Skip to main content

Transcription factor BACH1 in cancer: roles, mechanisms, and prospects for targeted therapy

Abstract

Transcription factor BTB domain and CNC homology 1 (BACH1) belongs to the Cap ‘n’ Collar and basic region Leucine Zipper (CNC-bZIP) family. BACH1 is widely expressed in mammalian tissues, where it regulates epigenetic modifications, heme homeostasis, and oxidative stress. Additionally, it is involved in immune system development. More importantly, BACH1 is highly expressed in and plays a key role in numerous malignant tumors, affecting cellular metabolism, tumor invasion and metastasis, proliferation, different cell death pathways, drug resistance, and the tumor microenvironment. However, few articles systematically summarized the roles of BACH1 in cancer. This review aims to highlight the research status of BACH1 in malignant tumor behaviors, and summarize its role in immune regulation in cancer. Moreover, this review focuses on the potential of BACH1 as a novel therapeutic target and prognostic biomarker. Notably, the mechanisms underlying the roles of BACH1 in ferroptosis, oxidative stress and tumor microenvironment remain to be explored. BACH1 has a dual impact on cancer, which affects the accuracy and efficiency of targeted drug delivery. Finally, the promising directions of future BACH1 research are prospected. A systematical and clear understanding of BACH1 would undoubtedly take us one step closer to facilitating its translation from basic research into the clinic.

Background

Transcription factor BACH1 belongs to the CNC-bZIP family and contains 736 amino acids [1, 2]. BACH1 was initially identified as a heterodimerization partner of MafK, which was discovered through a yeast two-hybrid screen [2]. BACH1 is expressed widely and plays an important role in coordinating the transcriptional activation and inhibition of MafK [2]. NotI restriction mapping and YAC contig mapping demonstrated that the BACH1 gene is present at 21q22.1, between NotI sites LA329 (D21S338) and LL60 (D21S389), approximately 400 kb away from LA329 [1, 3]. The mouse homolog shares 80.3% sequence similarity with human BACH1 [1]. In the last thirty years, the roles of BACH1 have been extensively studied. BACH1 regulates heme homeostasis and effectively mediates oxidative stress, thus playing an essential role in inflammatory diseases and cancer [4,5,6,7,8]. In addition, BACH1 also exerts essential effects on various malignant biological behaviors of tumors [9,10,11]. For example, results from our laboratory and others have shown that BACH1 promotes the metastasis of hepatocellular carcinoma, making it a promising novel biomarker predicting a poor prognosis [12].

This review dissects the molecular mechanisms underlying the roles of BACH1 in epigenetic modifications, heme homeostasis, and oxidative stress, with a focus on the hallmarks of cancer, including tumor cell metabolism, invasion and metastasis, proliferation, different cell death ways, drug resistance and tumor immunity. Finally, promising therapeutic strategies are presented by summarizing and analyzing the potential clinical roles of BACH1. At the same time, the current research status of BACH1 as a prognostic biomarker is discussed. BACH1 is expected to become an effective novel target for tumor therapy.

Structure of BACH1

BACH1 contains four functional regions, including a Broad-complex, Tramtrack, and Bric-à-brac (BTB) domain, six cysteine–proline (CP) motifs, bZIP domain, and cytoplasmic localization signal (CLS) (Fig. 1A, B). The N-terminal BTB domain is also commonly called the pox virus and zinc finger (POZ) domain and often appears in the transcription factors with the bZIP domain. The BTB domain interacts with non-BTB proteins and binds to the chromatin structure, an essential step for regulating gene transcription [13]. BACH1 must constitute dimers or oligomers to recognize particular sequences of target genes, and the N-hook motif plays a critical role in BACH1 dimerization [14]. As shown in Fig. 1C, D, the BTB domain of BACH1 regulates homodimerization in vitro, which enables BACH1 to interact with Maf protein and develop a divalent DNA-binding complex that induces a DNA loop [15]. The C-terminal bZIP domain comprises highly conserved fundamental regions, crucial for DNA interactions, as well as less conserved short amphipathic leucine zipper domains that develop a dimeric coiled-coil structure. Furthermore, they interact with other bZIP monomers via bZIP domains [16, 17]. The transcription factors with the bZIP domain dimerize around DNA via leucine zipper interactions, and each monomer’s basic regions combine similar sequences on the corresponding strands of DNA. The core sequences of DNA binding sites at which bZIP monomers develop dimeric interactions are named TPA response elements (TREs) or cAMP response elements (CREs) [18]. The dimers bind TREs or CREs to develop the palindromic sequences TGA(G/C)TCA and TGACGTCA [19]. For example, the bZIP domain mediates the heterodimerization of BACH1 and small Maf proteins around DNA, due to which the heterodimers reduce the expression of many redox homeostasis genes and bind the Maf recognition elements (MAREs) in the gene promoters. In addition, heme can directly bind to the six CP motifs of BACH1, leading to its inactivation, nuclear export, ubiquitination and degradation by hoil-1 [20,21,22]. This mechanism adjusts the intracellular heme levels. The CLS is a highly conserved structure in BACH1 that leads to its cytoplasmic accumulation, which depends on the nuclear exporter chromosome maintenance 1 (CRM1). The phosphorylation of BACH1 at tyrosine residues by antioxidants and cadmium induces the nuclear export of BACH1, playing a role in the activation of CLS by the extracellular signal-related kinase (ERK) [23, 24]. Similarly, inorganic arsenic modulates the intracellular localization of BACH1, but the details of the underlying mechanism still need to be confirmed [25].

Fig. 1
figure 1

Domain structure and binding profile of human BACH1 transcription factor. A Domain structure of BACH1. B Binding profile of human BACH1 transcription factor. C The interaction between BACH1 and small Maf protein in the presence of DNA via the bZIP domain. D A DNA loop is developed through multimeric interactions via the BTB domain

Epigenetic modifications

Epigenetic modifications affect gene expression by altering DNA and the bound proteins that constitute chromatin through chemical modifications [26]. Numerous studies demonstrate that BACH1 mediates the expression of target genes by inducing DNA methylation and histone modifications. A CpG island methylator phenotype (CIMP) refers to DNA hypermethylation of CpG islands in the promoters of tumor suppressor genes and DNA repair genes, which is a common biological phenomenon and plays a crucial role in chromosomal instability in carcinogenesis, as well as having a high correlation with tumor prognosis [27]. In BRAF (V600E) mutant skin and colon cancer, a complex comprising BACH1 and MAFG is formed, which is promoted by B-Raf proto-oncogene variant BRAF (V600E). The complex recruits DNMT3B and chromatin helicase DNA-binding protein-8 (a chromatin remodeling factor) to the promoters of the mismatch repair gene MLH1 and other target tumor repressor genes containing CIMP. In turn, the hypermethylation of promoters is induced to silence gene expression, thus promoting cancer progression [28, 29]. In addition, p53R175H recruits LSD2, a histone H3K4me1/2 demethylase, and binds to BACH1, thereby forming a p53R175H-BACH1-LSD2 complex, which mediates the selective regulation of BACH1 targets by affecting the methylation status of target genomic proteins [30]. BACH1 recruits PRC2 and directly interacts with subunit EZH2 to facilitate the trimethylation of lysine 27 in histone 3 (H3K27me3), thereby silencing the downstream mesendodermal genes such as TBXT, GATA6, and MSX2 [31]. BACH1 recruits NANOG and the MLL/SET1 complex to the chromatin loop. It maintains the trimethylated state of H3K4 and the activity of enhancer-related promoters, which induces the expression of stemness-associated genes in mESCs, such as NANOG, ZFP42, and LIF [32]. In addition to histone methylation, BACH1 competes with β-catenin for complex formation with transcription factor 4 (TCF4). It recruits histone deacetylase 1 (HDAC1) to the promoters of TCF4-targeted genes, thereby inhibiting the expression of VEGF and IL-8, which depends on the BTB domain [33, 34]. In MEFs, BACH1 forms protein complexes with p53, N-CoR, and HDAC1 to inhibit the activation of p53 and is recruited to the promoters of p53 target genes [35, 36]. However, the concrete molecular mechanisms remain unclear and merit further research. The BACH1/Maf heterodimer recruits the complex consisting of NuRD, SIN3A, and SWI/SNF to the locus control region, thereby inhibiting the transcription of the β-globin gene by remodeling chromatin through histone deacetylation [37].

Heme homeostasis

Heme is a common and critical cofactor involved in various cellular processes [38]. However, free heme that is not bound to proteins can damage cells and tissues by inducing oxidative stress, which underscores the importance of tightly regulating heme levels. Heme oxygenase-1 (HO-1) is a rate-limiting enzyme for heme synthesis that also facilitates its degradation into iron, carbon monoxide, and biliverdin. Typically, BACH1 generates a complex with small Maf proteins to bind to the enhancer of the HO-1 gene and represses its expression [4]. Deficiency of BACH1 disinhibits HO-1 expression, thereby decreasing heme degradation [4]. The regulatory effect of BACH1 on HO-1 is also controlled by nuclear factor erythroid 2-related factor 2 (Nrf2). Under chemical or oxidative stress, Keap1-mediated Nrf2 ubiquitination and proteasomal degradation are inhibited [39,40,41]. The free Nrf2 is the transported from the cytoplasm to the nucleus, where it competes with BACH1 for small Maf proteins as an activating transcription factor in MAREs and transactivates HO-1 expression (Fig. 2) [4, 42, 43]. At the same time, BACH1 accumulation is induced by the accumulation of free Nrf2 [44]. Conversely, heme-binding BACH1 is exported from the nucleus and degraded by ubiquitination [43, 45].

Fig. 2
figure 2

BACH1 plays a vital role in heme homeostasis. BACH1 regulates heme homeostasis by controlling the transcription of the HO-1 gene. The interaction between BACH1 and Nrf2 is also involved in the regulation of heme levels

Accordingly, BACH1 is sensitive to heme levels inside the cell, so that its transcription function and protein levels are mediated by heme. Increasing heme levels abolish the suppression of HO-1 by BACH1/MafK, causing a shift toward Nrf2 activation by repressing the DNA-binding activity of BACH1 [4, 20]. It is reported that heme also binds to CP motifs, thereby boosting the nuclear export of BACH1 and exposing it to ubiquitination by Crm1, which has been demonstrated in diverse cell lines [4, 22, 46]. Conversely, lower heme levels induced by a heme inhibitor (succinylacetone) facilitate the nuclear accumulation of BACH1 in murine embryonic fibroblasts [22]. Overall, the BACH1/HO-1/heme pathway maintains heme homeostasis via a negative feedback loop.

Oxidative stress

Oxidative stress increases the intracellular levels of reactive oxygen species (ROS), which are generally cytotoxic and can also have an oncogenic effect [47]. BACH1 acts as a dual transcription factor that directly or indirectly regulates oxidative stress under physiological and pathological conditions. Glutathione (GSH) mitigates ROS-induced tissue damage, and cystic fibrosis transmembrane conductance regulator (CFTR) is one of the cellular efflux pumps that export GSH to reduce oxidative stress [48]. Under physiological conditions, BACH1 directly suppresses the expression of the CFTR gene by binding to its -44kb antioxidant response element in complex with MafK [49]. Conversely, nuclear translocation of BACH1/Nrf2 under oxidative stress promotes CFTR expression by competitively binding to MafK [49]. At 20% ambient oxygen, BACH1 also activates CFTR expression by binding to other cis-regulatory elements [50]. Thus, the regulation of CFTR by BACH1 in response to environmental signaling is bidirectional. As a core regulator, the upstream regulation for BACH1 was widely reported in oxidative stress. The genetic ablation of BACH1 upregulates the expression of HO-1 and PGC-1α, thus reducing the production of ROS [51]. In addition, the BACH1/Nrf2 balance is mediated by signal stimulation. Following induction by the receptor activator of nuclear factor-κB ligand (RANKL), BACH1 is transported into the nucleus, thereby suppressing the induction of Nrf2 and facilitating ROS pathways [52]. Following induction by lipopolysaccharide (LPS), BACH1 regulates inflammation and oxidative stress in acute injury, which is associated with Nrf2/HO-1 signaling [53, 54]. Approximately 30% of human lung cancers gain mutations affecting Keap1 or NFE2L2 (encoding Nrf2), which lead to the stabilization of Nrf2, influence the nuclear transport of BACH1/Nrf2, and upregulate the expression of HO-1, thereby combating oxidative stress [22].

Immune system

BACH1 plays an essential role in the innate and adaptive immune system, especially mediating the development of B lymphocytes and differentiation of macrophages. Firstly, BACH1 expression is particularly promoted in mature lymphoid lineage cells. C/EBPβ (encoded by the Cebpb gene) is a critical transcription factor in differentiating myeloid lineage cells [55]. BACH1 directly suppresses the expression of myeloid lineage-related genes (e.g. Cebpb), thereby suppressing the differentiation of common lymphoid progenitors (CLPs) into myeloid lineage cells [56]. In addition, it is reported that both BACH1- and BACH2-deficient CLPs almost completely lack B cell differentiation in vitro. Similarly, combined BACH1- and BACH2-deficient mice have a decreased abundance of pro-B cells, which was not observed in cases of individual BACH1 or BACH2 deficiency [56]. This means that BACH1 has redundant functions during lymphoid development [56]. Early B cell factor 1 (Ebf1) promotes B lymphocyte lineage differentiation. In both BACH1- and BACH2-deficient CLPs, expression of Ebf1 is inhibited, while sufficient expression of Ebf1 rescues the normal differentiation to B lymphocytes. This suggests that BACH1 indirectly promotes the expression of the Ebf1 gene, thereby inducing pro-B lymphocyte development [56]. Overall, BACH1 makes a difference in the early stages of B cell development and consolidates early B cell features. As indispensable transcription factors in B lymphocytes, BACH1 and BACH2 are negatively modulated by miR-148a, suppressing B lymphocyte maturation and homeostasis together [56,57,58].

In addition, BACH1 plays a vital role in differentiating red pulp and bone marrow macrophages by controlling heme homeostasis [59]. BACH1 directly or indirectly inhibits Spic expression in monocytes and blocks their differentiation into red pulp and bone marrow macrophages, which is abrogated by heme. At the same time, the heme-mediated upregulation of HO-1 by BACH1 reduces cytotoxicity in red pulp and bone marrow macrophages. The increased abundance of red pulp and bone marrow macrophages greatly enhance red blood cell phagocytosis and lower heme levels. The BACH1/HO-1 pathway also promotes engulfment and the anti-inflammatory process of macrophages during efferocytosis by regulating the engulfment of low-heme apoptotic thymocytes and high-heme red blood cells [60]. Secondly, the BACH1/HO-1 pathway regulates the differentiation of inflammatory macrophages. BACH1 represses the differentiation of M2 macrophages, and one of the underlying mechanisms is probably the inhibition of HO-1 in macrophages by BACH1 [59]. In peritoneal macrophages, BACH1 suppresses the expression of genes involved in their polarization toward the M2 phenotypes, containing those encoding arginase 1, FIZZ1, CD206, and YM1 [61]. Thus, BACH1 controls diverse differentiation events in the macrophage lineage, but the precise mechanisms still need to be better understood and require further research. BACH1 also influences immunomodulatory activities and macrophage reprogramming by cellular bioenergetics. BACH1-deficient macrophages exhibit a mitochondrial metabolic shift, which presents as an increase in glycolysis and a reduction in oxidative phosphorylation (OXPHOS). In addition, BACH1 deficiency in macrophages promotes NLRP3 inflammasome activation by altering iNOS and COX-2 expression upon LPS stimulation [62]. Taken together, the available studies demonstrate that BACH1 has vital functions in the development, differentiation, and functioning of immune cells.

BACH1 in cancer

BACH1 is known as an essential transcriptional factor that regulates physiological processes. Thus, it is unsurprising that the dysregulation of BACH1 contributes to severe pathological changes, including tumorigenesis. Numerous studies have shown that BACH1 is highly expressed in multiple tumors and boosts their malignant biological behaviors. The following sections will elaborate on the roles of BACH1 in metabolism, invasion, metastasis, proliferation, and different cell death pathways of cancer cells. In addition, we discuss the potential roles of BACH1 in drug resistance and tumor immunity.

Metabolism

Mitochondrial metabolism and aerobic glycolysis

Cells typically gain energy through glycolysis in the cytoplasm, followed by mitochondrial OXPHOS with oxygen. However, cancer cells preferentially employ cytoplasmic glycolysis despite having sufficient oxygen. The phenomenon is identified as the “Warburg effect” or “aerobic glycolysis” [63, 64]. Notably, BACH1 regulates aerobic glycolysis to promote tumor growth, progression, and metastasis in breast cancer, lung cancer, hepatocellular cancer, and glioma by controlling the expression of genes encoding the mitochondrial electron transport chain (ETC) and glycolytic enzymes (Fig. 3). In breast cancer, BACH1 inhibits the expression of ETC complex (I-IV) genes and the activity of pyruvate dehydrogenase (PDH) via pyruvate dehydrogenase kinase (PDK), thereby decreasing mitochondrial metabolism (Fig. 3). BACH1 directly inhibits the mitochondrial ETC gene expression. Bioinformatics analyses support a negative association between the expression of the ETC gene and BACH1 expression in human breast cancer cells [10]. At the same time, deficiency of BACH1 stimulates the expression of the ETC gene, promotes mitochondrial respiration, and improves glucose utilization via the tricarboxylic acid (TCA) cycle [10]. Notably, the mitochondrial ECT complex is usually coupled to the TCA cycle. In addition, BACH1 increases the phosphorylation of both PDK and PDH on S293 but does not influence the PDH level to promote pyruvate decarboxylation. In lung cancer cells, BACH1 promotes the expression of glycolytic enzymes, such as hexokinase2 (HK2) and glyceraldehyde-3-phosphate dehydrogenase (GAPDH), to support aerobic glycolysis, which is accomplished by binding to the promoter regions of their encoding genes (Fig. 3). Further STRING database and RNA-seq analyses also support the positive association between BACH1 and glycolytic gene expression in lung cancer [9, 65]. Knockdown of BACH1 decreased the expression of HK2 and GAPDH to cancel out the rise of glucose metabolism-related indicators, and migration. Inhibitors such as 2-deoxyglucose (2-DG) and lonidamine (an HK2 inhibitor) block the glycolysis pathway at hexokinase, while 3-bromopyruvate (3-BP) represses GAPDH. Combined treatment with these drugs decreases glycolysis rates and effectively reverses the BACH1-induced migration of lung cancer cells. Moreover, BACH1 expression seems to be positively associated with PFKFB3 and SLC16A1 expression according to an analysis of human lung cancer datasets [9].

Fig. 3
figure 3

BACH1 affects cancer-cell metabolism thereby promoting tumor progression. BACH1 affects mitochondrial metabolism, aerobic glycolysis, and lactate metabolism in cancer cells by transcriptional regulation

In addition, BACH1 is regulated by other signals and thus also indirectly influences glycolysis in cancer cells. It is reported that overexpression of BACH1 stimulated by long non-coding RNAs (lncRNA) small nucleolar RNA host gene 5 (SNHG5) regulates the levels of proteins related to glucose metabolism by targeting miR-299, thus regulating glucose consumption and lactate secretion, thereby facilitating the growth and glycolysis of breast cancer cells [66]. Furthermore, the induction of BACH1 deubiquitination by USP47 promotes the Warburg effect and progression in non-small cell lung cancer, as demonstrated by metabolism-related indicators [67]. Similarly, the reduction of UBR7 levels mediated by methyltransferase ALKBH5 reduces the transcription of Keap1, which in turn upregulates the Nrf2/BACH1/HK2 pathway, promoting tumorigenesis in hepatocellular carcinoma by targeting glycolysis [68]. In addition, BACH1, stabilized by the regulator of chromatin condensation (RCC2) at the C-terminal, induces the upregulation of HK2, which drives glioma progression [69].

Lactate metabolism

Lactate derived from aerobic glycolysis, which is an energy substrate for aerobic cancer cells and glucose-deprived immune cells, facilitates tumor invasion and metastasis, making it a significant prognostic factor in cancer patients [70,71,72]. BACH1 inhibits the expression of monocarboxylate transporter 1 (MCT1, encoded by SLC16A) and lactate dehydrogenase B (LDHB), downregulating lactate utilization in triple-negative breast cancer (TNBC) cells (Fig. 3) [73]. MCTs are the first step in the cellular catabolism of extracellular lactate [74, 75]. BACH1 deficiency enhances the lactate utilization of TNBC cells and increases their sensitivity to MCT1 inhibitors. Secondly, intracellular lactate is transported into mitochondria by MCT2 and MCT3 or transformed into pyruvate by LDHB, while the depletion of BACH1 enhances LDHB activity [73]. Thus, BACH1 regulates lactate efflux or influx, presenting promising targets for cancer treatment. In addition, BACH1 promotes the secretion of lactic acid, which was reported to potentially serve as a nutrient for regulatory T cells, thereby promoting immunotherapy resistance [10, 76]. However, further study of concrete mechanisms may reveal novel targets for immunotherapy.

Tumor invasion and metastasis

Tumor invasion and metastasis include two stages: the dissemination of tumor cells from the primary tumor to distant tissues and organs and the adaptation of disseminated tumor cells to the distant tissue microenvironment. BACH1 facilitates tumor invasion and metastasis by regulating the EMT, ECM remodeling, and pro-metastatic factors (Fig. 4). At the same time, BACH1 upregulates the malignant features of cancer stem cells (CSCs).

Fig. 4
figure 4

BACH1 promotes tumor invasion and metastasis. BACH1 plays a core regulatory role in the EMT, ECM remodeling, and expression of pro-metastatic factors thereby promoting tumor invasion and metastasis

EMT

The epithelial-mesenchymal transition (EMT) is a critical step in the initiation of cancer metastasis by downregulating epithelial genes and upregulating mesenchymal genes, thereby enhancing the motility and invasiveness of tumor cells [77]. In pancreatic ductal adenocarcinoma (PDAC), BACH1 indirectly inhibits the expression of epithelial gene CDH1 by directly reducing FOXA1 expression and promoting SNAI2 expression [11]. In addition, BACH1 directly inhibits the expression of CLDN3 and CLDN4, which are essential for the adhesion of epithelial cells. Knockdown of BACH1 enhances the expression of plakophilin-2 (encoded by PKP2). At the same time, CHIP-seq demonstrated that PKP2 is the target gene of BACH1 [11]. However, the precise regulation for PKP2 by BACH1 was not studied further. It was reported that the overexpression of BACH1 induced by tank binding kinase 1 (TBK1) increases intracellular labile iron and decreases the expression of E-cadherin (encoded by CDH1), thereby promoting the metastasis of pancreatic cancer cells [78]. In renal cell carcinoma, overexpression of BACH1 downregulates E-cadherin to decrease cell adhesion and promote invasion [79]. In esophageal squamous cell carcinoma, BACH1 directly induces the expression of CDH2, SNAI2, and VIM to promote the EMT by binding to their promoter regions [80]. In ovarian cancer, BACH1 recruits HMGA2 to the promoter region of Snail, thereby promoting the EMT and motility of tumor cells [81, 82]. In glioma, overexpression of BACH1 enhances the expression of CDH2, SNAI2, and CD44 in U87 cells to promote the EMT [83]. In hepatocellular carcinoma (HCC), BACH1 upregulates the expression of cell motility-related genes, such as insulin-like growth factor 1 receptor (IGF1R) and protein tyrosine kinase 2 (PTK2), to facilitate the growth and metastasis of HCC. However, IGF2, the ligand of IGF1R, upregulates the expression of BACH1 via the IGFR/(ERK1/2)/EST1 axis, creating a positive feedback loop [12]. Additionally, the induction of BACH1 via the lncRNA TRG-AS1/ miR-4500 axis promotes the proliferation, migration, invasion, and EMT of HCC [84]. Thus, BACH1 plays a critical role in the EMT process of various cancers.

ECM remodeling

BACH1 modulates extracellular matrix (ECM) remodeling to facilitate glioma invasion [83]. Studies have demonstrated that high levels of collagens and fibronectin support the infiltration and invasion of glioma [85, 86]. BACH1 activates the transforming growth factor beta (TGF-β) signaling pathway to facilitate the expression of ECM components in tumors [83, 87]. Specifically, transforming growth factor beta receptor 2 (TGFBR2) is an initial regulator of the TGF-β signaling pathway. The combination of TGFBR2 and cytokines initiates the TGFBR1/ smad2/3 /smad4 axis, which promotes the transcription of fibrotic genes (collagens and fibronectin) [83, 87]. BACH1 enhances the expression of TGFBR2 and activates the smad2/3 signaling pathway. In addition, BACH1 induces the formation of invadopodia, thereby promoting the expression and secretion of MMP2, which can help degrade the ECM [83]. BACH1 regulates the degradation and deposition of the ECM, creating a favorable environment for tumor metastasis (Fig. 4). The regulation of ECM remodeling by BACH1 is a novel approach to cancer treatment, which merits further research.

Pro-metastatic factors

There is increasing evidence that aberrant expression of specific factors facilitates tumor progression and metastasis. MMP1 regulates the adhesion between tumor cells, promoting tumor invasion and metastasis [88]. High expression of CXCR4 is closely related to increased tumor metastasis, and CXCR4 is involved in multiple processes related to metastasis, including chemotaxis, colonization and proliferation [89]. Similarly, FHL1, DUSP1 and OPN are critical for tumor invasion and metastasis. BACH1 has been identified as a central regulator of these tumor-associated genes related to the promotion of metastasis (Fig. 4). In colorectal cancer, BACH1 facilitates the expression of CXCR4 as demonstrated by TCGA dataset analysis, western blotting, and immunohistochemistry [90]. In breast cancer, bioinformatic analysis indicated that BACH1 mediates four Bone Marrow Signature (BMS) genes, including MMP1, CXCR4, FHL1, and DUSP1 [91]. Specifically, BACH1 directly promotes the expression of MMP1 by binding to its promoter region to facilitate metastasis [91, 92]. MMP9 induces tumor metastasis in pre-metastatic sites and facilitates carcinogenesis [93,94,95]. Conversely, silencing of BACH1 decreases the expression of MMP9 [96]. Raf kinase inhibitory protein (RKIP) is a tumor metastasis inhibitor that mediates the growth and differentiation of various organisms and exhibits reduced expression in many solid tumors [97,98,99]. BACH1 is repressed by RKIP, which targets let-7, thereby inhibiting breast cancer bone metastasis by downregulating the expression of BMS genes, including MMP1, OPN, and CXCR4 [92]. Interestingly, Lee et al. reported that BACH1 downregulates the transcription of both RKIP and itself [100]. Furthermore, myc drives the transcription of oncogenic HOXB8 by binding to its super enhancers. HOXB8 occupies and activates the transcription of the BACH1 gene with BACH1 itself and interacts with BACH1 to cause a transcriptional cascade, thereby enhancing colorectal cancer invasion [101].

Cancer stem cells

Cancer stem cells (CSCs), which are involved in tumor progression and metastasis, have been discovered in many types of cancer, including lung, liver, and pancreatic cancer [102,103,104,105]. In lung cancer, BACH1 promotes the expression of CD44+, thereby inducing the proliferation and invasion of lung CSCs both in vivo and in vitro [106]. In addition, BACH1 activates MAPK signaling to promote the growth and stemness of CSCs [106]. Although the downstream of MAPK signaling pathway related to CD44+ remains unclear, it is known that multifunctional factors (OCT4, SOX2 and NANOG) mediate the biological activity of CSCs [107]. In human embryonic stem cells, BACH1 recruits three multifunctional factors, namely NANOG, SOX2 and OCT4, and induces their deubiquitylation by recruiting deubiquitinase USP7, thereby stabilizing the three factors [31]. Thus, it is possible that BACH1 upregulates CSC-like properties by stabilizing multifunctional transcription factors, which is expected to be further corroborated in the future.

Proliferation

A hallmark of cancer is replicative immortality, which is determined by many factors. BACH1 plays an essential role in regulating cancer cell proliferation. Activating mutations of Ras family proto-oncogenes are frequently found in human cancers. On the one hand, BACH1 inhibits HO-1 expression to maintain H-RasV12-induced ROS accumulation, which facilitates ERK signaling. On the other hand, H-RasV12 induces ERK signaling, which in turn suppresses the induction through negative feedback. As a resistor, BACH1 inhibits the feedback loop of ERK signaling for H-RasV12. Taken together, BACH1 effectively facilitates H-RasV12-induced ERK signaling, thereby promoting tumorigenesis and proliferation in the Ras-induced transformation MEFs model as a target of non-oncogene addiction by H-RasV12 [108]. In PDAC, a lower level of BACH1 upregulates HO-1, activates AKT, ERK and eNOS, while also promoting the expression of HIF1A and VEGF as well as downregulating PTEN, thereby enhancing the proliferation of PDAC cells [109]. In Ewing sarcoma, BACH1 downregulates EWSR1, an RNA-binding protein that mediates the transcription percent of cell cycle protein D1a and D1b by raising the transcription elongation rate of Ewing sarcoma cells [6, 110]. Notably, circRNAs, lncRNAs and miRNAs act as upstream regulators of BACH1 in tumor cell proliferation. In osteosarcoma, BACH1 is induced by circ_0081001 via targeting miR-494-3p to promote proliferation of osteosarcoma cells [111]. Similarly, the circ_0000337/miR-4458/BACH1 axis also promotes proliferation in osteosarcoma [112]. In colorectal cancer, BACH1 is upregulated by circ_0087862 via sponging miR-142-3p to facilitate proliferation of colorectal cancer cells [113]. In hepatocellular carcinoma, BACH1 is stimulated by lncRNA712 via sponging miR-142-3p, thereby boosting proliferation [114]. In esophageal cancer, BACH1 is enhanced by lncRNA SNHG8 via downregulation of miR-1270 to promote proliferation [115].

Intriguingly, the function of BACH1 is different during tumor formation. Sato et al. found that BACH1 does not influence the proliferation of AsPC-1 cells when they are seeded at high concentrations, while inhibiting it at lower initial seeding concentrations. Moreover, these results have been verified in other cells, such as pancreatic ductal adenocarcinoma cell lines, SW1990 and BxPC-3 [11]. However, the specific mechanism remains unknown. In addition, the function of BACH1 is also different among various cell lines. Han et al. revealed that the expression of BACH1 is positively correlated with the increasing proliferation of the EOC cell line. At the same time, lower expression of BACH1 is also correlated with increasing proliferation of human umbilical vein endothelial cells [116]. Thus, BACH1 plays different roles in multiple cell types under specific experimental conditions, in line with its dual roles as either a transcriptional activator or inhibitor.

Different cell death pathways influenced by BACH1

Ferroptosis

Ferroptosis is a programmed cell death pathway related to necrosis. Ferroptotic cells form lipid peroxides and lipid hydroxyl radicals due to labile ferrous iron (Fe2+), resulting in cell death. BACH1 represses the expression of gene groups involved in L-glutathione (GSH) synthesis and labile iron metabolism to promote ferroptosis (Fig. 5) [117]. Firstly, in the ferritin-ferroportin pathway, BACH1 suppresses the expression of FTL (encoding ferritin light chain) and FTH1 (encoding ferritin heavy chain 1) to promote the redox activity of Fe2+. Furthermore, BACH1 suppresses the expression of SLC40A1 (encoding ferroportin, which transports Fe2+ out of the cytoplasm), thereby increasing the cellular concentration of labile iron and facilitating ferroptosis [117,118,119,120,121,122]. Interestingly, BACH1 reduces HO-1 activation and represses the accumulation of Fe2+, thereby protecting against ferroptosis. However, although HO-1 is upregulated by lower BACH1 levels, enough ferritin isolates Fe2+ generated by heme degradation to rescue cells from ferroptosis. Thus, BACH1-mediated HO-1 upregulation may repress or promote ferroptosis, which depends on the balance of HO-1 and ferritin [4, 123]. Secondly, in the GSH-GPX4 pathway, BACH1 suppresses the expression of GCLC and GCLM genes to inhibit the key enzyme of the GSH synthetic pathway, thereby promoting ferroptosis. Furthermore, BACH1 suppresses the expression of SLC7A11 to downregulate the subunit of system Xc- and inhibit the activity of arachidonic acid 12-lipoxygenase (ALOX12), which promotes lipid peroxidation and ferroptosis as a lipoxygenase [6, 117, 124, 125]. Notably, hotspot mutant p53R175H abolishes BACH1-mediated downregulation of SLC7A11 thereby inhibiting ferroptosis to facilitate tumor growth. In addition, BACH1 expression induced by p53R175H upregulates CEMIP (encoding cell migration inducing hyaluronidase 1) to promote tumor metastasis. The bidirectional role of BACH1 in cancer is mediated by p53R175H by forming a ternary complex that regulates transcription [30]. Thirdly, in the Hippo pathway, BACH1 inhibits the expression of E-cadherin, which results in reduced intercellular contacts and deactivates the NF2-Hippo pathway to activate yes-associated protein (YAP, a ferroptosis-inducing transcriptional coregulator), thus promoting ferroptosis [126]. Fourthly, BACH1 may interact with Nrf2 and mediate the FSP1-CoQ pathway and tetrahydrobiopterin (BH4) system to promote ferroptosis, which is expected to be studied further [126].

Fig. 5
figure 5

BACH1 plays an essential regulatory role in different cell death pathways. A The regulatory role of BACH1 in ferroptosis. B BACH1 exerts divergent effects on apoptosis in various cancers

Studies have proved that Stearoyl-CoA desaturase-1 (SCD1) protects tumor cells from ferroptosis in lung, gastric, and ovarian cancer [127,128,129]. Significantly, BACH1 inhibits the transcription of SCD1 to reduce oleic acid biosynthesis, thereby decreasing ferroptosis resistance [130,131,132]. Thus, BACH1 leads to a concentration gradient between cell membranes with low oleic acid content and lymph fluid with high oleic acid, resulting in the chemoattraction of cancer cells and metastasis via lymph vessels. To sum up, BACH1-induced ferroptosis advances lymph node metastasis but represses subcutaneous growth and hematogenous metastasis [132].

Apoptosis

BACH1 has different functions in the regulation of apoptosis in the progression of cancer (Fig. 5). BACH1 is highly expressed in lung CSCs, where it increases the expression of BCL-2 protein and represses the expression of BAX as well as p53, thereby inhibiting apoptosis [106]. However, in breast cancer, C-X-C motif chemokine receptor 3-B (CXCR3-B) induced by its ligands facilitates the activation of p38-MAPK and inhibits ERK-1/2, while also increasing the nuclear localization of BACH1 and nuclear export of Nrf2, leading to inhibition of CXCR3-B-associated signaling pathways, promoting apoptosis while inhibiting proliferation [133].

Drug resistance in cancer

Drug resistance continues to be a persistent obstacle in cancer treatment. However, the mechanisms underlying chemotherapy resistance are still incompletely understood. BACH1 has been shown to confer resistance to temozolomide (TMZ) in glioblastoma. Specifically, BACH1 breaks the relationship between SP1 and p53 by competitively binding to p53 to abrogate the binding of SP1 to the promoter region of O-6-methylguanine-DNA methyltransferase (MGMT), thereby inducing MGMT expression and TMZ resistance. Thus, inhibitors targeting BACH1 may be a potential means of overcoming TMZ resistance [134]. In mantle cell lymphoma, BACH1 is stabilized by antioxidants blocking ROS, thereby aggravating bortezomib resistance instead of promoting cell death [135]. However, the downstream mechanism remains elusive. In breast cancer, BACH1 results in drug resistance related to mitochondrial metabolism. Metformin was reported to act as a potential antitumor drug by repressing the mitochondrial ETC complex I and other metabolic targets [136, 137]. Higher levels of BACH1 result in metformin resistance, and the effect is independent of pyruvate [10]. However, in shBACH1 cells, metformin resistance is only observed in the presence of pyruvate, which suggests that pyruvate levels affect metformin resistance by mediating the NAD+/NADH ratio. Furthermore, in BACH1-deficient cells, metformin, rotenone and antimycin A significantly repress cell growth and viability, meaning that BACH1 deficiency can effectively overcome cancer resistance to these drugs. Additionally, in BACH1-deficient cells, the induction of mitochondrial ETC genes such as COX15 or UQCRC1 commonly inhibits metformin resistance and cell growth. Notably, heme serves as an alternative trigger of BACH1 protein degradation, and heme administration emulates the knockdown of the BACH1 gene. In heme-treated TNBC cells, the sensitivity of metformin, rotenone and antimycin A increased significantly [10]. Thus, targeted inhibition of BACH1, as a promising strategy, increases the sensitivity to chemotherapy drugs that suppress mitochondrial metabolism in breast cancer and potentially other cancers [10]. Overall, BACH1 is involved in therapeutic resistance, and its downregulation provides a potential way to overcome drug resistance.

Regulation of the tumor immune microenvironment

BACH1 is widely expressed in dendritic cells, neutrophils, monocytes, and macrophages that are abundant in the tumor microenvironment (TME). It is known that these immune cells are involved in nearly all aspects of tumor biology, including proliferation, differentiation, and metastasis. Hence, targeting BACH1 regulation in the TME is considered as highly promising anti-tumor therapeutic strategy [138]. BACH1 expression in immune cells affects tumorigenesis and cancer cells produce high ROS levels. In alveolar macrophages (AMs), ROS-activated BACH1 downregulates the expression of PDLIM2 [139]. The downregulation of PDLIM2 inhibits the phagocytosis of AMs, activates STAT3, and facilitates the tumorigenic polarization/activation of AMs. At the same time, the downregulation of PDLIM2 facilitates the pulmonary recruitment of monocytes and their differentiation into AMs, thereby suppressing the effect of cytotoxic T lymphocytes (CTLs). Taken together, BACH1 represses the innate and adaptive immune response in lung oncogenesis, thus playing a cancer-promoting role.

Expression of BACH1 is related to tumor-infiltrating immune cells in different cancer types, and it is positively correlated with the level of tumor-infiltrating lymphocytes (TILs) in most cancer types according to the TIMER algorithm [76]. Further analysis indicated that BACH1-related genes are associated with ubiquitin-mediated proteolysis, T cell receptor, PD-1/PD-L1 expression, Th17 cell differentiation, endocytosis and other signaling pathways [76]. BACH1 is strongly associated with immune responses in glioblastoma, especially M0 and M2 tumor-associated macrophages (TAMs) [140]. Additionally, high BACH1 expression is positively correlated with increased expression of immune checkpoints (ICs, such as CD276, TIM-3, LAG3, TIGIT and LGALS9), TAM chemokines induced by glioblastoma cells (including MCP1, mGM-CSF, and EGF) and M2 TAM markers, which means that BACH1 contributes to an immunosuppressive TME in glioblastoma [140]. BACH1 is positively related to a high level of monocyte-myeloid-derived suppressor cells (Mo-MDSCs) in TNBC, which means that BACH1 may help cancer cells escape from immunosurveillance and repress immune overreaction, thus promoting tumorigenesis [141]. The specific mechanism is expected to be explored in vivo and in vitro in the future. In order to further understand the immunomodulatory role of BACH1 in various tumors, we deeply analyzed the correlation between BACH1 and various tumor-infiltrating immune cells using the TIMER 2.0 database (Fig. 6). In many tumors, BACH1 is positively correlated with multiple immune cells, especially macrophages, monocytes, neutrophils, cancer-associated fibroblasts, myeloid dendritic cells, and regulatory T cells. Among these immune cells, some play positive while others play negative roles in immunoregulation. Clarifying the implication of divergent effects contributes to our understanding of the roles of BACH1 in the tumor immune microenvironment. However, BACH1 is negatively associated with NK T cells, and it is worth investigating if it suppresses the NK-mediated killing of tumor cells. Moreover, we analyzed the correlation between BACH1 and TME at the single-cell level in the TISCH database. The results indicates that BACH1 is highly expressed in conventional CD4+ T cells, CD8+ T cells, B cells, monocyte-macrophages, and malignant cells in most tumors (Fig. 7). Notably, there was little difference between the results of TIMER 2.0 database analysis and single-cell level analysis in the TISCH database. The high dropout rate of present single-cell technologies leads to an abundance of incomplete gene tests. Overall, BACH1 has essential implications for the regulation of multiple immune cells. The roles of the significantly increased expression of BACH1 in various immune cells are yet to be explored, while the BACH1-mediated complex regulatory network in diverse cancer types and immune cells also merits further investigation.

Fig. 6
figure 6

Analysis of the correlation between BACH1 and tumor-infiltrating immune cells in various cancers. Correlation between BACH1 and the abundance of multiple infiltrating immune cells in various cancers using TIMER 2.0 database. The purity-corrected partial Spearman’s rho values are displayed with statistical significance (p<0.05) marked with solid squares. ACC: Adrenocortical Carcinoma; BLCA: Bladder Cancer; BRCA: Breast Cancer; CESC: Cervical Cancer; CHOL: Cholangiocarcinoma; COAD: Colon Adenocarcinoma; DLBC: Large B-cell Lymphoma; ESCA: Esophageal Cancer; GBM: Glioblastoma; HNSC: Head and Neck Cancer; KICH: Kidney Chromophobe; KIRC: Kidney Clear Cell Carcinoma; KIRP: Kidney Papillary Renal Cell Carcinoma; LGG: Lower Grade Glioma; LIHC: Hepatocellular Carcinoma; LUAD: Lung Adenocarcinoma; LUSC: Lung Squamous Cell Carcinoma; MESO: Mesothelioma; OV: Ovarian Cancer; PAAD: Pancreatic Adenocarcinoma; PCPG: Pheochromocytoma and Paraganglioma; PRAD: Prostate Adenocarcinoma; READ: Rectal Adenocarcinoma; SARC: Sarcoma; SKCM: Skin Cutaneous Melanoma; STAD: Stomach Adenocarcinoma; TGCT: Testicular Cancer; THCA: Thyroid Carcinoma; THYM: Thymoma; UCEC: Endometrioid Cancer; UCS: Uterine Carcinosarcoma; UVM: Ocular melanoma

Fig. 7
figure 7

Correlation between BACH1 and TME at the single-cell level in the TISCH database. Heatmap showing BACH1 expression in various cells from diverse datasets. AML: Acute Myeloid Leukemia; BRCA: Breast Cancer; CRC: Colorectal Cancer; HNSC: Head and Neck Cancer; KIRC: Kidney Clear Cell Carcinoma; LIHC: Hepatocellular Carcinoma; NSCLC: Non-Small Cell Lung Carcinoma; OV: Ovarian Cancer; PRAD: Prostate Adenocarcinoma; SKCM: Skin Cutaneous Melanoma; STAD: Stomach Adenocarcinoma

BACH1 as a therapeutic target

Based on its roles in benign and cancerous diseases, the blockade of BACH1 or its downstream effectors is a promising therapeutic strategy. In addition, some studies have demonstrated that their combination a contributes to the inhibition of tumor growth in cancer.

Inhibiting BACH1

Small molecule inhibitors of BACH1

Some small molecule inhibitors of BACH1 are being investigated in benign diseases. As a non-psychotropic phytocannabinoid, cannabidiol induces the nuclear export and cytoplasmic proteasomal degradation of BACH1 in keratinocytes [142]. Isomeric O-methyl cannabidiol quinone (a novel cannabidiol derivative) not only has similar effects on BACH1, but also activates Nrf2 in cell-culture models of neurodegenerative diseases [143]. The non-electrophilic compound HPP-4382 inhibits the ability of BACH1 to bind DNA by imitating heme. However, HPP-4382 does not appear to affect the levels and nuclear distribution of BACH1 [144]. Similarly, HPP-D (an oral derivative of HPP-4382) competitively inhibits BACH1 from binding antioxidant response elements and promotes Nrf2 binding, thereby enhancing the expression of γ-globin in in sickle erythroid progenitors [145]. A substituted benzimidazole (HPPE) facilitates the release of BACH1 from DNA binding elements and nuclear export in a mouse model of Parkinson’s disease [65]. HPPE is an effective mimic of heme and has higher safety, which makes it a potential treatment for Parkinson’s disease. Based on virtual structural screening, the small molecule 1-piperazineethanol, α-[(1,3-benzodioxol-5-yloxy) methyl] -4-(2-methoxyphenyl) (M2) was identified as an inhibitor of BACH1 and effectively reduced aflatoxin B1-induced oxidative damage [146]. However, the research on these BACH1 inhibitors is still at the stage of in vitro experiments or animal models, while their bioavailability and safety need to be further explored. HPP971 from VTV-Therapeutics and ML-0207/ASP8731 from Astellas-Pharma are in clinical development as BACH1 inhibitors [147,148,149]. HPP971 has completed two phase I studies and was well tolerated.

As a potential antitumor strategy, small molecule inhibitors of BACH1 are also being explored in cancer. Intravenous hemin (Panhematin®) is approved for the treatment of acute porphyria by the Food and Drug Administration [150]. Hemin has no significant toxicity and decreases BACH1 levels [10, 20, 22]. Consistently, hemin treatment was found to enhance the sensitivity of cancer cells to MCT1 and mitochondrial inhibitors in tumor models [10, 73]. TBE53 (the biotinylated form of TBE31) has been demonstrated to effectively and persistently promote the degradation of BACH1 by the E3 ligase FBXO22 (50-fold more potent than hemin), while TBE31 has no influence on BACH1 levels. TBE53 (but not TBE31) reduces breast cancer cell migration and invasion in a BACH1-dependent manner [149]. Therefore, TBE53 is a potentially more potent alternative to hemin for inducing BACH1 degradation in anti-tumor therapy. However, its pharmacokinetic and pharmacodynamic properties needs further exploration in vivo. In addition, CDDO-trifluoromethyl-amide (CDDO-TFEA) and CDDO-Bardoxolone-Methyl (CDDO-Me) both decrease the nuclear levels of BACH1 to reduce invasion of lung cancer cells [151].

SiRNA

In addition to the small molecule inhibitors, siRNA-mediated gene therapy is a novel and effective antitumor strategy. In this approach, tumor cells are transfected with specific siRNAs that downregulate the expression of BACH1 in a dose-dependent manner, thereby repressing the migration and proliferation of breast cancer cells, leading to cell cycle arrest and promoting apoptosis [152]. BACH1 siRNA produces similar effects in prostate and HT-29 colon cancer cells, including inhibition of invasion and migration [96, 153].

Inhibiting the downstream targets of BACH1

Antioxidants stabilize BACH1 to promote lung cancer metastasis, which depends on glycolysis and lactate secretion and is related to high ATP production rates. Small molecules such as 2-DG, lonidamine (a HK2 inhibitor), 3-BP (a GAPDH inhibitor), dichloroacetate (a pyruvate dehydrogenase kinase inhibitor that represses the production of lactate), and AZD3965 (MCT-1 inhibitor) repress antioxidant-induced cell migration in vitro. In addition, 3-BP and AZD3965 have a similar effect in vivo [9].

Inhibiting both BACH1 and its downstream targets together

The genetic knockdown or pharmacological inhibition of BACH1 increases the use of lactate in TNBC cells, thereby sensitizing them to MCT1 inhibitors. Thus, combining hemin and MCT inhibitors (SR13800 and AZD3965) also presents a promising therapeutic strategy [73]. In addition, regulation of mitochondrial metabolism inhibits tumor growth by targeting BACH1, thereby sensitizing tumors to mitochondrial inhibitors. Hemin sensitizes TNBC to metformin by promoting the degradation of BACH1, making combined hemin-metformin treatment a potential therapeutic strategy for other cancer types, such as papillary thyroid carcinoma [10, 154]. Targeted inhibition of HO-1 promotes the FBXO22-dependent degradation of BACH1, thereby reducing cell migration and metastasis, which has been demonstrated in animal models. Treatment with zinc protoporphyrin IX (ZnPPIX) as a HO-1 inhibitor significantly suppresses the growth of gastric cancer, hepatocellular carcinoma, and sarcoma [155,156,157,158]. In addition, docosahexaenoic acid (DHA) -induced lipid peroxidation leads to the degradation of BACH1, promoting the expression of the HO-1 gene in human cancer cells [159]. Thus, combining DHA and HO-1 inhibitors represents a promising treatment for controlling tumor growth.

BACH1 as a prognostic biomarker

Numerous studies have shown that BACH1 is a novel and promising prognostic biomarker. Pan-cancer analysis revealed that low BACH1 expression predicts better outcomes in patients with kidney papillary cell carcinoma, lower grade glioblastoma, hepatocellular carcinoma, pancreatic adenocarcinoma, sarcoma, and thyroid carcinoma. However, low BACH1 expression is associated with a worse prognosis in skin cutaneous melanoma [76]. The genetic variation rs372883C/T in the 3’-untranslated region of BACH1 is related to the risk of developing PDAC and gemcitabine sensitivity. The rs372883C variation leads to higher BACH1 levels than the rs372883T variation in normal pancreatic tissue and PDAC. Furthermore, patients with rs372883CC have a better response to gemcitabine and longer median survival time compared with rs372883TT patients [109]. Thus, the genetic variation rs372883C/T is a biomarker for gemcitabine treatment efficacy and prognosis. Lignitto et al. suggested that the levels and transcription characteristics of BACH1 are correlated with shorter survival, advanced clinical grade and stage, as well as metastasis in human lung adenocarcinoma [44]. BACH1 is a vital and independent predictor in early-stage lung adenocarcinoma. The expression level of BACH1 is negatively correlated with overall survival of patients [138]. In colorectal cancer, the high expression of BACH1 is positively correlated with a high clinicopathological stage [160]. The expression of BACH1 is markedly correlated with the pathological stage, tumor-node-metastasis (TNM) stage, distant metastasis, and survival outcomes. TNBC patients with higher BACH1 expression have shorter overall survival and disease-free survival, which suggests that BACH1 is a potential prognostic biomarker [161]. In addition, various polygenic risk models and compound nomogram models based on BACH1 have been established to assess the prognosis, such as the RKIP pathway metastasis signature (RPMS, for TNBC), BACH1 pathway metastasis signature (BPMS, an improvement from RPMS for TNBC), and a model for glioblastomas [140, 162].

Perspectives

Past studies have shown that BACH1 promotes tumor progression in multiple ways and is a promising prognostic biomarker. Table 1 lists the oncogenic roles of BACH1 in different cancers. Furthermore, BACH1 regulates the transcription of diverse genes (Table 2). At the same time, BACH1 is also a downstream target as a core transcription factor (Table 3). Thus, BACH1-targeted regulation may be a novel strategy for antitumor therapy. Nevertheless, the mechanisms underlying BACH1 regulation and function still remain unclear in many cases. As mentioned earlier, given that BACH1 promotes ferroptosis, it is worth exploring the possibility of targeting BACH1 to facilitate ferroptosis in the tumor microenvironment. ROS play a vital role in the tumor microenvironment and affect various signaling pathways [163]. Thus, it is pertinent to explore whether BACH1 regulates these pathways due to its known effects on oxidative stress. As mentioned above, BACH1 exerts a significant impact on carcinogenesis, making it possible that a single targeted agent may be effective. However, recent studies indicated that BACH1 nonmonotonically mediates the expression of multiple genes, including RKIP, CXCR4, MMP1 and BACH1 itself, serving alternatively as a transcriptional activator or a transcriptional repressor at different protein levels [164]. At the same time, some previous studies revealed its cancer-suppressing effect, and BACH1 may negatively affect cancer cell survival by mediating protein homeostasis [165,166,167]. Cancer cells preferentially use the ubiquitin-proteasome system. The FBXO22-induced downregulation of BACH1 promotes AML progression, which means that BACH1 has a negative effect on tumor growth in this context [168]. Hence, a simple suppression for BACH1 may have negative effects on tumor metastasis in some cases [164]. The nonmonotone fitness landscape of BACH1 improves the accuracy and efficiency of targeted drug delivery, thereby increasing clinical benefits, which is an exciting direction worth pursuing in the future. According to KEGG pathway enrichment analysis, BACH1-associated genes are involved in signaling pathways related to PD-1/PD-L1 expression, T cell receptor function, and Th17 cell differentiation [76]. These findings prompt the question of how BACH1 affects these signaling pathways in the tumor immune microenvironment. In addition, in silico analyses of many tumor types suggest that BACH1 has a significant association with immune checkpoints and the tumor immune environment. These mechanisms deserve to be explored further as they may provide novel therapeutic strategies for both benign diseases and cancers.

Table 1 The oncogenic roles of BACH1 in different types of cancer
Table 2 The regulatory effects of BACH1 on downstream target genes
Table 3 The effects of upstream regulators on BACH1.

Conclusions

This review systematically summarized the structure, regulatory mechanisms, and oncological significance of BACH1. However, the current research status provides little information on how BACH1 is post-translationally modified and dysregulated in cancer. In addition, the bioavailability and safety of BACH1 inhibitors remains to be investigated in clinical trials. Notably, BACH1 is highly expressed in almost all cancer types and plays a crucial role in the key processes of tumorigenesis including cancer metabolism, metastasis, proliferation, and drug resistance. Multiple reports illustrate that high BACH1 expression is associated with a poor prognosis. Thus, the crosstalk between BACH1 and other factors in the TME is a particularly fascinating area for future study. Overall, clarifying the mechanistic complexity of BACH1 will facilitate a deeper understanding of tumors on a fundamental level, as well as offering new avenues for translation from the lab bench into clinical applications, potentially leading to novel breakthroughs in targeted therapy.

Availability of data and materials

The datasets generated and/or analyzed during the current study are available in the UCSC database (https://xenabrowser.net/) and GDC database (https://portal.gdc.cancer.gov/).

Abbreviations

2-DG:

2-deoxyglucose

3-BP:

3-bromopyruvate

AML:

Acute myeloid leukemia

AMs:

Alveolar macrophages

BACH1:

BTB domain and CNC homology 1

BMS:

Bone marrow signature

BRAF (V600E):

B-Raf proto-oncogene variant

BTB:

Broad-complex, Tramtrack, and Brab-a-brac

CIMP:

A CpG island methylator phenotype

CLPs:

Common lymphoid progenitors

CLS:

Cytoplasmic localization signal

CNC-bZIP:

cap ‘n’ collar and basic region leucine zipper

CP:

Cysteine–proline

CREs:

cAMP response elements

CRM1:

Chromosome maintenance 1

CSCs:

Cancer stem cells

CXCR3-B:

C-X-C motif chemokine receptor 3-B

DHA:

Docosahexaenoic acid

Ebf1:

Early B-cell factor 1

ECM:

Extracellular matrix

EMT:

Epithelial-mesenchymal transition

ERK:

Extracellular regulated protein kinases

ETC:

Electron transport chains

GAPDH:

Glyceraldehyde-3-phosphate dehydrogenase

GSH:

L-glutathione

HCC:

Hepatocellular carcinoma

HDAC1:

Histone deacetylase 1

HK2:

Hexokinase2

HO-1:

Heme oxygenase-1

IGF1R:

Insulin-like growth factor 1 receptor

Keap1:

Kelch like ECH associated protein 1

LDHB:

Lactate dehydrogenase B

lncRNA:

Long non-coding RNA

MAREs:

Maf recognition elements

MCT1:

Monocarboxylate transporter 1

MEFs:

Mouse embryonic fibroblasts

MGMT:

O-6-methylguanine-DNA methyltransferase

Nrf2:

Nuclear factor erythroid 2-related factor 2

PDAC:

Pancreatic ductal adenocarcinoma

PDH:

Pyruvate dehydrogenase

PDK:

Pyruvate dehydrogenase kinase

RKIP:

Raf kinase inhibitory protein

ROS:

Reactive oxygen species

RPMS:

RKIP pathway metastasis signature

SCD1:

Stearoyl-CoA desaturase-1

TAM:

Tumor-associated macrophage

TCA:

Tricarboxylic acid

TCF4:

Transcription factor 4

TGFBR2:

Transforming growth factor beta receptor 2

TGF-β:

Transforming growth factor beta

TME:

Tumor microenvironment

TMZ:

Temozolomide

TNBC:

Triple-negative breast cancer

TREs:

TPA response elements

VEGF:

Vascular endothelial growth factor

Reference

  1. Blouin JL, Duriaux Saïl G, Guipponi M, Rossier C, Pappasavas MP, Antonarakis SE. Isolation of the human BACH1 transcription regulator gene, which maps to chromosome 21q22.1. Hum Genet. 1998;102(3):282–8.

    Article  CAS  PubMed  Google Scholar 

  2. Oyake T, Itoh K, Motohashi H, Hayashi N, Hoshino H, Nishizawa M, et al. Bach proteins belong to a novel family of BTB-basic leucine zipper transcription factors that interact with MafK and regulate transcription through the NF-E2 site. Mol Cell Biol. 1996;16(11):6083–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Ohira M, Seki N, Nagase T, Ishikawa K, Nomura N, Ohara O. Characterization of a human homolog (BACH1) of the mouse Bach1 gene encoding a BTB-basic leucine zipper transcription factor and its mapping to chromosome 21q22.1. Genomics. 1998;47(2):300–6.

    Article  CAS  PubMed  Google Scholar 

  4. Sun JY, Hoshino H, Takaku K, Nakajima O, Muto A, Suzuki H, et al. Hemoprotein Bach1 regulates enhancer availability of heme oxygenase-1 gene. EMBO J. 2002;21(19):5216–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Furuyama K, Kaneko K, Vargas PD. Heme as a magnificent molecule with multiple missions: heme determines its own fate and governs cellular homeostasis. Tohoku J Exp Med. 2007;213(1):1–16.

    Article  CAS  PubMed  Google Scholar 

  6. Warnatz HJ, Schmidt D, Manke T, Piccini I, Sultan M, Borodina T, et al. The BTB and CNC homology 1 (BACH1) target genes are involved in the oxidative stress response and in control of the cell cycle. J Biol Chem. 2011;286(26):23521–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Cohen B, Tempelhof H, Raz T, Oren R, Nicenboim J, Bochner F, et al. BACH family members regulate angiogenesis and lymphangiogenesis by modulating VEGFC expression. Life Sci Alliance. 2020;3(4):e202000666.

    Article  PubMed  PubMed Central  Google Scholar 

  8. Jia MP, Li QH, Guo JY, Shi WH, Zhu L, Huang YJ, et al. Deletion of BACH1 attenuates atherosclerosis by reducing endothelial inflammation. Circ Res. 2022;130(7):1038–55.

    Article  CAS  PubMed  Google Scholar 

  9. Wiel C, Le Gal K, Ibrahim MX, Jahangir CA, Kashif M, Yao HD, et al. BACH1 stabilization by antioxidants stimulates lung cancer metastasis. Cell. 2019;178(2):330-45.e22.

    Article  CAS  PubMed  Google Scholar 

  10. Lee J, Yesilkanal AE, Wynne JP, Frankenberger C, Liu J, Yan J, et al. Effective breast cancer combination therapy targeting BACH1 and mitochondrial metabolism. Nature. 2019;568(7751):254–8.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  11. Sato M, Matsumoto M, Saiki Y, Alam M, Nishizawa H, Rokugo M, et al. BACH1 promotes pancreatic cancer metastasis by repressing epithelial genes and enhancing epithelial-mesenchymal transition. Cancer Res. 2020;80(6):1279–92.

    Article  CAS  PubMed  Google Scholar 

  12. Xie M, Sun MY, Ji XY, Li D, Chen XP, Zhang BX, et al. Overexpression of BACH1 mediated by IGF2 facilitates hepatocellular carcinoma growth and metastasis via IGF1R and PTK2. Theranostics. 2022;12(3):1097–116.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Albagli O, Dhordain P, Deweindt C, Lecocq G, Leprince D. The BTB/POZ domain: a new protein-protein interaction motif common to DNA- and actin-binding proteins. Cell Growth Differ. 1995;6(9):1193–8.

    CAS  PubMed  Google Scholar 

  14. Katsani KR, Hajibagheri MA, Verrijzer CP. Co-operative DNA binding by GAGA transcription factor requires the conserved BTB/POZ domain and reorganizes promoter topology. EMBO J. 1999;18(3):698–708.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Igarashi K, Hoshino H, Muto A, Suwabe N, Nishikawa S, Nakauchi H, et al. Multivalent DNA binding complex generated by small Maf and Bach1 as a possible biochemical basis for beta-globin locus control region complex. J Biol Chem. 1998;273(19):11783–90.

    Article  CAS  PubMed  Google Scholar 

  16. Turner R, Tjian R. Leucine repeats and an adjacent DNA binding domain mediate the formation of functional cFos-cJun heterodimers. Science. 1989;243(4899):1689–94.

    Article  ADS  CAS  PubMed  Google Scholar 

  17. Glover JN, Harrison SC. Crystal structure of the heterodimeric bZIP transcription factor c-Fos-c-Jun bound to DNA. Nature. 1995;373(6511):257–61.

    Article  ADS  CAS  PubMed  Google Scholar 

  18. Reinke AW, Baek J, Ashenberg O, Keating AE. Networks of bZIP protein-protein interactions diversified over a billion years of evolution. Science. 2013;340(6133):730–4.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  19. Igarashi K, Kurosaki T, Roychoudhuri R. BACH transcription factors in innate and adaptive immunity. Nat Rev Immunol. 2017;17(7):437–50.

    Article  CAS  PubMed  Google Scholar 

  20. Ogawa K, Sun J, Taketani S, Nakajima O, Nishitani C, Sassa S, et al. Heme mediates derepression of Maf recognition element through direct binding to transcription repressor Bach1. EMBO J. 2001;20(11):2835–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Igarashi K, Watanabe Matsui M. Wearing red for signaling: the heme-bach axis in heme metabolism, oxidative stress response and iron immunology. Tohoku J Exp Med. 2014;232(4):229–53.

    Article  CAS  PubMed  Google Scholar 

  22. Zenke Kawasaki Y, Dohi Y, Katoh Y, Ikura T, Ikura M, Asahara T, et al. Heme induces ubiquitination and degradation of the transcription factor Bach1. Mol Cell Biol. 2007;27(19):6962–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Kaspar JW, Jaiswal AK. Antioxidant-induced phosphorylation of tyrosine 486 leads to rapid nuclear export of Bach1 that allows Nrf2 to bind to the antioxidant response element and activate defensive gene expression. J Biol Chem. 2010;285(1):153–62.

    Article  CAS  PubMed  Google Scholar 

  24. Suzuki H, Tashiro S, Sun JY, Doi H, Satomi S, Igarashi K. Cadmium induces nuclear export of Bach1, a transcriptional repressor of heme oxygenase-1 gene. J Biol Chem. 2003; 278(49):49246-53.

  25. Liu D, Duan XX, Dong DD, Bai CJ, Li X, Sun GF, et al. Activation of the Nrf2 pathway by inorganic arsenic in human hepatocytes and the role of transcriptional repressor Bach1. Oxid Med Cell Longev. 2013;2013:984546.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Esteve Puig R, Bueno Costa A, Esteller M. Writers, readers and erasers of RNA modifications in cancer. Cancer Lett. 2020;474:127–37.

    Article  CAS  PubMed  Google Scholar 

  27. Dahlin AM, Palmqvist R, Henriksson ML, Jacobsson M, Eklöf V, Rutegård J, et al. The role of the CpG island methylator phenotype in colorectal cancer prognosis depends on microsatellite instability screening status. Clin Cancer Res. 2010;16(6):1845–55.

    Article  CAS  PubMed  Google Scholar 

  28. Fang MG, Hutchinson L, Deng A, Green MR. Common BRAF(V600E)-directed pathway mediates widespread epigenetic silencing in colorectal cancer and melanoma. Proc Natl Acad Sci U S A. 2016;113(5):1250–5.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  29. Fang MG, Ou J, Hutchinson L, Green MR. The BRAF oncoprotein functions through the transcriptional repressor MAFG to mediate the CpG Island Methylator phenotype. Mol Cell. 2014;55(6):904–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Su ZY, Kon N, Yi JJ, Zhao HQ, Zhang WW, Tang QS, et al. Specific regulation of BACH1 by the hotspot mutant p53R175H reveals a distinct gain-of-function mechanism. Nat Cancer. 2023;4(4):564–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Wei XX, Guo JY, Li QH, Jia QQ, Jing Q, Li Y, et al. Bach1 regulates self-renewal and impedes mesendodermal differentiation of human embryonic stem cells. Sci Adv. 2019;5(3):eaau7887.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  32. Niu C, Wang SQ, Guo JY, Wei XX, Jia MP, Chen ZX, et al. BACH1 recruits NANOG and histone H3 lysine 4 methyltransferase MLL/SET1 complexes to regulate enhancer-promoter activity and maintains pluripotency. Nucleic Acids Res. 2021;49(4):1972–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Jiang L, Jia MP, Wei XX, Guo JY, Hao SY, Mei AH, et al. Bach1-induced suppression of angiogenesis is dependent on the BTB domain. EBioMedicine. 2020;51:102617.

    Article  PubMed  PubMed Central  Google Scholar 

  34. Jiang L, Yin M, Wei XX, Liu JX, Wang XH, Niu C, et al. Bach1 represses Wnt/β-catenin signaling and angiogenesis. Circ Res. 2015;117(4):364–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Dohi Y, Ikura T, Hoshikawa Y, Katoh Y, Ota K, Nakanome A, et al. Bach1 inhibits oxidative stress-induced cellular senescence by impeding p53 function on chromatin. Nat Struct Mol Biol. 2008;15(12):1246–54.

    Article  CAS  PubMed  Google Scholar 

  36. Igarashi K, Ota K, Nakame A. Regulation of cellular senescence by Bach1. Nihon Rinsho. 2009;67(7):1423–8.

    PubMed  Google Scholar 

  37. Brand M, Ranish JA, Kummer NT, Hamilton J, Igarashi K, Francastel C, et al. Dynamic changes in transcription factor complexes during erythroid differentiation revealed by quantitative proteomics. Nat Struct Mol Biol. 2004;11(1):73–80.

    Article  CAS  PubMed  Google Scholar 

  38. Reddi AR, Hamza I. Heme mobilization in animals: a metallolipid’s journey. Acc Chem Res. 2016;49(6):1104–10.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Cullinan SB, Gordan JD, Jin J, Harper JW, Diehl JA. The Keap1-BTB protein is an adaptor that bridges Nrf2 to a Cul3-based E3 ligase: oxidative stress sensing by a Cul3-Keap1 ligase. Mol Cell Biol. 2004;24(19):8477–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Kobayashi A, Kang MI, Okawa H, Ohtsuji M, Zenke Y, Chiba T, et al. Oxidative stress sensor Keap1 functions as an adaptor for Cul3-based E3 ligase to regulate proteasomal degradation of Nrf2. Mol Cell Biol. 2004;24(16):7130–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Zhang DD, Lo SC, Cross JV, Templeton DJ, Hannink M. Keap1 is a redox-regulated substrate adaptor protein for a Cul3-dependent ubiquitin ligase complex. Mol Cell Biol. 2004;24(24):10941–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Copple IM, Lister A, Obeng AD, Kitteringham NR, Jenkins RE, Layfield R, et al. Physical and functional interaction of sequestosome 1 with Keap1 regulates the Keap1-Nrf2 cell defense pathway. J Biol Chem. 2010;285(22):16782–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Sun J, Brand M, Zenke Y, Tashiro S, Groudine M, Igarashi K. Heme regulates the dynamic exchange of Bach1 and NF-E2-related factors in the Maf transcription factor network. Proc Natl Acad Sci U S A. 2004;101(6):1461–6.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  44. Lignitto L, LeBoeuf SE, Homer H, Jiang S, Askenazi M, Karakousi TR, et al. Nrf2 activation promotes lung cancer metastasis by inhibiting the degradation of Bach1. Cell. 2019;178(2):316-29 e18.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Chapple SJ, Keeley TP, Mastronicola D, Arno M, Vizcay Barrena G, Fleck R, et al. Bach1 differentially regulates distinct Nrf2-dependent genes in human venous and coronary artery endothelial cells adapted to physiological oxygen levels. Free Radic Biol Med. 2016;92:152–62.

    Article  CAS  PubMed  Google Scholar 

  46. Suzuki H, Tashiro S, Hira S, Sun J, Yamazaki C, Zenke Y, et al. Heme regulates gene expression by triggering Crm1-dependent nuclear export of Bach1. EMBO J. 2004;23(13):2544–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Hayes JD, Dinkova Kostova AT, Tew KD. Oxidative stress in cancer. Cancer Cell. 2020;38(2):167–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Kogan I, Ramjeesingh M, Li C, Kidd JF, Wang Y, Leslie EM, et al. CFTR directly mediates nucleotide-regulated glutathione flux. EMBO J. 2003;22(9):1981–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Zhang Z, Leir SH, Harris A. Oxidative stress regulates CFTR gene expression in human airway epithelial cells through a distal antioxidant response element. Am J Respir Cell Mol Biol. 2015;52(3):387–96.

    Article  PubMed  PubMed Central  Google Scholar 

  50. NandyMazumdar M, Paranjapye A, Browne J, Yin S, Leir SH, Harris A. BACH1, the master regulator of oxidative stress, has a dual effect on CFTR expression. Biochem J. 2021;478(20):3741–56.

    Article  CAS  PubMed  Google Scholar 

  51. Yusoff FM, Maruhashi T, Kawano KI, Nakashima A, Chayama K, Tashiro S, et al. Bach1 plays an important role in angiogenesis through regulation of oxidative stress. Microvasc Res. 2021;134:104126.

    Article  CAS  PubMed  Google Scholar 

  52. Kanzaki H, Shinohara F, Itohiya K, Yamaguchi Y, Katsumata Y, Matsuzawa M, et al. RANKL induces Bach1 nuclear import and attenuates Nrf2-mediated antioxidant enzymes, thereby augmenting intracellular reactive oxygen species signaling and osteoclastogenesis in mice. FASEB J. 2017;31(2):781–92.

    Article  CAS  PubMed  Google Scholar 

  53. Wang RX, Gu X, Zhang SX, Zhao YJ, Zhang HJ, Li FY. Deletion of BACH1 alleviates ferroptosis and protects against LPS-triggered acute lung injury by activating Nrf2/HO-1 signaling pathway. Biochem Biophys Res Commun. 2023;644:8–14.

    Article  CAS  PubMed  Google Scholar 

  54. Dorresteijn MJ, Paine A, Zilian E, Fenten MGE, Frenzel E, Janciauskiene S, et al. Cell-type-specific downregulation of heme oxygenase-1 by lipopolysaccharide via Bach1 in primary human mononuclear cells. Free Radic Biol Med. 2015;78:224–32.

    Article  CAS  PubMed  Google Scholar 

  55. Lin YC, Jhunjhunwala S, Benner C, Heinz S, Welinder E, Mansson R, et al. A global network of transcription factors, involving E2A, EBF1 and Foxo1, that orchestrates B cell fate. Nat Immunol. 2010;11(7):635–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Itoh Nakadai A, Hikota R, Muto A, Kometani K, Watanabe Matsui M, Sato Y, et al. The transcription repressors Bach2 and Bach1 promote B cell development by repressing the myeloid program. Nat Immunol. 2014;15(12):1171–80.

    Article  CAS  PubMed  Google Scholar 

  57. Porstner M, Winkelmann R, Daum P, Schmid J, Pracht K, Côrte Real J, et al. MiR-148a promotes plasma cell differentiation and targets the germinal center transcription factors Mitf and Bach2. Eur J Immunol. 2015;45(4):1206–15.

    Article  CAS  PubMed  Google Scholar 

  58. Yap DYH, Yung S, Lee P, Yam IYL, Tam C, Tang C, et al. B cell subsets and cellular signatures and disease relapse in lupus nephritis. Front Immunol. 2020;11:1732.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Haldar M, Kohyama M, So AYL, Kc W, Wu X, Briseño CG, et al. Heme-mediated SPI-C induction promotes monocyte differentiation into iron-recycling macrophages. Cell. 2014;156(6):1223–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Fige É, Szendrei J, Sós L, Kraszewska I, Potor L, Balla J, et al. Heme oxygenase-1 contributes to both the engulfment and the anti-Inflammatory program of macrophages during efferocytosis. Cells. 2021;10(3):652.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Harusato A, Naito Y, Takagi T, Uchiyama K, Mizushima K, Hirai Y, et al. BTB and CNC homolog 1 (Bach1) deficiency ameliorates TNBS colitis in mice: role of M2 macrophages and heme oxygenase-1. Inflamm Bowel Dis. 2013;19(4):740–53.

    Article  PubMed  Google Scholar 

  62. Pradhan P, Vijayan V, Cirksena K, Buettner FFR, Igarashi K, Motterlini R, et al. Genetic BACH1 deficiency alters mitochondrial function and increases NLRP3 inflammasome activation in mouse macrophages. Redox Biol. 2022;51:102265.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Warburg O. On respiratory impairment in cancer cells. Science. 1956;124(3215):269–70.

    Article  ADS  CAS  PubMed  Google Scholar 

  64. Li ZY, Zhang HF. Reprogramming of glucose, fatty acid and amino acid metabolism for cancer progression. Cell Mol Life Sci. 2016;73(2):377–92.

    Article  CAS  PubMed  Google Scholar 

  65. Ahuja M, Ammal Kaidery N, Attucks OC, McDade E, Hushpulian DM, Gaisin A, et al. Bach1 derepression is neuroprotective in a mouse model of Parkinson’s disease. Proc Natl Acad Sci U S A. 2021;118(45):e2111643118.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Huang SL, Huang ZC, Zhang CJ, Xie J, Lei SS, Wu YQ, et al. LncRNA SNHG5 promotes the glycolysis and proliferation of breast cancer cell through regulating BACH1 via targeting miR-299. Breast Cancer. 2022;29(1):65–76.

    Article  PubMed  Google Scholar 

  67. Peng J, Li WC, Tan NX, Lai XH, Jiang WL, Chen G. USP47 stabilizes BACH1 to promote the Warburg effect and non-small cell lung cancer development via stimulating Hk2 and Gapdh transcription. Am J Cancer Res. 2022;12(1):91–107.

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Zhao L, Kang M, Liu XM, Wang ZR, Wang Y, Chen HQ, et al. UBR7 inhibits HCC tumorigenesis by targeting Keap1/Nrf2/Bach1/HK2 and glycolysis. J Exp Clin Cancer Res. 2022;41(1):330.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Liu T, Wang YB, Wang YW, Cheung SKK, Or PMY, Wong CW, et al. The mitotic regulator RCC2 promotes glucose metabolism through BACH1-dependent transcriptional upregulation of hexokinase II in glioma. Cancer Lett. 2022;549:215914.

    Article  CAS  PubMed  Google Scholar 

  70. Faubert B, Solmonson A, DeBerardinis RJ. Metabolic reprogramming and cancer progression. Science. 2020; 368(6487):eaaw5473.

  71. Kennedy KM, Scarbrough PM, Ribeiro A, Richardson R, Yuan H, Sonveaux P, et al. Catabolism of exogenous lactate reveals it as a legitimate metabolic substrate in breast cancer. PLoS One. 2013;8(9):e75154.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  72. Faubert B, Li KY, Cai L, Hensley CT, Kim J, Zacharias LG, et al. Lactate metabolism in human lung tumors. Cell. 2017;171(2):358-71 e9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Padilla J, Lee BS, Zhai K, Rentz B, Bobo T, Dowling NM, et al. A heme-binding transcription factor BACH1 regulates lactate catabolism suggesting a combined therapy for triple-negative breast cancer. Cells. 2022;11(7):1177.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Halestrap AP. The SLC16 gene family - structure, role and regulation in health and disease. Mol Aspects Med. 2013;34(2–3):337–49.

    Article  CAS  PubMed  Google Scholar 

  75. Hong CS, Graham NA, Gu W, Espindola Camacho C, Mah V, Maresh EL, et al. MCT1 modulates cancer cell pyruvate export and growth of tumors that co-express MCT1 and MCT4. Cell Rep. 2016;14(7):1590–601.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Liu ZM, Wang J, Chen HY, Wu ZK, Liao FB, Wang S, et al. Uncovering BTB and CNC homology1 (BACH1) as a novel cancer therapeutic target. Front Genet. 2022;13:920911.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Dongre A, Weinberg RA. New insights into the mechanisms of epithelial-mesenchymal transition and implications for cancer. Nat Rev Mol Cell Biol. 2019;20(2):69–84.

    Article  CAS  PubMed  Google Scholar 

  78. Liu L, Matsumoto M, Matsui Watanabe M, Ochiai K, Callens BKK, Nguyen LC, et al. BACH1 expression is promoted by tank binding kinase 1 (TBK1) in pancreatic cancer cells to increase iron and reduce the expression of E-cadherin. Antioxidants (Basel). 2022;11(8):1460.

    Article  CAS  PubMed  Google Scholar 

  79. Takemoto K, Kobatake K, Miura K, Fukushima T, Babasaki T, Miyamoto S, et al. BACH1 promotes clear cell renal cell carcinoma progression by upregulating oxidative stress-related tumorigenicity. Cancer Sci. 2023;114(2):436–48.

    Article  CAS  PubMed  Google Scholar 

  80. Zhao Y, Gao JJ, Xie XF, Nan P, Liu F, Sun YL, et al. BACH1 promotes the progression of esophageal squamous cell carcinoma by inducing the epithelial-mesenchymal transition and angiogenesis. Cancer Med. 2021;10(10):3413–26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Han WY, Zhang YQ, Niu C, Guo JY, Li JJ, Wei XX, et al. BTB and CNC homology 1 (Bach1) promotes human ovarian cancer cell metastasis by HMGA2-mediated epithelial-mesenchymal transition. Cancer Lett. 2019;445:45–56.

    Article  CAS  PubMed  Google Scholar 

  82. Mansoori B, Mohammadi A, Asadzadeh Z, Shirjang S, Minouei M, Abedi Gaballu F, et al. HMGA2 and Bach-1 cooperate to promote breast cancer cell malignancy. J Cell Physiol. 2019;234(10):17714–26.

    Article  CAS  PubMed  Google Scholar 

  83. Cong ZX, Yuan F, Wang HD, Cai XM, Zhu JH, Tang T, et al. BTB domain and CNC homolog 1 promotes glioma invasion mainly through regulating extracellular matrix and increases ferroptosis sensitivity. Biochim Biophys Acta Mol Basis Dis. 2022;1868(12):166554.

    Article  CAS  PubMed  Google Scholar 

  84. Sun X, Qian Y, Wang X, Cao R, Zhang J, Chen W, et al. LncRNA TRG-AS1 stimulates hepatocellular carcinoma progression by sponging miR-4500 to modulate BACH1. Cancer Cell Int. 2020;20:367.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Pencheva N, de Gooijer MC, Vis DJ, Wessels LFA, Würdinger T, van Tellingen O, et al. Identification of a druggable pathway controlling glioblastoma invasiveness. Cell Rep. 2017;20(1):48–60.

    Article  CAS  PubMed  Google Scholar 

  86. Yin W, Zhu HC, Tan J, Xin ZQ, Zhou QW, Cao YD, et al. Identification of collagen genes related to immune infiltration and epithelial-mesenchymal transition in glioma. Cancer Cell Int. 2021;21(1):276.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Miwa S, Yokota M, Ueyama Y, Maeda K, Ogoshi Y, Seki N, et al. Discovery of selective transforming growth factor β type II receptor inhibitors as antifibrosis agents. ACS Med Chem Lett. 2021;12(5):745–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Liu M, Hu Y, Zhang MF, Luo KJ, Xie XY, Wen J, et al. MMP1 promotes tumor growth and metastasis in esophageal squamous cell carcinoma. Cancer Lett. 2016;377(1):97–104.

    Article  CAS  PubMed  Google Scholar 

  89. Mortezaee K. CXCL12/CXCR4 axis in the microenvironment of solid tumors: a critical mediator of metastasis. Life Sci. 2020;249:117534.

    Article  CAS  PubMed  Google Scholar 

  90. Zhu GD, Liu F, OuYang S, Zhou R, Jiang FN, Zhang B, et al. BACH1 promotes the progression of human colorectal cancer through BACH1/CXCR4 pathway. Biochem Biophys Res Commun. 2018;499(2):120–7.

    Article  CAS  PubMed  Google Scholar 

  91. Liang YJ, Wu H, Lei R, Chong RA, Wei Y, Lu X, et al. Transcriptional network analysis identifies BACH1 as a master regulator of breast cancer bone metastasis. J Biol Chem. 2012;287(40):33533–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Yun J, Frankenberger CA, Kuo WL, Boelens MC, Eves EM, Cheng N, et al. Signalling pathway for RKIP and Let-7 regulates and predicts metastatic breast cancer. EMBO J. 2011;30(21):4500–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Coussens LM, Tinkle CL, Hanahan D, Werb Z. MMP-9 supplied by bone marrow-derived cells contributes to skin carcinogenesis. Cell. 2000;103(3):481–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Bergers G, Brekken R, McMahon G, Vu TH, Itoh T, Tamaki K, et al. Matrix metalloproteinase-9 triggers the angiogenic switch during carcinogenesis. Nat Cell Biol. 2000;2(10):737–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Hiratsuka S, Nakamura K, Iwai S, Murakami M, Itoh T, Kijima H, et al. MMP9 induction by vascular endothelial growth factor receptor-1 is involved in lung-specific metastasis. Cancer Cell. 2002;2(4):289–300.

    Article  CAS  PubMed  Google Scholar 

  96. Shajari N, Davudian S, Kazemi T, Mansoori B, Salehi S, Khaze Shahgoli V, et al. Silencing of BACH1 inhibits invasion and migration of prostate cancer cells by altering metastasis-related gene expression. Artif Cells Nanomed Biotechnol. 2018;46(7):1495–504.

    Article  CAS  PubMed  Google Scholar 

  97. Granovsky AE, Rosner MR. Raf kinase inhibitory protein: a signal transduction modulator and metastasis suppressor. Cell Res. 2008;18(4):452–7.

    Article  CAS  PubMed  Google Scholar 

  98. Chatterjee D, Sabo E, Tavares R, Resnick MB. Inverse association between Raf kinase inhibitory protein and signal transducers and activators of transcription 3 expression in gastric adenocarcinoma patients: implications for clinical outcome. Clin Cancer Res. 2008;14(10):2994–3001.

    Article  CAS  PubMed  Google Scholar 

  99. Hagan S, Al Mulla F, Mallon E, Oien K, Ferrier R, Gusterson B, et al. Reduction of Raf-1 kinase inhibitor protein expression correlates with breast cancer metastasis. Clin Cancer Res. 2005;11(20):7392–7.

    Article  CAS  PubMed  Google Scholar 

  100. Lee J, Lee J, Farquhar KS, Yun J, Frankenberger CA, Bevilacqua E, et al. Network of mutually repressive metastasis regulators can promote cell heterogeneity and metastatic transitions. Proc Natl Acad Sci U S A. 2014;111(3):E364–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Ying Y, Wang YJ, Huang XY, Sun YM, Zhang JB, Li MQ, et al. Oncogenic HOXB8 is driven by MYC-regulated super-enhancer and potentiates colorectal cancer invasiveness via BACH1. Oncogene. 2020;39(5):1004–17.

    Article  CAS  PubMed  Google Scholar 

  102. Jordan CT, Guzman ML, Noble M. Cancer stem cells. N Engl J Med. 2006;355(12):1253–61.

    Article  CAS  PubMed  Google Scholar 

  103. Eramo A, Lotti F, Sette G, Pilozzi E, Biffoni M, Di Virgilio A, et al. Identification and expansion of the tumorigenic lung cancer stem cell population. Cell Death Differ. 2008;15(3):504–14.

    Article  CAS  PubMed  Google Scholar 

  104. Wang Q, Liang N, Yang T, Li Y, Li J, Huang Q, et al. DNMT1-mediated methylation of BEX1 regulates stemness and tumorigenicity in liver cancer. J Hepatol. 2021;75(5):1142–53.

    Article  CAS  PubMed  Google Scholar 

  105. Li C, Heidt DG, Dalerba P, Burant CF, Zhang L, Adsay V, et al. Identification of pancreatic cancer stem cells. Cancer Res. 2007;67(3):1030–7.

    Article  CAS  PubMed  Google Scholar 

  106. Jiang P, Li F, Liu ZL, Hao SY, Gao J, Li SQ. BTB and CNC homology 1 (Bach1) induces lung cancer stem cell phenotypes by stimulating CD44 expression. Respir Res. 2021;22(1):320.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Yang L, Shi P, Zhao G, Xu J, Peng W, Zhang J, et al. Targeting cancer stem cell pathways for cancer therapy. Signal Transduct Target Ther. 2020;5(1):8.

    Article  PubMed  PubMed Central  Google Scholar 

  108. Nakanome A, Brydun A, Matsumoto M, Ota K, Funayama R, Nakayama K, et al. Bach1 is critical for the transformation of mouse embryonic fibroblasts by Ras(V12) and maintains ERK signaling. Oncogene. 2013;32(27):3231–45.

    Article  CAS  PubMed  Google Scholar 

  109. Huang XD, Zheng J, Li J, Che X, Tan WL, Tan W, et al. Functional role of BTB and CNC Homology 1 gene in pancreatic cancer and its association with survival in patients treated with gemcitabine. Theranostics. 2018;8(12):3366–79.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Sanchez G, Bittencourt D, Laud K, Barbier J, Delattre O, Auboeuf D, et al. Alteration of cyclin D1 transcript elongation by a mutated transcription factor up-regulates the oncogenic D1b splice isoform in cancer. Proc Natl Acad Sci U S A. 2008;105(16):6004–9.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  111. Liu S, Duan K, Zhang X, Cao X, Wang X, Meng F, et al. Circ_0081001 down-regulates miR-494-3p to enhance BACH1 expression and promotes osteosarcoma progression. Aging (Albany NY). 2021;13(13):17274–84.

    Article  CAS  PubMed  Google Scholar 

  112. Fang Y, Long F. Circular RNA circ_0000337 contributes to osteosarcoma via the miR-4458/BACH1 pathway. Cancer Biomark. 2020;28(4):411–9.

    Article  CAS  PubMed  Google Scholar 

  113. Huang B, Cui DJ, Yan F, Yang LC, Zhang MM, Zhao X. Circ_0087862 promotes the progression of colorectal cancer by sponging miR-142-3p and up-regulating BACH1 expression. Kaohsiung J Med Sci. 2021;37(12):1048–57.

    Article  CAS  PubMed  Google Scholar 

  114. Cui D, Ni C. LncRNA lnc712 promotes tumorigenesis in hepatocellular carcinoma by targeting miR-142-3p/Bach1 Axis. Cancer Manag Res. 2020;12:11285–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Wu YH, Liang Y, Li M, Zhang HD. Knockdown of long non-coding RNA SNHG8 suppresses the progression of esophageal cancer by regulating miR-1270/BACH1 axis. Bioengineered. 2022;13(2):3384–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Wang XH, Liu JX, Jiang L, Wei XX, Niu C, Wang R, et al. Bach1 induces endothelial cell apoptosis and cell-cycle arrest through ROS generation. Oxid Med Cell Longev. 2016;2016:6234043.

    Article  PubMed  PubMed Central  Google Scholar 

  117. Nishizawa H, Matsumoto M, Shindo T, Saigusa D, Kato H, Suzuki K, et al. Ferroptosis is controlled by the coordinated transcriptional regulation of glutathione and labile iron metabolism by the transcription factor BACH1. J Biol Chem. 2020;295(1):69–82.

    Article  CAS  PubMed  Google Scholar 

  118. Namgaladze D, Fuhrmann DC, Brüne B. Interplay of Nrf2 and BACH1 in inducing ferroportin expression and enhancing resistance of human macrophages towards ferroptosis. Cell Death Discov. 2022;8(1):327.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Hintze KJ, Katoh Y, Igarashi K, Theil EC. Bach1 repression of ferritin and thioredoxin reductase1 is heme-sensitive in cells and in vitro and coordinates expression with heme oxygenase1, beta-globin, and NADP(H) quinone (oxido) reductase1. J Biol Chem. 2007;282(47):34365–71.

    Article  CAS  PubMed  Google Scholar 

  120. Marro S, Chiabrando D, Messana E, Stolte J, Turco E, Tolosano E, et al. Heme controls ferroportin1 (FPN1) transcription involving Bach1, Nrf2 and a MARE/ARE sequence motif at position -7007 of the FPN1 promoter. Haematologica. 2010;95(8):1261–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Gatti RA, Shaked R, Mohandas TK, Salser W. Human ferritin genes: chromosomal assignments and polymorphisms. Am J Hum Genet. 1987;41(4):654–67.

    CAS  PubMed  PubMed Central  Google Scholar 

  122. Lymboussaki A, Pignatti E, Montosi G, Garuti C, Haile DJ, Pietrangelo A. The role of the iron responsive element in the control of ferroportin1/IREG1/MTP1 gene expression. J Hepatol. 2003;39(5):710–5.

    Article  CAS  PubMed  Google Scholar 

  123. Hassannia B, Wiernicki B, Ingold I, Qu F, Van Herck S, Tyurina YY, et al. Nano-targeted induction of dual ferroptotic mechanisms eradicates high-risk neuroblastoma. J Clin Invest. 2018;128(8):3341–55.

    Article  PubMed  PubMed Central  Google Scholar 

  124. Yang WS, SriRamaratnam R, Welsch ME, Shimada K, Skouta R, Viswanathan VS, et al. Regulation of ferroptotic cancer cell death by GPX4. Cell. 2014;156(1–2):317–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Chu B, Kon N, Chen D, Li T, Liu T, Jiang L, et al. ALOX12 is required for p53-mediated tumour suppression through a distinct ferroptosis pathway. Nat Cell Biol. 2019;21(5):579–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Nishizawa H, Yamanaka M, Igarashi K. Ferroptosis: regulation by competition between NRF2 and BACH1 and propagation of the death signal. FEBS J. 2022;290(7):1688–704.

    Article  PubMed  Google Scholar 

  127. Tesfay L, Paul BT, Konstorum A, Deng Z, Cox AO, Lee J, et al. Stearoyl-CoA desaturase 1 protects ovarian cancer cells from ferroptotic cell death. Cancer Res. 2019;79(20):5355–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Jiang YQ, Mao C, Yang R, Yan B, Shi Y, Liu XL, et al. EGLN1/c-Myc induced lymphoid-specific helicase inhibits ferroptosis through lipid metabolic gene expression changes. Theranostics. 2017;7(13):3293–305.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Wang C, Shi M, Ji J, Cai Q, Zhao QF, Jiang JL, et al. Stearoyl-CoA desaturase 1 (SCD1) facilitates the growth and anti-ferroptosis of gastric cancer cells and predicts poor prognosis of gastric cancer. Aging (Albany NY). 2020;12(15):15374–91.

    Article  CAS  PubMed  Google Scholar 

  130. Ubellacker JM, Tasdogan A, Ramesh V, Shen B, Mitchell EC, Martin Sandoval MS, et al. Lymph protects metastasizing melanoma cells from ferroptosis. Nature. 2020;585(7823):113–8.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  131. Magtanong L, Ko PJ, To M, Cao JY, Forcina GC, Tarangelo A, et al. Exogenous monounsaturated fatty acids promote a ferroptosis-resistant cell state. Cell Chem Biol. 2019;26(3):420-32 e9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Xie XF, Tian LS, Zhao Y, Liu F, Dai SY, Gu XL, et al. BACH1-induced ferroptosis drives lymphatic metastasis by repressing the biosynthesis of monounsaturated fatty acids. Cell Death Dis. 2023;14(1):48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Balan M, Pal S. A novel CXCR3-B chemokine receptor-induced growth-inhibitory signal in cancer cells is mediated through the regulation of Bach1 protein and Nrf2 protein nuclear translocation. J Biol Chem. 2014;289(6):3126–37.

    Article  CAS  PubMed  Google Scholar 

  134. Nie E, Jin X, Wu WN, Yu TF, Zhou X, Zhi TL, et al. BACH1 promotes temozolomide resistance in glioblastoma through antagonizing the function of p53. Sci Rep. 2016;6:39743.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  135. Feng M, Wang J, Sun M, Li GL, Li BX, Zhang H. 3-Methyladenine but not antioxidants to overcome BACH2-mediated bortezomib resistance in mantle cell lymphoma. Cancer Cell Int. 2021;21(1):279.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Wheaton WW, Weinberg SE, Hamanaka RB, Soberanes S, Sullivan LB, Anso E, et al. Metformin inhibits mitochondrial complex I of cancer cells to reduce tumorigenesis. Elife. 2014;3:e02242.

    Article  PubMed  PubMed Central  Google Scholar 

  137. Liu X, Romero IL, Litchfield LM, Lengyel E, Locasale JW. Metformin targets central carbon metabolism and reveals mitochondrial requirements in human cancers. Cell Metab. 2016;24(5):728–39.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Zhou J, Liu ZM, Zhang HB, Lei TY, Liu JH, Zhao YQ, et al. BACH1: a potential predictor of survival in early-stage lung adenocarcinoma. Oxid Med Cell Longev. 2022;2022:3921095.

    PubMed  PubMed Central  Google Scholar 

  139. Li L, Sun F, Han L, Liu X, Xiao Y, Gregory AD, et al. PDLIM2 repression by ROS in alveolar macrophages promotes lung tumorigenesis. JCI Insight. 2021;6(5):e144394.

    Article  PubMed  PubMed Central  Google Scholar 

  140. Yuan F, Cong Z, Cai X, Zhu J, Yuan L, Wang Y, et al. BACH1 as a potential target for immunotherapy in glioblastomas. Int Immunopharmacol. 2022;103:108451.

    Article  CAS  PubMed  Google Scholar 

  141. Elbasateeny SS, Yassin MA, Mokhtar MM, Ismail AM, Ebian HF, Hussein S, et al. Prognostic implications of MALAT1 and BACH1 expression and their correlation with CTCs and Mo-MDSCs in triple negative breast cancer and surgical management options. Int J Breast Cancer. 2022;2022:8096764.

    Article  PubMed  PubMed Central  Google Scholar 

  142. Casares L, García V, Garrido Rodríguez M, Millán E, Collado JA, García-Martín A, et al. Cannabidiol induces antioxidant pathways in keratinocytes by targeting BACH1. Redox Biol. 2020;28:101321.

    Article  CAS  PubMed  Google Scholar 

  143. Casares L, Unciti Broceta JD, Prados ME, Caprioglio D, Mattoteia D, Higgins M, et al. Isomeric O-methyl cannabidiolquinones with dual BACH1/NRF2 activity. Redox Biol. 2020;37:101689.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Attucks OC, Jasmer KJ, Hannink M, Kassis J, Zhong Z, Gupta S, et al. Induction of heme oxygenase I (HMOX1) by HPP-4382: a novel modulator of Bach1 activity. PLoS One. 2014;9(7):e101044.

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  145. Palani CD, Zhu X, Alagar M, Attucks OC, Pace BS. Bach1 inhibitor HPP-D mediates γ-globin gene activation in sickle erythroid progenitors. Blood Cells Mol Dis. 2024;104:102792.

    Article  CAS  PubMed  Google Scholar 

  146. Zhang J, Hu S, Zhao C, Zhou Y, Zhang L, Liu H, et al. Genome-scale CRISPR knockout screening identifies BACH1 as a key regulator of aflatoxin B1-induced oxidative damage. Antioxidants (Basel). 2022;11(9):1787.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Nataraja S, Singh M, Demes S, Olson L, Stanwix J, Biddle M, et al. ML-0207/ASP8731: a novel BACH1 inhibitor that induces fetal hemoglobin in treatment of sickle cell disease. Blood. 2021;138(Supplement 1):854–54.

    Article  Google Scholar 

  148. Belcher JD, Nataraja S, Abdulla F, Zhang P, Chen C, Nguyen J, et al. The BACH1 inhibitor ASP8731 inhibits inflammation and vaso-occlusion and induces fetal hemoglobin in sickle cell disease. Front Med (Lausanne). 2023;10:1101501.

    Article  PubMed  Google Scholar 

  149. Moreno R, Casares L, Higgins M, Ali KX, Honda T, Wiel C, et al. Biotinylation of an acetylenic tricyclic bis(cyanoenone) lowers its potency as an NRF2 activator while creating a novel activity against BACH1. Free Radic Biol Med. 2022;191:203–11.

    Article  CAS  PubMed  Google Scholar 

  150. Bissell DM, Anderson KE, Bonkovsky HL. Porphyria. N Engl J Med. 2017;377(9):862–72.

    Article  CAS  PubMed  Google Scholar 

  151. Casares L, Moreno R, Ali KX, Higgins M, Dayalan Naidu S, Neill G, et al. The synthetic triterpenoids CDDO-TFEA and CDDO-Me, but not CDDO, promote nuclear exclusion of BACH1 impairing its activity. Redox Biol. 2022;51:102291.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Aletaha M, Mansoori B, Mohammadi A, Fazeli M, Baradaran B. Therapeutic effects of bach1 siRNA on human breast adenocarcinoma cell line. Biomed Pharmacother. 2017;88:34–42.

    Article  CAS  PubMed  Google Scholar 

  153. Davudian S, Shajari N, Kazemi T, Mansoori B, Salehi S, Mohammadi A, et al. BACH1 silencing by siRNA inhibits migration of HT-29 colon cancer cells through reduction of metastasis-related genes. Biomed Pharmacother. 2016;84:191–8.

    Article  CAS  PubMed  Google Scholar 

  154. Yu Y, Feng C, Kuang J, Guo LX, Guan HX. Metformin exerts an antitumoral effect on papillary thyroid cancer cells through altered cell energy metabolism and sensitized by BACH1 depletion. Endocrine. 2022;76(1):116–31.

    Article  CAS  PubMed  Google Scholar 

  155. Shang FT, Hui LL, An XS, Zhang XC, Guo SG, Kui Z. ZnPPIX inhibits peritoneal metastasis of gastric cancer via its antiangiogenic activity. Biomed Pharmacother. 2015;71:240–6.

    Article  CAS  PubMed  Google Scholar 

  156. Fang J, Sawa T, Akaike T, Akuta T, Sahoo SK, Khaled G, et al. In vivo antitumor activity of pegylated zinc protoporphyrin: targeted inhibition of heme oxygenase in solid tumor. Cancer Res. 2003;63(13):3567–74.

    CAS  PubMed  Google Scholar 

  157. Tanaka S, Akaike T, Fang J, Beppu T, Ogawa M, Tamura F, et al. Antiapoptotic effect of haem oxygenase-1 induced by nitric oxide in experimental solid tumour. Br J Cancer. 2003;88(6):902–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Doi K, Akaike T, Fujii S, Tanaka S, Ikebe N, Beppu T, et al. Induction of haem oxygenase-1 nitric oxide and ischaemia in experimental solid tumours and implications for tumour growth. Br J Cancer. 1999; 80(12):1945-54.

  159. Wang S, Hannafon BN, Wolf RF, Zhou J, Avery JE, Wu J, et al. Characterization of docosahexaenoic acid (DHA)-induced heme oxygenase-1 (HO-1) expression in human cancer cells: the importance of enhanced BTB and CNC homology 1 (Bach1) degradation. J Nutr Biochem. 2014;25(5):515–25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. El Deek HEM, Ahmed AM, Mohammed RAA. Aberration of Nrf2-Bach1 pathway in colorectal carcinoma; role in carcinogenesis and tumor progression. Ann Diagn Pathol. 2019;38:138–44.

    Article  PubMed  Google Scholar 

  161. Ou XQ, Gao GF, Bazhabayi M, Zhang KM, Liu F, Xiao XS. MALAT1 and BACH1 are prognostic biomarkers for triple-negative breast cancer. J Cancer Res Ther. 2019;15(7):1597–602.

    Article  CAS  PubMed  Google Scholar 

  162. Lee U, Frankenberger C, Yun J, Bevilacqua E, Caldas C, Chin SF, et al. A prognostic gene signature for metastasis-free survival of triple negative breast cancer patients. PLoS One. 2013;8(12):e82125.

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  163. Kuo CL, Ponneri Babuharisankar A, Lin YC, Lien HW, Lo YK, Chou HY, et al. Mitochondrial oxidative stress in the tumor microenvironment and cancer immunoescape: foe or friend? J Biomed Sci. 2022;29(1):74.

    Article  PubMed  PubMed Central  Google Scholar 

  164. Wan Y, Cohen J, Szenk M, Farquhar KS, Coraci D, Krzysztoń R, et al. Nonmonotone invasion landscape by noise-aware control of metastasis activator levels. Nat Chem Biol. 2023;19(7):887–99.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Igarashi K, Nishizawa H, Saiki Y, Matsumoto M. The transcription factor BACH1 at the crossroads of cancer biology: From epithelial-mesenchymal transition to ferroptosis. J Biol Chem. 2021;297(3):101032.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Jiang TY, Pan YF, Wan ZH, Lin YK, Zhu B, Yuan ZG, et al. PTEN status determines chemosensitivity to proteasome inhibition in cholangiocarcinoma. Sci Transl Med. 2020;12(562):eaay0152.

    Article  CAS  PubMed  Google Scholar 

  167. Jiang TY, Feng XF, Fang Z, Cui XW, Lin YK, Pan YF, et al. PTEN deficiency facilitates the therapeutic vulnerability to proteasome inhibitor bortezomib in gallbladder cancer. Cancer Lett. 2021;501:187–99.

    Article  CAS  PubMed  Google Scholar 

  168. Zhu XN, Wei YS, Yang Q, Liu HR, Zhi Z, Zhu D, et al. FBXO22 promotes leukemogenesis by targeting BACH1 in MLL-rearranged acute myeloid leukemia. J Hematol Oncol. 2023;16(1):9.

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This research was supported by grants from the National Natural Science Foundation of China No. U23A20451 (L.X.), No. 82273310 (L.X.), No. 82173313 (W.H.), No. 82372917 (W.H.), the Natural Science Foundation of Hubei Province 2022CFA016 (L.X.), and the Basic Research Support Program of Huazhong University of Science and Technology 2023BR038 (L.X.).

Author information

Authors and Affiliations

Authors

Contributions

D.H. and S.L. collected the related references and drafted the manuscript. Z.Z. collected and analyzed the related data. X.L., Y.W. and M.S. edited the figures. J.J., J.Z. and Z.W. edited the tables. X.C., B.Z., X.X., S.W. and S.X. revised the manuscript. L.X., W.H. and Y.W. designed the study. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Yufei Wang, Wenjie Huang or Limin Xia.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher's Note

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

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, 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 changes were made. 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/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hu, D., Zhang, Z., Luo, X. et al. Transcription factor BACH1 in cancer: roles, mechanisms, and prospects for targeted therapy. Biomark Res 12, 21 (2024). https://doi.org/10.1186/s40364-024-00570-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40364-024-00570-4

Keywords