Skip to main content
Erschienen in:

Open Access 10.09.2024 | Review

BHLHE41, a transcriptional repressor involved in physiological processes and tumor development

verfasst von: Caroline Bret, Fabienne Desmots-Loyer, Jérôme Moreaux, Thierry Fest

Erschienen in: Cellular Oncology | Ausgabe 1/2025

Abstract

BHLHE41 is a nuclear transcriptional repressor that belongs to the basic helix-loop-helix protein superfamily. BHLHE41 expression tends to be restricted to specific tissues and is regulated by environmental cues and biological events. BHLHE41 homodimerizes or heterodimerizes with various partners, influencing its transcription factor function. BHLHE41 is involved in the regulation of many physiological processes implicated in tissue/organ homeostasis, such as myogenesis, adipogenesis, circadian rhythms and DNA repair. At cellular level, BHLHE41 is involved in the regulation of mesenchymal stem cell properties, tissue-specific macrophage functions and lymphoid lineage physiology. In several cancer types, BHLHE41 modulates the expression of different transcriptional programs influencing cell cycle control, apoptosis, invasiveness, epithelial to mesenchymal transition and hypoxia response in the tumor environment. Depending on the cancer cell type, BHLHE41 can act as a tumor suppressor or an oncogene, and could be a target for innovative therapies. This review summarizes the available knowledge on BHLHE41 structure, biological functions, regulation and potential partners, as well as its role in physiological processes, and its implication in major cancer steps.
Hinweise

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Abkürzungen
AML
Acute Myeloid Leukemia
ARID
AT-rich interaction domain
ARNT1
Aryl hydrocarbon Receptor Nuclear Translocator 1
bHLH
basic Helix-Loop-Helix
CASP8
Caspase 8
CCDN1
Cyclin D1
CEBPA
CCAAT/Enhancer-Binding Protein Alpha
CEBPB
CCAAT/Enhancer-Binding Protein Beta
CRY1
CRYptochrome circadian regulator 1
CRY2
CRYptochrome circadian regulator 2
DEC2
Differentiated Embryo Chondrocyte 2
DOT1L
Disruptor of telomeric silencing 1-like
EAP
Elongation Assisting Protein
EC
Endometrial Carcinoma
EHMT2
Euchromatic Histone lysine Methyltransferase 2 (G9a)
EMT
Epithelial to Mesenchymal Transition
HBS
HIF-1 Binding Site
HDAC
Histone DeACetylase
HIF
Hypoxia Inducible Factor-1
HRE
Hypoxia Response Element
IgH
Immunoglobulin Heavy locus
HDAC1
Histone DeACetylase 1
KO
Knock-Down
MBC
Memory B Cell
MM
Multiple Myeloma
MSC
Mesenchymal Stem Cell
MYOD1
Myogenic Differentiation 1 (also known as MyoD)
NBC
Naive B Cell
NFIL3
Nuclear Factor, Interleukin 3 regulated
NGF
Nerve Growth Factor
PARP
Poly (ADP-Ribose) Polymerase
PAS
Drosophila Period, human Aryl hydrocarbon receptor nuclear translocator and Drosophila Single-minded
PC
Plasma cell
PER1
PERiod circadian regulator 1
PER2
PERiod circadian regulator 2
PER3
PERiod circadian regulator 3
PPARgamma
Peroxisome Proliferator Activated Receptor gamma
RCC
Renal Cell Carcinoma
RXR
Retinoid X Receptor alpha
SACC
Salivary Adenoid Cystic Carcinoma
SENP
SENtrin/SUMO-specific Protease
SHARP1
Split and HAiry Related Protein 1
siRNA
small interfering RNA
SIRT1
sirtuin 1
SNAI2
SNAIl family transcriptional repressor 2
SREBP-1
Sterol Regulatory Element Binding transcription factor 1
TF
Transcription Factor
TIMP1
Tissue Inhibitor of MetalloProteinase 1
TNM
Tumor, lymph Node, Metastasis
TNF
Tumor-necrosis factor
TGF
Transforming growth factor
VEGF
Vascular Endothelial Growth Factor

1 Background

BHLHE41, also called differentiated embryo chondrocyte 2 (DEC2), split and hairy related protein 1 (SHARP1) or BHLHB3, is a nuclear transcription factor (TF) that belongs to the basic helix-loop-helix (bHLH) protein superfamily. Six phylogenic groups (group A to group F) of bHLH proteins have been described [1]. Their molecular structure includes helix loop helix (HLH) domains, that act as protein oligomerization interfaces facilitating protein-protein interactions and the formation of homo or heterodimeric complexes. They are juxtaposed to basic amino acid regions that bind to DNA at consensus hexanucleotide sequences, such as E-box sequences and to a lesser degree, N-box sequences. These sequences are found in the transcriptional regulatory region of many genes. Additional domains act as dimerization regions, including Drosophila Period, human Aryl hydrocarbon receptor nuclear translocator and Drosophila Single-minded (PAS) domains in group C bHLH proteins [2], orange domains in group E bHLH proteins [3] and leucine zipper domains in group B bHLH members [4]. The different combinations of dimeric complexes and the DNA binding sequence heterogeneity determine how bHLH proteins control critical biological processes, through transcriptional regulation from yeast to humans [1, 5].
In this review, we summarize the biological characteristics of BHLHE41, including its structure and expression, its interacting partners and its transcriptional activities. We also review its involvement in physiological processes and its role in solid cancer and hematological malignancies.

2 Biological characteristics of BHLHE41

2.1 Structure

The human BHLHE41 gene is located on chromosome 12, at 12p12.1 [6]. It includes five exons and four introns, with two E-boxes at −0.3 and −1.8 kb [7]. Rat Bhlhe41 was the first cloned in 1997 [8]. The human gene, initially named DEC2, and its mouse homologue were characterized in 2001 [6].
The human BHLHE41 protein is 482 amino acids long, and has a molecular weight of 50 kDa. It contains a bHLH domain at the N terminus that binds to the E box motif CANNTG in the target promoters [9], an orange motif in the central region, and an alanine/glycine-rich sequence in the C-terminal region (Fig. 1).
The bHLH domain also contains an LxxLL motif for the interaction with nuclear retinoid receptors [11]. The bHLH region is crucial for BHLHE41 functionality, because it allows the formation of homo/heterodimers before binding to the E-box sequence [5].
BHLHE41 protein structure shares 42% of homology with BHLHE40 (also known as DEC1 or BHLHB2), another member of the bHLH superfamily. Their bHLH domains are highly conserved (97% of homology) and allow the formation of BHLHE40-BHLHE41 heterodimers that regulate the expression of common downstream target genes [10, 12].

2.2 Expression

BHLHE40 is ubiquitously expressed in human tissues. Conversely, the expression of its paralogue BHLHE41 is tissue-restricted (Fig. 2).
Specifically, BHLHE41 is highly expressed in skeletal muscle, brain and retina, found at intermediate levels in the heart, pancreas and thyroid, and detectable, but at low levels, in the lung and placenta [6, 8, 13, 14]. The expression of this gene is downregulated through a regulatory feedback loop involving BHLHE41 and BHLHE40 [7]. Indeed, it has been observed that an overexpression of BHLHE41 or BHLHE40 results in BHLHE41 downregulation [15]. This effect is mediated through the binding of BHLHE40 to the E box sequence in BHLHE41 promoter [16]. Additionally, BHLHE41 can repress BHLHE40 transcription by interacting with Specificity Protein 1 (SP1) [9]. BHLHE41 expression is also regulated by different factors including members of the clock family proteins, growth factors, cytokines, transcription factors, nutrients or hormones (Table 1).
Table 1
Factors modulating BHLHE41 expression. Several conditions and biological processes modulate BHLHE41 expression, including hypoxia through clock proteins, metabolism through nutrients or hormones, growth factors, cytokines and transcription factors
Molecule type
Molecule(s)
Effect on BHLHE41 expression
Cellular context/model
Reference(s)
Clock proteins
CLOCK/BMAL1-2
Upregulation
Endothelial cells, NIH/3T3 cell line (fibroblasts isolated from a mouse NIH/Swiss embryo), Clock mutant mouse (Ck/Ck)
[7, 17]
BHLHE40
Downregulation
NIH/3T3 cell line
[15]
BHLHE41
Autoregulation: downregulation
C2C12 cell line (mouse myoblasts)
[7]
PER1, PER2
Downregulation
C2C12 cell line
[7]
CRY1, CRY2
Downregulation
C2C12 cell line
[7]
Growth factors, cytokines
NGF
Upregulation
PC12 cell line (derived from a transplantable rat pheochromocytoma)
[8]
TGF beta
Upregulation
Head and neck squamous cell carcinoma model
[18]
TNF alpha
Upregulation
Synovial fibroblasts
[19]
Transcription factors
HIF-1 alpha
Upregulation
ATDC5 cell line (mouse chondrogenic cell line), 293T cell line (human embryonic kidney cells), and HeLa cell line (human epithelial cells obtained from a cervical adenocarcinoma)
[20]
ROR alpha
Upregulation
3T3-L1 cell line (murine adipocyte cells)
[21]
p63
Upregulation
MDA-MB-231 cell line (human breast cancer cells)
[22]
GLI1, GLI2
Upregulation
Pancreatic cancer cells
[23]
GATA3
Upregulation
Naive CD4+ T cells from Il4+/+ or Il4−/− mice
[24]
SREBP-1
Upregulation
Muscle cells
[25]
Nutrients/Hormones
Glucose and amino acids
Upregulation
Mouse liver
[26]
Insulin
Upregulation
Rat H4IIE hepatoma cells
[27]
Of note, most of these factors can also up regulate the expression of BHLHE40, as it has been described for Nerve Growth Factor (NGF) [8], Transforming growth factor (TGF) beta [28] or Tumor-Necrosis Factor (TNF) alpha [29]. These parallel modifications of expression potentially influence the fate of BHLHE41/BHLHE40 heterodimer.
Hypoxia is a regulator of BHLHE41 expression. Hypoxia-inducible factor-1 (HIF-1) is a TF regulating oxygen homeostasis and involved in the adaptation to oxygen deprivation [30]. HIF-1 is a protein complex composed of two subunits: HIF-1 alpha and HIF-1 beta (also known as aryl hydrocarbon receptor nuclear translocator 1). HIF-1 beta is constitutively expressed. In normoxic conditions, HIF-1 alpha is rapidly degraded after hydroxylation by prolyl-hydroxylase and ubiquitination by the von Hippel–Lindau protein for proteasomal destruction. In contrast, under hypoxic conditions, HIF-1 alpha is stabilized, activated and translocated to the nuclear compartment, where it heterodimerizes with HIF-1 beta [31]. The resulting complex interacts with the HIF-1 binding site in the Hypoxia Response Element (HRE) of the promoters of hypoxia-inducible genes to modulate their expression. Miyazaki and colleagues reported that BHLHE41 harbors a HRE in its promoter region and that its expression is induced by hypoxia [20]. Moreover, BHLHE41 can interact with HIF-1 alpha and contribute to its proteasomal degradation [32]. In addition, upon induction in hypoxic conditions, BHLHE41 downregulates vascular endothelial growth factor (VEGF), a major cytokine involved in angiogenesis [33].
Besides hypoxia, BHLHE41 expression is also modulated throughout circadian rhythms. In mammals, they are controlled by molecular clockwork systems that involve a series of clock genes, including CLOCK, BMAL1, BMAL2, PER1, PER2, PER3, CRY1, CRY2, CRY3, BHLHE40 and BHLHE41. The protein CLOCK is a transcription factor heterodimerizing with BMAL1, binding to E-box elements upstream of the period (PER1, PER2, PER3) and cryptochrome (CRY1, CRY2) genes and upregulating their transcription. In turn, PER and CRY heterodimerize and repress their own transcription through a feedback loop that involves CLOCK-BMAL1 complexes which also upregulate BHLHE41 [7, 20, 34]. Of note, mouse mutants that were designed to explore the role of BHLHE41 in circadian rhythms regulation and that do not express BHLHE41 were viable [35], suggesting the potential redundancy of bHLH family members.
Several soluble factors also modulate BHLHE41 expression. In an in vitro model to study neurite outgrowth, BHLHE41 expression can be induced by NGF, an essential factor for neuronal development. It has been proposed that BHLHE41 could mediate cellular changes at late stages of the central nervous system development [8]. TGF beta signaling in the bone marrow environment leads to BHLHE41 induction in mice xenografted with HEp3 cells, a model of head and neck squamous cell carcinoma [18]. Moreover, in an in vitro model of synovial fibroblasts, TNF alpha increases BHLHE41 expression in a NF-kappa B dependent manner [19].
BHLHE41 expression is also positively modulated by different TFs and intracellular proteins depending on the cellular context, including RAR-related orphan receptor (ROR) alpha [21], p63 (a member of the p53 family) [22], GLI1 and GLI2 (two members of the family of zinc finger proteins acting as mediators of the sonic hedgehog signal transduction cascade) [23] and GATA3 [24].
In addition, it has been observed that a combination of glucose and amino acids can rapidly modulate the expression of BHLHE41 and clock genes in mouse liver, suggesting an influence of nutrients in the regulation of the liver clock [26]. In rat H4IIE hepatoma cells, insulin rapidly increases BHLHE41 production in a dose-dependent manner, through different signaling pathways, including PI3K/Rac/JNK pathways [27].

2.3 Interacting partners

BHLHE41 homo- or hetero-dimerizes with its paralogue BHLHE40 or other proteins, in particular several TFs, including CCAAT/enhancer-binding protein alpha (CEBPA) and beta (CEBPB) [36], HIF-1 alpha [32, 33], myogenic differentiation 1 (MYOD1) [5], RORA [37], retinoid X receptor (RXR) alpha [10], sterol regulatory element binding protein 1c (SREBP-1c) [38], SP1 [9], and the RNA binding protein NONO [37] (Table 2). Furthermore, BHLHE41 interacts with a variety of proteins including cell cycle controlling factors casein kinase 2 alpha-1 (CSNK2A1), CSNK2A2, CSNK2B, proteins involved in the control of cell growth and division (i.e. PPP2CB, PPP2R1B, PPP2R5E [37], LLGL2 and KLC3) [39], cell signaling molecules (e.g. RASD1, a member of the Ras superfamily of small GTPases and MKNK2) [37, 39], proteins involved in epigenetic mechanisms (e.g. HDAC1, SIRT1 or WDR5) [5, 29], and DNA damage response proteins (ATM, CDCA5, CSNK1E, RTF2 and POLK) [32]. BHLHE41 also interacts with factors implicated in protein trafficking (e.g. VPS26A, a component of the retromer complex involved in the retrograde transport of proteins from endosomes to the trans-Golgi network) [39], protein degradation factors (e.g. FBXW11, a member of the F-box protein family that is a subunit of the ubiquitin protein ligase complex [37] and PSMA4, the core alpha subunit of the 20 S proteosome [28]), GSK3B (a protein implicated in the control of glucose homeostasis [37]), CLOCK proteins (CRY1, CRY2, PER2 [37]), and with MYH13 (a myosin chain involved in muscle contraction [39]). BHLHE41 also forms heterodimers with nuclear factor interleukin 3 regulated (NFIL3, also known as E4BP4), a basic leucine zipper TF. This heterodimer can bind to the PER2 EE-element (E-box and E-box-like sequences) to modulate the circadian oscillation [40].
Table 2
BHLHE41 known or predicted interacting partners. A large diversity of interacting partners has been described for BHLHE41, including other TFs, factors involved in transcriptional regulation and RNA splicing, cell cycle, cell growth and division, cell signaling, epigenetic mechanisms, DNA damage response, protein trafficking, ubiquitination, proteasomal degradation, metabolism, circadian rhythms, muscular activity
Biological process or molecule type
Partner
Function(s) of the partner
Cellular context of the demonstrated interaction(s)
Method(s) used to demonstrate the interaction
Reference(s)
Transcription factor (TF)
BHLHE40
(basic helix-loop-helix family member e40)
Basic helix-loop-helix protein expressed in various tissues, interacting with ARNTL or competing for E-box binding sites in PER1 promoter.
NIH/3T3 cell line (fibroblasts isolated from a mouse NIH/Swiss embryo)
Pull-down assay
[12]
 
BHLHE41
(basic helix-loop-helix family member e41)
Interaction as homodimers.
Neuro 2a cells (murine neuroblastoma cell line)
Immunoprecipitation assay
[41]
 
BMAL1
(basic helix-loop-helix ARNT like 1)
Basic helix-loop-helix protein forming a heterodimer with CLOCK. This heterodimer binds to E-box enhancer elements upstream of Period (PER1, PER2, PER3) and Cryptochrome (CRY1, CRY2) genes and induces their transcription.
Mouse model
Yeast two-hybrid assay
[42]
 
CEBPA
(CCAAT enhancer binding protein alpha)
TF with a basic leucine zipper domain that recognizes the CCAAT motif in the promoter of target genes. This TF forms homodimers and also heterodimers with CCAAT/enhancer-binding proteins beta and gamma. This protein modulates the expression of genes involved in cell cycle regulation and in body weight homeostasis.
3T3-L1 cell line (murine adipocyte cell line)
Immunoprecipitation assay
[36]
 
CEBPB (CCAAT enhancer binding protein beta)
TF containing a basic leucine zipper domain. It forms homodimers and also heterodimers with CCAAT/enhancer-binding proteins alpha, delta, and gamma.
3T3L1 cell line
Immunoprecipitation assay
[36]
 
CLOCK
(clock circadian regulator)
TF displaying a leading role in circadian rhythms regulation. The protein contains a bHLH domain and a histone acetyltransferase activity. It forms a heterodimer with ARNTL (BMAL1) that binds to E-box enhancer elements upstream of Period (PER1, PER2, PER3) and Cryptochrome (CRY1, CRY2) genes and activates transcription of these genes. PER and CRY heterodimerize and repress their own transcription by interacting in a feedback loop with CLOCK/ARNTL complexes.
HEK293 cell line (human embryonic kidney cell line
Yeast two-hybrid interaction screen
[37]
 
HIF1A
(Hypoxia inducible factor 1 subunit alpha)
Alpha subunit of Hypoxia-Inducible Factor-1 (HIF-1). This subunit forms heterodimers with the beta subunit (1:1). HIF-1 is a master regulator of the cellular and systemic homeostatic response to hypoxia.
MDA-MB-231 cell line (human breast cancer cells)
Immunoprecipitation assay
[32, 33]
 
MYOD1 (myogenic differentiation 1)
Nuclear protein belonging to the bHLH family of TF and the myogenic factor subfamily. It regulates muscle cell differentiation and muscle regeneration.
C2C12 cell line (mouse myoblasts)
Glutathione S transferase pull-down assay
[5]
 
RORA
(RAR related orphan receptor A)
Member of the NR1 subfamily of nuclear hormone receptors. It can bind as a monomer or as a homodimer to hypoxia response elements upstream of several genes to enhance the expression of those genes.
HEK293 cell line
Yeast two-hybrid interaction screen
[37]
 
RXRA
(retinoid X receptor alpha)
Retinoid X receptors (RXRs) are nuclear receptors that act as TF and mediate the biological effects of retinoids by their involvement in retinoic acid-mediated gene activation.
HEK293 cell line, HepG2 (human hepatocellular carcinoma) cell line
Glutathione S transferase pull-down assay
[11]
 
SREBP-1c
(sterol regulatory element binding transcription factor 1)
SREBP-1c is a bHLH-leucine zipper TF that binds to the sterol regulatory element-1 (SRE1), a motif found in the promoter of the low-density lipoprotein receptor gene and other genes involved in sterol biosynthesis.
NIH/3T3 cell line
Glutathione S transferase pull-down assay
[38]
 
SP1
(Sp1 transcription factor)
SP1 is a zinc finger TF that binds to GC-rich motifs of promoters and that is involved in many cellular processes, including cell differentiation, growth, apoptosis, immune responses, response to DNA damage, and chromatin remodeling.
10T1/2 cell line (mouse embryo cells)
Immunoprecipitation assay
[9]
 
TCF3
(Transcription factor 3)
TCF3 is a member of the E protein (class I) family of helix-loop-helix transcription factors. E proteins activate transcription by binding to regulatory E-box sequences on target genes as heterodimers or homodimers. E proteins are involved in lymphopoiesis, and TCF3 is required for B and T lymphocyte development.
TCF3 alterations may play a role in lymphoid malignancies; chromosomal translocations involving this gene have been associated with lymphoid malignancies, including pre-B-cell acute lymphoblastic leukemia (t(1;19), with PBX1), childhood leukemia (t(19;19), with TFPT) and acute leukemia (t(12;19), with ZNF384).
10T1/2 cell line
Immunoprecipitation assay
[9]
 
NPAS4
(neuronal PAS domain protein 2neuronal PAS domain protein 2)
NPAS4 is a member of the basic helix-loop-helix-PAS family of TF. A similar mouse protein has a putative regulatory role in the acquisition of memory and may function as a part of a molecular clock system.
Putative interaction
Data from the Reactome platform (http://​www.​reactome.​org) and from the GeneMania website (https://​genemania.​org)
[43]
Transcriptional regulation and RNA splicing
NONO
(non-POU domain containing octamer binding)
NONO is a RNA-binding protein that plays various roles in the nuclear compartment, including transcriptional regulation and RNA splicing.
HEK293 cell line
Yeast two-hybrid interaction screen
[37]
Cell cycle
CSNK2A1
(casein kinase 2 alpha 1)
This protein corresponds to the alpha subunit of casein kinase II, a serine/threonine protein kinase that phosphorylates acidic proteins such as casein. It is involved in various cellular processes, including cell cycle control, apoptosis and circadian rhythms.
HEK293 cell line
Yeast two-hybrid interaction screen
[37]
CSNK2A2
(casein kinase 2 alpha 2)
This protein corresponds to the alpha 2 subunit of casein kinase II.
HEK293 cell line
Yeast two-hybrid interaction screen
[37]
CSNK2B
(casein kinase 2 beta)
This protein corresponds to the beta subunit of casein kinase II.
HEK293 cell line
Yeast two-hybrid interaction screen
[37]
Cell growth and division
PPP2CB
(protein phosphatase 2 catalytic subunit beta)
PPP2CB is the beta isoform of the catalytic subunit of the phosphatase 2 A catalytic subunit. Protein phosphatase 2 A is one of the four major Ser/Thr phosphatases, and it is implicated in the negative control of cell growth and division.
HEK293 cell line
Yeast two-hybrid interaction screen
[37]
PPP2R1B
(protein phosphatase 2 scaffold subunit Abeta)
PPP2R1B is the beta isoform of the constant regulatory subunit A of protein phosphatase 2. The constant regulatory subunit A serves as a scaffolding molecule to coordinate the assembly of the catalytic subunit and a variable regulatory B subunit.
HEK293 cell line
Yeast two-hybrid interaction screen
[37]
PPP2R5E
(protein phosphatase 2 regulatory subunit B epsilon)
PPP2R5E is the epsilon isoform of the regulatory subunit B56 of phosphatase 2 A. The B regulatory subunit might modulate substrate selectivity and catalytic activity.
HEK293 cell line
Yeast two-hybrid interaction screen
[37]
LLGL2
(LLGL scribble cell polarity complex component 2)
LLGL2 is a protein similar to Drosophila lethal (2) giant larvae, which plays a role in asymmetric cell division, epithelial cell polarity, and cell migration.
HeLa cell line (human epithelial cells obtained from a cervical adenocarcinoma)
Quantitative proteomics
[39]
KLC3
(kinesin light chain 3)
KLC3 is a member of the kinesin light chain gene family involved in transporting cargos along microtubules.
HeLa cell line
Quantitative proteomics
[39]
Cell signaling
RASD1
(ras related dexamethasone induced 1)
RASD1 is a member of the Ras superfamily of small GTPases and is induced by dexamethasone. The protein is an activator of G-protein signaling and functions as a direct nucleotide exchange factor for Gi-Go proteins.
HEK293 cell line
Yeast two-hybrid interaction screen
[37]
MKNK2
(MAPK interacting serine/threonine kinase 2)
MKNK2 is a member of the calcium/calmodulin-dependent protein kinases (CAMK) Ser/Thr protein kinase family. This protein is one of the downstream kinases activated by mitogen-activated protein (MAP) kinases. It phosphorylates the eukaryotic initiation factor 4E (eIF4E) and plays major roles in the initiation of mRNA translation, oncogenic transformation, and malignant cell proliferation.
HeLa cell line
Quantitative proteomics
[39]
Epigenetic mechanisms
HDAC1
(histone deacetylase 1)
This protein belongs to the histone deacetylase family and is a component of the histone deacetylase complex. Histone acetylation and deacetylation play a leading role in the regulation of gene expression.
NIH/3T3 cell line, C2C12 cell line, COS-7 cell line (monkey kidney fibroblast-like cells)
Glutathione S transferase pull-down assay
[5]
G9a (EHMT2, euchromatic histone lysine methyltransferase 2)
G9a is a methyltransferase that methylates lysine residues of histone H3. Methylation of H3 at lysine 9 results in the recruitment of additional epigenetic regulators and transcription repression.
C2C12 cell line, HEK293 cell line, C3H10T1/2 cell line isolated from a the C3H mouse embryo cell line
Immunoprecipitation assay
[44]
SIRT1
(sirtuin 1)
SIRT1 is a member of the sirtuin family of proteins that are homologues to the yeast Sir2 protein. Yeast sirtuin proteins regulate epigenetic gene silencing and suppress DNA recombination.
C2C12 cell line
Immunoprecipitation assay
[5]
WDR5
(WD repeat domain 5)
WDR5 is a member of the WD repeat protein family. WD motifs are minimally conserved regions of ~40 amino acids, typically containing the Gly-His and Trp-Asp (GH-WD) repeat motif. They play a crucial role in mediating protein-protein interactions, contributing to the formation of multiprotein complexes. Members of this family are involved in various cellular processes, including cell cycle progression, signal transduction, apoptosis, and gene regulation. WDR5 is a methyltransferase adaptor protein that acts as a core subunit of the human MLL1-4 H3K4 methyltransferase complexes, contributing to enhancing the activity of H3K4 methyltransferases.
HeLa cell line
Yeast two-hybrid interaction screen
[37]
DNA damage response
ATM
(ATM serine/threonine kinase)
This protein belongs to the PI3/PI4-kinase family and is a major cell cycle checkpoint kinase that regulates a wide variety of downstream proteins, including the tumor suppressor proteins p53 and BRCA1, the checkpoint kinase CHK2, the checkpoint proteins RAD17 and RAD9, and the DNA repair protein NBS1. This protein is a master controller of cell cycle checkpoint signaling pathways and is required for the cell response to DNA damage and genome stability.
HeLa cell line
Quantitative proteomics
[39]
CDCA5
(cell division cycle-associated 5)
CDCA5 is a major regulator of sister chromatid cohesion and segregation during meiosis and mitosis. CDCA5 also plays an essential role in DNA repair and regulates the activity of cell cycle-associated proteins and transcription factors, thereby promoting proliferation and participating in cancer cell apoptosis.
HeLa cell line
Quantitative proteomics
[39]
CSNK1E
(casein kinase 1 epsilon)
This protein is a serine/threonine protein kinase and a member of the casein kinase I protein family that is implicated in the control of cytoplasmic and nuclear processes, including DNA replication and repair.
HEK293 cell line
Yeast two-hybrid interaction screen
[37]
RTF2
(replication termination factor 2)
RTF2 is a protein involved in DNA binding activity, the cellular response to DNA damage and DNA stability regulation.
HeLa cell line
Quantitative proteomics
[39]
POLK
(DNA polymerase kappa)
POLK is a DNA polymerase that catalyzes translesion DNA synthesis, allowing DNA replication in the presence of DNA lesions.
HeLa cell line
Quantitative proteomics
[39]
Protein trafficking
VPS26A
(VPS26 retromer complex component A)
The gene VPS26A belongs to a group of vacuolar protein sorting (VPS) genes. The encoded protein is a component of a large multimeric complex, the retromer complex, involved in retrograde transport of proteins from endosomes to the trans-Golgi network.
HeLa cell line
Quantitative proteomics
[39]
Ubiquitination
FBXW11
(F-box and WD repeat domain containing 11)
FBXW11 is a member of the F-box protein family characterized by the F-box motif. F-box proteins are one of the four subunits of the ubiquitin protein ligase complex called SCF (SKP1-cullin-F-box) that functions in phosphorylation-dependent ubiquitination.
HEK293 cell line
Yeast two-hybrid interaction screen
[37]
Proteasome
PSMA4 (proteasome 20 S subunit alpha 4)
PSMA4 is a core alpha subunit of the 20 S proteosome, a highly ordered ring-shaped structure with four rings of 28 non-identical subunits.
MDA-MB-231 cell
Immunoprecipitation assay
[32]
Control of metabolism
GSK3B
(glycogen synthase kinase 3 beta)
This protein is a serine-threonine kinase belonging to the glycogen synthase kinase subfamily. It acts as a negative regulator of glucose homeostasis and is involved in energy metabolism, inflammation, endoplasmic reticulum stress, mitochondrial dysfunction, and apoptotic pathways.
HEK293 cell line
Yeast two-hybrid interaction screen
[37]
Clock proteins and related proteins
CRY1
(cryptochrome circadian regulator 1)
CRY1 is a flavin adenine dinucleotide-binding protein that is a key component of the circadian core oscillator complex regulating the circadian clock.
HEK293 cell line
Yeast two-hybrid interaction screen
[37]
CRY2
(cryptochrome circadian regulator 2)
CRY2 is a flavin adenine dinucleotide-binding protein that is a key component of the circadian core oscillator complex regulating the circadian clock.
HEK293 cell line
Yeast two-hybrid interaction screen
[37]
PER2
(period circadian regulator 2)
PER2 is a member of the Period family of genes involved in circadian rhythms regulation and is expressed in a circadian pattern in the suprachiasmatic nucleus.
HEK293 cell line
Yeast two-hybrid interaction screen
[37]
NFIL3 (nuclear factor, interleukin 3 regulated), also known as E4BP4
NFIL3 is a basic leucine zipper transcription factor that binds to activating transcription factor (ATF) sites in many cellular and viral promoters. The encoded protein represses PER1 and PER2 expression and is therefore, involved in circadian rhythms regulation.
NIH/3T3 cell line
Glutathione S transferase pull-down assay and immunoprecipitation assay
[40]
Muscular activity
MYH13
(myosin heavy chain 13)
The gene MYH13 is predicted to encode a protein for microfilament motor activity and involved in muscle contraction.
HeLa cell line
Quantitative proteomics
[39]

2.4 Transcriptional activities

Although most bHLH members act as transcriptional activators, BHLHE41 functions as a nuclear transcriptional repressor through different mechanisms [9, 41] (Fig. 3).
First, this repressive activity relies on the direct binding of BHLHE41 to the E-box sequence of target promoters. A competition with E-box binding trans-activators then occurs. Finally, BHLHE41 recruits co-repressors, including Histone DeACetylase 1 (HDAC1), sirtuin 1 (SIRT1) and Euchromatic Histone lysine Methyltransferase 2 (EHMT2, also known as G9a) that mediate the deposition of repressive chromatin marks [45].
Second, this inhibitory activity can occur through direct protein-protein interactions, independently of DNA binding. These interactions involve other TFs, such as SP1 [9]. BHLHE41 can heterodimerize with bHLH factors to form non-functional complexes or to sequester these factors, as observed for MYOD1 [5]. It also antagonizes the transcriptional activity of other TFs, such as HIF-1 alpha, by promoting their proteasome-mediated degradation [32]. In addition, BHLHE41 regulates the expression of other TFs of the bHLH family that display repressor activity, such as TWIST1, involved in the regulation of myogenesis and osteogenesis [46]. Furthermore, in HEK293 cells, BHLHE41 suppresses ligand-induced RXR alpha transactivation activity, repressing target gene expression mediated by the RXR-LXR heterodimer [11].

2.5 Involvement of BHLHE41 in physiological processes

TFs of the bHLH superfamily contribute to the regulation of embryo development, homeostasis of different organs/tissues and differentiation of various cell types, including neuronal, skeletal muscle and hematopoietic cells [5, 4750]. BHLHE41 also is implicated in various biological processes (Fig. 4).

2.5.1 Myogenesis regulation

BHLHE41 is expressed in developing skeletal muscles where it heterodimerizes with the myogenic regulatory factor MYOD1. This factor belongs to bHLH family, is expressed by myoblasts and is involved in skeletal muscle differentiation, driving transcription of muscle-specific genes. BHLHE41 is expressed in proliferating myoblasts and is down-regulated during myogenic differentiation. It thus acts as a negative regulator of myogenesis, its overexpression inhibiting myogenic differentiation [51].
In addition, BHLHE41 interacts with the methyltransferase G9a that belongs to the SET domain-containing Su(var)3–9 family of proteins including Suv39h1/h2, SETDB1, SETDB2 and DIM-5. G9a mediates transcriptional repression activities by mono- and di-methylation of histone H3 lysine-9. It can also methylate MYOD1, resulting in the inhibition of its transcriptional activity and the negative regulation of myoblast differentiation [52].
BHLHE41-G9a interactions enhance BHLHE41 inhibition of skeletal myogenesis [44]. Interestingly, sumoylation of BHLHE41 by SUMO1 is required for the interaction with G9a and its full transcriptional repression activity during skeletal muscle differentiation [53, 54]. BHLHE41 also modulates skeletal muscle regeneration by regulating TGF-beta signaling and by counteracting SMAD-3-dependent expression of collagens and TIMP1 (tissue inhibitor of metalloproteinase 1) [55]. SREBP-1, which is involved in myogenesis regulation, can induce BHLHE41 expression that in turn mediates the repression of muscle genes [25]. BHLHE41 is a HDAC4 target in satellite cells, a specific type of muscle stem cells. During proliferative stage of myogenesis HDAC4 represses BHLHE41 expression, allowing satellite cell differentiation and fusion [56].

2.5.2 Adipogenesis regulation

BHLHE41 interacts with CEBPA and CEBPB, two transcriptional master regulators of adipogenic differentiation. It inhibits their transcriptional activity by retaining HDAC1 and G9a at the regulatory sites of their promoters [45], as well as for Peroxisome Proliferator-Activated Receptor gamma (PPARG), another master transcriptional regulator of adipogenesis [36]. SUMOylation enhances BHLHE41 repression of PPARG expression. This effect can be reversed by SENP1, a protease of the Sentrin/SUMO-specific Protease (SENP) family that de-SUMOylates BHLHE41, thus suppressing BHLHE41 inhibition of PPARG expression and enhancing adipogenesis [57].

2.5.3 Spermatogenesis

Spermatogenesis is a continuous process that relies on self-renewal and differentiation of spermatogonial stem cells [58]. Several factors, including SOHLH1, a testis-specific bHLH TF, are needed during the gonocyte-to-spermatogonia transition. BHLHE41 can inhibit spermatogonial differentiation in neonatal murine germ cells by repressing SOHLH1 expression [59]. The mechanisms that might inhibit the activity of BHLHE41 to allow the progress of spermatogenesis remain unexplored.

2.5.4 Regulation of circadian rhythms

In mammals, circadian rhythms are regulated by the suprachiasmatic nucleus of the hypothalamus and by other areas in the brain and peripheral tissues. At cellular level, several clock gene families are involved in circadian rhythms generation. They act in cooperation in the so-called “molecular clock”, through an autoregulatory transcription-translation feedback loop. CLOCK and BMAL1, two major clock genes, encode proteins that heterodimerize in the cytoplasm and then translocate to the nucleus, where they bind to E-box enhancer sequences and activate the transcription of other clock genes [60], including BHLHE40 and BHLHE41 [7, 12]. Honma et al. showed, in a murine model, that BHLHE40 and BHLHE41 are expressed in the suprachiasmatic nucleus in a circadian way. They mighty repress the CLOCK-BMAL1-induced transactivation of the PER1 promoter via direct protein interactions with BMAL1 or through E-box motif competition. Therefore, BHLHE40 and BHLHE41 are two regulators of the mammalian molecular clock and are thus considered as clock genes [42]. Of note, BHLHE41 gene mutations affect sleep duration in humans [61, 62]. A missense mutation in exon 5 of BHLHE41 inducing a replacement of proline at position 384 by an arginine was described in individuals with short sleep duration. This mutation resulted in the inability of the BHLHE41 protein to interact properly with the circadian clock transcription factors CLOCK and BMAL1 [61].
BHLHE41 is also implicated in the interplay between circadian and metabolic rhythms. BHLHE41 contributes to the circadian control of liver metabolism by regulating the expression of P450 cytochromes [63], through its interaction with CEBPA and the formation of a complex with HDAC1 [64]. Moreover, BHLHE41 expression is decreased by glucose deprivation [65], whereas it is increased by insulin [27].

2.5.5 DNA repair mechanisms

BHLHE41 influences DNA repair mechanisms. Liu and colleagues observed that BHLHE41 is upregulated in response to genotoxic agents in mouse NIH3T3 cells, resulting in S and G2/M cell cycle arrest. BHLHE41 upregulation led to increased expression of BRCA1, which is required for cell cycle arrest and DNA repair, and GADD45A, a p53 target gene also involved in cell cycle arrest. In addition, upon BHLHE41 overexpression, cells are protected from DNA damage-induced apoptosis and p53, p21 and BAX expression levels are reduced [66]. This indicates that BHLHE41 is involved in survival mechanisms when cells are treated by DNA-damaging agents. Hypoxia can reduce the expression of different genes involved in the DNA damage response, leading to impaired DNA repair and genome instability [6772]. In hypoxic conditions, BHLHE41 downregulates DNA damage response genes through the repression of their promoter activities [71].

2.5.6 Homeostasis of mesenchymal stem cells

BHLHE41 is highly expressed in mesenchymal stem cells (MSC) that can differentiate into several cell types, including adipocytes and smooth muscle cells. BHLHE41 suppresses MSC differentiation into muscle cells and adipocytes by inhibiting the transcriptional activities of MYOD1 and CEPB. This suggests that BHLHE41 may be involved in maintaining MSCs in an undifferentiated state [36, 73].

2.5.7 Regulation of tissue-specific macrophage properties

Macrophages are innate immune cells that contribute to the front-line immune defense and tissue homeostasis. Macrophages also display tissue-specific profiles through tissue-specific molecular programs, including transcriptional networks [74]. Rauschmeier et al. observed that BHLHE41 and its paralogue BHLHE40 can repress transcriptional activities that are crucial for defining the identity of lung alveolar macrophages, likely occurring at least in part through histone deacetylation. In this context, BHLHE41 is involved in their final stages of differentiation and maturation [75]. In addition, macrophages can be involved during repair and remodeling after myocardial infarction. Xu et al. reported that macrophages expressing BHLHE41 could limit fibrosis and expansion of developing infarct area [76].

2.5.8 BHLHE41 in the lymphoid lineage

The differentiation of B lymphocytes into plasma cells (PC) is driven by a complex network of TFs. Among them, the transcriptional repressor PRDM1/BLIMP1 is the master regulator of terminal B cell development [77]. During PC differentiation, the expression of PRDM1/BLIMP1 is significantly increased from the proliferating plasmablast stage. PRDM1/BLIMP1 directly represses several genes, including BCL6, PAX5, SPIB/Spi-B, ID3, CTIIA and MYC [78, 79]. PAX5 is essential for B-lineage commitment and activates the expression of B cell identity genes, including CD19. It represses the expression of genes activated in PCs, such as IGH immunoglobulin heavy locus and XBP1. PRDM1/BLIMP1 blocks immunoglobulin class switching by suppressing the expression of AID, KU70, KU86, DNA-PKCs and STAT6 [5]. In 2006, De Vos et al. reported that BHLHE41 was highly induced during PC differentiation, particularly at the plasmablast and PC stages. More recently, using an in vitro model of naive B cell differentiation, Santamaria et al. found that BHLHE40 is preferentially associated with non-differentiating CD23+ cells, whereas BHLHE41 is part of a terminal differentiation gene signature expressed in a subset of CD23- pre-plasmablast cells [80]. In addition, BHLHE41 appears in signatures of memory B cell (MBC) differentiation and activation [81, 82]. It is expressed in memory B cells (MBCs) that are poised for activation [83], as well as in more mature quiescent MBCs [84]. In addition, it might regulate the pre-MBC cell fate [85].
Furthermore, Kreslavsky et al. described a high expression of Bhlhe41 in B-1a murine cells. These long-lived, self-renewing innate-like B cells are localized in peritoneal and pleural cavities and contribute to the first-line defense against pathogens and self-antigens. Interestingly, during murine B cell development, Bhlhe41 is weakly expressed in PC, compared with B-1a cells. In Bhlhe41-/- mice, B cell development is unaffected, but the number of B-1a cells is significantly reduced. In these cells, Bhlhe41 acts as a transcriptional repressor, targeting cell cycle regulators, restricting B-1a cell proliferation and affecting their self-renewal. Therefore, Bhlhe41 is considered essential for B-1a cell development and functionality [86].
During B-1a differentiation, Bhlhe41 expression is under the control of ARID3A, a member of the ARID (AT-rich interaction domain) superfamily of DNA binding proteins involved in large chromatin-modulating complexes, with major roles in embryogenesis [87].
Upon activation after antigen recognition, naive CD4+ T cells can differentiate into distinct effector subsets, including T helper 2 (Th2) cells. These T lymphocytes secrete interleukin (IL)-4, IL-5 and IL-13, which play a significant role in the defense against parasites, allergic diseases and immunoglobulin production by B cells [88]. The differentiation into Th2 cells is tightly regulated and IL-2 has a key role in this process [89]. BHLHE41 is highly and selectively expressed in Th2 cells among all T helper cell subsets [24]. Its induction is particularly high during the late phase of Th2 cell differentiation, and seems to require IL-4, which activates STAT6, which in turn induces GATA3 expression [90]. In addition, BHLHE41 can induce the upregulation of CD25, the alpha subunit of the IL-2 receptor, resulting in hyper-responsiveness to IL-2 stimulation [91].

2.6 BHLHE41 in cancer

According to the role of BHLHE41 as a transcriptional repressor and its interactions with epigenetic enzymes participating in normal cell differentiation, deregulation of BHLHE41 may contribute to tumorigenesis [92]. Epigenetic modifications, miRNA, epitranscriptomic regulation and post translational modification may regulate BHLHE41 expression [93]. Furthermore, since BHLHE41 is involved in DNA damage response and of immune cell differentiation, defects may contribute to DNA damage response abnormalities and immune cell dysfunction promoting tumor formation [94].
Several studies have shown that, in solid tumors, BHLHE41 is involved in the regulation of cell proliferation, apoptosis, hypoxia response, invasion and metastasis. However, depending on the tumor type, BHLHE41 acts as an oncogene or a tumor suppressor (Fig. 5).

2.6.1 Endometrial carcinoma

In solid tumors, cancer cells are often exposed to hypoxia that can modulate the expression of various genes and contributes to aggressive cancer phenotypes [95]. Yunokawa et al. described in endometrial carcinoma cell lines cultured in hypoxic conditions an increase in the expression levels of HIF and its target genes, including BHLHE41. In addition, they observed that BHLHE41 expression was higher in complex atypical endometrial hyperplasia and endometrial carcinoma samples, compared with normal endometrium, suggesting a contribution of this TF to carcinogenesis [96]. However, other findings are in favor of BHLHE41 anti-tumor activity. Specifically, in the human endometrial carcinoma cell lines Ishiwaka and HEC-1B, overexpression of BHLHE41 inhibited their migration, invasion, and metastatic potential by attenuating NOTCH1 signaling [97]. Another study reported that BHLHE41 expression was significantly lower in endometrial carcinoma than in normal endometrium samples. In addition, BHLHE41 expression was inversely correlated with both myometrial invasion and lymph node invasion. In agreement, BHLHE41 expression in endometrial carcinoma samples progressively decreased with the increasing stage and histological grade. In addition, BHLHE41 upregulation inhibited tumor angiogenesis by decreasing HIF-1 alpha, leading to tumor growth inhibition [98]. Through the association with SP1, BHLHE41 might suppress the expression of TWIST1, a gene involved in Epithelial to Mesenchymal Transition (EMT) [99]. Of note, MIR130 microRNA family members can suppress the expression of BHLHE41 and BHLHE40 and promote EMT and tumor cell invasion in the HHUA endometrial cancer cell line [100].

2.6.2 Breast cancer

Several studies performed in breast cancer cell lines suggest that BHLHE41 has anti-apoptotic effects. Liu et al. observed that BHLHE41 knock-down (KO) in the human MCF-7 cell line enhanced apoptosis, affecting FAS, MYC, CASP8 (caspase 8), PARP (poly (ADP-ribose) polymerase) and BAX gene expression [101]. Similarly, Wu et al. described an increase in cleaved PARP in this cell line after BHLHE41 deletion [102]. BHLHE41 might also suppress the cell invasion potential of MCF-7 and MDA-MB-231 cells [103]. Montagner et al. showed that in MDA-231 cells, BHLHE41 reduces HIF activity and limits the effects of hypoxia, thus controlling the migratory and invasive properties. This effect is based on BHLHE41 binding to HIF-1 alpha, leading to its proteasomal degradation [32]. Conversely, Wu et al. found that BHLHE41 contributes to hypoxia-induced MCF-7 cell proliferation [104]. Bexarotene, a selective RXR agonist, used as chemopreventive agent, represses CCDN1 (cyclin D1) transcription, leading to cell cycle progression arrest. In MCF-7 cells, BHLHE41, but not BHLHE40, was induced by incubation with bexarotene. Subsequently, BHLHE41 bound the CCDN1 promoter and recruited HDAC1, resulting in the repression of CCDN1 transcription [105]. Indeed, tumors appeared earlier and were significantly larger in mice xenografted with MCF-7 cells in which BHLHE41 expression was inhibited by siRNAs compared to control siRNA. Interestingly, BHLHE41 was found in a breast cancer dormancy gene signature based on expression profile studies performed in dormant breast cancer cells [106]. In a large cohort of patients with triple-negative breast cancer, a low BHLHE41 expression level was correlated with a higher probability of metastasis development and reduced survival. It was also associated with a gene expression signature linked to TGF-beta activity, p53 mutations, and high HIF activity [32]. Fang et al. observed a lower BHLHE41 expression in primary breast cancer samples compared with normal tissues. They also found that a lower BHLHE41 expression level was correlated with a more aggressive phenotype and worse overall survival [107].

2.6.3 Oral cancers

Bhawal et al. observed that BHLHE41 gene expression increased after BHLHE40 silencing in the human tongue squamous cell carcinoma HSC-3 cell line. In turn, BHLHE41 inactivation enhances BHLHE40 expression. In addition, siRNA-mediated BHLHE41 silencing decreased the cyclin D1 level [16], a major regulator of the cell cycle G1/S transition [108]. Conversely, BHLHE41 overexpression in the HSC-3 cell line resulted in the inhibition of the pro-apoptotic factor BIM and of cisplatin-induced apoptosis, suggesting an anti-apoptotic effect [109]. Similar results were obtained in TE-11 cells (human esophageal squamous cell carcinoma) [110].
Salivary adenoid cystic carcinoma (SACC) is a cancer characterized by a substantial risk of recurrence and metastasis. It was observed that BHLHE41 might contribute to tumor cell dormancy in the SACC-83 cell line and primary SACC cells [111].

2.6.4 Gastric cancer

In a series of 49 human gastric cancer samples evaluated by immunohistochemistry, Li et al. observed a lower BHLHE41 expression compared with adjacent non-tumoral tissues. Moreover, this expression level was negatively correlated with clinical data, including lymph node invasion, TNM staging (Tumor, lymph Node, Metastasis) and survival. In vitro, the overexpression of BHLHE41 in the gastric cancer cell lines MGC803 and MKN-45 inhibited cell proliferation and increased apoptosis [112]. BHLHE41 also increased the sensitivity of gastric cancer cells to 5-fluorouracil through STAT5A inhibition [113]. In patient-derived xenograft (PDX) mouse models, BHLHE41 suppressed tumor growth and metastasis of human gastric cancer cells [112].

2.6.5 Colon cancer

In HCT116 and Lovo colon cancer cell lines, BHLHE41 overexpression decreased cell viability, migration and invasion. It also blocked the cell cycle and induced apoptosis. In mice xenografted with BHLHE41-overexpressing HCT116 cells, tumor weight was decreased compared to control cells. Of note, BHLHE41 reduced the level of HIF-1 alpha and EMT-related proteins in tumor tissues, suggesting that a high BHLHE41 expression attenuates the malignant behavior of colon cancer cells [114].

2.6.6 Pancreatic cancer

In BxPC-3 human pancreatic cancer cell line, TGF-beta increased the expression of BHLHE41, resulting in impaired cancer cell migration and invasive properties. BHLHE41 could limit pancreatic cancer progression by inhibiting the transcriptional repressor SNAIL family transcriptional repressor 2 (SNAI2, also known as SLUG) and BHLHE40 expression [115].
In a pancreatic ductal adenocarcinoma cell line, the expression of MLH1, a gene encoding a protein member of the DNA mismatch repair system, was suppressed by BHLHE41 following its induction by GLI1, a zinc finger TF that inhibits the activity of the mismatch repair system in these cells [23], this finding linking BHLHE41 to the DNA damage response.

2.6.7 Thyroid cancer

At mRNA and protein levels, BHLHE41 was significantly downregulated in thyroid cancer samples compared with normal tissues. In addition, in thyroid cancer cell lines, the overexpression of BHLHE41 demonstrated inhibitory effects on their viability, migration and invasive properties, concurrently reducing the expression of HIF-1 alpha [14].

2.6.8 Lung cancer

In lung adenocarcinoma, BHLHE41 expression was downregulated compared with normal lung tissue [116]. Moreover, a higher BHLHE41 expression was associated with a favorable prognosis [117]. In the A549, NCI-H520 and NCI-H596 lung cancer cell lines, colony formation was inhibited by BHLHE41 overexpression, mainly through the downregulation of CCDN1, leading to the inhibition of cell proliferation in NCI-H520 cells [116]. BHLHE41 also enhanced autophagy in the A549 and H2030 adenocarcinoma cell lines [117].

2.6.9 Osteosarcoma

While BHLHE41 has been reported to promote HIF-1 alpha degradation in several cancer models, it exhibits a contrasting role by stabilizing this factor in the osteosarcoma cell lines U2OS, MNNG and 143B. Elevated levels of both BHLHE41 and HIF-1 alpha in osteosarcoma cells have been associated with heightened invasiveness and metastatic potential. In addition, the activation of HIF-1 alpha upregulated the expression of BHLHE41, inducing a feedback loop in these cells [118]. It is noteworthy that, in osteosarcoma cells, BHLHE41 expression was negatively modulated by microRNA-138, which also promoted cell apoptosis [119].

2.6.10 Prostate cancer

In the PC-3 human prostate cancer cell line, Liu et al. observed opposite effects of BHLHE41 and BHLHE40. In the presence of TGF-beta, a pivotal factor for epithelial-mesenchymal transition (EMT) induction, BHLHE41 expression decreased, whereas BHLHE40 expression increased. Furthermore, BHLHE40 enhanced migratory capacities, whereas BHLHE41 impaired them in the PC-3 cell line. BHLHE41 and BHLHE40 also differentially regulated the expression of EMT-related factors in PC-3 cells [120]. Notably, BHLHE41 inhibited paclitaxel-induced apoptosis of human prostate cancer cells [121].

2.6.11 Renal cell carcinoma

Renal cell carcinoma (RCC) is one of the most common urologic cancer types, and clear cell RCC is the main subtype. In a series of 50 clear-cell RCC samples, Shen et al. observed that BHLHE41 was highly expressed in 94% of tumor tissue samples compared to normal adjacent tissues. In vitro, BHLHE41 gene silencing in A498 and in CAKI-1 RCC cell lines impaired cell proliferation and migration, indicating that it might act as a tumor-promoting factor in RCC [122]. However, other studies showed that BHLHE41 expression decreased during RCC stage progression [123]. A study reported that BHLHE41 promoted RCC cell growth without HIF1 expression [124]. However, there is no reported link between BHLHE41 expression and overall survival in patients with RCC using TCGA database investigations [125].

2.6.12 Glioblastoma

In the U87 malignant glioma cell line and U251 glioblastoma cell line, BHLHE41 overexpression promoted cell proliferation through the ERK/CCDN1 pathway and colony formation. Its deletion had the opposite effect in these tumor cells [126].

2.6.13 Expression and role of BHLHE41 in hematological malignancies

2.6.13.1 Acute Myeloid Leukemias
The KMT2A gene (previously known as MLL for Mixed Lineage Leukemia) is located on chromosome 11q23 and encodes a large histone methyltransferase. In acute leukemia, the 11q23 translocation, in which a region of KMT2A is fused with a partner gene, is one of the most common chromosomal abnormalities. Fusion with AF6 results in the fusion protein KMT2A-AF6 that can recruit elongation assisting protein (EAP) and disruptor of telomeric silencing 1-like (DOT1L, a histone methyltransferase that methylates lysine-79 of histone H3) to activate the transcriptional elongation of target genes [127129]. Interestingly, Numata et al. found that BHLHE41 was overexpressed in KMT2A::AF6 acute myeloid leukemia (AML) cells, and that BHLHE41 was a direct transcriptional target of KMT2A-AF6. They did not observe BHLHE41 overexpression in other AML subtypes and in normal CD34+ bone marrow cells. In addition, PDX human BHLHE41-deleted KMT2A::AF6 AML cells showed a prolonged survival compared to control condition. This effect was linked to increased apoptosis, reduced growth and decreased colony formation of blast cells in the absence of BHLHE41 expression. Of note, BHLHE41 deletion did not affect normal hematopoiesis. Therefore, BHLHE41 is described as crucial to maintain clonogenic growth and to prevent apoptosis in KMT2A::AF6 AML cells [130].
2.6.13.2 Plasma cell disorders
Multiple myeloma is characterized by accumulation of malignant plasma cells in the bone marrow compartment. A high BHLHE41 expression was observed in malignant plasma cells [131]. In addition, in a Korean cohort of 67 patients with newly diagnosed multiple myeloma, BHLHE41 somatic mutations were associated with poor outcome [132]. Similarly, BHLHE41 was found upregulated in Waldenström macroglobulinemia samples [133], Waldenström macroglobulinemia being a rare indolent B-cell lymphoma with bone marrow infiltration by lymphoplasmacytic tumoral cells and IgM monoclonal gammopathy.
2.6.13.3 BHLHE41 is an essential gene in hematological malignancies
We used public datasets of RNAi viability assay screening (Dependency Map data, Broad Institute, www.​depmap.​org) [134] to determine whether BHLHE41 is an essential gene in several cancers. We observed that BHLHE41 was associated with a significantly lower DEMETER2 score (a measure of gene dependency) in Ewing sarcoma (P=1E-9, n=10), multiple myeloma (P=1.6E-5, n=16), non-Hodgkin lymphoma (P=6.1E-6, n=26) and AML (P=6E-5, n=22) cell lines (Fig. 6A). Furthermore, using the Genomicscape webtool (http://​www.​genomicscape.​com) and publicly available cohorts of patients with hematological malignancies, we identified that a high expression of BHLHE41 is associated with a poor survival in patients with AML and MM (Fig. 6B), supporting the results of the RNAi screening (Fig. 6A).
Given that BHLHE41 could govern the expression of genes required for the development of germinal center B cells and plasma cells, its deregulation could be involved in the biology of non-Hodgkin lymphoma and multiple myeloma. In mice, loss of BHLHE41 impedes plasma cell differentiation at an activated B cell stage, regulating IGH and key plasma cell TFs, including PRDM1/BLIMP1 and XBP1. A study reported that BHLHE41 is required for the induction and maintenance of open chromatin regions [135], suggesting a potential role in the pathophysiology of malignant cells. A more recent study reported that, in Ewing sarcoma, ID2, a member of the ID1-4 family which binds and sequesters bHLH TFs, is a critical regulator of developmental-related genes and tumor growth in vitro and in vivo [136]. Altogether, these findings underscore the need for further studies to decipher the precise role of BHLHE41 in cancer, and its interest as a potential therapeutic target.

3 Conclusion

The complexity of BHLHE41 activity as a transcriptional repressor relies on the multitude of pathways that regulate its expression and on the diversity of interaction partners. Its activity contributes to the regulation of basic physiological mechanisms including transcriptional regulation, DNA damage response and immune cell differentiation. Defects in DNA damage response and immune response play a major role in cancer development. Recent investigations underlined a BHLHE41 overexpression in several cancers including ovarian serous adenocarcinoma, stomach adenocarcinoma and thyroid cancer [125]. Furthermore, genetic changes involving BHLHE41 with missense mutation, amplification, gene fusion and deletion have been identified in solid cancer, suggesting that BHLHE41 deregulations may be associated with processes related to cancer development [125]. Accordingly, BHLHE41 was identified as a factor involved in many solid cancer types and in some hematological malignancies. Its contribution to tumor mechanisms, as a pro-tumoral factor or as an anti-tumoral factor, is not fully characterized in the different cancer types and require further investigations. However, RNAi screening uncovers BHLHE41 as an essential gene in different cancers including Ewing sarcoma, multiple myeloma, Hodgkin lymphoma and acute myeloid leukemia. Since BHLHE41 could govern the expression of genes required for the development of germinal center B cells and plasma cells, its deregulation could play a role in the pathophysiology of non-Hodgkin lymphoma and multiple myeloma. Using publicly available cohorts of patients, a high expression of BHLHE41 is associated with a poor outcome in multiple myeloma and acute myeloid leukemia hematological malignancies, supporting the results of RNAi screening and suggesting it could represent a potential therapeutic target. The biological functions of BHLHE41 and their interactions with other proteins are complex in normal and malignant cells. BHLHE41 might be a new candidate biomarker in cancer, and possibly a new therapeutic target.

Acknowledgements

Not applicable.

Declarations

Not applicable.
Not applicable.

Competing interests

The authors declare no competing interests.
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by-nc-nd/​4.​0/​.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Unsere Produktempfehlungen

e.Med Interdisziplinär

Kombi-Abonnement

Für Ihren Erfolg in Klinik und Praxis - Die beste Hilfe in Ihrem Arbeitsalltag

Mit e.Med Interdisziplinär erhalten Sie Zugang zu allen CME-Fortbildungen und Fachzeitschriften auf SpringerMedizin.de.

e.Med Innere Medizin

Kombi-Abonnement

Mit e.Med Innere Medizin erhalten Sie Zugang zu CME-Fortbildungen des Fachgebietes Innere Medizin, den Premium-Inhalten der internistischen Fachzeitschriften, inklusive einer gedruckten internistischen Zeitschrift Ihrer Wahl.

Literatur
2.
Zurück zum Zitat R.J. Kewley, M.L. Whitelaw, A. Chapman-Smith, The mammalian basic helix-loop-helix/PAS family of transcriptional regulators. Int. J. Biochem. Cell. Biol. 36(2), 189–204 (2004)PubMedCrossRef R.J. Kewley, M.L. Whitelaw, A. Chapman-Smith, The mammalian basic helix-loop-helix/PAS family of transcriptional regulators. Int. J. Biochem. Cell. Biol. 36(2), 189–204 (2004)PubMedCrossRef
3.
Zurück zum Zitat R.L. Davis, D.L. Turner, Vertebrate hairy and enhancer of split related proteins: transcriptional repressors regulating cellular differentiation and embryonic patterning. Oncogene. 20(58), 8342–8357 (2001)PubMedCrossRef R.L. Davis, D.L. Turner, Vertebrate hairy and enhancer of split related proteins: transcriptional repressors regulating cellular differentiation and embryonic patterning. Oncogene. 20(58), 8342–8357 (2001)PubMedCrossRef
4.
Zurück zum Zitat C.R. Vinson, K.C. Garcia, Molecular model for DNA recognition by the family of basic-helix-loop-helix-zipper proteins. New. Biol. 4(4), 396–403 (1992)PubMed C.R. Vinson, K.C. Garcia, Molecular model for DNA recognition by the family of basic-helix-loop-helix-zipper proteins. New. Biol. 4(4), 396–403 (1992)PubMed
5.
Zurück zum Zitat K. Fujimoto, H. Hamaguchi, T. Hashiba, T. Nakamura, T. Kawamoto, F. Sato et al., Transcriptional repression by the basic helix-loop-helix protein Dec2: multiple mechanisms through E-box elements. Int. J. Mol. Med. 19(6), 925–932 (2007)PubMed K. Fujimoto, H. Hamaguchi, T. Hashiba, T. Nakamura, T. Kawamoto, F. Sato et al., Transcriptional repression by the basic helix-loop-helix protein Dec2: multiple mechanisms through E-box elements. Int. J. Mol. Med. 19(6), 925–932 (2007)PubMed
6.
Zurück zum Zitat K. Fujimoto, M. Shen, M. Noshiro, K. Matsubara, S. Shingu, K. Honda et al., Molecular cloning and characterization of DEC2, a new member of basic helix-loop-helix proteins. Biochem. Biophys. Res. Commun. 280(1), 164–171 (2001)PubMedCrossRef K. Fujimoto, M. Shen, M. Noshiro, K. Matsubara, S. Shingu, K. Honda et al., Molecular cloning and characterization of DEC2, a new member of basic helix-loop-helix proteins. Biochem. Biophys. Res. Commun. 280(1), 164–171 (2001)PubMedCrossRef
7.
Zurück zum Zitat H. Hamaguchi, K. Fujimoto, T. Kawamoto, M. Noshiro, K. Maemura, N. Takeda et al., Expression of the gene for Dec2, a basic helix-loop-helix transcription factor, is regulated by a molecular clock system. Biochem. J. 382(Pt 1), 43–50 (2004)PubMedPubMedCentralCrossRef H. Hamaguchi, K. Fujimoto, T. Kawamoto, M. Noshiro, K. Maemura, N. Takeda et al., Expression of the gene for Dec2, a basic helix-loop-helix transcription factor, is regulated by a molecular clock system. Biochem. J. 382(Pt 1), 43–50 (2004)PubMedPubMedCentralCrossRef
8.
Zurück zum Zitat M.J. Rossner, J. Dorr, P. Gass, M.H. Schwab, K.A. Nave, SHARPs: mammalian enhancer-of-split- and hairy-related proteins coupled to neuronal stimulation. Mol. Cell. Neurosci. 10(3–4), 460–475 (1997)PubMedCrossRef M.J. Rossner, J. Dorr, P. Gass, M.H. Schwab, K.A. Nave, SHARPs: mammalian enhancer-of-split- and hairy-related proteins coupled to neuronal stimulation. Mol. Cell. Neurosci. 10(3–4), 460–475 (1997)PubMedCrossRef
9.
Zurück zum Zitat S. Azmi, H. Sun, A. Ozog, R. Taneja, mSharp-1/DEC2, a basic helix-loop-helix protein functions as a transcriptional repressor of E box activity and Stra13 expression. J. Biol. Chem. 278(22), 20098–20109 (2003)PubMedCrossRef S. Azmi, H. Sun, A. Ozog, R. Taneja, mSharp-1/DEC2, a basic helix-loop-helix protein functions as a transcriptional repressor of E box activity and Stra13 expression. J. Biol. Chem. 278(22), 20098–20109 (2003)PubMedCrossRef
10.
Zurück zum Zitat K. Yamada, K. Miyamoto, Basic helix-loop-helix transcription factors, BHLHB2 and BHLHB3; their gene expressions are regulated by multiple extracellular stimuli. Front. Biosci. 10, 3151–3171 (2005)PubMedCrossRef K. Yamada, K. Miyamoto, Basic helix-loop-helix transcription factors, BHLHB2 and BHLHB3; their gene expressions are regulated by multiple extracellular stimuli. Front. Biosci. 10, 3151–3171 (2005)PubMedCrossRef
11.
Zurück zum Zitat Y. Cho, M. Noshiro, M. Choi, K. Morita, T. Kawamoto, K. Fujimoto et al., The basic helix-loop-helix proteins differentiated embryo chondrocyte (DEC) 1 and DEC2 function as corepressors of retinoid X receptors. Mol. Pharmacol. 76(6), 1360–1369 (2009)PubMedCrossRef Y. Cho, M. Noshiro, M. Choi, K. Morita, T. Kawamoto, K. Fujimoto et al., The basic helix-loop-helix proteins differentiated embryo chondrocyte (DEC) 1 and DEC2 function as corepressors of retinoid X receptors. Mol. Pharmacol. 76(6), 1360–1369 (2009)PubMedCrossRef
12.
Zurück zum Zitat F. Sato, T. Kawamoto, K. Fujimoto, M. Noshiro, K.K. Honda, S. Honma et al., Functional analysis of the basic helix-loop-helix transcription factor DEC1 in circadian regulation. Interaction with BMAL1. Eur. J. Biochem. 271(22), 4409–4419 (2004)PubMedCrossRef F. Sato, T. Kawamoto, K. Fujimoto, M. Noshiro, K.K. Honda, S. Honma et al., Functional analysis of the basic helix-loop-helix transcription factor DEC1 in circadian regulation. Interaction with BMAL1. Eur. J. Biochem. 271(22), 4409–4419 (2004)PubMedCrossRef
13.
Zurück zum Zitat J.H. Chuang, A.A. Yarmishyn, D.K. Hwang, C.C. Hsu, M.L. Wang, Y.P. Yang et al., Expression profiling of cell-intrinsic regulators in the process of differentiation of human iPSCs into retinal lineages. Stem Cell. Res. Ther. 9(1), 140 (2018)PubMedPubMedCentralCrossRef J.H. Chuang, A.A. Yarmishyn, D.K. Hwang, C.C. Hsu, M.L. Wang, Y.P. Yang et al., Expression profiling of cell-intrinsic regulators in the process of differentiation of human iPSCs into retinal lineages. Stem Cell. Res. Ther. 9(1), 140 (2018)PubMedPubMedCentralCrossRef
14.
Zurück zum Zitat Z.H. Zhou, B. Wang, X.B. Cheng, X.E. Zhang, J. Tang, W.J. Tang et al., Roles of SHARP1 in thyroid cancer. Mol. Med. Rep. 13(6), 5365–5371 (2016)PubMedCrossRef Z.H. Zhou, B. Wang, X.B. Cheng, X.E. Zhang, J. Tang, W.J. Tang et al., Roles of SHARP1 in thyroid cancer. Mol. Med. Rep. 13(6), 5365–5371 (2016)PubMedCrossRef
15.
Zurück zum Zitat T. Kawamoto, M. Noshiro, F. Sato, K. Maemura, N. Takeda, R. Nagai et al., A novel autofeedback loop of Dec1 transcription involved in circadian rhythm regulation. Biochem. Biophys. Res. Commun. 313(1), 117–124 (2004)PubMedCrossRef T. Kawamoto, M. Noshiro, F. Sato, K. Maemura, N. Takeda, R. Nagai et al., A novel autofeedback loop of Dec1 transcription involved in circadian rhythm regulation. Biochem. Biophys. Res. Commun. 313(1), 117–124 (2004)PubMedCrossRef
16.
Zurück zum Zitat U.K. Bhawal, F. Sato, Y. Arakawa, K. Fujimoto, T. Kawamoto, K. Tanimoto et al., Basic helix-loop-helix transcription factor DEC1 negatively regulates cyclin D1. J. Pathol. 224(3), 420–429 (2011)PubMedCrossRef U.K. Bhawal, F. Sato, Y. Arakawa, K. Fujimoto, T. Kawamoto, K. Tanimoto et al., Basic helix-loop-helix transcription factor DEC1 negatively regulates cyclin D1. J. Pathol. 224(3), 420–429 (2011)PubMedCrossRef
17.
Zurück zum Zitat M.P. Butler, S. Honma, T. Fukumoto, T. Kawamoto, K. Fujimoto, M. Noshiro et al., Dec1 and Dec2 expression is disrupted in the suprachiasmatic nuclei of clock mutant mice. J. Biol. Rhythms. 19(2), 126–134 (2004)PubMedCrossRef M.P. Butler, S. Honma, T. Fukumoto, T. Kawamoto, K. Fujimoto, M. Noshiro et al., Dec1 and Dec2 expression is disrupted in the suprachiasmatic nuclei of clock mutant mice. J. Biol. Rhythms. 19(2), 126–134 (2004)PubMedCrossRef
18.
Zurück zum Zitat P. Bragado, Y. Estrada, F. Parikh, S. Krause, C. Capobianco, H.G. Farina et al., TGF-beta2 dictates disseminated tumour cell fate in target organs through TGF-beta-RIII and p38alpha/beta signalling. Nat. Cell. Biol. 15(11), 1351–1361 (2013)PubMedPubMedCentralCrossRef P. Bragado, Y. Estrada, F. Parikh, S. Krause, C. Capobianco, H.G. Farina et al., TGF-beta2 dictates disseminated tumour cell fate in target organs through TGF-beta-RIII and p38alpha/beta signalling. Nat. Cell. Biol. 15(11), 1351–1361 (2013)PubMedPubMedCentralCrossRef
19.
Zurück zum Zitat J. Olkkonen, V.P. Kouri, J. Hynninen, Y.T. Konttinen, J. Mandelin, Differentially expressed in chondrocytes 2 (DEC2) increases the expression of IL-1beta and is abundantly Present in Synovial membrane in rheumatoid arthritis. PLoS One. 10(12), e0145279 (2015)PubMedPubMedCentralCrossRef J. Olkkonen, V.P. Kouri, J. Hynninen, Y.T. Konttinen, J. Mandelin, Differentially expressed in chondrocytes 2 (DEC2) increases the expression of IL-1beta and is abundantly Present in Synovial membrane in rheumatoid arthritis. PLoS One. 10(12), e0145279 (2015)PubMedPubMedCentralCrossRef
20.
Zurück zum Zitat K. Miyazaki, T. Kawamoto, K. Tanimoto, M. Nishiyama, H. Honda, Y. Kato, Identification of functional hypoxia response elements in the promoter region of the DEC1 and DEC2 genes. J. Biol. Chem. 277(49), 47014–47021 (2002)PubMedCrossRef K. Miyazaki, T. Kawamoto, K. Tanimoto, M. Nishiyama, H. Honda, Y. Kato, Identification of functional hypoxia response elements in the promoter region of the DEC1 and DEC2 genes. J. Biol. Chem. 277(49), 47014–47021 (2002)PubMedCrossRef
21.
Zurück zum Zitat N. Ozaki, M. Noshiro, T. Kawamoto, A. Nakashima, K. Honda, U. Fukuzaki-Dohi et al., Regulation of basic helix-loop-helix transcription factors Dec1 and Dec2 by RORalpha and their roles in adipogenesis. Genes Cells. 17(2), 109–121 (2012)PubMedCrossRef N. Ozaki, M. Noshiro, T. Kawamoto, A. Nakashima, K. Honda, U. Fukuzaki-Dohi et al., Regulation of basic helix-loop-helix transcription factors Dec1 and Dec2 by RORalpha and their roles in adipogenesis. Genes Cells. 17(2), 109–121 (2012)PubMedCrossRef
22.
Zurück zum Zitat M. Adorno, M. Cordenonsi, M. Montagner, S. Dupont, C. Wong, B. Hann et al., A Mutant-p53/Smad complex opposes p63 to empower TGFbeta-induced metastasis. Cell. 137(1), 87–98 (2009)PubMedCrossRef M. Adorno, M. Cordenonsi, M. Montagner, S. Dupont, C. Wong, B. Hann et al., A Mutant-p53/Smad complex opposes p63 to empower TGFbeta-induced metastasis. Cell. 137(1), 87–98 (2009)PubMedCrossRef
23.
Zurück zum Zitat S. Inaguma, M. Riku, M. Hashimoto, H. Murakami, S. Saga, H. Ikeda et al., GLI1 interferes with the DNA mismatch repair system in pancreatic cancer through BHLHE41-mediated suppression of MLH1. Cancer Res. 73(24), 7313–7323 (2013)PubMedCrossRef S. Inaguma, M. Riku, M. Hashimoto, H. Murakami, S. Saga, H. Ikeda et al., GLI1 interferes with the DNA mismatch repair system in pancreatic cancer through BHLHE41-mediated suppression of MLH1. Cancer Res. 73(24), 7313–7323 (2013)PubMedCrossRef
24.
Zurück zum Zitat X.O. Yang, P. Angkasekwinai, J. Zhu, J. Peng, Z. Liu, R. Nurieva et al., Requirement for the basic helix-loop-helix transcription factor Dec2 in initial TH2 lineage commitment. Nat. Immunol. 10(12), 1260–1266 (2009)PubMedPubMedCentralCrossRef X.O. Yang, P. Angkasekwinai, J. Zhu, J. Peng, Z. Liu, R. Nurieva et al., Requirement for the basic helix-loop-helix transcription factor Dec2 in initial TH2 lineage commitment. Nat. Immunol. 10(12), 1260–1266 (2009)PubMedPubMedCentralCrossRef
25.
Zurück zum Zitat V. Lecomte, E. Meugnier, V. Euthine, C. Durand, D. Freyssenet, G. Nemoz et al., A new role for sterol regulatory element binding protein 1 transcription factors in the regulation of muscle mass and muscle cell differentiation. Mol. Cell. Biol. 30(5), 1182–1198 (2010)PubMedCrossRef V. Lecomte, E. Meugnier, V. Euthine, C. Durand, D. Freyssenet, G. Nemoz et al., A new role for sterol regulatory element binding protein 1 transcription factors in the regulation of muscle mass and muscle cell differentiation. Mol. Cell. Biol. 30(5), 1182–1198 (2010)PubMedCrossRef
26.
Zurück zum Zitat H. Oike, K. Nagai, T. Fukushima, N. Ishida, M. Kobori, Feeding cues and injected nutrients induce acute expression of multiple clock genes in the mouse liver. PLoS One. 6(8), e23709 (2011)PubMedPubMedCentralCrossRef H. Oike, K. Nagai, T. Fukushima, N. Ishida, M. Kobori, Feeding cues and injected nutrients induce acute expression of multiple clock genes in the mouse liver. PLoS One. 6(8), e23709 (2011)PubMedPubMedCentralCrossRef
27.
Zurück zum Zitat K. Takagi, K. Asano, A. Haneishi, M. Ono, Y. Komatsu, T. Yamamoto et al., Insulin stimulates the expression of the SHARP-1 gene via multiple signaling pathways. Horm. Metab. Res. 46(6), 397–403 (2014)PubMedCrossRef K. Takagi, K. Asano, A. Haneishi, M. Ono, Y. Komatsu, T. Yamamoto et al., Insulin stimulates the expression of the SHARP-1 gene via multiple signaling pathways. Horm. Metab. Res. 46(6), 397–403 (2014)PubMedCrossRef
28.
Zurück zum Zitat L. Zawel, J. Yu, C.J. Torrance, S. Markowitz, K.W. Kinzler, B. Vogelstein et al., DEC1 is a downstream target of TGF-beta with sequence-specific transcriptional repressor activities. Proc. Natl. Acad. Sci. U S A 99(5), 2848–2853 (2002)PubMedPubMedCentralCrossRef L. Zawel, J. Yu, C.J. Torrance, S. Markowitz, K.W. Kinzler, B. Vogelstein et al., DEC1 is a downstream target of TGF-beta with sequence-specific transcriptional repressor activities. Proc. Natl. Acad. Sci. U S A 99(5), 2848–2853 (2002)PubMedPubMedCentralCrossRef
29.
Zurück zum Zitat A.V. Ivanova, S.V. Ivanov, X. Zhang, V.N. Ivanov, O.A. Timofeeva, M.I. Lerman, STRA13 interacts with STAT3 and modulates transcription of STAT3-dependent targets. J. Mol. Biol. 340(4), 641–653 (2004)PubMedCrossRef A.V. Ivanova, S.V. Ivanov, X. Zhang, V.N. Ivanov, O.A. Timofeeva, M.I. Lerman, STRA13 interacts with STAT3 and modulates transcription of STAT3-dependent targets. J. Mol. Biol. 340(4), 641–653 (2004)PubMedCrossRef
30.
Zurück zum Zitat G.L. Semenza, HIF-1: mediator of physiological and pathophysiological responses to hypoxia. J. Appl. Physiol. (1985). 88(4), 1474–1480 (2000)PubMedCrossRef G.L. Semenza, HIF-1: mediator of physiological and pathophysiological responses to hypoxia. J. Appl. Physiol. (1985). 88(4), 1474–1480 (2000)PubMedCrossRef
31.
Zurück zum Zitat K. Tanimoto, Y. Makino, T. Pereira, L. Poellinger, Mechanism of regulation of the hypoxia-inducible factor-1 alpha by the Von Hippel-Lindau tumor suppressor protein. EMBO J. 19(16), 4298–4309 (2000)PubMedPubMedCentralCrossRef K. Tanimoto, Y. Makino, T. Pereira, L. Poellinger, Mechanism of regulation of the hypoxia-inducible factor-1 alpha by the Von Hippel-Lindau tumor suppressor protein. EMBO J. 19(16), 4298–4309 (2000)PubMedPubMedCentralCrossRef
32.
Zurück zum Zitat M. Montagner, E. Enzo, M. Forcato, F. Zanconato, A. Parenti, E. Rampazzo et al., SHARP1 suppresses breast cancer metastasis by promoting degradation of hypoxia-inducible factors. Nature. 487(7407), 380–384 (2012)PubMedCrossRef M. Montagner, E. Enzo, M. Forcato, F. Zanconato, A. Parenti, E. Rampazzo et al., SHARP1 suppresses breast cancer metastasis by promoting degradation of hypoxia-inducible factors. Nature. 487(7407), 380–384 (2012)PubMedCrossRef
33.
Zurück zum Zitat F. Sato, U.K. Bhawal, T. Kawamoto, K. Fujimoto, T. Imaizumi, T. Imanaka et al., Basic-helix-loop-helix (bHLH) transcription factor DEC2 negatively regulates vascular endothelial growth factor expression. Genes Cells. 13(2), 131–144 (2008)PubMedCrossRef F. Sato, U.K. Bhawal, T. Kawamoto, K. Fujimoto, T. Imaizumi, T. Imanaka et al., Basic-helix-loop-helix (bHLH) transcription factor DEC2 negatively regulates vascular endothelial growth factor expression. Genes Cells. 13(2), 131–144 (2008)PubMedCrossRef
34.
Zurück zum Zitat Y. Kato, T. Kawamoto, K. Fujimoto, M. Noshiro, DEC1/STRA13/SHARP2 and DEC2/SHARP1 coordinate physiological processes, including circadian rhythms in response to environmental stimuli. Curr. Top. Dev. Biol. 110, 339–372 (2014)PubMedCrossRef Y. Kato, T. Kawamoto, K. Fujimoto, M. Noshiro, DEC1/STRA13/SHARP2 and DEC2/SHARP1 coordinate physiological processes, including circadian rhythms in response to environmental stimuli. Curr. Top. Dev. Biol. 110, 339–372 (2014)PubMedCrossRef
35.
Zurück zum Zitat M.J. Rossner, H. Oster, S.P. Wichert, L. Reinecke, M.C. Wehr, J. Reinecke et al., Disturbed clockwork resetting in Sharp-1 and Sharp-2 single and double mutant mice. PLoS One. 3(7), e2762 (2008)PubMedPubMedCentralCrossRef M.J. Rossner, H. Oster, S.P. Wichert, L. Reinecke, M.C. Wehr, J. Reinecke et al., Disturbed clockwork resetting in Sharp-1 and Sharp-2 single and double mutant mice. PLoS One. 3(7), e2762 (2008)PubMedPubMedCentralCrossRef
36.
Zurück zum Zitat N.T. Gulbagci, L. Li, B. Ling, S. Gopinadhan, M. Walsh, M. Rossner et al., SHARP1/DEC2 inhibits adipogenic differentiation by regulating the activity of C/EBP. EMBO Rep. 10(1), 79–86 (2009)PubMedCrossRef N.T. Gulbagci, L. Li, B. Ling, S. Gopinadhan, M. Walsh, M. Rossner et al., SHARP1/DEC2 inhibits adipogenic differentiation by regulating the activity of C/EBP. EMBO Rep. 10(1), 79–86 (2009)PubMedCrossRef
37.
Zurück zum Zitat T. Wallach, K. Schellenberg, B. Maier, R.K. Kalathur, P. Porras, E.E. Wanker et al., Dynamic circadian protein-protein interaction networks predict temporal organization of cellular functions. PLoS Genet. 9(3), e1003398 (2013)PubMedPubMedCentralCrossRef T. Wallach, K. Schellenberg, B. Maier, R.K. Kalathur, P. Porras, E.E. Wanker et al., Dynamic circadian protein-protein interaction networks predict temporal organization of cellular functions. PLoS Genet. 9(3), e1003398 (2013)PubMedPubMedCentralCrossRef
38.
Zurück zum Zitat S.M. Choi, H.J. Cho, H. Cho, K.H. Kim, J.B. Kim, H. Park, Stra13/DEC1 and DEC2 inhibit sterol regulatory element binding protein-1c in a hypoxia-inducible factor-dependent mechanism. Nucleic Acids Res. 36(20), 6372–6385 (2008)PubMedPubMedCentralCrossRef S.M. Choi, H.J. Cho, H. Cho, K.H. Kim, J.B. Kim, H. Park, Stra13/DEC1 and DEC2 inhibit sterol regulatory element binding protein-1c in a hypoxia-inducible factor-dependent mechanism. Nucleic Acids Res. 36(20), 6372–6385 (2008)PubMedPubMedCentralCrossRef
39.
Zurück zum Zitat M.Y. Hein, N.C. Hubner, I. Poser, J. Cox, N. Nagaraj, Y. Toyoda et al., A human interactome in three quantitative dimensions organized by stoichiometries and abundances. Cell. 163(3), 712–723 (2015)PubMedCrossRef M.Y. Hein, N.C. Hubner, I. Poser, J. Cox, N. Nagaraj, Y. Toyoda et al., A human interactome in three quantitative dimensions organized by stoichiometries and abundances. Cell. 163(3), 712–723 (2015)PubMedCrossRef
40.
Zurück zum Zitat S. Tanoue, K. Fujimoto, J. Myung, F. Hatanaka, Y. Kato, T. Takumi, DEC2-E4BP4 heterodimer represses the transcriptional enhancer activity of the EE element in the Per2 promoter. Front. Neurol. 6, 166 (2015)PubMedPubMedCentralCrossRef S. Tanoue, K. Fujimoto, J. Myung, F. Hatanaka, Y. Kato, T. Takumi, DEC2-E4BP4 heterodimer represses the transcriptional enhancer activity of the EE element in the Per2 promoter. Front. Neurol. 6, 166 (2015)PubMedPubMedCentralCrossRef
41.
Zurück zum Zitat M. Garriga-Canut, A. Roopra, N.J. Buckley, The basic helix-loop-helix protein, sharp-1, represses transcription by a histone deacetylase-dependent and histone deacetylase-independent mechanism. J. Biol. Chem. 276(18), 14821–14828 (2001)PubMedCrossRef M. Garriga-Canut, A. Roopra, N.J. Buckley, The basic helix-loop-helix protein, sharp-1, represses transcription by a histone deacetylase-dependent and histone deacetylase-independent mechanism. J. Biol. Chem. 276(18), 14821–14828 (2001)PubMedCrossRef
42.
Zurück zum Zitat S. Honma, T. Kawamoto, Y. Takagi, K. Fujimoto, F. Sato, M. Noshiro et al., Dec1 and Dec2 are regulators of the mammalian molecular clock. Nature. 419(6909), 841–844 (2002)PubMedCrossRef S. Honma, T. Kawamoto, Y. Takagi, K. Fujimoto, F. Sato, M. Noshiro et al., Dec1 and Dec2 are regulators of the mammalian molecular clock. Nature. 419(6909), 841–844 (2002)PubMedCrossRef
43.
Zurück zum Zitat I. Vastrik, P. D’Eustachio, E. Schmidt, G. Gopinath, D. Croft, de B. Bono et al., Reactome: a knowledge base of biologic pathways and processes. Genome Biol. 8(3), R39 (2007)PubMedPubMedCentralCrossRef I. Vastrik, P. D’Eustachio, E. Schmidt, G. Gopinath, D. Croft, de B. Bono et al., Reactome: a knowledge base of biologic pathways and processes. Genome Biol. 8(3), R39 (2007)PubMedPubMedCentralCrossRef
44.
Zurück zum Zitat B.M. Ling, S. Gopinadhan, W.K. Kok, S.R. Shankar, P. Gopal, N. Bharathy et al., G9a mediates Sharp-1-dependent inhibition of skeletal muscle differentiation. Mol. Biol. Cell. 23(24), 4778–4785 (2012)PubMedPubMedCentralCrossRef B.M. Ling, S. Gopinadhan, W.K. Kok, S.R. Shankar, P. Gopal, N. Bharathy et al., G9a mediates Sharp-1-dependent inhibition of skeletal muscle differentiation. Mol. Biol. Cell. 23(24), 4778–4785 (2012)PubMedPubMedCentralCrossRef
45.
Zurück zum Zitat J.R. Ow, Y.H. Tan, Y. Jin, A.G. Bahirvani, R. Taneja, Stra13 and Sharp-1, the non-grouchy regulators of development and disease. Curr. Top. Dev. Biol. 110, 317–338 (2014)PubMedCrossRef J.R. Ow, Y.H. Tan, Y. Jin, A.G. Bahirvani, R. Taneja, Stra13 and Sharp-1, the non-grouchy regulators of development and disease. Curr. Top. Dev. Biol. 110, 317–338 (2014)PubMedCrossRef
46.
Zurück zum Zitat M. Suzuki, F. Sato, U.K. Bhawal, The basic helix-loop-helix (bHLH) transcription factor DEC2 negatively regulates Twist1 through an E-box element. Biochem. Biophys. Res. Commun. 455(3–4), 390–395 (2014)PubMedCrossRef M. Suzuki, F. Sato, U.K. Bhawal, The basic helix-loop-helix (bHLH) transcription factor DEC2 negatively regulates Twist1 through an E-box element. Biochem. Biophys. Res. Commun. 455(3–4), 390–395 (2014)PubMedCrossRef
47.
Zurück zum Zitat S.E. Ross, M.E. Greenberg, C.D. Stiles, Basic helix-loop-helix factors in cortical development. Neuron. 39(1), 13–25 (2003)PubMedCrossRef S.E. Ross, M.E. Greenberg, C.D. Stiles, Basic helix-loop-helix factors in cortical development. Neuron. 39(1), 13–25 (2003)PubMedCrossRef
48.
Zurück zum Zitat H.H. Arnold, B. Winter, Muscle differentiation: more complexity to the network of myogenic regulators. Curr. Opin. Genet. Dev. 8(5), 539–544 (1998)PubMedCrossRef H.H. Arnold, B. Winter, Muscle differentiation: more complexity to the network of myogenic regulators. Curr. Opin. Genet. Dev. 8(5), 539–544 (1998)PubMedCrossRef
49.
Zurück zum Zitat C. Porcher, E.C. Liao, Y. Fujiwara, L.I. Zon, S.H. Orkin, Specification of hematopoietic and vascular development by the bHLH transcription factor SCL without direct DNA binding. Development. 126(20), 4603–4615 (1999)PubMedCrossRef C. Porcher, E.C. Liao, Y. Fujiwara, L.I. Zon, S.H. Orkin, Specification of hematopoietic and vascular development by the bHLH transcription factor SCL without direct DNA binding. Development. 126(20), 4603–4615 (1999)PubMedCrossRef
50.
Zurück zum Zitat M. Boudjelal, R. Taneja, S. Matsubara, P. Bouillet, P. Dolle, P. Chambon, Overexpression of Stra13, a novel retinoic acid-inducible gene of the basic helix-loop-helix family, inhibits mesodermal and promotes neuronal differentiation of P19 cells. Genes Dev. 11(16), 2052–2065 (1997)PubMedPubMedCentralCrossRef M. Boudjelal, R. Taneja, S. Matsubara, P. Bouillet, P. Dolle, P. Chambon, Overexpression of Stra13, a novel retinoic acid-inducible gene of the basic helix-loop-helix family, inhibits mesodermal and promotes neuronal differentiation of P19 cells. Genes Dev. 11(16), 2052–2065 (1997)PubMedPubMedCentralCrossRef
51.
Zurück zum Zitat S. Azmi, A. Ozog, R. Taneja, Sharp-1/DEC2 inhibits skeletal muscle differentiation through repression of myogenic transcription factors. J. Biol. Chem. 279(50), 52643–52652 (2004)PubMedCrossRef S. Azmi, A. Ozog, R. Taneja, Sharp-1/DEC2 inhibits skeletal muscle differentiation through repression of myogenic transcription factors. J. Biol. Chem. 279(50), 52643–52652 (2004)PubMedCrossRef
52.
Zurück zum Zitat S.R. Shankar, A.G. Bahirvani, V.K. Rao, N. Bharathy, J.R. Ow, R. Taneja, G9a, a multipotent regulator of gene expression. Epigenetics. 8(1), 16–22 (2013)PubMedPubMedCentralCrossRef S.R. Shankar, A.G. Bahirvani, V.K. Rao, N. Bharathy, J.R. Ow, R. Taneja, G9a, a multipotent regulator of gene expression. Epigenetics. 8(1), 16–22 (2013)PubMedPubMedCentralCrossRef
53.
Zurück zum Zitat Y. Wang, S.R. Shankar, D. Kher, B.M. Ling, R. Taneja, Sumoylation of the basic helix-loop-helix transcription factor sharp-1 regulates recruitment of the histone methyltransferase G9a and function in myogenesis. J. Biol. Chem. 288(24), 17654–17662 (2013)PubMedPubMedCentralCrossRef Y. Wang, S.R. Shankar, D. Kher, B.M. Ling, R. Taneja, Sumoylation of the basic helix-loop-helix transcription factor sharp-1 regulates recruitment of the histone methyltransferase G9a and function in myogenesis. J. Biol. Chem. 288(24), 17654–17662 (2013)PubMedPubMedCentralCrossRef
54.
Zurück zum Zitat K. Kunz, K. Wagner, L. Mendler, S. Holper, N. Dehne, S. Muller, SUMO signaling by hypoxic inactivation of SUMO-Specific isopeptidases. Cell. Rep. 16(11), 3075–3086 (2016)PubMedCrossRef K. Kunz, K. Wagner, L. Mendler, S. Holper, N. Dehne, S. Muller, SUMO signaling by hypoxic inactivation of SUMO-Specific isopeptidases. Cell. Rep. 16(11), 3075–3086 (2016)PubMedCrossRef
55.
Zurück zum Zitat S. Acharjee, T.K. Chung, S. Gopinadhan, S.R. Shankar, Y. Wang, L. Li et al., Sharp-1 regulates TGF-beta signaling and skeletal muscle regeneration. J. Cell. Sci. 127(Pt 3), 599–608 (2014)PubMed S. Acharjee, T.K. Chung, S. Gopinadhan, S.R. Shankar, Y. Wang, L. Li et al., Sharp-1 regulates TGF-beta signaling and skeletal muscle regeneration. J. Cell. Sci. 127(Pt 3), 599–608 (2014)PubMed
56.
Zurück zum Zitat N. Marroncelli, M. Bianchi, M. Bertin, S. Consalvi, V. Saccone, De M. Bardi et al., HDAC4 regulates satellite cell proliferation and differentiation by targeting P21 and Sharp1 genes. Sci. Rep. 8(1), 3448 (2018)PubMedPubMedCentralCrossRef N. Marroncelli, M. Bianchi, M. Bertin, S. Consalvi, V. Saccone, De M. Bardi et al., HDAC4 regulates satellite cell proliferation and differentiation by targeting P21 and Sharp1 genes. Sci. Rep. 8(1), 3448 (2018)PubMedPubMedCentralCrossRef
57.
Zurück zum Zitat B. Liu, T. Wang, W. Mei, D. Li, R. Cai, Y. Zuo et al., Small ubiquitin-like modifier (SUMO) protein-specific protease 1 de-SUMOylates Sharp-1 protein and controls adipocyte differentiation. J. Biol. Chem. 289(32), 22358–22364 (2014)PubMedPubMedCentralCrossRef B. Liu, T. Wang, W. Mei, D. Li, R. Cai, Y. Zuo et al., Small ubiquitin-like modifier (SUMO) protein-specific protease 1 de-SUMOylates Sharp-1 protein and controls adipocyte differentiation. J. Biol. Chem. 289(32), 22358–22364 (2014)PubMedPubMedCentralCrossRef
58.
59.
Zurück zum Zitat Y. Makino, N.H. Jensen, N. Yokota, M.J. Rossner, H. Akiyama, K. Shirahige et al., Single cell RNA-sequencing identified Dec2 as a suppressive factor for spermatogonial differentiation by inhibiting Sohlh1 expression. Sci. Rep. 9(1), 6063 (2019)PubMedPubMedCentralCrossRef Y. Makino, N.H. Jensen, N. Yokota, M.J. Rossner, H. Akiyama, K. Shirahige et al., Single cell RNA-sequencing identified Dec2 as a suppressive factor for spermatogonial differentiation by inhibiting Sohlh1 expression. Sci. Rep. 9(1), 6063 (2019)PubMedPubMedCentralCrossRef
60.
Zurück zum Zitat N. Gekakis, D. Staknis, H.B. Nguyen, F.C. Davis, L.D. Wilsbacher, D.P. King et al., Role of the CLOCK protein in the mammalian circadian mechanism. Science. 280(5369), 1564–1569 (1998)PubMedCrossRef N. Gekakis, D. Staknis, H.B. Nguyen, F.C. Davis, L.D. Wilsbacher, D.P. King et al., Role of the CLOCK protein in the mammalian circadian mechanism. Science. 280(5369), 1564–1569 (1998)PubMedCrossRef
61.
Zurück zum Zitat Y. He, C.R. Jones, N. Fujiki, Y. Xu, B. Guo, J.L. Jr. Holder et al., The transcriptional repressor DEC2 regulates sleep length in mammals. Science. 325(5942), 866–870 (2009)PubMedPubMedCentralCrossRef Y. He, C.R. Jones, N. Fujiki, Y. Xu, B. Guo, J.L. Jr. Holder et al., The transcriptional repressor DEC2 regulates sleep length in mammals. Science. 325(5942), 866–870 (2009)PubMedPubMedCentralCrossRef
62.
Zurück zum Zitat R. Pellegrino, I.H. Kavakli, N. Goel, C.J. Cardinale, D.F. Dinges, S.T. Kuna et al., A novel BHLHE41 variant is associated with short sleep and resistance to sleep deprivation in humans. Sleep. 37(8), 1327–1336 (2014)PubMedPubMedCentralCrossRef R. Pellegrino, I.H. Kavakli, N. Goel, C.J. Cardinale, D.F. Dinges, S.T. Kuna et al., A novel BHLHE41 variant is associated with short sleep and resistance to sleep deprivation in humans. Sleep. 37(8), 1327–1336 (2014)PubMedPubMedCentralCrossRef
63.
Zurück zum Zitat M. Noshiro, T. Kawamoto, M. Furukawa, K. Fujimoto, Y. Yoshida, E. Sasabe et al., Rhythmic expression of DEC1 and DEC2 in peripheral tissues: DEC2 is a potent suppressor for hepatic cytochrome P450s opposing DBP. Genes Cells. 9(4), 317–329 (2004)PubMedCrossRef M. Noshiro, T. Kawamoto, M. Furukawa, K. Fujimoto, Y. Yoshida, E. Sasabe et al., Rhythmic expression of DEC1 and DEC2 in peripheral tissues: DEC2 is a potent suppressor for hepatic cytochrome P450s opposing DBP. Genes Cells. 9(4), 317–329 (2004)PubMedCrossRef
64.
Zurück zum Zitat N. Matsunaga, M. Inoue, N. Kusunose, K. Kakimoto, K. Hamamura, Y. Hanada et al., Time-dependent interaction between differentiated embryo chondrocyte-2 and CCAAT/enhancer-binding protein alpha underlies the circadian expression of CYP2D6 in serum-shocked HepG2 cells. Mol. Pharmacol. 81(5), 739–747 (2012)PubMedCrossRef N. Matsunaga, M. Inoue, N. Kusunose, K. Kakimoto, K. Hamamura, Y. Hanada et al., Time-dependent interaction between differentiated embryo chondrocyte-2 and CCAAT/enhancer-binding protein alpha underlies the circadian expression of CYP2D6 in serum-shocked HepG2 cells. Mol. Pharmacol. 81(5), 739–747 (2012)PubMedCrossRef
65.
Zurück zum Zitat F. Sato, Y. Muragaki, T. Kawamoto, K. Fujimoto, Y. Kato, Y. Zhang, Rhythmic expression of DEC2 protein in vitro and in vivo. Biomed. Rep. 4(6), 704–710 (2016)PubMedPubMedCentralCrossRef F. Sato, Y. Muragaki, T. Kawamoto, K. Fujimoto, Y. Kato, Y. Zhang, Rhythmic expression of DEC2 protein in vitro and in vivo. Biomed. Rep. 4(6), 704–710 (2016)PubMedPubMedCentralCrossRef
66.
Zurück zum Zitat J.J. Liu, T.K. Chung, J. Li, R. Taneja, Sharp-1 modulates the cellular response to DNA damage. FEBS Lett. 584(3), 619–624 (2010)PubMedCrossRef J.J. Liu, T.K. Chung, J. Li, R. Taneja, Sharp-1 modulates the cellular response to DNA damage. FEBS Lett. 584(3), 619–624 (2010)PubMedCrossRef
67.
Zurück zum Zitat V.T. Mihaylova, R.S. Bindra, J. Yuan, D. Campisi, L. Narayanan, R. Jensen et al., Decreased expression of the DNA mismatch repair gene Mlh1 under hypoxic stress in mammalian cells. Mol. Cell. Biol. 23(9), 3265–3273 (2003)PubMedPubMedCentralCrossRef V.T. Mihaylova, R.S. Bindra, J. Yuan, D. Campisi, L. Narayanan, R. Jensen et al., Decreased expression of the DNA mismatch repair gene Mlh1 under hypoxic stress in mammalian cells. Mol. Cell. Biol. 23(9), 3265–3273 (2003)PubMedPubMedCentralCrossRef
68.
Zurück zum Zitat R.S. Bindra, P.M. Glazer, Repression of RAD51 gene expression by E2F4/p130 complexes in hypoxia. Oncogene. 26(14), 2048–2057 (2007)PubMedCrossRef R.S. Bindra, P.M. Glazer, Repression of RAD51 gene expression by E2F4/p130 complexes in hypoxia. Oncogene. 26(14), 2048–2057 (2007)PubMedCrossRef
69.
Zurück zum Zitat M. Koshiji, K.K. To, S. Hammer, K. Kumamoto, A.L. Harris, P. Modrich et al., HIF-1alpha induces genetic instability by transcriptionally downregulating MutSalpha expression. Mol. Cell. 17(6), 793–803 (2005)PubMedCrossRef M. Koshiji, K.K. To, S. Hammer, K. Kumamoto, A.L. Harris, P. Modrich et al., HIF-1alpha induces genetic instability by transcriptionally downregulating MutSalpha expression. Mol. Cell. 17(6), 793–803 (2005)PubMedCrossRef
70.
Zurück zum Zitat R.S. Bindra, S.L. Gibson, A. Meng, U. Westermark, M. Jasin, A.J. Pierce et al., Hypoxia-induced down-regulation of BRCA1 expression by E2Fs. Cancer Res. 65(24), 11597–11604 (2005)PubMedCrossRef R.S. Bindra, S.L. Gibson, A. Meng, U. Westermark, M. Jasin, A.J. Pierce et al., Hypoxia-induced down-regulation of BRCA1 expression by E2Fs. Cancer Res. 65(24), 11597–11604 (2005)PubMedCrossRef
71.
Zurück zum Zitat H. Nakamura, H. Bono, K. Hiyama, T. Kawamoto, Y. Kato, T. Nakanishi et al., Differentiated embryo chondrocyte plays a crucial role in DNA damage response via transcriptional regulation under hypoxic conditions. PLoS One. 13(2), e0192136 (2018)PubMedPubMedCentralCrossRef H. Nakamura, H. Bono, K. Hiyama, T. Kawamoto, Y. Kato, T. Nakanishi et al., Differentiated embryo chondrocyte plays a crucial role in DNA damage response via transcriptional regulation under hypoxic conditions. PLoS One. 13(2), e0192136 (2018)PubMedPubMedCentralCrossRef
72.
Zurück zum Zitat H. Nakamura, K. Tanimoto, K. Hiyama, M. Yunokawa, T. Kawamoto, Y. Kato et al., Human mismatch repair gene, MLH1, is transcriptionally repressed by the hypoxia-inducible transcription factors, DEC1 and DEC2. Oncogene. 27(30), 4200–4209 (2008)PubMedCrossRef H. Nakamura, K. Tanimoto, K. Hiyama, M. Yunokawa, T. Kawamoto, Y. Kato et al., Human mismatch repair gene, MLH1, is transcriptionally repressed by the hypoxia-inducible transcription factors, DEC1 and DEC2. Oncogene. 27(30), 4200–4209 (2008)PubMedCrossRef
73.
Zurück zum Zitat M.F. Pittenger, A.M. Mackay, S.C. Beck, R.K. Jaiswal, R. Douglas, J.D. Mosca et al., Multilineage potential of adult human mesenchymal stem cells. Science. 284(5411), 143–147 (1999)PubMedCrossRef M.F. Pittenger, A.M. Mackay, S.C. Beck, R.K. Jaiswal, R. Douglas, J.D. Mosca et al., Multilineage potential of adult human mesenchymal stem cells. Science. 284(5411), 143–147 (1999)PubMedCrossRef
74.
Zurück zum Zitat Y. Lavin, A. Mortha, A. Rahman, M. Merad, Regulation of macrophage development and function in peripheral tissues. Nat. Rev. Immunol. 15(12), 731–744 (2015)PubMedPubMedCentralCrossRef Y. Lavin, A. Mortha, A. Rahman, M. Merad, Regulation of macrophage development and function in peripheral tissues. Nat. Rev. Immunol. 15(12), 731–744 (2015)PubMedPubMedCentralCrossRef
75.
Zurück zum Zitat R. Rauschmeier, C. Gustafsson, A. Reinhardt, A.G. N, L. Tortola, D. Cansever et al., Bhlhe40 and Bhlhe41 transcription factors regulate alveolar macrophage self-renewal and identity. EMBO J. 38(19), e101233 (2019)PubMedPubMedCentralCrossRef R. Rauschmeier, C. Gustafsson, A. Reinhardt, A.G. N, L. Tortola, D. Cansever et al., Bhlhe40 and Bhlhe41 transcription factors regulate alveolar macrophage self-renewal and identity. EMBO J. 38(19), e101233 (2019)PubMedPubMedCentralCrossRef
76.
Zurück zum Zitat Y. Xu, K. Jiang, F. Su, R. Deng, Z. Cheng, D. Wang et al., A transient wave of Bhlhe41(+) resident macrophages enables remodeling of the developing infarcted myocardium. Cell. Rep. 42(10), 113174 (2023)PubMedCrossRef Y. Xu, K. Jiang, F. Su, R. Deng, Z. Cheng, D. Wang et al., A transient wave of Bhlhe41(+) resident macrophages enables remodeling of the developing infarcted myocardium. Cell. Rep. 42(10), 113174 (2023)PubMedCrossRef
77.
Zurück zum Zitat J.F. Piskurich, K.I. Lin, Y. Lin, Y. Wang, J.P. Ting, K. Calame, BLIMP-I mediates extinction of major histocompatibility class II transactivator expression in plasma cells. Nat. Immunol. 1(6), 526–532 (2000)PubMedCrossRef J.F. Piskurich, K.I. Lin, Y. Lin, Y. Wang, J.P. Ting, K. Calame, BLIMP-I mediates extinction of major histocompatibility class II transactivator expression in plasma cells. Nat. Immunol. 1(6), 526–532 (2000)PubMedCrossRef
78.
Zurück zum Zitat A.L. Shaffer, K.I. Lin, T.C. Kuo, X. Yu, E.M. Hurt, A. Rosenwald et al., Blimp-1 orchestrates plasma cell differentiation by extinguishing the mature B cell gene expression program. Immunity. 17(1), 51–62 (2002)PubMedCrossRef A.L. Shaffer, K.I. Lin, T.C. Kuo, X. Yu, E.M. Hurt, A. Rosenwald et al., Blimp-1 orchestrates plasma cell differentiation by extinguishing the mature B cell gene expression program. Immunity. 17(1), 51–62 (2002)PubMedCrossRef
79.
Zurück zum Zitat Y. Lin, K. Wong, K. Calame, Repression of c-myc transcription by Blimp-1, an inducer of terminal B cell differentiation. Science. 276(5312), 596–599 (1997)PubMedCrossRef Y. Lin, K. Wong, K. Calame, Repression of c-myc transcription by Blimp-1, an inducer of terminal B cell differentiation. Science. 276(5312), 596–599 (1997)PubMedCrossRef
80.
Zurück zum Zitat K. Santamaria, F. Desmots, S. Leonard, G. Caron, M. Haas, C. Delaloy et al., Committed human CD23-Negative light-zone Germinal Center B cells delineate transcriptional program supporting plasma cell differentiation. Front. Immunol. 12, 744573 (2021)PubMedPubMedCentralCrossRef K. Santamaria, F. Desmots, S. Leonard, G. Caron, M. Haas, C. Delaloy et al., Committed human CD23-Negative light-zone Germinal Center B cells delineate transcriptional program supporting plasma cell differentiation. Front. Immunol. 12, 744573 (2021)PubMedPubMedCentralCrossRef
81.
Zurück zum Zitat A.B. Holmes, C. Corinaldesi, Q. Shen, R. Kumar, N. Compagno, Z. Wang et al., Single-cell analysis of germinal-center B cells informs on lymphoma cell of origin and outcome. J. Exp. Med. 2020;217(10) A.B. Holmes, C. Corinaldesi, Q. Shen, R. Kumar, N. Compagno, Z. Wang et al., Single-cell analysis of germinal-center B cells informs on lymphoma cell of origin and outcome. J. Exp. Med. 2020;217(10)
82.
Zurück zum Zitat H.W. King, N. Orban, J.C. Riches, A.J. Clear, G. Warnes, S.A. Teichmann et al., Single-cell analysis of human B cell maturation predicts how antibody class switching shapes selection dynamics. Sci. Immunol. 2021;6(56) H.W. King, N. Orban, J.C. Riches, A.J. Clear, G. Warnes, S.A. Teichmann et al., Single-cell analysis of human B cell maturation predicts how antibody class switching shapes selection dynamics. Sci. Immunol. 2021;6(56)
83.
Zurück zum Zitat J.S. He, S. Subramaniam, V. Narang, K. Srinivasan, S.P. Saunders, D. Carbajo et al., IgG1 memory B cells keep the memory of IgE responses. Nat. Commun. 8(1), 641 (2017)PubMedPubMedCentralCrossRef J.S. He, S. Subramaniam, V. Narang, K. Srinivasan, S.P. Saunders, D. Carbajo et al., IgG1 memory B cells keep the memory of IgE responses. Nat. Commun. 8(1), 641 (2017)PubMedPubMedCentralCrossRef
84.
Zurück zum Zitat B.J. Laidlaw, L. Duan, Y. Xu, S.E. Vazquez, J.G. Cyster, The transcription factor Hhex cooperates with the corepressor Tle3 to promote memory B cell development. Nat. Immunol. 21(9), 1082–1093 (2020)PubMedPubMedCentralCrossRef B.J. Laidlaw, L. Duan, Y. Xu, S.E. Vazquez, J.G. Cyster, The transcription factor Hhex cooperates with the corepressor Tle3 to promote memory B cell development. Nat. Immunol. 21(9), 1082–1093 (2020)PubMedPubMedCentralCrossRef
85.
Zurück zum Zitat V. Glaros, R. Rauschmeier, A.V. Artemov, A. Reinhardt, S. Ols, A. Emmanouilidi et al., Limited access to antigen drives generation of early B cell memory while restraining the plasmablast response. Immunity. 54(9), 2005–2023 (2021). e10PubMedPubMedCentralCrossRef V. Glaros, R. Rauschmeier, A.V. Artemov, A. Reinhardt, S. Ols, A. Emmanouilidi et al., Limited access to antigen drives generation of early B cell memory while restraining the plasmablast response. Immunity. 54(9), 2005–2023 (2021). e10PubMedPubMedCentralCrossRef
86.
Zurück zum Zitat T. Kreslavsky, B. Vilagos, H. Tagoh, D.K. Poliakova, T.A. Schwickert, M. Wohner et al., Essential role for the transcription factor Bhlhe41 in regulating the development, self-renewal and BCR repertoire of B-1a cells. Nat. Immunol. 18(4), 442–455 (2017)PubMedPubMedCentralCrossRef T. Kreslavsky, B. Vilagos, H. Tagoh, D.K. Poliakova, T.A. Schwickert, M. Wohner et al., Essential role for the transcription factor Bhlhe41 in regulating the development, self-renewal and BCR repertoire of B-1a cells. Nat. Immunol. 18(4), 442–455 (2017)PubMedPubMedCentralCrossRef
87.
Zurück zum Zitat K. Hayakawa, Y.S. Li, S.A. Shinton, S.R. Bandi, A.M. Formica, J. Brill-Dashoff et al., Crucial role of increased Arid3a at the Pre-B and immature B cell stages for B1a cell generation. Front. Immunol. 10, 457 (2019)PubMedPubMedCentralCrossRef K. Hayakawa, Y.S. Li, S.A. Shinton, S.R. Bandi, A.M. Formica, J. Brill-Dashoff et al., Crucial role of increased Arid3a at the Pre-B and immature B cell stages for B1a cell generation. Front. Immunol. 10, 457 (2019)PubMedPubMedCentralCrossRef
88.
Zurück zum Zitat K.M. Murphy, S.L. Reiner, The lineage decisions of helper T cells. Nat. Rev. Immunol. 2(12), 933–944 (2002)PubMedCrossRef K.M. Murphy, S.L. Reiner, The lineage decisions of helper T cells. Nat. Rev. Immunol. 2(12), 933–944 (2002)PubMedCrossRef
89.
Zurück zum Zitat J. Cote-Sierra, G. Foucras, L. Guo, L. Chiodetti, H.A. Young, J. Hu-Li et al., Interleukin 2 plays a central role in Th2 differentiation. Proc. Natl. Acad. Sci. U S A 101(11), 3880–3885 (2004)PubMedPubMedCentralCrossRef J. Cote-Sierra, G. Foucras, L. Guo, L. Chiodetti, H.A. Young, J. Hu-Li et al., Interleukin 2 plays a central role in Th2 differentiation. Proc. Natl. Acad. Sci. U S A 101(11), 3880–3885 (2004)PubMedPubMedCentralCrossRef
91.
Zurück zum Zitat Z. Liu, Z. Li, K. Mao, J. Zou, Y. Wang, Z. Tao et al., Dec2 promotes Th2 cell differentiation by enhancing IL-2R signaling. J. Immunol. 183(10), 6320–6329 (2009)PubMedCrossRef Z. Liu, Z. Li, K. Mao, J. Zou, Y. Wang, Z. Tao et al., Dec2 promotes Th2 cell differentiation by enhancing IL-2R signaling. J. Immunol. 183(10), 6320–6329 (2009)PubMedCrossRef
92.
Zurück zum Zitat H. Qi, Q. Cao, Q. Liu, MicroRNA-16 directly binds to DEC2 and inactivates the TLR4 signaling pathway to inhibit lupus nephritis-induced kidney tissue hyperplasia and mesangial cell proliferation. Int. Immunopharmacol. 88, 106859 (2020)PubMedCrossRef H. Qi, Q. Cao, Q. Liu, MicroRNA-16 directly binds to DEC2 and inactivates the TLR4 signaling pathway to inhibit lupus nephritis-induced kidney tissue hyperplasia and mesangial cell proliferation. Int. Immunopharmacol. 88, 106859 (2020)PubMedCrossRef
93.
Zurück zum Zitat H. Chen, Y. Pan, Q. Zhou, C. Liang, C.C. Wong, Y. Zhou et al., METTL3 inhibits Antitumor Immunity by Targeting m(6)A-BHLHE41-CXCL1/CXCR2 Axis to promote Colorectal Cancer. Gastroenterology. 163(4), 891–907 (2022)PubMedCrossRef H. Chen, Y. Pan, Q. Zhou, C. Liang, C.C. Wong, Y. Zhou et al., METTL3 inhibits Antitumor Immunity by Targeting m(6)A-BHLHE41-CXCL1/CXCR2 Axis to promote Colorectal Cancer. Gastroenterology. 163(4), 891–907 (2022)PubMedCrossRef
94.
Zurück zum Zitat Y. Wang, V.K. Rao, W.K. Kok, D.N. Roy, S. Sethi, B.M. Ling et al., SUMO modification of Stra13 is required for repression of cyclin D1 expression and cellular growth arrest. PLoS One. 7(8), e43137 (2012)PubMedPubMedCentralCrossRef Y. Wang, V.K. Rao, W.K. Kok, D.N. Roy, S. Sethi, B.M. Ling et al., SUMO modification of Stra13 is required for repression of cyclin D1 expression and cellular growth arrest. PLoS One. 7(8), e43137 (2012)PubMedPubMedCentralCrossRef
95.
Zurück zum Zitat A. Emami Nejad, S. Najafgholian, A. Rostami, A. Sistani, S. Shojaeifar, M. Esparvarinha et al., The role of hypoxia in the tumor microenvironment and development of cancer stem cell: a novel approach to developing treatment. Cancer Cell. Int. 21(1), 62 (2021)PubMedPubMedCentralCrossRef A. Emami Nejad, S. Najafgholian, A. Rostami, A. Sistani, S. Shojaeifar, M. Esparvarinha et al., The role of hypoxia in the tumor microenvironment and development of cancer stem cell: a novel approach to developing treatment. Cancer Cell. Int. 21(1), 62 (2021)PubMedPubMedCentralCrossRef
96.
Zurück zum Zitat M. Yunokawa, K. Tanimoto, H. Nakamura, N. Nagai, Y. Kudo, T. Kawamoto et al., Differential regulation of DEC2 among hypoxia-inducible genes in endometrial carcinomas. Oncol. Rep. 17(4), 871–878 (2007)PubMed M. Yunokawa, K. Tanimoto, H. Nakamura, N. Nagai, Y. Kudo, T. Kawamoto et al., Differential regulation of DEC2 among hypoxia-inducible genes in endometrial carcinomas. Oncol. Rep. 17(4), 871–878 (2007)PubMed
97.
Zurück zum Zitat Y. Liao, X. He, H. Qiu, Q. Che, F. Wang, W. Lu et al., Suppression of the epithelial-mesenchymal transition by SHARP1 is linked to the NOTCH1 signaling pathway in metastasis of endometrial cancer. BMC Cancer. 14, 487 (2014)PubMedPubMedCentralCrossRef Y. Liao, X. He, H. Qiu, Q. Che, F. Wang, W. Lu et al., Suppression of the epithelial-mesenchymal transition by SHARP1 is linked to the NOTCH1 signaling pathway in metastasis of endometrial cancer. BMC Cancer. 14, 487 (2014)PubMedPubMedCentralCrossRef
98.
Zurück zum Zitat Y. Liao, W. Lu, Q. Che, T. Yang, H. Qiu, H. Zhang et al., SHARP1 suppresses angiogenesis of endometrial cancer by decreasing hypoxia-inducible factor-1alpha level. PLoS One. 9(6), e99907 (2014)PubMedPubMedCentralCrossRef Y. Liao, W. Lu, Q. Che, T. Yang, H. Qiu, H. Zhang et al., SHARP1 suppresses angiogenesis of endometrial cancer by decreasing hypoxia-inducible factor-1alpha level. PLoS One. 9(6), e99907 (2014)PubMedPubMedCentralCrossRef
99.
Zurück zum Zitat K. Asanoma, G. Liu, T. Yamane, Y. Miyanari, T. Takao, H. Yagi et al., Regulation of the mechanism of TWIST1 transcription by BHLHE40 and BHLHE41 in Cancer cells. Mol. Cell. Biol. 35(24), 4096–4109 (2015)PubMedPubMedCentralCrossRef K. Asanoma, G. Liu, T. Yamane, Y. Miyanari, T. Takao, H. Yagi et al., Regulation of the mechanism of TWIST1 transcription by BHLHE40 and BHLHE41 in Cancer cells. Mol. Cell. Biol. 35(24), 4096–4109 (2015)PubMedPubMedCentralCrossRef
100.
Zurück zum Zitat K. Asanoma, E. Hori, S. Yoshida, H. Yagi, I. Onoyama, K. Kodama et al., Mutual suppression between BHLHE40/BHLHE41 and the MIR301B-MIR130B cluster is involved in epithelial-to-mesenchymal transition of endometrial cancer cells. Oncotarget. 10(45), 4640–4654 (2019)PubMedPubMedCentralCrossRef K. Asanoma, E. Hori, S. Yoshida, H. Yagi, I. Onoyama, K. Kodama et al., Mutual suppression between BHLHE40/BHLHE41 and the MIR301B-MIR130B cluster is involved in epithelial-to-mesenchymal transition of endometrial cancer cells. Oncotarget. 10(45), 4640–4654 (2019)PubMedPubMedCentralCrossRef
101.
Zurück zum Zitat Y. Liu, F. Sato, T. Kawamoto, K. Fujimoto, S. Morohashi, H. Akasaka et al., Anti-apoptotic effect of the basic helix-loop-helix (bHLH) transcription factor DEC2 in human breast cancer cells. Genes Cells. 15(4), 315–325 (2010)PubMedCrossRef Y. Liu, F. Sato, T. Kawamoto, K. Fujimoto, S. Morohashi, H. Akasaka et al., Anti-apoptotic effect of the basic helix-loop-helix (bHLH) transcription factor DEC2 in human breast cancer cells. Genes Cells. 15(4), 315–325 (2010)PubMedCrossRef
102.
Zurück zum Zitat Y. Wu, F. Sato, U.K. Bhawal, T. Kawamoto, K. Fujimoto, M. Noshiro et al., Basic helix-loop-helix transcription factors DEC1 and DEC2 regulate the paclitaxel-induced apoptotic pathway of MCF-7 human breast cancer cells. Int. J. Mol. Med. 27(4), 491–495 (2011)PubMed Y. Wu, F. Sato, U.K. Bhawal, T. Kawamoto, K. Fujimoto, M. Noshiro et al., Basic helix-loop-helix transcription factors DEC1 and DEC2 regulate the paclitaxel-induced apoptotic pathway of MCF-7 human breast cancer cells. Int. J. Mol. Med. 27(4), 491–495 (2011)PubMed
103.
Zurück zum Zitat D. Zhang, Q. Zheng, C. Wang, N. Zhao, Y. Liu, E. Wang, BHLHE41 suppresses MCF-7 cell invasion via MAPK/JNK pathway. J. Cell. Mol. Med. 24(7), 4001–4010 (2020)PubMedPubMedCentralCrossRef D. Zhang, Q. Zheng, C. Wang, N. Zhao, Y. Liu, E. Wang, BHLHE41 suppresses MCF-7 cell invasion via MAPK/JNK pathway. J. Cell. Mol. Med. 24(7), 4001–4010 (2020)PubMedPubMedCentralCrossRef
104.
Zurück zum Zitat Y. Wu, H. Sato, T. Suzuki, T. Yoshizawa, S. Morohashi, H. Seino et al., Involvement of c-Myc in the proliferation of MCF-7 human breast cancer cells induced by bHLH transcription factor DEC2. Int. J. Mol. Med. 35(3), 815–820 (2015)PubMedCrossRef Y. Wu, H. Sato, T. Suzuki, T. Yoshizawa, S. Morohashi, H. Seino et al., Involvement of c-Myc in the proliferation of MCF-7 human breast cancer cells induced by bHLH transcription factor DEC2. Int. J. Mol. Med. 35(3), 815–820 (2015)PubMedCrossRef
105.
Zurück zum Zitat Y. Li, Q. Shen, H.T. Kim, R.P. Bissonnette, W.W. Lamph, B. Yan et al., The rexinoid bexarotene represses cyclin D1 transcription by inducing the DEC2 transcriptional repressor. Breast Cancer Res. Treat. 128(3), 667–677 (2011)PubMedCrossRef Y. Li, Q. Shen, H.T. Kim, R.P. Bissonnette, W.W. Lamph, B. Yan et al., The rexinoid bexarotene represses cyclin D1 transcription by inducing the DEC2 transcriptional repressor. Breast Cancer Res. Treat. 128(3), 667–677 (2011)PubMedCrossRef
106.
Zurück zum Zitat R.S. Kim, A. Avivar-Valderas, Y. Estrada, P. Bragado, M.S. Sosa, J.A. Aguirre-Ghiso et al., Dormancy signatures and metastasis in estrogen receptor positive and negative breast cancer. PLoS One. 7(4), e35569 (2012)PubMedPubMedCentralCrossRef R.S. Kim, A. Avivar-Valderas, Y. Estrada, P. Bragado, M.S. Sosa, J.A. Aguirre-Ghiso et al., Dormancy signatures and metastasis in estrogen receptor positive and negative breast cancer. PLoS One. 7(4), e35569 (2012)PubMedPubMedCentralCrossRef
107.
108.
Zurück zum Zitat M. Fu, C. Wang, Z. Li, T. Sakamaki, R.G. Pestell, Minireview, Cyclin D1: normal and abnormal functions. Endocrinology. 145(12), 5439–5447 (2004)PubMedCrossRef M. Fu, C. Wang, Z. Li, T. Sakamaki, R.G. Pestell, Minireview, Cyclin D1: normal and abnormal functions. Endocrinology. 145(12), 5439–5447 (2004)PubMedCrossRef
109.
Zurück zum Zitat Y. Wu, F. Sato, U.K. Bhawal, T. Kawamoto, K. Fujimoto, M. Noshiro et al., BHLH transcription factor DEC2 regulates pro-apoptotic factor Bim in human oral cancer HSC-3 cells. Biomed. Res. 33(2), 75–82 (2012)PubMedCrossRef Y. Wu, F. Sato, U.K. Bhawal, T. Kawamoto, K. Fujimoto, M. Noshiro et al., BHLH transcription factor DEC2 regulates pro-apoptotic factor Bim in human oral cancer HSC-3 cells. Biomed. Res. 33(2), 75–82 (2012)PubMedCrossRef
110.
Zurück zum Zitat H. Sato, Y. Wu, Y. Kato, Q. Liu, H. Hirai, T. Yoshizawa et al., DEC2 expression antagonizes cisplatin–induced apoptosis in human esophageal squamous cell carcinoma. Mol. Med. Rep. 16(1), 43–48 (2017)PubMedPubMedCentralCrossRef H. Sato, Y. Wu, Y. Kato, Q. Liu, H. Hirai, T. Yoshizawa et al., DEC2 expression antagonizes cisplatin–induced apoptosis in human esophageal squamous cell carcinoma. Mol. Med. Rep. 16(1), 43–48 (2017)PubMedPubMedCentralCrossRef
111.
Zurück zum Zitat X. Yang, J.S. Wu, M. Li, W.L. Zhang, X.L. Gao, H.F. Wang et al., Inhibition of DEC2 is necessary for exiting cell dormancy in salivary adenoid cystic carcinoma. J. Exp. Clin. Cancer Res. 40(1), 169 (2021)PubMedPubMedCentralCrossRef X. Yang, J.S. Wu, M. Li, W.L. Zhang, X.L. Gao, H.F. Wang et al., Inhibition of DEC2 is necessary for exiting cell dormancy in salivary adenoid cystic carcinoma. J. Exp. Clin. Cancer Res. 40(1), 169 (2021)PubMedPubMedCentralCrossRef
112.
Zurück zum Zitat P. Li, Y.F. Jia, X.L. Ma, Y. Zheng, Y. Kong, Y. Zhang et al., DEC2 suppresses tumor proliferation and metastasis by regulating ERK/NF-kappaB pathway in gastric cancer. Am. J. Cancer Res. 6(8), 1741–1757 (2016)PubMedPubMedCentralCrossRef P. Li, Y.F. Jia, X.L. Ma, Y. Zheng, Y. Kong, Y. Zhang et al., DEC2 suppresses tumor proliferation and metastasis by regulating ERK/NF-kappaB pathway in gastric cancer. Am. J. Cancer Res. 6(8), 1741–1757 (2016)PubMedPubMedCentralCrossRef
113.
Zurück zum Zitat H. Li, X. Ma, D. Xiao, Y. Jia, Y. Wang, Expression of DEC2 enhances chemosensitivity by inhibiting STAT5A in gastric cancer. J. Cell. Biochem. 120(5), 8447–8456 (2019)PubMedCrossRef H. Li, X. Ma, D. Xiao, Y. Jia, Y. Wang, Expression of DEC2 enhances chemosensitivity by inhibiting STAT5A in gastric cancer. J. Cell. Biochem. 120(5), 8447–8456 (2019)PubMedCrossRef
114.
Zurück zum Zitat S. Chen, Q.J. Dong, Z.A. Wan, S. Gao, S.L. Tu, R. Chai, BHLHE41 overexpression alleviates the malignant behavior of Colon Cancer cells Induced by Hypoxia via modulating HIF-1alpha/EMT pathway. Gastroenterol. Res. Pract. 2022, 6972331 (2022)PubMedPubMedCentralCrossRef S. Chen, Q.J. Dong, Z.A. Wan, S. Gao, S.L. Tu, R. Chai, BHLHE41 overexpression alleviates the malignant behavior of Colon Cancer cells Induced by Hypoxia via modulating HIF-1alpha/EMT pathway. Gastroenterol. Res. Pract. 2022, 6972331 (2022)PubMedPubMedCentralCrossRef
115.
Zurück zum Zitat F. Sato, H. Kawamura, Y. Wu, H. Sato, D. Jin, U.K. Bhawal et al., The basic helix-loop-helix transcription factor DEC2 inhibits TGF-beta-induced tumor progression in human pancreatic cancer BxPC-3 cells. Int. J. Mol. Med. 30(3), 495–501 (2012)PubMedCrossRef F. Sato, H. Kawamura, Y. Wu, H. Sato, D. Jin, U.K. Bhawal et al., The basic helix-loop-helix transcription factor DEC2 inhibits TGF-beta-induced tumor progression in human pancreatic cancer BxPC-3 cells. Int. J. Mol. Med. 30(3), 495–501 (2012)PubMedCrossRef
116.
Zurück zum Zitat F.S. Falvella, F. Colombo, M. Spinola, M. Campiglio, U. Pastorino, T.A. Dragani, BHLHB3: a candidate tumor suppressor in lung cancer. Oncogene. 27(26), 3761–3764 (2008)PubMedCrossRef F.S. Falvella, F. Colombo, M. Spinola, M. Campiglio, U. Pastorino, T.A. Dragani, BHLHB3: a candidate tumor suppressor in lung cancer. Oncogene. 27(26), 3761–3764 (2008)PubMedCrossRef
117.
Zurück zum Zitat T. Nagata, K. Minami, M. Yamamoto, T. Hiraki, M. Idogawa, K. Fujimoto et al., BHLHE41/DEC2 expression induces autophagic cell death in Lung Cancer cells and is Associated with favorable prognosis for patients with Lung Adenocarcinoma. Int. J. Mol. Sci. 2021;22(21) T. Nagata, K. Minami, M. Yamamoto, T. Hiraki, M. Idogawa, K. Fujimoto et al., BHLHE41/DEC2 expression induces autophagic cell death in Lung Cancer cells and is Associated with favorable prognosis for patients with Lung Adenocarcinoma. Int. J. Mol. Sci. 2021;22(21)
118.
Zurück zum Zitat T. Hu, N. He, Y. Yang, C. Yin, N. Sang, Q. Yang, DEC2 expression is positively correlated with HIF-1 activation and the invasiveness of human osteosarcomas. J. Exp. Clin. Cancer Res. 34(1), 22 (2015)PubMedPubMedCentralCrossRef T. Hu, N. He, Y. Yang, C. Yin, N. Sang, Q. Yang, DEC2 expression is positively correlated with HIF-1 activation and the invasiveness of human osteosarcomas. J. Exp. Clin. Cancer Res. 34(1), 22 (2015)PubMedPubMedCentralCrossRef
119.
Zurück zum Zitat B. Jiang, W. Mu, J. Wang, J. Lu, S. Jiang, L. Li et al., MicroRNA-138 functions as a tumor suppressor in osteosarcoma by targeting differentiated embryonic chondrocyte gene 2. J. Exp. Clin. Cancer Res. 35, 69 (2016)PubMedPubMedCentralCrossRef B. Jiang, W. Mu, J. Wang, J. Lu, S. Jiang, L. Li et al., MicroRNA-138 functions as a tumor suppressor in osteosarcoma by targeting differentiated embryonic chondrocyte gene 2. J. Exp. Clin. Cancer Res. 35, 69 (2016)PubMedPubMedCentralCrossRef
120.
Zurück zum Zitat Q. Liu, Y. Wu, H. Seino, T. Haga, T. Yoshizawa, S. Morohashi et al., Correlation between DEC1/DEC2 and epithelial–mesenchymal transition in human prostate cancer PC–3 cells. Mol. Med. Rep. 18(4), 3859–3865 (2018)PubMedPubMedCentral Q. Liu, Y. Wu, H. Seino, T. Haga, T. Yoshizawa, S. Morohashi et al., Correlation between DEC1/DEC2 and epithelial–mesenchymal transition in human prostate cancer PC–3 cells. Mol. Med. Rep. 18(4), 3859–3865 (2018)PubMedPubMedCentral
121.
Zurück zum Zitat Q. Liu, Y. Wu, T. Yoshizawa, X. Yan, S. Morohashi, H. Seino et al., Basic helix-loop-helix transcription factor DEC2 functions as an anti-apoptotic factor during paclitaxel-induced apoptosis in human prostate cancer cells. Int. J. Mol. Med. 38(6), 1727–1733 (2016)PubMedPubMedCentralCrossRef Q. Liu, Y. Wu, T. Yoshizawa, X. Yan, S. Morohashi, H. Seino et al., Basic helix-loop-helix transcription factor DEC2 functions as an anti-apoptotic factor during paclitaxel-induced apoptosis in human prostate cancer cells. Int. J. Mol. Med. 38(6), 1727–1733 (2016)PubMedPubMedCentralCrossRef
122.
Zurück zum Zitat Z. Shen, L. Zhu, C. Zhang, X. Cui, J. Lu, Overexpression of BHLHE41, correlated with DNA hypomethylation in 3’UTR region, promotes the growth of human clear cell renal cell carcinoma. Oncol. Rep. 41(4), 2137–2147 (2019)PubMedPubMedCentral Z. Shen, L. Zhu, C. Zhang, X. Cui, J. Lu, Overexpression of BHLHE41, correlated with DNA hypomethylation in 3’UTR region, promotes the growth of human clear cell renal cell carcinoma. Oncol. Rep. 41(4), 2137–2147 (2019)PubMedPubMedCentral
123.
Zurück zum Zitat N. Apanovich, P. Apanovich, D. Mansorunov, A. Kuzevanova, V. Matveev, A. Karpukhin, The choice of candidates in survival markers based on coordinated gene expression in Renal Cancer. Front. Oncol. 11, 615787 (2021)PubMedPubMedCentralCrossRef N. Apanovich, P. Apanovich, D. Mansorunov, A. Kuzevanova, V. Matveev, A. Karpukhin, The choice of candidates in survival markers based on coordinated gene expression in Renal Cancer. Front. Oncol. 11, 615787 (2021)PubMedPubMedCentralCrossRef
124.
Zurück zum Zitat P. Bigot, L.M. Colli, M.J. Machiela, L. Jessop, T.A. Myers, J. Carrouget et al., Functional characterization of the 12p12.1 renal cancer-susceptibility locus implicates BHLHE41. Nat. Commun. 7, 12098 (2016)PubMedPubMedCentralCrossRef P. Bigot, L.M. Colli, M.J. Machiela, L. Jessop, T.A. Myers, J. Carrouget et al., Functional characterization of the 12p12.1 renal cancer-susceptibility locus implicates BHLHE41. Nat. Commun. 7, 12098 (2016)PubMedPubMedCentralCrossRef
125.
Zurück zum Zitat T. Furukawa, K. Mimami, T. Nagata, M. Yamamoto, M. Sato, A. Tanimoto, Approach to functions of BHLHE41/DEC2 in Non-small Lung Cancer Development. Int. J. Mol. Sci. 2023;24(14) T. Furukawa, K. Mimami, T. Nagata, M. Yamamoto, M. Sato, A. Tanimoto, Approach to functions of BHLHE41/DEC2 in Non-small Lung Cancer Development. Int. J. Mol. Sci. 2023;24(14)
126.
Zurück zum Zitat C. Wang, N. Zhao, Q. Zheng, D. Zhang, Y. Liu, BHLHE41 promotes U87 and U251 cell proliferation via ERK/cyclinD1 signaling pathway. Cancer Manag Res. 11, 7657–7672 (2019)PubMedPubMedCentralCrossRef C. Wang, N. Zhao, Q. Zheng, D. Zhang, Y. Liu, BHLHE41 promotes U87 and U251 cell proliferation via ERK/cyclinD1 signaling pathway. Cancer Manag Res. 11, 7657–7672 (2019)PubMedPubMedCentralCrossRef
127.
Zurück zum Zitat K.M. Bernt, N. Zhu, A.U. Sinha, S. Vempati, J. Faber, A.V. Krivtsov et al., MLL-rearranged leukemia is dependent on aberrant H3K79 methylation by DOT1L. Cancer Cell. 20(1), 66–78 (2011)PubMedPubMedCentralCrossRef K.M. Bernt, N. Zhu, A.U. Sinha, S. Vempati, J. Faber, A.V. Krivtsov et al., MLL-rearranged leukemia is dependent on aberrant H3K79 methylation by DOT1L. Cancer Cell. 20(1), 66–78 (2011)PubMedPubMedCentralCrossRef
128.
Zurück zum Zitat M. Liedtke, P.M. Ayton, T.C. Somervaille, K.S. Smith, M.L. Cleary, Self-association mediated by the Ras Association 1 domain of AF6 activates the oncogenic potential of MLL-AF6. Blood. 116(1), 63–70 (2010)PubMedPubMedCentralCrossRef M. Liedtke, P.M. Ayton, T.C. Somervaille, K.S. Smith, M.L. Cleary, Self-association mediated by the Ras Association 1 domain of AF6 activates the oncogenic potential of MLL-AF6. Blood. 116(1), 63–70 (2010)PubMedPubMedCentralCrossRef
129.
Zurück zum Zitat A.J. Deshpande, L. Chen, M. Fazio, A.U. Sinha, K.M. Bernt, D. Banka et al., Leukemic transformation by the MLL-AF6 fusion oncogene requires the H3K79 methyltransferase Dot1l. Blood. 121(13), 2533–2541 (2013)PubMedPubMedCentralCrossRef A.J. Deshpande, L. Chen, M. Fazio, A.U. Sinha, K.M. Bernt, D. Banka et al., Leukemic transformation by the MLL-AF6 fusion oncogene requires the H3K79 methyltransferase Dot1l. Blood. 121(13), 2533–2541 (2013)PubMedPubMedCentralCrossRef
130.
Zurück zum Zitat A. Numata, H.S. Kwok, A. Kawasaki, J. Li, Q.L. Zhou, J. Kerry et al., The basic helix-loop-helix transcription factor SHARP1 is an oncogenic driver in MLL-AF6 acute myelogenous leukemia. Nat. Commun. 9(1), 1622 (2018)PubMedPubMedCentralCrossRef A. Numata, H.S. Kwok, A. Kawasaki, J. Li, Q.L. Zhou, J. Kerry et al., The basic helix-loop-helix transcription factor SHARP1 is an oncogenic driver in MLL-AF6 acute myelogenous leukemia. Nat. Commun. 9(1), 1622 (2018)PubMedPubMedCentralCrossRef
131.
Zurück zum Zitat De J. Vos, D. Hose, T. Reme, K. Tarte, J. Moreaux, K. Mahtouk et al., Microarray-based understanding of normal and malignant plasma cells. Immunol. Rev. 210, 86–104 (2006)PubMedPubMedCentralCrossRef De J. Vos, D. Hose, T. Reme, K. Tarte, J. Moreaux, K. Mahtouk et al., Microarray-based understanding of normal and malignant plasma cells. Immunol. Rev. 210, 86–104 (2006)PubMedPubMedCentralCrossRef
132.
Zurück zum Zitat N. Lee, S.M. Kim, Y. Lee, D. Jeong, J. Yun, S. Ryu et al., Prognostic value of integrated cytogenetic, somatic variation, and copy number variation analyses in Korean patients with newly diagnosed multiple myeloma. PLoS One. 16(2), e0246322 (2021)PubMedPubMedCentralCrossRef N. Lee, S.M. Kim, Y. Lee, D. Jeong, J. Yun, S. Ryu et al., Prognostic value of integrated cytogenetic, somatic variation, and copy number variation analyses in Korean patients with newly diagnosed multiple myeloma. PLoS One. 16(2), e0246322 (2021)PubMedPubMedCentralCrossRef
133.
Zurück zum Zitat A. Trojani, A. Greco, A. Tedeschi, M. Lodola, Di B. Camillo, F. Ricci et al., Microarray demonstrates different gene expression profiling signatures between Waldenstrom macroglobulinemia and IgM monoclonal gammopathy of undetermined significance. Clin. Lymphoma Myeloma Leuk. 13(2), 208–210 (2013)PubMedCrossRef A. Trojani, A. Greco, A. Tedeschi, M. Lodola, Di B. Camillo, F. Ricci et al., Microarray demonstrates different gene expression profiling signatures between Waldenstrom macroglobulinemia and IgM monoclonal gammopathy of undetermined significance. Clin. Lymphoma Myeloma Leuk. 13(2), 208–210 (2013)PubMedCrossRef
134.
Zurück zum Zitat A. Tsherniak, F. Vazquez, P.G. Montgomery, B.A. Weir, G. Kryukov, G.S. Cowley et al., Defining Cancer Dependency Map Cell. 170(3), 564–76e16 (2017)PubMed A. Tsherniak, F. Vazquez, P.G. Montgomery, B.A. Weir, G. Kryukov, G.S. Cowley et al., Defining Cancer Dependency Map Cell. 170(3), 564–76e16 (2017)PubMed
135.
Zurück zum Zitat M. Wohner, H. Tagoh, I. Bilic, M. Jaritz, D.K. Poliakova, M. Fischer et al., Molecular functions of the transcription factors E2A and E2-2 in controlling germinal center B cell and plasma cell development. J. Exp. Med. 213(7), 1201–1221 (2016)PubMedPubMedCentralCrossRef M. Wohner, H. Tagoh, I. Bilic, M. Jaritz, D.K. Poliakova, M. Fischer et al., Molecular functions of the transcription factors E2A and E2-2 in controlling germinal center B cell and plasma cell development. J. Exp. Med. 213(7), 1201–1221 (2016)PubMedPubMedCentralCrossRef
136.
Zurück zum Zitat S.L. Koppenhafer, K.L. Goss, E. Voigt, E. Croushore, W.W. Terry, J. Ostergaard et al., Inhibitor of DNA binding 2 (ID2) regulates the expression of developmental genes and tumorigenesis in ewing sarcoma. Oncogene. 41(20), 2873–2884 (2022)PubMedPubMedCentralCrossRef S.L. Koppenhafer, K.L. Goss, E. Voigt, E. Croushore, W.W. Terry, J. Ostergaard et al., Inhibitor of DNA binding 2 (ID2) regulates the expression of developmental genes and tumorigenesis in ewing sarcoma. Oncogene. 41(20), 2873–2884 (2022)PubMedPubMedCentralCrossRef
Metadaten
Titel
BHLHE41, a transcriptional repressor involved in physiological processes and tumor development
verfasst von
Caroline Bret
Fabienne Desmots-Loyer
Jérôme Moreaux
Thierry Fest
Publikationsdatum
10.09.2024
Verlag
Springer Netherlands
Erschienen in
Cellular Oncology / Ausgabe 1/2025
Print ISSN: 2211-3428
Elektronische ISSN: 2211-3436
DOI
https://doi.org/10.1007/s13402-024-00973-3

Neu im Fachgebiet Pathologie

„KI“ in der Rechtsmedizin – von der Forschung in die Praxis: Welche Herausforderungen ergeben sich?

Der Einsatz künstlicher Intelligenz (KI) in der Rechtsmedizin ist absehbar, KI-Anwendungen könnten im Ermittlungsverfahren oder in foro bald eine zentrale Rolle in Entscheidungsprozessen einnehmen. Dann werden insbesondere Transparenz …

Latent spaces of generative models for forensic age estimation

  • Open Access
  • Original reports

Similar to other parts of our society, machine learning has emerged as a popular tool within different areas of forensic medicine and will soon fuel more and more research and practice niches of our disciplines. Given the rapid advances, the …

Artificial intelligence in forensic pathology: an Australian and New Zealand perspective

  • Open Access
  • Leitthema

Artificial intelligence application has gained popularity in the last decade. Its application is implemented into multiple industries including the health sector; however, discipline-specific artificial intelligence applications are not widely …

Künstliche Intelligenz in der forensisch-radiologischen Altersdiagnostik

Fragen zu Implementierung und Nutzbarkeit von künstlicher Intelligenz (KI) spielen eine immer größere Rolle in der Forensischen Altersdiagnostik bei Lebenden, insbesondere im Rahmen forensisch-radiologischer Ansätze. Bis dato liegen bereits …