boil. Evol. 2005, 22, 1569–1578. [Google Scholar] [Green Version]
  244. Putnam, N.H.; Butts, T.; Ferrier, D.E.K.; Furlong, R.F.; Hellsten, U.; Kawashima, T.; Robinson-Rechavi, M.; Shoguchi, E.; Terry, A.; Yu, J.-K.; et al. The amphioxus genome and the evolution of the chordate karyotype. Nature 2008, 453, 1064–1071. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Dehal, P.; Boore, J.L. Two rounds of whole genome duplication in the ancestral vertebrate. PLoS Boil. 2005, 3, e314. [Google Scholar] [CrossRef] [PubMed]
  246. Goode, D.K.; Elgar, G. ThePAX258gene subfamily: A comparative perspective. Dev. Dyn. 2009, 238, 2951–2974. [Google Scholar] [CrossRef]
  247. Bassham, S.; Cañestro, C.; Postlethwait, J.H. Evolution of developmental roles of Pax2/5/8 paralogs after independent duplication in urochordate and vertebrate lineages. BMC Boil. 2008, 6, 35. [Google Scholar] [CrossRef]
  248. Paixão-Côrtes, V.R.; Salzano, F.M.; Bortolini, M.C. Evolutionary History of Chordate PAX Genes: Dynamics of Change in a Complex Gene Family. PLoS ONE 2013, 8, e73560. [Google Scholar] [CrossRef]
  249. Miller, D.J.; Hayward, D.C.; Reece-Hoyes, J.S.; Scholten, I.; Catmull, J.; Gehring, W.J.; Callaerts, P.; Larsen, J.E.; Ball, E.E. Pax gene diversity in the basal cnidarian Acropora millepora (Cnidaria, Anthozoa): Implications for the evolution of the Pax gene family. Proc. Natl. Acad. Sci. USA 2000, 97, 4475–4480. [Google Scholar] [CrossRef]
  250. Galliot, B.; De Vargas, C.; Miller, D. Evolution of homeobox genes: Q 50 Paired-like genes founded the Paired class. Dev. Genes Evol. 1999, 209, 186–197. [Google Scholar] [CrossRef]
  251. Hoshiyama, D.; Suga, H.; Iwabe, N.; Koyanagi, M.; Nikoh, N.; Kuma, K.-I.; Matsuda, F.; Honjo, T.; Miyata, T. Sponge Pax cDNA Related to Pax-2/5/8 and Ancient Gene Duplications in the Pax Family. J. Mol. Evol. 1998, 47, 640–648. [Google Scholar] [CrossRef] [PubMed]
  252. Balczarek, K.A.; Lai, Z.C.; Kumar, S. Evolution of functional diversification of the paired box (Pax) DNA-binding domains. Mol. Boil. Evol. 1997, 14, 829–842. [Google Scholar] [CrossRef] [PubMed]
  253. Bernasconi, M.; Remppis, A.; Fredericks, W.J.; Rauscher, F.J.; Schäfer, B.W. Induction of apoptosis in rhabdomyosarcoma cells through down-regulation of PAX proteins. Proc. Natl. Acad. Sci. USA 1996, 93, 13164–13169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Gnarra, J.R.; Dressler, G.R. Expression of Pax-2 in human renal cell carcinoma and growth inhibition by antisense oligonucleotides. Cancer Res. 1995, 55, 4092–4098. [Google Scholar]
  255. Luu, V.-D.; Boysen, G.; Struckmann, K.; Casagrande, S.; Von Teichman, A.; Wild, P.J.; Sulser, T.; Schraml, P.; Moch, H. Loss of VHL and Hypoxia Provokes PAX2 Up-Regulation in Clear Cell Renal Cell Carcinoma. Clin. Cancer Res. 2009, 15, 3297–3304. [Google Scholar] [CrossRef] [Green Version]
  256. Buttiglieri, S.; Deregibus, M.C.; Bravo, S.; Cassoni, P.; Chiarle, R.; Bussolati, B.; Camussi, G. Role of Pax2 in apoptosis resistance and proinvasive phenotype of Kaposi’s sarcoma cells. J. Biol. Chem. 2004, 279, 4136–4143. [Google Scholar] [CrossRef]
  257. Relaix, F.; Polimeni, M.; Rocancourt, D.; Ponzetto, C.; Schäfer, B.W.; Buckingham, M. The transcriptional activator PAX3-FKHR rescues the defects of Pax3 mutant mice but induces a myogenic gain-of-function phenotype with ligand-independent activation of Met signaling in vivo. Genome Res. 2003, 17, 2950–2965. [Google Scholar] [CrossRef]
  258. Miyamoto, T.; Kakizawa, T.; Ichikawa, K.; Nishio, S.; Kajikawa, S.; Hashizume, K. Expression of Dominant Negative Form of PAX4 in Human Insulinoma. Biochem. Biophys. Res. Commun. 2001, 282, 34–40. [Google Scholar] [CrossRef]
  259. Chang, J.Y.; Hu, Y.; Siegel, E.; Stanley, L.; Zhou, Y.-H. PAX6 increases glioma cell susceptibility to detachment and oxidative stress. J. Neuro-Oncol. 2007, 84, 9–19. [Google Scholar] [CrossRef]
  260. Kroll, T.G.; Sarraf, P.; Pecciarini, L.; Chen, C.J.; Mueller, E.; Spiegelman, B.M.; A Fletcher, J. PAX8-PPARgamma1 fusion oncogene in human thyroid carcinoma [corrected]. Science 2000, 289, 1357–1360. [Google Scholar] [CrossRef]
  261. Li, C.G.; Eccles, M.R. PAX Genes in Cancer; Friends or Foes? Front. Genet. 2012, 3, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  262. Bertrand, N.; Castro, D.S.; Guillemot, F. Proneural genes and the specification of neural cell types. Nat. Rev. Neurosci. 2002, 3, 517–530. [Google Scholar] [CrossRef] [PubMed]
  263. Murre, C.; Massari, M.E. Helix-Loop-Helix Proteins: Regulators of Transcription in Eucaryotic Organisms. Mol. Cell. Boil. 2000, 20, 429–440. [Google Scholar] [Green Version]
  264. Jones, S. An overview of the basic helix-loop-helix proteins. Genome Boil. 2004, 5, 226. [Google Scholar] [CrossRef] [PubMed]
  265. Thiery, J.P.; Morgan, M. Breast cancer progression with a Twist. Nat. Med. 2004, 10, 777–778. [Google Scholar] [CrossRef] [PubMed]
  266. Dang, C.V. MYC on the path to cancer. Cell 2012, 149, 22–35. [Google Scholar] [CrossRef]
  267. Dennis, D.J.; Han, S.; Schuurmans, C. bHLH transcription factors in neural development, disease, and reprogramming. Brain Res. 2019, 1705, 48–65. [Google Scholar] [CrossRef]
  268. Fode, C.; Gradwohl, G.; Morin, X.; Dierich, A.; LeMeur, M.; Goridis, C.; Guillemot, F. The bHLH Protein NEUROGENIN 2 Is a Determination Factor for Epibranchial Placode–Derived Sensory Neurons. Neuron 1998, 20, 483–494. [Google Scholar] [CrossRef]
  269. Guillemot, F.; Lo, L.-C.; Johnson, J.E.; Auerbach, A.; Anderson, D.J.; Joyner, A.L. Mammalian achaete-scute homolog 1 is required for the early development of olfactory and autonomic neurons. Cell 1993, 75, 463–476. [Google Scholar] [CrossRef]
  270. Scardigli, R.; Schuurmans, C.; Gradwohl, G.; Guillemot, F. Crossregulation between Neurogenin2 and Pathways Specifying Neuronal Identity in the Spinal Cord. Neuron 2001, 31, 203–217. [Google Scholar] [CrossRef] [Green Version]
  271. Casarosa, S.; Fode, C.; Guillemot, F. Mash1 regulates neurogenesis in the ventral telencephalon. Development 1999, 126, 525–534. [Google Scholar] [PubMed]
  272. Horton, S.; Meredith, A.; Richardson, J.A.; Johnson, J.E. Correct Coordination of Neuronal Differentiation Events in Ventral Forebrain Requires the bHLH Factor MASH1. Mol. Cell. Neurosci. 1999, 14, 355–369. [Google Scholar] [CrossRef] [PubMed]
  273. Ma, Q.; Sommer, L.; Cserjesi, P.; Anderson, D.J. Mash1 and neurogenin1 expression patterns define complementary domains of neuroepithelium in the developing CNS and are correlated with regions expressing notch ligands. J. Neurosci. 1997, 17, 3644–3652. [Google Scholar] [CrossRef] [PubMed]
  274. Nieto, M.; Schuurmans, C.; Britz, O.; Guillemot, F. Neural bHLH Genes Control the Neuronal versus Glial Fate Decision in Cortical Progenitors. Neuron 2001, 29, 401–413. [Google Scholar] [CrossRef] [Green Version]
  275. Tomita, K.; Moriyoshi, K.; Nakanishi, S.; Guillemot, F.; Kageyama, R. Mammalian achaete–scute and atonal homologs regulate neuronal versus glial fate determination in the central nervous system. EMBO J. 2000, 19, 5460–5472. [Google Scholar] [CrossRef]
  276. Caiazzo, M.; Dell’Anno, M.T.; Dvoretskova, E.; Lazarevic, D.; Taverna, S.; Leo, D.; Sotnikova, T.D.; Menegon, A.; Roncaglia, P.; Colciago, G.; et al. Direct generation of functional dopaminergic neurons from mouse and human fibroblasts. Nature 2011, 476, 224–227. [Google Scholar] [CrossRef]
  277. Pfisterer, U.; Kirkeby, A.; Torper, O.; Wood, J.; Nelander, J.; Dufour, A.; Björklund, A.; Lindvall, O.; Jakobsson, J.; Parmar, M. Direct conversion of human fibroblasts to dopaminergic neurons. Proc. Natl. Acad. Sci. USA 2011, 108, 10343–10348. [Google Scholar] [CrossRef] [Green Version]
  278. Pang, Z.P.; Yang, N.; Vierbuchen, T.; Ostermeier, A.; Fuentes, D.R.; Yang, T.Q.; Citri, A.; Sebastiano, V.; Marro, S.; Südhof, T.C.; et al. Induction of human neuronal cells by defined transcription factors. Nature 2011, 476, 220–223. [Google Scholar] [CrossRef]
  279. Vierbuchen, T.; Ostermeier, A.; Pang, Z.P.; Kokubu, Y.; Südhof, T.C.; Wernig, M. Direct conversion of fibroblasts to functional neurons by defined factors. Nature 2010, 463, 1035–1041. [Google Scholar] [CrossRef] [Green Version]
  280. Mazzoni, E.; Mahony, S.; Closser, M.; A Morrison, C.; Nedelec, S.; Williams, D.J.; An, D.; Gifford, D.K.; Wichterle, H. Synergistic binding of transcription factors to cell-specific enhancers programs motor neuron identity. Nat. Neurosci. 2013, 16, 1219–1227. [Google Scholar] [CrossRef] [Green Version]
  281. Schwab, M.H.; Bartholomae, A.; Heimrich, B.; Feldmeyer, D.; Druffel-Augustin, S.; Goebbels, S.; Naya, F.J.; Zhao, S.; Frotscher, M.; Tsai, M.-J.; et al. Neuronal Basic Helix-Loop-Helix Proteins (NEX and BETA2/Neuro D) Regulate Terminal Granule Cell Differentiation in the Hippocampus. J. Neurosci. 2000, 20, 3714–3724. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  282. Takebayashi, H.; Nabeshima, Y.; Yoshida, S.; Chisaka, O.; Ikenaka, K.; Nabeshima, Y.-I. The basic helix-loop-helix factor olig2 is essential for the development of motoneuron and oligodendrocyte lineages. Curr. Boil. 2002, 12, 1157–1163. [Google Scholar] [CrossRef]
  283. Lu, Q.R.; Cai, L.; Rowitch, D.; Cepko, C.L.; Stiles, C.D. Ectopic expression of Olig1 promotes oligodendrocyte formation and reduces neuronal survival in developing mouse cortex. Nat. Neurosci. 2001, 4, 973–974. [Google Scholar] [CrossRef] [PubMed]
  284. Hartenstein, V.; Takashima, S.; Hartenstein, P.; Asanad, S.; Asanad, K. bHLH proneural genes as cell fate determinants of entero-endocrine cells, an evolutionarily conserved lineage sharing a common root with sensory neurons. Dev. Boil. 2017, 431, 36–47. [Google Scholar] [CrossRef]
  285. VanDussen, K.L.; Samuelson, L.C. Mouse atonal homolog 1 directs intestinal progenitors to secretory cell rather than absorptive cell fate. Dev. Boil. 2010, 346, 215–223. [Google Scholar] [CrossRef] [Green Version]
  286. Li, H.; De Faria, J.P.; Andrew, P.; Nitarska, J.; Richardson, W.D. Phosphorylation regulates OLIG2 cofactor choice and the motor neuron-oligodendrocyte fate switch. Neuron 2011, 69, 918–929. [Google Scholar] [CrossRef]
  287. Huang, H.-P.; Tsai, M.-J.; Naya, F.J.; Qiu, Y.; Mutoh, H.; DeMayo, F.J.; Leiter, A.B. Diabetes, defective pancreatic morphogenesis, and abnormal enteroendocrine differentiation in BETA2/NeuroD-deficient mice. Genes Dev. 1997, 11, 2323–2334. [Google Scholar] [Green Version]
  288. Risebro, C.A.; Smart, N.; Dupays, L.; Breckenridge, R.; Mohun, T.J.; Riley, P.R. Hand1 regulates cardiomyocyte proliferation versus differentiation in the developing heart. Development 2006, 133, 4595–4606. [Google Scholar] [CrossRef] [Green Version]
  289. Quarto, N.; Shailendra, S.; Meyer, N.P.; Menon, S.; Renda, A.; Longaker, M.T. Twist1-Haploinsufficiency Selectively Enhances the Osteoskeletal Capacity of Mesoderm-Derived Parietal Bone Through Downregulation of Fgf23. Front. Physiol. 2018, 9, 1426. [Google Scholar] [CrossRef]
  290. Bildsoe, H.; Loebel, D.A.; Jones, V.J.; Chen, Y.-T.; Behringer, R.R.; Tam, P.P. Requirement for Twist1 in frontonasal and skull vault development in the mouse embryo. Dev. Boil. 2009, 331, 176–188. [Google Scholar] [CrossRef] [Green Version]
  291. Segev, E.; Halachmi, N.; Salzberg, A.; Ben-Arie, N. Nato3 is an evolutionarily conserved bHLH transcription factor expressed in the CNS of Drosophila and mouse. Mech. Dev. 2001, 106, 197–202. [Google Scholar] [CrossRef]
  292. Han, Z.; Yi, P.; Li, X.; Olson, E.N. Hand, an evolutionarily conserved bHLH transcription factor required for Drosophila cardiogenesis and hematopoiesis. Development 2006, 133, 1175–1182. [Google Scholar] [CrossRef] [PubMed]
  293. Ledent, V.; Vervoort, M. The Basic Helix-Loop-Helix Protein Family: Comparative Genomics and Phylogenetic Analysis. Genome Res. 2001, 11, 754–770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  294. Riechmann, J.L.; Heard, J.; Martin, G.; Reuber, L.; Jiang, C.-Z.; Keddie, J.; Adam, L.; Pineda, O.; Ratcliffe, O.J.; Samaha, R.R.; et al. Arabidopsis Transcription Factors: Genome-Wide Comparative Analysis Among Eukaryotes. Science 2000, 290, 2105–2110. [Google Scholar] [CrossRef]
  295. Goossens, J.; Mertens, J.; Goossens, A. Role and functioning of bHLH transcription factors in jasmonate signalling. J. Exp. Bot. 2017, 68, 1333–1347. [Google Scholar] [CrossRef]
  296. Zhu, Z.; Chen, G.; Guo, X.; Yin, W.; Yu, X.; Hu, J.; Hu, Z. Overexpression of SlPRE2, an atypical bHLH transcription factor, affects plant morphology and fruit pigment accumulation in tomato. Sci. Rep. 2017, 7, 5786. [Google Scholar] [CrossRef]
  297. Morgenstern, B.; Atchley, W.R. Evolution of bHLH transcription factors: Modular evolution by domain shuffling? Mol. Biol. Evol. 1999, 16, 1654–1663. [Google Scholar] [CrossRef]
  298. Li, X.; Duan, X.; Jiang, H.; Sun, Y.; Tang, Y.; Yuan, Z.; Guo, J.; Liang, W.; Chen, L.; Yin, J.; et al. Genome-Wide Analysis of Basic/Helix-Loop-Helix Transcription Factor Family in Rice and Arabidopsis. Plant Physiol. 2006, 141, 1167–1184. [Google Scholar] [CrossRef] [Green Version]
  299. Toledo-Ortiz, G.; Huq, E.; Quail, P.H. The Arabidopsis basic/helix-loop-helix transcription factor family. Plant Cell 2003, 15, 1749–1770. [Google Scholar] [CrossRef]
  300. Amoutzias, G.D.; Robertson, D.L.; Oliver, S.G.; Bornberg-Bauer, E. Convergent evolution of gene networks by single-gene duplications in higher eukaryotes. EMBO Rep. 2004, 5, 274–279. [Google Scholar] [CrossRef]
  301. Simionato, E.; Ledent, V.; Richards, G.; Thomas-Chollier, M.; Kerner, P.; Coornaert, D.; Degnan, B.M.; Vervoort, M. Origin and diversification of the basic helix-loop-helix gene family in metazoans: Insights from comparative genomics. BMC Evol. Boil. 2007, 7, 33. [Google Scholar] [CrossRef] [PubMed]
  302. Carretero-Paulet, L.; Galstyan, A.; Roig-Villanova, I.; Martínez-García, J.F.; Bilbao-Castro, J.R.; Robertson, D.L. Genome-Wide Classification and Evolutionary Analysis of the bHLH Family of Transcription Factors in Arabidopsis, Poplar, Rice, Moss, and Algae. Plant Physiol. 2010, 153, 1398–1412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Pires, N.; Dolan, L. Origin and diversification of basic-helix-loop-helix proteins in plants. Mo. Biol. Evol. 2010, 27, 862–874. [Google Scholar] [CrossRef] [PubMed]
  304. Schild, C.; Wirth, M.; Reichert, M.; Schmid, R.M.; Saur, D.; Schneider, G. PI3K signaling maintains c-myc expression to regulate transcription of E2F1 in pancreatic cancer cells. Mol. Carcinog. 2009, 48, 1149–1158. [Google Scholar] [CrossRef]
  305. Palomero, T.; Lim, W.K.; Odom, D.T.; Sulis, M.L.; Real, P.J.; Margolin, A.; Barnes, K.C.; O’Neil, J.; Neuberg, D.; Weng, A.P.; et al. NOTCH1 directly regulates c-MYC and activates a feed-forward-loop transcriptional network promoting leukemic cell growth. Proc. Natl. Acad. Sci. USA 2006, 103, 18261–18266. [Google Scholar] [CrossRef]
  306. He, T.-C.; Sparks, A.B.; Rago, C.; Hermeking, H.; Zawel, L.; Da Costa, L.T.; Morin, P.J.; Vogelstein, B.; Kinzler, K.W. Identification of c-MYC as a Target of the APC Pathway. Science 1998, 281, 1509–1512. [Google Scholar] [CrossRef]
  307. Gysin, S.; Salt, M.; Young, A.; McCormick, F. Therapeutic Strategies for Targeting Ras Proteins. Genes Cancer 2011, 2, 359–372. [Google Scholar] [CrossRef] [Green Version]
  308. Adams, J.M.; Harris, A.W.; Pinkert, C.A.; Corcoran, L.M.; Alexander, W.S.; Cory, S.; Palmiter, R.D.; Brinster, R.L. The c-myc oncogene driven by immunoglobulin enhancers induces lymphoid malignancy in transgenic mice. Nature 1985, 318, 533–538. [Google Scholar] [CrossRef]
  309. Belandia, B.; Powell, S.M.; García-Pedrero, J.M.; Walker, M.M.; Bevan, C.L.; Parker, M.G. Hey1, a Mediator of Notch Signaling, Is an Androgen Receptor Corepressor. Mol. Cell. Boil. 2005, 25, 1425–1436. [Google Scholar] [CrossRef] [Green Version]
  310. Wu, Y.; Sato, F.; Yamada, T.; Bhawal, U.K.; Kawamoto, T.; Fujimoto, K.; Noshiro, M.; Seino, H.; Morohashi, S.; Hakamada, K.; et al. The BHLH transcription factor DEC1 plays an important role in the epithelial-mesenchymal transition of pancreatic cancer. Int. J. Oncol. 2012, 41, 1337–1346. [Google Scholar] [CrossRef] [Green Version]
  311. Patel, D.; Chaudhary, J. Increased expression of bHLH Transcription Factor E2A (TCF3) in prostate cancer promotes proliferation and confers resistance to doxorubicin induced apoptosis. Biochem. Biophys. Res. Commun. 2012, 422, 146–151. [Google Scholar] [CrossRef] [PubMed]
  312. Holmquist-Mengelbier, L.; Fredlund, E.; Lofstedt, T.; Noguera, R.; Navarro, S.; Nilsson, H.; Pietras, A.; Vallon-Christersson, J.; Borg, A.; Gradin, K.; et al. Recruitment of HIF-1alpha and HIF-2alpha to common target genes is differentially regulated in neuroblastoma: HIF-2alpha promotes an aggressive phenotype. Cancer Cell 2006, 10, 413–423. [Google Scholar] [CrossRef] [PubMed]
  313. Villarino, N.; Signaevskaia, L.; Van Niekerk, J.; Medal, R.; Kim, H.; Lahmy, R.; Scully, K.; Pinkerton, A.; Kim, S.; Lowy, A.; et al. A screen for inducers of bHLH activity identifies pitavastatin as a regulator of p21, Rb phosphorylation and E2F target gene expression in pancreatic cancer. Oncotarget 2017, 8, 53154–53167. [Google Scholar] [CrossRef] [PubMed]
  314. Lambert, S.A.; Jolma, A.; Campitelli, L.F.; Das, P.K.; Yin, Y.; Albu, M.; Chen, X.; Taipale, J.; Hughes, T.R.; Weirauch, M.T. The Human Transcription Factors. Cell 2018, 172, 650–665. [Google Scholar] [CrossRef]
  315. Ravasi, T.; Suzuki, H.; Cannistraci, C.V.; Katayama, S.; Bajic, V.B.; Tan, K.; Akalin, A.; Schmeier, S.; Kanamori-Katayama, M.; Bertin, N.; et al. An Atlas of Combinatorial Transcriptional Regulation in Mouse and Man. Cell 2010, 140, 744–752. [Google Scholar] [CrossRef] [Green Version]
  316. Chou, S.-J.; Tole, S. Lhx2, an evolutionarily conserved, multifunctional regulator of forebrain development. Brain Res. 2019, 1705, 1–14. [Google Scholar] [CrossRef]
  317. Matthews, J.M.; Lester, K.; Joseph, S.; Curtis, D.J. LIM-domain-only proteins in cancer. Nat. Rev. Cancer 2013, 13, 111–122. [Google Scholar] [CrossRef]
  318. Safe, S.; Abbruzzese, J.; Abdelrahim, M.; Hedrick, E. Specificity Protein Transcription Factors and Cancer: Opportunities for Drug Development. Cancer Prev. Res. 2018, 11, 371–382. [Google Scholar] [CrossRef]
  319. Gilding, L.N.; Somervaille, T.C.P. The Diverse Consequences of FOXC1 Deregulation in Cancer. Cancers 2019, 11, 184. [Google Scholar] [CrossRef]
  320. Gartel, A.L. FOXM1 in Cancer: Interactions and Vulnerabilities. Cancer Res. 2017, 77, 3135–3139. [Google Scholar] [CrossRef]
  321. Golson, M.L.; Kaestner, K.H. Fox transcription factors: From development to disease. Development 2016, 143, 4558–4570. [Google Scholar] [CrossRef] [PubMed]
  322. Hannenhalli, S.; Kaestner, K.H. The evolution of Fox genes and their role in development and disease. Nat. Rev. Genet. 2009, 10, 233–240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  323. Moens, C.B.; Selleri, L. Hox cofactors in vertebrate development. Dev. Boil. 2006, 291, 193–206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  324. Bhatlekar, S.; Addya, S.; Salunek, M.; Orr, C.R.; Surrey, S.; McKenzie, S.; Fields, J.Z.; Boman, B.M. Identification of a developmental gene expression signature, including HOX genes, for the normal human colonic crypt stem cell niche: Overexpression of the signature parallels stem cell overpopulation during colon tumorigenesis. Stem Cells Dev. 2014, 23, 167–179. [Google Scholar] [CrossRef] [PubMed]
  325. Feng, F.Y.; Brenner, J.C.; Hussain, M.; Chinnaiyan, A.M. Molecular pathways: Targeting ETS gene fusions in cancer. Clin. Cancer Res. 2014, 20, 4442–4448. [Google Scholar] [CrossRef] [PubMed]
  326. Tuteja, G.; Chung, T.; Bejerano, G. Changes in the enhancer landscape during early placental development uncover a trophoblast invasion gene-enhancer network. Placenta 2016, 37, 45–55. [Google Scholar] [CrossRef] [PubMed]
  327. Sizemore, G.M.; Pitarresi, J.R.; Balakrishnan, S.; Ostrowski, M.C. The ETS family of oncogenic transcription factors in solid tumours. Nat. Rev. Cancer 2017, 17, 337–351. [Google Scholar] [CrossRef]
  328. Ahmad, N.; Kumar, R. Steroid hormone receptors in cancer development: A target for cancer therapeutics. Cancer Lett. 2011, 300, 1–9. [Google Scholar] [CrossRef]
  329. Mendell, A.L.; MacLusky, N.J. Neurosteroid Metabolites of Gonadal Steroid Hormones in Neuroprotection: Implications for Sex Differences in Neurodegenerative Disease. Front. Mol. Neurosci. 2018, 11, 359. [Google Scholar] [CrossRef] [Green Version]
  330. Lee, S.-U.; Maeda, T. POK/ZBTB proteins: An emerging family of proteins that regulate lymphoid development and function. Immunol. Rev. 2012, 247, 107–119. [Google Scholar] [CrossRef]
  331. Romano, K.A.; Campo, A.M.-D.; Kasahara, K.; Chittim, C.L.; Vivas, E.I.; Amador-Noguez, D.; Balskus, E.P.; Rey, F.E. Metabolic, Epigenetic, and Transgenerational Effects of Gut Bacterial Choline Consumption. Cell Host Microbe 2017, 22, 279–290. [Google Scholar] [CrossRef] [PubMed]
  332. Diamantopoulos, P.T.; Kotsianidis, I.; Symeonidis, A.; Pappa, V.; Galanopoulos, A.; Gogos, D.; Karakatsanis, S.; Papadaki, H.; Palla, A.; Hatzimichael, E.; et al. Chronic myelomonocytic leukemia treated with 5-azacytidine—Results from the Hellenic 5-Azacytidine Registry: Proposal of a new risk stratification system. Leuk. Lymphoma 2018, 60, 1721–1730. [Google Scholar] [CrossRef] [PubMed]
  333. Yan, C.; Higgins, P.J. Drugging the undruggable: Transcription therapy for cancer. Biochim. Biophys. Acta 2013, 1835, 76–85. [Google Scholar] [CrossRef] [PubMed]
  334. Liu, L.; Jin, G.; Zhou, X. Modeling the relationship of epigenetic modifications to transcription factor binding. Nucleic Acids Res. 2015, 43, 3873–3885. [Google Scholar] [CrossRef] [Green Version]
  335. Freed-Pastor, W.A.; Prives, C. Mutant p53: One name, many proteins. Genes Dev. 2012, 26, 1268–1286. [Google Scholar] [CrossRef]
  336. Baliou, S.; Adamaki, M.; Kyriakopoulos, A.M.; Spandidos, D.A.; Panayiotidis, M.; Christodoulou, I.; Zoumpourlis, V. CRISPR therapeutic tools for complex genetic disorders and cancer (Review). Int. J. Oncol. 2018, 53, 443–468. [Google Scholar] [CrossRef]
  337. Ghosh, D.; Venkataramani, P.; Nandi, S.; Bhattacharjee, S. CRISPR-Cas9 a boon or bane: The bumpy road ahead to cancer therapeutics. Cancer Cell Int. 2019, 19, 12. [Google Scholar] [CrossRef]
  338. Chen, B.; Gilbert, L.A.; Cimini, B.A.; Schnitzbauer, J.; Zhang, W.; Li, G.-W.; Park, J.; Blackburn, E.H.; Weissman, J.S.; Qi, L.S.; et al. Dynamic imaging of genomic loci in living human cells by an optimized CRISPR/Cas system. Cell 2013, 155, 1479–1491. [Google Scholar] [CrossRef]
  339. Friedland, A.E.; Tzur, Y.B.; Esvelt, K.M.; Colaiácovo, M.P.; Church, G.M.; Calarco, J.A. Heritable genome editing in C. elegans via a CRISPR-Cas9 system. Nat. Methods 2013, 10, 741–743. [Google Scholar] [CrossRef] [Green Version]
  340. Shachaf, C.M.; Kopelman, A.M.; Arvanitis, C.; Karlsson, Å.; Beer, S.; Mandl, S.; Bachmann, M.H.; Borowsky, A.D.; Ruebner, B.; Cardiff, R.D.; et al. MYC inactivation uncovers pluripotent differentiation and tumour dormancy in hepatocellular cancer. Nature 2004, 431, 1112–1117. [Google Scholar] [CrossRef]
  341. Polstein, L.R.; Gersbach, C.A. A light-inducible CRISPR/Cas9 system for control of endogenous gene activation. Nat. Methods 2015, 11, 198–200. [Google Scholar] [CrossRef] [PubMed]
  342. Doudna, J.A.; Charpentier, E. Genome editing. The new frontier of genome engineering with CRISPR-Cas9. Science 2014, 346, 1258096. [Google Scholar] [CrossRef] [PubMed]
  343. Hsu, P.D.; Lander, E.S.; Zhang, F. Development and applications of CRISPR-Cas9 for genome engineering. Cell 2014, 157, 1262–1278. [Google Scholar] [CrossRef] [PubMed]
  344. Guernet, A.; Mungamuri, S.K.; Cartier, D.; Sachidanandam, R.; Jayaprakash, A.; Adriouch, S.; Vezain, M.; Charbonnier, F.; Rohkin, G.; Coutant, S.; et al. CRISPR-Barcoding for Intratumor Genetic Heterogeneity Modeling and Functional Analysis of Oncogenic Driver Mutations. Mol. Cell 2016, 63, 526–538. [Google Scholar] [CrossRef] [Green Version]
  345. Bhattacharjee, S.; Nandi, S. Choices have consequences: The nexus between DNA repair pathways and genomic instability in cancer. Clin. Transl. Med. 2016, 5, 45. [Google Scholar] [CrossRef]
  346. Bhattacharjee, S.; Nandi, S. DNA damage response and cancer therapeutics through the lens of the Fanconi Anemia DNA repair pathway. Cell Commun. Signal. 2017, 15, 41. [Google Scholar] [CrossRef]
  347. Bhattacharjee, S.; Nandi, S. Rare Genetic Diseases with Defects in DNA Repair: Opportunities and Challenges in Orphan Drug Development for Targeted Cancer Therapy. Cancers 2018, 10, 298. [Google Scholar] [CrossRef]
  348. Bhattacharjee, S.; Nandi, S. Synthetic lethality in DNA repair network: A novel avenue in targeted cancer therapy and combination therapeutics. IUBMB Life 2017, 69, 929–937. [Google Scholar] [CrossRef] [Green Version]
  349. Philipsen, S. A new twist to the GATA switch. Blood 2013, 122, 3391–3392. [Google Scholar] [CrossRef] [Green Version]
  350. Smith, A.N.; Miller, L.-A.; Radice, G.; Ashery-Padan, R.; Lang, R.A. Stage-dependent modes of Pax6-Sox2 epistasis regulate lens development and eye morphogenesis. Development 2009, 136, 3377. [Google Scholar] [CrossRef]
  351. Ali, F.; Hindley, C.; McDowell, G.; Deibler, R.; Jones, A.; Kirschner, M.; Guillemot, F.; Philpott, A. Cell cycle-regulated multi-site phosphorylation of Neurogenin 2 coordinates cell cycling with differentiation during neurogenesis. Development 2011, 138, 4267–4277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  352. Lim, S.-C.; Kim, S.M.; Choi, J.E.; Kim, C.H.; Duong, H.-Q.; Han, S.I.; Kang, H.S. Sodium salicylate switches glucose depletion-induced necrosis to autophagy and inhibits high mobility group box protein 1 release in A549 lung adenocarcinoma cells. Oncol. Rep. 2008, 19, 1165–1171. [Google Scholar] [CrossRef] [PubMed]
  353. Luo, Y.; Ohmori, H.; Fujii, K.; Moriwaka, Y.; Sasahira, T.; Kurihara, M.; Tatsumoto, N.; Sasaki, T.; Yamashita, Y.; Kuniyasu, H. HMGB1 attenuates anti-metastatic defence of the liver in colorectal cancer. Eur. J. Cancer 2010, 46, 791–799. [Google Scholar] [CrossRef] [PubMed]
  354. Jothi, M.; Mal, M.; Keller, C.; Mal, A.K. Small molecule inhibition of PAX3-FOXO1 through AKT activation suppresses malignant phenotypes of alveolar rhabdomyosarcoma. Mol. Cancer Ther. 2013, 12, 2663–2674. [Google Scholar] [CrossRef]
  355. Wang, H.; Hammoudeh, D.I.; Follis, A.V.; Reese, B.E.; Lazo, J.S.; Metallo, S.J.; Prochownik, E.V. Improved low molecular weight Myc-Max inhibitors. Mol. Cancer Ther. 2007, 6, 2399–2408. [Google Scholar] [CrossRef]
  356. Carabet, L.A.; Rennie, P.S.; Cherkasov, A. Therapeutic Inhibition of Myc in Cancer. Structural Bases and Computer-Aided Drug Discovery Approaches. Int. J. Mol. Sci. 2018, 20, 120. [Google Scholar] [CrossRef]
  357. Wang, X.-N.; Su, X.-X.; Cheng, S.-Q.; Sun, Z.-Y.; Huang, Z.-S.; Ou, T.-M. MYC modulators in cancer: A patent review. Expert Opin. Ther. Patents 2019, 29, 353–367. [Google Scholar] [CrossRef]
  358. Peto, R.; Roe, F.J.; Lee, P.N.; Levy, L.; Clack, J. Cancer and ageing in mice and men. Br. J. Cancer 1975, 32, 411–426. [Google Scholar] [CrossRef]
  359. Nunney, L. Lineage selection and the evolution of multistage carcinogenesis. Proc. R. Soc. B Boil. Sci. 1999, 266, 493–498. [Google Scholar] [CrossRef] [Green Version]
  360. Caulin, A.F.; Maley, C.C. Peto’s Paradox: evolution’s prescription for cancer prevention. Trends Ecol. Evol. 2011, 26, 175–182. [Google Scholar] [CrossRef]
  361. Gaughran, S.J.; Evlyn, P.; Stearns, S.C. How elephants beat cancer. eLife 2016. [Google Scholar] [CrossRef] [PubMed]
  362. Abegglen, L.M.; Caulin, A.F.; Chan, A.; Lee, K.; Robinson, R.; Campbell, M.S.; Kiso, W.K.; Schmitt, D.L.; Waddell, P.J.; Bhaskara, S.; et al. Potential Mechanisms for Cancer Resistance in Elephants and Comparative Cellular Response to DNA Damage in Humans. JAMA 2015, 314, 1850–1860. [Google Scholar] [CrossRef] [PubMed]
  363. Vazquez, J.M.; Sulak, M.; Chigurupati, S.; Lynch, V.J. A Zombie LIF Gene in Elephants Is Upregulated by TP53 to Induce Apoptosis in Response to DNA Damage. Cell Rep. 2018, 24, 1765–1776. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  364. Haupt, S.; Haupt, Y. P53 at the start of the 21st century: Lessons from elephants. F1000Research 2017, 6, 2041. [Google Scholar] [CrossRef] [PubMed]
  365. Nagy, J.D.; Victor, E.M.; Cropper, J.H. Why don’t all whales have cancer? A novel hypothesis resolving Peto’s paradox. Integr. Comp. Boil. 2007, 47, 317–328. [Google Scholar] [CrossRef] [PubMed]
  366. Martineau, D.; Lemberger, K.; Dallaire, A.; Labelle, P.; Lipscomb, T.P.; Michel, P.; Mikaelian, I. Cancer in wildlife, a case study: Beluga from the St. Lawrence estuary, Quebec, Canada. Environ. Health Perspect. 2002, 110, 285–292. [Google Scholar] [CrossRef]
  367. George, J.C.; Bada, J.; Zeh, J.; Scott, L.; Brown, S.E.; O’Hara, T.; Suydam, R. Age and growth estimates of bowhead whales (Balaena mysticetus) via aspartic acid racemization. Can. J. Zool. 1999, 77, 571–580. [Google Scholar] [CrossRef]
  368. Keane, M.; Semeiks, J.; Webb, A.E.; Li, Y.I.; Quesada, V.; Craig, T.; Madsen, L.B.; Van Dam, S.; Brawand, D.; Marques, P.I.; et al. Insights into the Evolution of Longevity from the Bowhead Whale Genome. Cell Rep. 2015, 10, 112–122. [Google Scholar] [CrossRef] [Green Version]
  369. Ma, S.; Gladyshev, V.N. Molecular signatures of longevity: Insights from cross-species comparative studies. Semin. Cell Dev. Boil. 2017, 70, 190–203. [Google Scholar] [CrossRef]
  370. Smith, E.F.; Townsend, C.O. A Plant-Tumor of Bacterial Origin. Science 1907, 25, 671–673. [Google Scholar] [CrossRef] [Green Version]
  371. Sharp, W.R.; Gunckel, J.E. Physiological Comparisons of Pith Callus with Crown-Gall and Genetic Tumors of Nicotiana glauca, N. langsdorffii, and N. glauca-langsdorffii Grown in Vitro. I. Tumor Induction and Proliferation. Plant Physiol. 1969, 44, 1069–1072. [Google Scholar] [CrossRef] [PubMed]
  372. Sakakibara, H.; Kasahara, H.; Ueda, N.; Kojima, M.; Takei, K.; Hishiyama, S.; Asami, T.; Okada, K.; Kamiya, Y.; Yamaya, T.; et al. Agrobacterium tumefaciens increases cytokinin production in plastids by modifying the biosynthetic pathway in the host plant. Proc. Natl. Acad. Sci. USA 2005, 102, 9972–9977. [Google Scholar] [CrossRef] [PubMed]
  373. Gutzat, R.; Borghi, L.; Gruissem, W. Emerging roles of RETINOBLASTOMA-RELATED proteins in evolution and plant development. Trends Plant Sci. 2012, 17, 139–148. [Google Scholar] [CrossRef] [PubMed]
  374. Kong, L.-J.; Hanley-Bowdoin, L. A Geminivirus Replication Protein Interacts with a Protein Kinase and a Motor Protein That Display Different Expression Patterns during Plant Development and Infection. Plant Cell 2002, 14, 1817–1832. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Development and Cancer: Two sides of the same coin. Schematic representation of parallel processes in development and cancer. Three examples of these processes have been outlined: i) EMT, ii) progenitor proliferation and, iii) cell migration. In the context of development, EMT is involved in neural crest development; progenitor proliferation is associated with stem cell maturation and commitment while fertilization; zygote formation and migration of the blastocyst to the uterine wall involves cell migration. In the context of cancer, EMT is involved in metastasis, progenitor proliferation, increased self-renewal and immune evasion while cell migration occurs when cancer cells migrate from organs/ blood vessels to the surrounding tissues. Examples of key transcription factors that orchestrate physiological processes in both embryonic development and cancer are included alongside.
Figure 1. Development and Cancer: Two sides of the same coin. Schematic representation of parallel processes in development and cancer. Three examples of these processes have been outlined: i) EMT, ii) progenitor proliferation and, iii) cell migration. In the context of development, EMT is involved in neural crest development; progenitor proliferation is associated with stem cell maturation and commitment while fertilization; zygote formation and migration of the blastocyst to the uterine wall involves cell migration. In the context of cancer, EMT is involved in metastasis, progenitor proliferation, increased self-renewal and immune evasion while cell migration occurs when cancer cells migrate from organs/ blood vessels to the surrounding tissues. Examples of key transcription factors that orchestrate physiological processes in both embryonic development and cancer are included alongside.
Genes 10 00794 g001
Figure 2. Targeting Transcription Factors in Cancer. Four different Transcription Factor (TF) families, namely HMG, GATA, PAX and bHLH in development and cancer. HMGA1 binds DNA through ‘AT-hook’ motifs to induce or stabilize DNA and/or protein conformations. This triggers enhanced transcription by RNA polymerase II. GATA switch occurs when GATA-1 displaces GATA-2 from FOG-1 when hematopoietic stem/progenitor cells (HSPC) undergo first steps of erythrocytic/megakaryocytic differentiation [349]. PAX6 and Sox2 cooperate functionally and regulate lens development and eye morphogenesis [350]. Two different phosphorylation states of Neurogenin 2 (Ngn2), a bHLH TF, leads to either differentiation or neurogenesis [351]. Examples of candidate drugs targeting each TF are highlighted (Table 2).
Figure 2. Targeting Transcription Factors in Cancer. Four different Transcription Factor (TF) families, namely HMG, GATA, PAX and bHLH in development and cancer. HMGA1 binds DNA through ‘AT-hook’ motifs to induce or stabilize DNA and/or protein conformations. This triggers enhanced transcription by RNA polymerase II. GATA switch occurs when GATA-1 displaces GATA-2 from FOG-1 when hematopoietic stem/progenitor cells (HSPC) undergo first steps of erythrocytic/megakaryocytic differentiation [349]. PAX6 and Sox2 cooperate functionally and regulate lens development and eye morphogenesis [350]. Two different phosphorylation states of Neurogenin 2 (Ngn2), a bHLH TF, leads to either differentiation or neurogenesis [351]. Examples of candidate drugs targeting each TF are highlighted (Table 2).
Genes 10 00794 g002
Table 1. The role of transcription factors in embryonic development and cancer.
Table 1. The role of transcription factors in embryonic development and cancer.
Transcription
Factor Family
SubtypeRole in DevelopmentRole in Cancer
High Mobility Group Proteins (HMG)HMGN1Corneal epithelium development and maintenance [87,88,89]Regulates transcription of proto-oncogenes and pro-metastatic genes like c-fos, BCL3, N-cadherin, JunB and c-Jun [103]
HMGA1Regulator of adipogenesis [104], stem cell state [105] and lymphohematopoietic differentiation; crucial for normal sperm production in mouseOverexpressed in colon, breast and invasive ovarian carcinomas, pancreatic and non-small cell lung adenocarcinomas [106]
HMGA2Neural crest cells specification in Xenopus (essential for animal growth) [93]; governs the exit of embryonic stem cells from pluripotent ground state; cell proliferation and distal epithelium differentiation during embryonic lung developmentOverexpressed in pancreatic and non-small cell lung adenocarcinomas [106]
HMGB1Neural stem cell proliferation, differentiation, and maintenance [96]Overexpressed in pancreatic (PDAC), gastric, colon, hepatocellular, and non-small cell lung adenocarcinomas [107]
GATAGATA1Development of erythrocytes, megakaryocytes, mast cells, and eosinophils [108,109]Mutations in GATA1 in Down syndrome patients associated with DS-AMKL [110]
GATA2Hematopoiesis [111]Mediates Kras-driven tumorigenesis in NSCLC; GATA2 mutated in a subset of human CML [112,113,114]
GATA3T-cell lymphopoiesis, self-renewal, and differentiation of long-term HSCs [115]Tumor suppressor and strong prognostic marker in breast cancer [116]
GATA4Cardiac angiogenesis and bile homoeostasis [117,118]Downregulated in gastric, lung, ovarian, colorectal, esophageal, glioblastoma, and large B-cell lymphoma [114,119,120]
GATA5Cardiac development [121,122]Downregulated in gastric, lung, ovarian, colorectal, esophageal, glioblastoma, and large B-cell lymphoma [119,123]
GATA6Hepatic and cardiac development [124,125]Tumor suppressor in astrocytoma; overexpressed in colon and pancreatic cancer [126,127,128]
PAXPAX1Maturation of thymocytes [129]Hypermethylated in cervical cancer [130]
PAX2Prevention of tubular cells from apoptosis post-injury [131,132]Overexpressed in ovarian, renal cell, and bladder carcinomas. Regulates ERBB2 expression in breast cancer [133,134]
PAX3Early neurogenesis; regulation of sensory neuron generation from precursor cells. Maintenance of undifferentiated state of muscle stem cells [135,136]PAX3-FKHR fusion protein acts as an oncogene in alveolar rhabdomyosarcomas. Overexpressed in primary melanomas [137,138,139]
PAX4Protection of pancreaticβ-cells from apoptosis [140]Upregulated in human insulinomas [141]
PAX5B lymphopoiesis [142]Tumor suppressor in hepatocellular carcinomas; overexpressed in B-cell neoplasms; good prognostic marker in breast cancer [143,144,145]
PAX6Eye organogenesis and neural stem cell self-renewal, neuroectoderm cell fate determination [146]Oncogenic role in pancreatic adenocarcinoma and glioblastoma [147,148,149]
PAX7Proliferation and maintenance of postnatal and muscle satellite cells [150]PAX7-FKHR fusion protein acts as an oncogene in alveolar rhabdomyosarcomas [137]
PAX8Thyroglobulin regulation; maintenance of thyroid progenitor cells [151,152]Oncogenic role in renal, ovarian, lung, and thyroid cancers and certain glioblastoma subtypes [153]
PAX9Development of permanent teeth [154]Oncogenic role in lung cancer and oral squamous cell carcinomas [155,156]
bHLHTWIST 1Osteogenesis and craniofacial development [157,158]Induces EMT; activated during tumor progression [159,160,161]
TWIST 2Osteogenesis and bone proliferation [157]Induces EMT; activated during tumor progression [159]
MYCSkeletal development, osteogenesis, stem and progenitor cell maintenance and self-renewal, organogenesis [162]Oncogenic role in various cancer signaling pathways; tumor maintenance; copy number variations observed in pancreatic ductal adenocarcinoma [163,164]
ATOH1Differentiation of granule cells of the cerebellum and inner ear hair cells [165]Tumor suppressor; silenced in most colorectal cancers; induces differentiation of gastric cancer stem cells; drives metastasis of medulloblastoma; lineage-dependency oncogene in Merkel cell carcinoma.
NEUROD1Differentiation of inner ear sensory neurons, cerebellum, and the hippocampus [166]Survival and migration of neuroendocrine lung carcinomas; cell motility and tumor formation of neuroblastoma; in cooperation with Otx2, controls Group 3 medulloblastoma active enhancer landscape [167]
NEUROD2Formation of corpus callosum, essential for communication between the two cerebral hemispheres [168]Tumor suppressor and prognostic biomarker in Glioblastoma; copy number gains of NEUROD2 in male breast cancer (prognostic value) [169]
HAND1Proliferation, differentiation, and morphogenesis of embryonic ventricle cardiomyocytes [170,171]Downregulated in medulloblastoma; facilitates proliferation and metastasis in gastrointestinal stromal tumor; silenced in over 90% of human primary colorectal tumors. Methylation of HAND1 associated with poor survival in gastric cancer; involved in thyroid carcinogenesis [172]
HAND2Proliferation, differentiation, and morphogenesis of embryonic ventricle cardiomyocytes [170,171]Tumor suppressor in endometroid endometrial carcinoma. HAND2 suppression upregulates Fgfs in endometriosis [173].
OLIG1Oligodendrocyte differentiation in the neocortex [174]Aberrant DNA methylation in non-small cell lung cancer [175]
OLIG2Oligodendrocyte differentiation in the spinal cord [174]Universally expressed in gliomas [176]
DEC1Embryonic endochondral bone development [177]; upregulated in growth plate cartilage and chondrocytes; cartilage terminal differentiation; blocks myogenesis in bovine cells [178]Critical in cell cycle regulation and cell death in breast and oral cancer; DEC1 induces EMT in pancreatic cancer [179]
DEC2Proliferation and differentiation of chondrocytes;
neuronal differentiation; adipogenesis. Negative regulator of proliferation and differentiation of chondrocyte-lineage committed mesenchymal stem cells [180]
Critical in cell cycle regulation and cell death in breast and oral cancer [181]
HES1Cell fate determination and epidermal development [182];
epidermal development [183]; heterogenous ES cell differentiation [184]; proneural gene expression and neuronal differentiation [185]; brain morphogenesis [186]; development of the arterial pole of the heart; thyroid gland development [187].
Deregulated in several cancers and positively regulate levels of the tumor suppressor gene p53 [188]
HEY1 & HEY2Embryonic vascular development [189]; maintenance of neural precursor cells; spatial-temporal pattern of mammalian auditory hair cell differentiation [190]. HEY1 is involved in odontogenic/osteogenic differentiation and cardiac development [191].Deregulated in several cancers and positively regulates levels of the tumor suppressor gene p53 [188]
Table 2. Ongoing preclinical and clinical trials on transcription factor targets in different types of cancer.
Table 2. Ongoing preclinical and clinical trials on transcription factor targets in different types of cancer.
Molecular TargetCandidate DrugCondition or DiseaseStage of TestingOther Targets & Disease ConditionsDirect or Nonselective InhibitionReference/ClinicalTrial.gov Identifier
HMGB1Sodium salicylateLung adenocarcinomaPreclinicalTargets—Mitogen-activated protein kinases (MAPK), Caspase 3, NF-κb, p38 kinase, AP-1
Disease—Acute Myeloid Leukemia
Nonselective[352]
HMGB1Anti-HMGB1 antibodyColorectal cancerPreclinicalDisease—Stroke, Epilepsy, Neudegenerative diseases, neuropathic painDirect[353]
GATA-3MLN9708LymphomaPhase IITargets—p38 kinase, Janus Kinase (JNK), NF-κb
Disease—Breast Cancer
NonselectiveNCT02158975
GATA-2Busulfan, Fludarabine,
Busulfan and Cyclophosphamide
Myelodysplastic SyndromesPhase IIDisease—Chronic Myelogenous Leukemia, LymphomasChemo-therapyNCT01861106
Pax3-Foxo1ThapsigarginAlveolar RhabdomyosarcomaPreclinicalTargets—Sarco/endoplasmic reticulum Ca2+ ATPase (SERCA), Nicotinic acetylcholine receptorsNonselective[354]
bHLHPitavastatinPancreatic cancerPreclinicalTargets—3-hydroxy-3-methyl glutaryl coenzyme A reductase
Disease—Hypercholestrolemia and dyslipidemia
Nonselective[313]
Reverse the association between Myc and its obligate bHLH heterodimerization partner, Max10058-F4Promyelocytic leukemiaPreclinicalTargets—MYCN, Myc/Max dimerization
DiseaseMYCN-amplified neuroblastoma, Acute Myeloid Leukemia
Direct[355]
MycMycro1, Mycro2 and Mycro3LeukemiaPreclinicalTargets—Myc/Max dimerizationDirect[314,356,357]
MYCLenalidomide
and Combination chemotherapy
B-cell lymphomaPhase I/IITargets—CRL4 E3 Ubiquitin ligase
Disease—Multiple Myeloma
NonselectiveNCT02213913
MYCIbrutinibGastrooesophageal CancerPhase IITargets—Bruton’s tyrosine kinase (BTK), CD20
Disease—B-cell cancers such as mantle cell lymphoma, chronic lymphocytic leukemia and Waldenstrom’s macroglobulinemia
NonselectiveNCT02884453
MYCDA-EPOCH-R followed by NivolumabB-cell lymphomaPhase IITargets (Nivolumab)—PD-L1
Disease—Squamous non-small cell lung cancer, renal-cell carcinoma, small cell lung cancer
NonselectiveNCT03620578

Share and Cite

MDPI and ACS Style

Huilgol, D.; Venkataramani, P.; Nandi, S.; Bhattacharjee, S. Transcription Factors That Govern Development and Disease: An Achilles Heel in Cancer. Genes 2019, 10, 794. https://doi.org/10.3390/genes10100794

AMA Style

Huilgol D, Venkataramani P, Nandi S, Bhattacharjee S. Transcription Factors That Govern Development and Disease: An Achilles Heel in Cancer. Genes. 2019; 10(10):794. https://doi.org/10.3390/genes10100794

Chicago/Turabian Style

Huilgol, Dhananjay, Prabhadevi Venkataramani, Saikat Nandi, and Sonali Bhattacharjee. 2019. "Transcription Factors That Govern Development and Disease: An Achilles Heel in Cancer" Genes 10, no. 10: 794. https://doi.org/10.3390/genes10100794

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