Skip to main content

Multifunctional acyltransferase HBO1: a key regulatory factor for cellular functions

Abstract

HBO1, also known as KAT7 or MYST2, is a crucial histone acetyltransferase with diverse cellular functions. It typically forms complexes with protein subunits or cofactors such as MEAF6, ING4, or ING5, and JADE1/2/3 or BRPF1/2/3, where the BRPF or JADE proteins serve as the scaffold targeting histone H3 or H4, respectively. The histone acetylation mediated by HBO1 plays significant roles in DNA replication and gene expression regulation. Additionally, HBO1 catalyzes the modification of proteins through acylation with propionyl, butyryl, crotonyl, benzoyl, and acetoacetyl groups. HBO1 undergoes ubiquitination and degradation by two types of ubiquitin complexes and can also act as an E3 ubiquitin ligase for the estrogen receptor α (ERα). Moreover, HBO1 participates in the expansion of medullary thymic epithelial cells (mTECs) and regulates the expression of peripheral tissue genes (PTGs) mediated by autoimmune regulator (AIRE), thus inducing immune tolerance. Furthermore, HBO1 influences the renewal of hematopoietic stem cells and the development of neural stem cells significantly. Importantly, the overexpression of HBO1 in various cancers suggests its carcinogenic role and potential as a therapeutic target. This review summarizes recent advancements in understanding HBO1’s involvement in acylation modification, DNA replication, ubiquitination, immunity, and stem cell renewal.

Introduction

HBO1, also known as KAT7 or MYST2, belongs to the MYST acetyltransferase family and is primarily responsible for histone H3 and H4 acetylation [1,2,3,4]. It exerts diverse roles in crucial cellular processes, including DNA replication and repair, gene transcription, protein ubiquitination, immune regulation, stem cell pluripotency and self-renewal maintenance, and embryonic development. Initially identified as an interaction partner for the largest subunit ORC1 of the Origin Recognition Complex (ORC) [4], HBO1 is a 611-amino-acid protein comprising a unique N-terminal serine-rich region (22% in aa 1–167) and a conserved C-terminal domain of 270 amino acids specific to the MYST protein family. The MYST domain encompasses an acetyl-CoA binding region, facilitating histone acetylation, and an atypical C2HC zinc finger [5]. Studies suggest that the N-terminal domain (NTD) of HBO1 harbors a transcriptional inhibitory domain, capable of downregulating androgen receptor expression and sequestering an essential cofactor from the NF-κB transcription complex, thereby reducing NF-κB activity [6, 7]. However, the crystal structure of NTD remains largely unexplored. Intriguingly, HBO1 (311–611) displays higher activity than HBO1 (1–611) in vitro, implying a potential negative regulation of the histone acetyltransferase (HAT) activity of the C-terminal MYST domain by the NTD of HBO1 [8]. Therefore, protein modifications such as phosphorylation or cofactor/substrate binding may serve as pathways to alleviate this negative regulation and fully unleash the HAT activity of HBO1.

Histone acetylation stands as a central function of HBO1. Acetylation, a prevalent protein modification in cells, holds pivotal roles in diverse cellular processes including cell proliferation, gene transcription, and signal transduction. Specifically, in the context of genomic histones, acetylation induces structural relaxation, thereby exposing DNA sites for crucial processes such as replication and transcription [9, 10]. The addition or removal of acetyl groups from histone lysine residues is mediated by lysine acetyltransferases (KATs) and deacetyltransferases, respectively. As a member of the KAT family, the functional mechanism and specific impact of HBO1 are still not fully elucidated. Moreover, the observed high expression of HBO1 in various cancer cells suggests its potential as a target for cancer treatment [11]. In summary, our understanding of HBO1 remains incomplete, underscoring its substantial research value and warranting further investigation.

This article presents a comprehensive overview of HBO1 as a multifunctional acyltransferase, which is a key factor of cell functions, emphasizing its regulatory roles in DNA replication and participation in DNA repair. Additionally, HBO1’s involvement in ubiquitination, where it can be ubiquitinated itself and also acts as a ubiquitin ligase, is explored. The crucial role of HBO1 in immune regulation and T-cell development is highlighted, alongside its contribution to the regulation of stem cell pluripotency and self-renewal. Moreover, the article delves into the association between HBO1 and various diseases, including malignant tumors and chronic obstructive pulmonary disease, with the aim of providing a fresh perspective for a comprehensive and systematic understanding of the multifaceted functions and mechanisms of the acyltransferase HBO1.

HBO1 is a multifunctional acyltransferase

As a member of the histone acetyltransferase family, HBO1 typically forms protein complexes with various cofactors or partner proteins, serving as the core catalytic subunit to exert its function. The identified HBO1 protein complex primarily consists of HBO1, ING4/5, MEAF6, and BRPF1/2/3 or JADE1/2/3 [12]. ING4 and ING5, belonging to the ING tumor suppressor family, regulate the cell cycle and apoptosis [13, 14]. H3K4me3 is a mark that is found near the transcription start site (TSS) of actively transcribed genes [15]. The N-terminal domain of ING4/5 forms homodimers or heterodimers, recognizing the histone H3 lysine 4 trimethylation (H3K4me3) site through the C-terminal PHD domain, subsequently recruiting the HBO1 complex to promote histone acetylation [13, 14, 16,17,18,19]. BRPF or JADE serve as scaffold proteins for the HBO1 complex, enhancing its acetylation function [2, 16, 20, 21]. BRPF primarily targets histone H3, while JADE predominantly targets H4 [3, 17]. The BRPF protein typically comprises an N-terminal PHD-C2H2 zinc finger-PHD domain (PZP), a central bromodomain, and a C-terminal PWWP domain [17, 22]. When HBO1 binds to BRPF1, the complex’s action on chromatin acetylation is confined to histone H3, resulting in the increases of H3K23ac and H3K14ac [17]. The interaction between BRPF2 and HBO1 is primarily localized to a short N-terminal region of the former and the MYST domain of the latter. The simultaneous binding of both to histone proteins allows for the correct positioning of the N-terminal tails of the histones at the acetyltransferase active site of HBO1, thereby enhancing the acetyltransferase activity of HBO1. The binding of the N-terminal region of BRPF2 may also stabilize HBO1 in a more physiological conformation, thereby enhancing its interaction with histone substrates, ultimately boosting the acetyltransferase activity of HBO1 [21]. Additionally, compared with HBO1 alone, the BRPF3 complex enhances the levels of H3K9ac, H3K14ac, and H4K16ac [22]. JADE-1 is the most prevalent member of the JADE family and serves as the essential cofactor for the HBO1 acetyltransferase complex [23]. Similar to BRPF, JADE also possesses two PHD domains, capable of synergistic action with ING4/5 to enhance HBO1-mediated histone acetylation [2, 14, 16]. Additionally, phosphorylation of JADE1 during the cell cycle may regulate the removal of HBO1 complexes from chromatin, facilitating histone deacetylation during mitosis [24].

Remarkably, these subunits are not exclusive to the HBO1 complex, and some subunits are involved in the composition of other HAT complexes. Studies have shown that JADE1 physically associates to Tip60 HAT and enhances the acetylation targeting H4 [25]. In addition, BRPF, ING5, and MEAF6 are involved in the composition of MOZ/MORF complex. Similar to its role in HBO1 complex, BRPF acts as a scaffold in MOZ/MORF complex to connect MOZ/MORF with ING5 and MEAF6. Additionally, BRPF is also important for activating the MOZ/MORF complex. Just like its function in HBO1 complex, BRPF is able to upregulate the acetyltransferase activity of MOZ/MORF [26]. ING5 targets H3K4me3 at the promoter of active transcription region and recruits the MOZ/MORF complex for histone acetylation [27]. In conclusion, by studying these subunits shared between the HBO1 complex and other HAT complexes, we have a clearer understanding that their roles in the HBO1 complex are universal.

HBO1 significantly influences gene expression, signal transduction, and cellular growth by affecting acetylation function. HBO1 is located at the transcriptional start site (TSS) of active genes, where it directly acetylates histones, thereby regulating gene transcription. The strength of acetylation signal HBO1 signaling correlates closely with gene expression levels [14, 28]. For instance, HBO1 is recruited to the promoter of glycolysis-related genes by the transcription factor SIX1, facilitating their transcription through H4K5 acetylation. HBO1 knockdown or knockout leads to decreased glucose uptake, pyruvate levels, lactate production, adenosine triphosphate (ATP) levels, and extracellular acidification rate (ECAR), while oxygen consumption rate (OCR) increases. In cancer cells, SIX1 enhances the Warburg effect via this pathway [29]. Additionally, HBO1 promotes CTNNB1 gene transcription by acetylating H3K14, H4K8, and H4K12, thereby activating the Wnt/β-catenin signaling pathway [30]. In leukemia, HBO1 upregulates HOXA9 and HOXA10 expression through H3K14 acetylation, maintaining leukemia stem cell characteristics [31, 32]. Moreover, the NUP98–HBO1 fusion protein induces abnormal histone acetylation, leading to increased acetylation levels at the HOXA9 promoter on H4K8, H4K12, and H3K14 and activation of carcinogenic features in chronic myeloid mononuclear leukemia (CMML) [33]. HBO1 also participates in p53-mediated transcriptional activation of p21/CDKN1A and GADD45A through its HAT active site [34]( Fig. 1).

Fig. 1
figure 1

HBO1 complexes and gene expression. HBO1 forms complexes with other proteins to acetylate histones and promote the expression of multiple genes. The HBO1 complex with BRPF as the scaffold mainly targets histone H3 acetylation, while the HBO1 complex with JADE as the scaffold mainly targets histone H4 acetylation. The histone acetylation will promote the expression of many genes including SIX1, CNTTB1, HOXA9, HOXA10, p21/CDKN1A, and GADD45A

However, recent research on BRPF2–HBO1 and JADE1–HBO1 complexes has challenged the traditional view of HBO1’s substrate specificity. It has been found that HBO1 is not restricted to histone H3 and H4 acetylation, as previously believed. In addition to acetylation, HBO1 can catalyze propionylation, butyrylation, and crotonylation both in vivo and in vitro. Furthermore, HBO1 can extend its catalytic activity to histone H2. Interestingly, the specific targeting function of BRPF and JADE appears to be less distinct, leading to considerable overlap in catalytic sites on histones H3 and H4 within the HBO1 complex [28]. This intriguing observation warrants further investigation to fully understand its implications. Additionally, lysine benzoylation (Kbz) has emerged as a novel posttranslational modification involved in chromatin remodeling, transcriptional regulation, and tumor growth. HBO1 has been identified as a participant in Kbz in mammals, further expanding its functional repertoire [35]. Lastly, HBO1 is also involved in histone acetyl-acetylation (Kacac) processes, adding another layer of complexity to its regulatory mechanisms [36].

In summary, HBO1 emerges as a multifunctional acyltransferase with significant involvement in histone acetylation and potentially nonhistone substrates. However, the interaction mechanism between HBO1 and the MEAF6 protein within the HBO1 complex remains unexplored, along with its biological implications. Furthermore, the distinctions among the BRPF1/2/3 and JADE1/2/3 families have not been systematically elucidated. It remains unclear whether their anchoring sites on histones, as scaffold proteins, exhibit consistency across these families. These unresolved questions underscore the need for further research in this field.

HBO1 has a regulatory effect on DNA replication and is involved in DNA repair after ultraviolet irradiation

Eukaryotic DNA replication is a continuous process encompassing various tightly regulated steps, including replication origin recognition, pre-replication complex loading, and replication fork initiation. Key factors such as ORC1, CDT1, CDC6, and MCM are essential for orchestrating these steps [37]. HBO1 is widely recognized for its indispensable role in DNA replication, as evidenced by numerous studies.

During the assembly of the pre-replication complex (pre-RC), ORC initially recruits CDC6 and CDT1 to the replication origin in the early stage of the G1 phase [38]. Subsequently, CDT1 directly interacts with HBO1, facilitating the recruitment of HBO1 to the replication origin [39]. The HBO1–JADE complex promotes H4K5/8/12 acetylation, leading to the relaxation of chromatin conformation. This process facilitates the loading of the MCM complex onto replication origins and promotes the assembly of pre-replication complexes [39, 40]. Depletion of Xenopus laevis HBO1 in Xenopus laevis egg extracts results in the loss of MCM2-7 chromatin binding and elimination of DNA replication, indicating the necessity of HBO1 for MCM2-7 complex chromatin binding during the G1 phase, a critical step in DNA replication licensing [41]. Studies have demonstrated that HBO1 and MCM2 functions rely on the N-terminal domain of MCM2 and the C2HC zinc finger of HBO1 [5]. HBO1 directly interacts with CDT1 and enhances CDT1-dependent replication, although it is not indispensable for CDT1’s association with replication origins [39]. In the context of stress response, phosphorylation of CDT1 can inhibit the recruitment of HBO1 histone acetylase, consequently blocking replication licensing [42]. Moreover, the BRPF3 scaffold specifically guides HBO1 to H3K14ac, promoting the loading of CDC45 to activate DNA replication in the S phase [1]. Thus, through collaboration with different scaffolds, HBO1-mediated chromatin acetylation facilitates two consecutive steps in replication initiation: licensing and activation. In the S phase, the regulatory protein Geminin prevents the second round of DNA replication by inhibiting the essential replication factor CDT1. Notably, HBO1 may be inhibited to affect this process, because Geminin does not inhibit MCM loading through simple spatial interference of the CDT1–MCM2-7 interaction, but plays a role through nonspatial mechanisms [40, 43, 44]. HBO1 has been shown to acetylate ORC2, MCM2, CDC6, and Geminin in vitro, indicating its potential role in regulating the initiation of DNA replication by acetylating these factors [41]( Fig. 2).

Fig. 2
figure 2

HBO1-mediated histone acetylation in DNA replication. HBO1-mediated histone acetylation promotes G1 phase DNA replication licensing and S phase DNA replication activation. In the G1 phase, the HBO1–JADE complex promotes the acetylation of H4K5/8/12 to relax the chromatin conformation, and facilitates the loading of the MCM complex to the replication starting point, promoting the assembly of the pre-replication complex. In S phase, the HBO1–BRPF3 complex specifically directs H3K14ac, thereby promoting CDC45 loading to activate S-phase DNA replication. Additionally, the regulatory protein Geminin prevents the second round of DNA replication by inhibiting the basic replication factor CDT1, possibly by inhibiting HBO1

The role of HBO1 in the origin of replication is also subject to regulation by other factors. For instance, FAD24 (adipocyte differentiation factor 24) has been identified to interact with HBO1 during the process of pre-adipocytes transitioning into adipocytes through mitotic clonal expansion (MCE). FAD24 co-localizes with HBO1 in chromatin during pre-replication complex assembly. Inhibition of FAD24 expression during adipocyte differentiation leads to reduced recruitment of HBO1 to the origin of DNA replication, while knockout of the HBO1 gene inhibits MCE and adipogenesis. These findings suggest that FAD24 acts as an auxiliary factor in recruiting HBO1 to the origin of DNA replication [45,46,47]. Furthermore, under conditions of hyperosmotic stress, HBO1 can directly bind to p53, thereby inhibiting HBO1-HAT activity and subsequently impeding the loading of the MCM2-7 complex. This results in the stalling of pre-replication complex assembly. Treatment with hydroxyurea (HU), which blocks DNA replication fork progression, leads to downregulation of p53-dependent HBO1 activity [8, 11]. Interestingly, after the activity of HBO1-HAT decreased, the MCM2-7 complex still binds to chromatin. One possible explanation is that HBO1 may have other functions in the S phase, although the specific mechanisms remain unclear.

The experiments indicating that “HBO1 plays an important role in DNA replication and cell proliferation” have predominantly utilized human tumor cell lines or other immortalized cells, such as HeLa cells, C33A cells, MCF7 cells, Saos2 cells, A549 cells, and 293 T cells. Interestingly, experiments conducted using mouse embryos revealed that fibroblasts lacking HBO1, isolated from embryos, can proliferate normally and exhibit normal MCM localization. Moreover, embryos lacking HBO1 can progress through normal development up to the gastrulation stage, with developmental abnormalities and mortality occurring thereafter. Specifically, organs such as blood vessels and mesenchyme fail to differentiate and develop normally in embryos lacking HBO1, leading to a significant decrease in total RNA extraction from cells. Notably, crucial regulatory genes such as Notch1 during gastrulation development were undetected in HBO1-deficient embryos. These findings suggest that embryonic death may be attributed to inadequate gene expression products rather than defects in cell proliferation. It suggests that HBO1’s primary role in embryonic development appears to be as a gene expression activator rather than a participant in DNA replication and cell proliferation [48]. Subsequent experiments conducted by the same research team using immortalized cells such as HeLa and 293T cells revealed that loss of HBO1 had minimal impact on cell proliferation and DNA replication. Instead, loss of HBO1 primarily affected genes involved in cell adhesion, resulting in reduced cell adhesion, particularly in 293T cells, with relatively minor effects on other cellular processes [49]. Additionally, research has demonstrated that HBO1 plays a specific role at the centromere, where it interacts with M18BP1 to positively regulate the assembly of CENP-A and counteract heterochromatin-mediated centromere inactivation [50]. In summary, there remains significant scope for research and clarification regarding HBO1’s role in cell proliferation and DNA replication, as well as its overall impact. Questions persist regarding whether HBO1’s function is truly cell-intrinsic and the underlying mechanism of its identification, prompting deeper inquiry into its biological significance.

HBO1 participates in DNA repair following ultraviolet irradiation, primarily through its involvement in global genome nucleotide excision repair (GG-NER) [51]. Upon UV-induced DNA damage, the protein DDB2 recognizes sites of cyclobutyl purine dimers (CPD) and swiftly localizes to the damaged site [52, 53]. However, the densely packed chromatin structure poses a barrier to the entry of repair proteins. Therefore, histone modifications, such as acetylation and ATP-dependent chromatin remodeling, are crucial during NER to overcome these structural obstacles [54]. HBO1 plays a key role in this process by phosphorylation at Ser50 and Ser53 by ATM/ARM, facilitating its binding to DDB2 and subsequent histone acetylation [51]. Additionally, HBO1 interacts with chromatin remodeling proteins ACF1 and SNF2H [55, 56], aiding in the maintenance of ACF1–SNF2H at the damage site to induce chromatin remodeling [51]. Furthermore, HBO1 mediates the phosphorylation of methyltransferase MLL1 at Ser516, leading to its localization at UV damage sites and subsequent methylation of histone H3K4 [57]. BAZ1A, a subunit of the chromatin remodeling factor ISWI family, targets trimethylated histone H3K4 (H3K4me3), disrupting the interaction between DNA and histones and facilitating the recruitment of NER factors, including XPC, for DNA repair [51, 58, 59]. This coordinated action underscores the critical role of HBO1 in orchestrating chromatin modifications essential for efficient GG-NER (Fig. 3).

Fig. 3
figure 3

HBO1 involvement in DNA repair after UV irradiation. HBO1 plays a crucial role in the DNA repair process after UV irradiation, primarily associated with global genome nucleotide excision repair (GG-NER). Following UV-induced DNA damage, DDB2 recognizes the cyclobutyl pyrimidine dimer (CPD) site, and phosphorylated HBO1 binds to DDB2, mediating histone acetylation. HBO1 also maintains the chromatin remodeling agent ACF1–SNF2H at the damage site, inducing chromatin remodeling. Additionally, the methyltransferase MLL1 interacts with HBO1 and localizes at the UV damage site to methylate histone H3K4. BAZ1A, a subunit of the SWI/SNF chromatin remodeling factor, targets trimethylated histone H3K4 (H3K4me3). These mechanisms collectively disrupt the interaction between DNA and histones, facilitating the loading of NER factors including XPC for DNA repair

HBO1 can be ubiquitinated and also act as a ubiquitin ligase

Two ubiquitin complexes, SCF (SKP1/Cullin-1/Fbxw15) and CRL4 (DDB1-CUL4A-RBX1), are involved in the ubiquitination of HBO1. Fbxw15 interacts with HBO1 to mediate its ubiquitination, leading to degradation primarily in the cytoplasm despite the presence of histone acetyltransferase activity associated with HBO1 in the nucleus. Lys338 has been identified as the receptor site for SCF-mediated ubiquitination of HBO1. Mitogen-activated protein kinase (MEK1) phosphorylates HBO1, promoting its degradation via the Fbxw15-mediated ubiquitin proteasome pathway. Studies have shown that overexpression of MEK1 increases HBO1 degradation, while MEK1 silencing stabilizes HBO1 acetyltransferase. Knockout of Fbxw15 abolishes MEK1-induced HBO1 degradation, indicating Fbxw15 dependence in this process. Furthermore, MEK1 knockout disrupts the interaction between HBO1 and Cullin1/Fbxw15 and reduces HBO1 ubiquitination in cells, suggesting MEK1’s role in HBO1 phosphorylation and subsequent degradation by Fbxw15-mediated ubiquitination. In mouse lung epithelial cells (MLE-12), endotoxin lipopolysaccharide (LPS) induces HBO1 degradation via MEK1 phosphorylation and the Fbxw15-mediated ubiquitin proteasome pathway, resulting in reduced H3K14ac levels and cell proliferation [60]. Conversely, LPS stimulation in THP-1 monocytes and human primary macrophages leads to increased HBO1 protein levels owing to elevated deubiquitinase USP25 levels, promoting HBO1 deubiquitination and stabilization. USP25-mediated deubiquitination enhances HBO1’s response to the endotoxin-induced inflammatory response, thereby boosting the transcription of interleukin (IL)-1β, IL-6, and IL-10 mediated by HBO1 [61]. Moreover, after UV irradiation-induced DNA damage, HBO1 is degraded by the DDB2-mediated CRL4 complex. Ser50 and Ser53 phosphorylation of HBO1 in an ATM/ATR-dependent manner facilitates its preferential ubiquitination by CRL4DDB2, essential for appropriate cell cycle arrest to complete DNA repair. Mutating Ser50 and Ser53 inhibits HBO1 phosphorylation, leading to failure in repairing DNA damage post-UV irradiation and inhibiting cell proliferation [62] (Fig. 4).

Fig. 4
figure 4

Ubiquitination and degradation of HBO1. HBO1 can undergo ubiquitination by the SCF (SKP1/Cullin-1/Fbxw15) and CRL4 (DDB1-CUL4A-RBX1) complexes. Protein kinase MEK1 phosphorylates HBO1, promoting its degradation via the Fbxw15-mediated ubiquitin proteasome pathway. In mouse lung epithelial cells (MLE-12), LPS induces HBO1 degradation through this pathway. Conversely, LPS stimulation in THP-1 monocytes and human primary macrophages inhibits HBO1 ubiquitination by increasing the level of deubiquitinase USP25 protein, resulting in varying degrees of HBO1 protein elevation. Furthermore, DNA damage caused by UV irradiation leads to the degradation of HBO1 by the DDB2-mediated CRL4 (DDB1–CUL4A–RBX1) complex

Indeed, HBO1 exhibits an intriguing dual role as an E3 ubiquitin ligase, targeting not only itself but also other proteins. In breast cancer, HBO1 functions as an E3 ubiquitin ligase to negatively regulate the stability of estrogen receptor α (ERα) [63]. Its MYST domain possesses E3 ligase activity, facilitating the proteasome-dependent degradation of ERα. Interestingly, estradiol-17β can inhibit HBO1’s E3 ligase activity on ERα in vitro, thereby attenuating ERα ubiquitination, whereas highly active ERα mutants are more susceptible to HBO1’s E3 ligase activity [64]. Whether HBO1 can exert its ubiquitin ligase function on additional proteins remains a topic worthy of further investigation.

HBO1 is essential for immune regulation and T cell development

Thymic epithelial cells (TECs) govern the differentiation and selection of thymic T lymphocytes [65], with medullary thymic epithelial cells (mTECs) being particularly influential in negative selection of autoreactive thymocytes and the differentiation of regulatory T cells (Tregs) [66]. The autoimmune regulator (AIRE) orchestrates the transcription of numerous peripheral tissue genes (PTGs) in mTECs [66, 67]. Deficiencies in AIRE, observed in both humans and mice, lead to impaired expression of relevant PTGs in mTECs, resulting in the escape of autoreactive T cells from the thymus or the failure to induce Treg cells, thereby fostering impaired thymic negative selection and the development of multiple organ autoimmune diseases [68,69,70]. Recent research underscores the pivotal role of HBO1 in thymus development and the mediation of immune tolerance. Notably, highly transcribed HBO1 and abundant H3K14ac are evident across all TEC subsets. HBO1-deficient mice exhibit hypoplastic thymuses in young adulthood, characterized by significantly reduced numbers of thymic epithelial cells, particularly in mTECs. Flow cytometry analyses reveal a diminished proportion and count of AIRE+ cells, suggesting that HBO1 in TECs is crucial for AIRE+ and thymic medulla expansion. Moreover, in mice lacking HBO1, the expression of AIRE-dependent PTGs, including AIRE-induced lung antigen BPIFB9A essential for lung immune tolerance, substantially decreases compared with control groups [71, 72], while AIRE-independent PTGs remain mostly unaffected. Using the small molecule inhibitor WM-3835, which targets HBO1 function, or employing the HBO1 gene deletion system, acute inhibition of HBO1 activity demonstrates no interference with AIRE expression, nor does it alter the AIRE protein level in TECs. However, it does impede the transcriptional activation of AIRE target genes. Mechanistically, HBO1 may enhance chromatin accessibility around the promoter of AIRE-regulated genes through its HAT activity, thereby facilitating the normal transcription of PTGs. In summary, HBO1 plays a critical role in promoting the expansion of mTECs and serves as a major regulator in AIRE-mediated PTG expression, consequently contributing to the induction of immune tolerance [73].

During T cell development, the expression of CD8 genes undergoes regulation through the concerted action of at least five different CD8 enhancers [74]. Recent investigations have highlighted the involvement of BRPF2 and HBO1 in this regulatory process. Specifically, BRPF2 and HBO1 form a complex responsible for H3K14 acetylation of the CD8 locus, with BRPF2 binding to the enhancer and HBO1 binding to the promoter of the CD8 gene. Microarray analysis and other findings have underscored the critical role of the BRPF2–HBO1 complex in acetylating H3K14 on the transcriptional regulatory elements of the CD8 gene, a process necessary for the effective activation of the CD8 gene. Moreover, the BRPF2–HBO1 complex has been shown to directly interact with key regulatory factors involved in CD8 gene activation, such as RUNX family transcription factors and Ikaros [75, 76]. This complex may facilitate chromatin relaxation through H3K14ac, subsequently recruiting transcription complexes to the CD8 enhancer to fully activate the CD8 locus. Importantly, these findings highlight the role of the BRPF2–HBO1 complex in activating CD8 expression rather than merely maintaining its expression [77]. Furthermore, HBO1 has been implicated in regulating the functional activity of CD8+ tissue-resident memory T cells (Trm) and tumor-infiltrating lymphocytes (TIL). SCML4, a transcription factor critical for Trm and TIL survival and activation, has been found to bind to components in the HBO1–BRPF2–ING4 complex through its C-terminal domain. Treatment with a BRPF2 inhibitor (NI-57) significantly reduced the expression of intracellular T cell effector molecules (IFNG and GZMB) in Jurkat cells, while treatment with an H3K14ac deacetylase inhibitor (HDAC-IN-38) notably increased their expression. Mechanistically, SCML4 recruits the HBO1–BRPF2–ING4 complex to mediate H3K14ac, thereby enhancing chromatin accessibility during T cell activation and increasing the expression of relevant genes associated with T cell function [78](Fig. 5).

Fig. 5
figure 5

HBO1 promotes immune-related gene expression via histone acetylation. BRPF2 and HBO1 form complexes that bind to the enhancer and promoter of the CD8 gene, respectively. Together, they acetylate histone H3 at lysine 14 (H3K14ac) at the CD8 locus, leading to full activation of the CD8 gene. The BRPF2–HBO1 complex also interacts directly with key regulatory factors, such as the RUNX family transcription factors and Ikaros, to activate CD8 gene expression. SCML4, a transcription factor crucial for CD8+ resident memory T cells (Trm) and tumor-infiltrating lymphocytes (TIL), recruits the HBO1–BRPF2–ING4 complex to mediate H3K14ac, thereby enhancing chromatin accessibility during T cell activation and increasing the expression of T cell effector molecules, such as interferon-gamma (IFNG) and granzyme B (GZMB)

HBO1 is involved in regulating pluripotency and self-renewal of stem cells

HBO1 is involved in regulating the self-renewal of hematopoietic stem cells. Adult hematopoiesis is a tightly regulated process [79], and HBO1 has emerged as a crucial factor in maintaining hematopoietic stem cells (HSCs). Studies utilizing HBO1 gene-deficient mice have revealed that HBO1 deficiency leads to pancytopenia in both blood and bone marrow within 2–6 weeks of HBO1 gene deletion, ultimately resulting in death due to hematopoietic failure. HBO1-deficient mice exhibit significantly reduced numbers of hematopoietic stem cells and progenitor cells, as well as diminished levels of peripheral blood cells, lymphocytes, and monocytes. HSCs in vivo can undergo three types of division: symmetric self-renewal, asymmetric division producing one HSC and one multipotent progenitor cell, and symmetric differentiation yielding two multipotent progenitor cells [80, 81]. Interestingly, HBO1 deficiency disrupts symmetrical self-renewal of HSCs, with all divisions potentially leading to symmetrical differentiation. Competitive transplantation experiments further demonstrate the involvement of HBO1 in maintaining the repopulating ability of HSCs, underscoring its critical role in HSC pool stability. Genome analysis of HBO1 knockout mice has revealed downregulation of genes crucial for HSC function, including Hoxa9, Pbx1, GATA2, Mpl, Itga2b, and Irf8. These genes play pivotal roles in HSC quiescence, proliferation, and development [82, 83]. For instance, Mpl is involved in HSC quiescence and proliferation, while GATA2 is essential for hematopoietic stem/progenitor cell (HSPC) development [84, 85]. Moreover, HOX proteins and their cofactors, such as Pbx1 and Meis1, are critical for cell identity during embryonic development and adult hematopoiesis, with Hoxa9 deficiency resulting in reduced long-term repopulating ability of HSCs [86, 87]. Notably, the expression of genes essential for stem cell function is dependent on elevated levels of H3K14ac at their respective loci [88]. HBO1 deficiency leads to decreased H3K14ac levels, particularly at genes crucial for hematopoietic stem cell function. This suggests that HBO1 promotes the expression of a transcription factor network through its histone acetyltransferase (HAT) activity, which is indispensable for the maintenance and self-renewal of HSCs during adult hematopoiesis [89].

HBO1 is also involved in the normal differentiation of neural stem cells. Neural stem cells in the forebrain possess the remarkable ability to differentiate into neurons, astrocytes, and oligodendrocytes [90, 91]. HBO1 has been identified as a critical factor required for the differentiation of neural stem/progenitor cells (NSPCs). Studies utilizing mouse NSPCs lacking HBO1 have revealed that these cells exhibit slow proliferation rates for at least 15 generations and are unable to differentiate into neurons and oligodendrocytes, but can only differentiate into astrocytes. Deletion of HBO1 results in a decrease in the level of H3K14ac in cells, accompanied by the downregulation of more than 1000 genes. Interestingly, these downregulated genes are not necessary for NSPC proliferation in vitro but are crucial for nervous system development, neuronal differentiation, synaptic assembly, and behavioral regulation. Additionally, genes that are normally upregulated during normal differentiation are not activated in HBO1 knockout cells. These genes are known to play specific roles in neuronal differentiation, including axon guidance, neuroactive ligand–receptor interaction, synaptic function, and cognition. For instance, SOX2, a transcription factor proposed to initiate neuronal processes by activating genes such as Neurod1 [92], requires HBO1 for its activation of target genes during differentiation. The absence of HBO1 during the middle stage of neurogenesis leads to abnormal cortical development and increased cell death. HBO1 knockout mice exhibit reduced cerebral cortex depth, increased cell density, enlarged lateral ventricles, smaller corpus callosum diameter, underdeveloped hippocampal structure, and underdeveloped dentate gyrus compared with control mice. Interestingly, reexpression of HBO1 after a short duration of deletion rapidly restores NSPC differentiation potential. However, delayed reexpression only partially restores NSPC plasticity, requiring long-term reexpression for full restoration [93]. In conclusion, HBO1-mediated H3K14ac plays pivotal roles in the normal differentiation and brain development of NSPCs, highlighting their importance in neurogenesis and brain function.

HBO1 is closely related to many diseases

The expression of HBO1 has been closely associated with the development of various diseases. In several primary human tumor types, including testicular cancer, ovarian cancer, breast cancer, gastric/esophageal cancer, and bladder cancer, HBO1 protein expression was found to be strongly upregulated [11]. In non-small cell lung cancer (NSCLC), the transcription level of HBO1 is increased, and HBO1 silencing or knockout has been shown to strongly inhibit cancer cell viability, proliferation, and migration, while its ectopic overexpression enhances these processes. H3–H4 histone acetylation and the expression of several potential oncogenes (CCR2, MYLK, VEGFR2, and OCIAD2) were significantly reduced in NSCLC cells with HBO1 silencing or knockout, suggesting that HBO1 may promote cancer cell growth through its HAT activity [94]. Similar patterns of HBO1 expression and function have been observed in osteosarcoma and liver cancer cells, indicating a potential role in promoting cancer development via similar mechanisms [95, 96]. Conversely, downregulation of HBO1 has been found to alleviate the activation of hepatic stellate cells, inhibiting liver fibrosis [97]. HBO1 has also been implicated in the activation of the Wnt/β-catenin signaling pathway, contributing to the development of human glioblastoma, B-cell acute lymphoblastic leukemia, and bladder cancer [30, 98, 99]. Interestingly, HBO1 expression is significantly reduced in bronchial epithelial cells (HBEC) of patients with chronic obstructive pulmonary disease (COPD). Experimental studies in emphysema model mice have demonstrated that HBO1 can mitigate HBEC apoptosis and emphysema induced by cigarette smoke extract (CSE), suggesting a protective role for HBO1 in COPD pathogenesis [100].

WM-3835 (N′-(4-fluoro-5-methyl-[1,1′-biphenyl]-3-carbonyl)-3-hydroxybenzenesulfonohydrazide), a specific HBO1 inhibitor, has been developed. WM-3835 can reduce the activity of acute myeloid leukemia tumor cells by inhibiting the level of H3K14Ac regulated by HBO1 and further reducing the transcription of HOXA9 and HOXA10 [31]. In addition, WM-3835 also targets HBO1 to inhibit the development of castration-resistant prostate cancer (CRPC) [101], NSCLC [94], osteosarcoma [95], and other tumors. We hope that the advent of some drugs based on WM-3835 will bring a new dawn to treatment of HBO1-related diseases by inhibiting HBO1.

Summary and prospects

HBO1 exhibits a wide array of functions in cell biology, ranging from cell proliferation and gene expression to immune regulation, stem cell development, and cancer. However, several aspects of HBO1’s activity and regulation remain to be fully understood. One key area for investigation is the substrate selectivity of HBO1 acetyltransferase. Understanding which histone residues and nonhistone proteins are targeted by HBO1 will provide insights into its diverse cellular functions. Additionally, the specific mechanism by which the N-terminal domain regulates HBO1 activity requires further elucidation, as it likely plays a crucial role in modulating HBO1’s function. Moreover, HBO1’s involvement in DNA replication presents an intriguing area for exploration. Clarifying HBO1’s precise role in this process and its interaction with other replication factors will enhance our understanding of DNA replication regulation. Structural studies aimed at deciphering the complete structure of HBO1 are essential for comprehensively understanding its function. Such studies will shed light on the interaction between HBO1 domains, cofactor binding, and aid in the design of HBO1-targeting molecules for therapeutic purposes. The interaction between HBO1 and MEAF6, as well as MEAF6’s specific role within the HBO1 complex, remains poorly understood and warrants further investigation. Furthermore, the relationship between HBO1’s acetyltransferase activity and disease pathogenesis requires special attention, particularly in cancer where HBO1 is highly expressed. Elucidating the specific functions of HBO1 in cancer cells could uncover novel therapeutic targets for cancer treatment.

Data availability

Not applicable.

References

  1. Feng Y, Vlassis A, Roques C, Lalonde ME, González-Aguilera C, Lambert JP, Lee SB, Zhao X, Alabert C, Johansen JV, et al. BRPF3-HBO1 regulates replication origin activation and histone H3K14 acetylation. EMBO J. 2016;35:176–92. https://doiorg.publicaciones.saludcastillayleon.es/10.15252/embj.201591293.

    Article  CAS  PubMed  Google Scholar 

  2. Han J, Lachance C, Ricketts MD, McCullough CE, Gerace M, Black BE, Côté J, Marmorstein R. The scaffolding protein JADE1 physically links the acetyltransferase subunit HBO1 with its histone H3–H4 substrate. J Biol Chem. 2018;293:4498–509. https://doiorg.publicaciones.saludcastillayleon.es/10.1074/jbc.RA117.000677.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Havasi A, Haegele JA, Gall JM, Blackmon S, Ichimura T, Bonegio RG, Panchenko MV. Histone acetyl transferase (HAT) HBO1 and JADE1 in epithelial cell regeneration. Am J Pathol. 2013;182:152–62. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.ajpath.2012.09.017.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Iizuka M, Stillman B. Histone acetyltransferase HBO1 interacts with the ORC1 subunit of the human initiator protein. J Biol Chem. 1999;274:23027–34. https://doiorg.publicaciones.saludcastillayleon.es/10.1074/jbc.274.33.23027.

    Article  CAS  PubMed  Google Scholar 

  5. Burke TW, Cook JG, Asano M, Nevins JR. Replication factors MCM2 and ORC1 interact with the histone acetyltransferase HBO1. J Biol Chem. 2001;276:15397–408. https://doiorg.publicaciones.saludcastillayleon.es/10.1074/jbc.M011556200.

    Article  CAS  PubMed  Google Scholar 

  6. Sharma M, Zarnegar M, Li X, Lim B, Sun Z. Androgen receptor interacts with a novel MYST protein, HBO1. J Biol Chem. 2000;275:35200–8. https://doiorg.publicaciones.saludcastillayleon.es/10.1074/jbc.M004838200.

    Article  CAS  PubMed  Google Scholar 

  7. Contzler R, Regamey A, Favre B, Roger T, Hohl D, Huber M. Histone acetyltransferase HBO1 inhibits NF-kappaB activity by coactivator sequestration. Biochem Biophys Res Commun. 2006;350:208–13. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.bbrc.2006.09.030.

    Article  CAS  PubMed  Google Scholar 

  8. Iizuka M, Sarmento OF, Sekiya T, Scrable H, Allis CD, Smith MM. Hbo1 links p53-dependent stress signaling to DNA replication licensing. Mol Cell Biol. 2008;28:140–53. https://doiorg.publicaciones.saludcastillayleon.es/10.1128/mcb.00662-07.

    Article  CAS  PubMed  Google Scholar 

  9. Choudhary C, Kumar C, Gnad F, Nielsen ML, Rehman M, Walther TC, Olsen JV, Mann M. Lysine acetylation targets protein complexes and co-regulates major cellular functions. Science. 2009;325:834–40. https://doiorg.publicaciones.saludcastillayleon.es/10.1126/science.1175371.

    Article  CAS  PubMed  Google Scholar 

  10. Shvedunova M, Akhtar A. Modulation of cellular processes by histone and non-histone protein acetylation. Nat Rev Mol Cell Biol. 2022;23:329–49. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41580-021-00441-y.

    Article  CAS  PubMed  Google Scholar 

  11. Iizuka M, Takahashi Y, Mizzen CA, Cook RG, Fujita M, Allis CD, Frierson HF Jr, Fukusato T, Smith MM. Histone acetyltransferase Hbo1: catalytic activity, cellular abundance, and links to primary cancers. Gene. 2009;436:108–14. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.gene.2009.01.020.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Doyon Y, Cayrou C, Ullah M, Landry AJ, Côté V, Selleck W, Lane WS, Tan S, Yang XJ, Côté J. ING tumor suppressor proteins are critical regulators of chromatin acetylation required for genome expression and perpetuation. Mol Cell. 2006;21:51–64. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.molcel.2005.12.007.

    Article  CAS  PubMed  Google Scholar 

  13. Hung T, Binda O, Champagne KS, Kuo AJ, Johnson K, Chang HY, Simon MD, Kutateladze TG, Gozani O. ING4 mediates crosstalk between histone H3 K4 trimethylation and H3 acetylation to attenuate cellular transformation. Mol Cell. 2009;33:248–56. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.molcel.2008.12.016.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Avvakumov N, Lalonde ME, Saksouk N, Paquet E, Glass KC, Landry AJ, Doyon Y, Cayrou C, Robitaille GA, Richard DE, et al. Conserved molecular interactions within the HBO1 acetyltransferase complexes regulate cell proliferation. Mol Cell Biol. 2012;32:689–703. https://doiorg.publicaciones.saludcastillayleon.es/10.1128/mcb.06455-11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Shilatifard A. Chromatin modifications by methylation and ubiquitination: implications in the regulation of gene expression. Annu Rev Biochem. 2006;75:243–69. https://doiorg.publicaciones.saludcastillayleon.es/10.1146/annurev.biochem.75.103004.142422.

    Article  CAS  PubMed  Google Scholar 

  16. Saksouk N, Avvakumov N, Champagne KS, Hung T, Doyon Y, Cayrou C, Paquet E, Ullah M, Landry AJ, Côté V, et al. HBO1 HAT complexes target chromatin throughout gene coding regions via multiple PHD finger interactions with histone H3 tail. Mol Cell. 2009;33:257–65. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.molcel.2009.01.007.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Lalonde ME, Avvakumov N, Glass KC, Joncas FH, Saksouk N, Holliday M, Paquet E, Yan K, Tong Q, Klein BJ, et al. Exchange of associated factors directs a switch in HBO1 acetyltransferase histone tail specificity. Genes Dev. 2013;27:2009–24. https://doiorg.publicaciones.saludcastillayleon.es/10.1101/gad.223396.113.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Ormaza G, Rodríguez JA, Ibáñez de Opakua A, Merino N, Villate M, Gorroño I, Rábano M, Palmero I, Vilaseca M, Kypta R, et al. The tumor suppressor ING5 is a dimeric, bivalent recognition molecule of the histone H3K4me3 mark. J Mol Biol. 2019;431:2298–319. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.jmb.2019.04.018.

    Article  CAS  PubMed  Google Scholar 

  19. Ormaza G, Medagli B, Ibáñez de Opakua A, Rodríguez JA, Merino N, Villate M, Onesti S, Blanco FJ. The tumor suppressor inhibitor of growth 4 binds double-stranded DNA through its disordered central region. FEBS Lett. 2017;591:425–32. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/1873-3468.12514.

    Article  CAS  PubMed  Google Scholar 

  20. Mishima Y, Miyagi S, Saraya A, Negishi M, Endoh M, Endo TA, Toyoda T, Shinga J, Katsumoto T, Chiba T, et al. The Hbo1-Brd1/Brpf2 complex is responsible for global acetylation of H3K14 and required for fetal liver erythropoiesis. Blood. 2011;118:2443–53. https://doiorg.publicaciones.saludcastillayleon.es/10.1182/blood-2011-01-331892.

    Article  CAS  PubMed  Google Scholar 

  21. Tao Y, Zhong C, Zhu J, Xu S, Ding J. Structural and mechanistic insights into regulation of HBO1 histone acetyltransferase activity by BRPF2. Nucleic Acids Res. 2017;45:5707–19. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/nar/gkx142.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Yan K, You L, Degerny C, Ghorbani M, Liu X, Chen L, Li L, Miao D, Yang XJ. The chromatin regulator BRPF3 preferentially activates the HBO1 acetyltransferase but is dispensable for mouse development and survival. J Biol Chem. 2016;291:2647–63. https://doiorg.publicaciones.saludcastillayleon.es/10.1074/jbc.M115.703041.

    Article  CAS  PubMed  Google Scholar 

  23. Foy RL, Song IY, Chitalia VC, Cohen HT, Saksouk N, Cayrou C, Vaziri C, Côté J, Panchenko MV. Role of Jade-1 in the histone acetyltransferase (HAT) HBO1 complex. J Biol Chem. 2008;283:28817–26. https://doiorg.publicaciones.saludcastillayleon.es/10.1074/jbc.M801407200.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Siriwardana NS, Meyer R, Havasi A, Dominguez I, Panchenko MV. Cell cycle-dependent chromatin shuttling of HBO1-JADE1 histone acetyl transferase (HAT) complex. Cell Cycle. 2014;13:1885–901. https://doiorg.publicaciones.saludcastillayleon.es/10.4161/cc.28759.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Panchenko MV, Zhou MI, Cohen HT. von Hippel-Lindau partner Jade-1 is a transcriptional co-activator associated with histone acetyltransferase activity. J Biol Chem. 2004;279:56032–41. https://doiorg.publicaciones.saludcastillayleon.es/10.1074/jbc.M410487200.

    Article  CAS  PubMed  Google Scholar 

  26. Ullah M, Pelletier N, Xiao L, Zhao SP, Wang K, Degerny C, Tahmasebi S, Cayrou C, Doyon Y, Goh SL, et al. Molecular architecture of quartet MOZ/MORF histone acetyltransferase complexes. Mol Cell Biol. 2008;28:6828–43. https://doiorg.publicaciones.saludcastillayleon.es/10.1128/mcb.01297-08.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Carlson S, Glass KC. The MOZ histone acetyltransferase in epigenetic signaling and disease. J Cell Physiol. 2014;229:1571–4. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/jcp.24617.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Xiao Y, Li W, Yang H, Pan L, Zhang L, Lu L, Chen J, Wei W, Ye J, Li J, et al. HBO1 is a versatile histone acyltransferase critical for promoter histone acylations. Nucleic Acids Res. 2021;49:8037–59. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/nar/gkab607.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Li L, Liang Y, Kang L, Liu Y, Gao S, Chen S, Li Y, You W, Dong Q, Hong T, et al. Transcriptional regulation of the Warburg effect in cancer by SIX1. Cancer Cell. 2018;33:368-385.e367. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.ccell.2018.01.010.

    Article  CAS  PubMed  Google Scholar 

  30. Wang H, Qiu Y, Zhang H, Chang N, Hu Y, Chen J, Hu R, Liao P, Li Z, Yang Y, et al. Histone acetylation by HBO1 (KAT7) activates Wnt/β-catenin signaling to promote leukemogenesis in B-cell acute lymphoblastic leukemia. Cell Death Dis. 2023;14:498. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41419-023-06019-0.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. MacPherson L, Anokye J, Yeung MM, Lam EYN, Chan YC, Weng CF, Yeh P, Knezevic K, Butler MS, Hoegl A, et al. HBO1 is required for the maintenance of leukaemia stem cells. Nature. 2020;577:266–70. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41586-019-1835-6.

    Article  CAS  PubMed  Google Scholar 

  32. Aryal S, Zhang Y, Wren S, Li C, Lu R. Molecular regulators of HOXA9 in acute myeloid leukemia. FEBS J. 2023;290:321–39. https://doiorg.publicaciones.saludcastillayleon.es/10.1111/febs.16268.

    Article  CAS  PubMed  Google Scholar 

  33. Hayashi Y, Harada Y, Kagiyama Y, Nishikawa S, Ding Y, Imagawa J, Shingai N, Kato N, Kitaura J, Hokaiwado S, et al. NUP98-HBO1-fusion generates phenotypically and genetically relevant chronic myelomonocytic leukemia pathogenesis. Blood Adv. 2019;3:1047–60. https://doiorg.publicaciones.saludcastillayleon.es/10.1182/bloodadvances.2018025007.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Wright DG, Marchal C, Hoang K, Ankney JA, Nguyen ST, Rushing AW, Polakowski N, Miotto B, Lemasson I. Human T-cell leukemia virus type-1-encoded protein HBZ represses p53 function by inhibiting the acetyltransferase activity of p300/CBP and HBO1. Oncotarget. 2016;7:1687–706. https://doiorg.publicaciones.saludcastillayleon.es/10.18632/oncotarget.6424.

    Article  PubMed  Google Scholar 

  35. Tan D, Wei W, Han Z, Ren X, Yan C, Qi S, Song X, Zheng YG, Wong J, Huang H. HBO1 catalyzes lysine benzoylation in mammalian cells. iScience. 2022;25:105443. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.isci.2022.105443.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Gao Y, Sheng X, Tan D, Kim S, Choi S, Paudel S, Lee T, Yan C, Tan M, Kim KM, et al. Identification of histone lysine acetoacetylation as a dynamic post-translational modification regulated by HBO1. Adv Sci (Weinh). 2023;10: e2300032. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/advs.202300032.

    Article  CAS  PubMed  Google Scholar 

  37. Chen X, Liu G, Leffak M. Activation of a human chromosomal replication origin by protein tethering. Nucleic Acids Res. 2013;41:6460–74. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/nar/gkt368.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Bell SP, Dutta A. DNA replication in eukaryotic cells. Annu Rev Biochem. 2002;71:333–74. https://doiorg.publicaciones.saludcastillayleon.es/10.1146/annurev.biochem.71.110601.135425.

    Article  CAS  PubMed  Google Scholar 

  39. Miotto B, Struhl K. HBO1 histone acetylase is a coactivator of the replication licensing factor Cdt1. Genes Dev. 2008;22:2633–8. https://doiorg.publicaciones.saludcastillayleon.es/10.1101/gad.1674108.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Miotto B, Struhl K. HBO1 histone acetylase activity is essential for DNA replication licensing and inhibited by Geminin. Mol Cell. 2010;37:57–66. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.molcel.2009.12.012.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Iizuka M, Matsui T, Takisawa H, Smith MM. Regulation of replication licensing by acetyltransferase Hbo1. Mol Cell Biol. 2006;26:1098–108. https://doiorg.publicaciones.saludcastillayleon.es/10.1128/mcb.26.3.1098-1108.2006.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Miotto B, Struhl K. JNK1 phosphorylation of Cdt1 inhibits recruitment of HBO1 histone acetylase and blocks replication licensing in response to stress. Mol Cell. 2011;44:62–71. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.molcel.2011.06.021.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Wong PG, Glozak MA, Cao TV, Vaziri C, Seto E, Alexandrow M. Chromatin unfolding by Cdt1 regulates MCM loading via opposing functions of HBO1 and HDAC11-geminin. Cell Cycle. 2010;9:4351–63. https://doiorg.publicaciones.saludcastillayleon.es/10.4161/cc.9.21.13596.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Suchyta M, Miotto B, McGarry TJ. An inactive geminin mutant that binds cdt1. Genes (Basel). 2015;6:252–66. https://doiorg.publicaciones.saludcastillayleon.es/10.3390/genes6020252.

    Article  CAS  PubMed  Google Scholar 

  45. Johmura Y, Osada S, Nishizuka M, Imagawa M. FAD24 acts in concert with histone acetyltransferase HBO1 to promote adipogenesis by controlling DNA replication. J Biol Chem. 2008;283:2265–74. https://doiorg.publicaciones.saludcastillayleon.es/10.1074/jbc.M707880200.

    Article  CAS  PubMed  Google Scholar 

  46. Imagawa M. Molecular mechanisms of early-stage adipocyte differentiation and multi-functional roles of newly isolated adipogenic factors. Yakugaku Zasshi. 2016;136:649–58. https://doiorg.publicaciones.saludcastillayleon.es/10.1248/yakushi.15-00260.

    Article  CAS  PubMed  Google Scholar 

  47. Johmura Y, Osada S, Nishizuka M, Imagawa M. FAD24, a regulator of adipogenesis, is required for the regulation of DNA replication in cell proliferation. Biol Pharm Bull. 2008;31:1092–5. https://doiorg.publicaciones.saludcastillayleon.es/10.1248/bpb.31.1092.

    Article  CAS  PubMed  Google Scholar 

  48. Kueh AJ, Dixon MP, Voss AK, Thomas T. HBO1 is required for H3K14 acetylation and normal transcriptional activity during embryonic development. Mol Cell Biol. 2011;31:845–60. https://doiorg.publicaciones.saludcastillayleon.es/10.1128/mcb.00159-10.

    Article  CAS  PubMed  Google Scholar 

  49. Kueh AJ, Eccles S, Tang L, Garnham AL, May RE, Herold MJ, Smyth GK, Voss AK, Thomas T. HBO1 (KAT7) does not have an essential role in cell proliferation, DNA replication, or histone 4 acetylation in human cells. Mol Cell Biol. 2020. https://doiorg.publicaciones.saludcastillayleon.es/10.1128/mcb.00506-19.

    Article  PubMed  PubMed Central  Google Scholar 

  50. Ohzeki J, Shono N, Otake K, Martins NM, Kugou K, Kimura H, Nagase T, Larionov V, Earnshaw WC, Masumoto H. KAT7/HBO1/MYST2 regulates CENP-A chromatin assembly by antagonizing Suv39h1-mediated centromere inactivation. Dev Cell. 2016;37:413–27. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.devcel.2016.05.006.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Niida H, Matsunuma R, Horiguchi R, Uchida C, Nakazawa Y, Motegi A, Nishimoto K, Sakai S, Ohhata T, Kitagawa K, et al. Phosphorylated HBO1 at UV irradiated sites is essential for nucleotide excision repair. Nat Commun. 2017;8:16102. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/ncomms16102.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Luijsterburg MS, Goedhart J, Moser J, Kool H, Geverts B, Houtsmuller AB, Mullenders LH, Vermeulen W, van Driel R. Dynamic in vivo interaction of DDB2 E3 ubiquitin ligase with UV-damaged DNA is independent of damage-recognition protein XPC. J Cell Sci. 2007;120:2706–16. https://doiorg.publicaciones.saludcastillayleon.es/10.1242/jcs.008367.

    Article  CAS  PubMed  Google Scholar 

  53. Alekseev S, Luijsterburg MS, Pines A, Geverts B, Mari PO, Giglia-Mari G, Lans H, Houtsmuller AB, Mullenders LH, Hoeijmakers JH, Vermeulen W. Cellular concentrations of DDB2 regulate dynamic binding of DDB1 at UV-induced DNA damage. Mol Cell Biol. 2008;28:7402–13. https://doiorg.publicaciones.saludcastillayleon.es/10.1128/mcb.01108-08.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Waters R, van Eijk P, Reed S. Histone modification and chromatin remodeling during NER. DNA Repair (Amst). 2015;36:105–13. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.dnarep.2015.09.013.

    Article  CAS  PubMed  Google Scholar 

  55. Aydin ÖZ, Marteijn JA, Ribeiro-Silva C, Rodríguez López A, Wijgers N, Smeenk G, van Attikum H, Poot RA, Vermeulen W, Lans H. Human ISWI complexes are targeted by SMARCA5 ATPase and SLIDE domains to help resolve lesion-stalled transcription. Nucleic Acids Res. 2014;42:8473–85. https://doiorg.publicaciones.saludcastillayleon.es/10.1093/nar/gku565.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Erdel F, Rippe K. Binding kinetics of human ISWI chromatin-remodelers to DNA repair sites elucidate their target location mechanism. Nucleus. 2011;2:105–12. https://doiorg.publicaciones.saludcastillayleon.es/10.4161/nucl.2.2.15209.

    Article  PubMed  PubMed Central  Google Scholar 

  57. Koyauchi T, Niida H, Motegi A, Sakai S, Uchida C, Ohhata T, Iijima K, Yokoyama A, Suda T, Kitagawa M. Chromatin-remodeling factor BAZ1A/ACF1 targets UV damage sites in an MLL1-dependent manner to facilitate nucleotide excision repair. Biochim Biophys Acta Mol Cell Res. 2022;1869: 119332. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.bbamcr.2022.119332.

    Article  CAS  PubMed  Google Scholar 

  58. Li H, Ilin S, Wang W, Duncan EM, Wysocka J, Allis CD, Patel DJ. Molecular basis for site-specific read-out of histone H3K4me3 by the BPTF PHD finger of NURF. Nature. 2006;442:91–5. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/nature04802.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Zhou BO, Zhou JQ. Recent transcription-induced histone H3 lysine 4 (H3K4) methylation inhibits gene reactivation. J Biol Chem. 2011;286:34770–6. https://doiorg.publicaciones.saludcastillayleon.es/10.1074/jbc.M111.273128.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Zou C, Chen Y, Smith RM, Snavely C, Li J, Coon TA, Chen BB, Zhao Y, Mallampalli RK. SCF(Fbxw15) mediates histone acetyltransferase binding to origin recognition complex (HBO1) ubiquitin-proteasomal degradation to regulate cell proliferation. J Biol Chem. 2013;288:6306–16. https://doiorg.publicaciones.saludcastillayleon.es/10.1074/jbc.M112.426882.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Long C, Lai Y, Li J, Huang J, Zou C. LPS promotes HBO1 stability via USP25 to modulate inflammatory gene transcription in THP-1 cells. Biochim Biophys Acta Gene Regul Mech. 2018;1861:773–82. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.bbagrm.2018.08.001.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Matsunuma R, Niida H, Ohhata T, Kitagawa K, Sakai S, Uchida C, Shiotani B, Matsumoto M, Nakayama KI, Ogura H, et al. UV damage-induced phosphorylation of HBO1 triggers CRL4DDB2-mediated degradation to regulate cell proliferation. Mol Cell Biol. 2016;36:394–406. https://doiorg.publicaciones.saludcastillayleon.es/10.1128/mcb.00809-15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Iizuka M, Susa T, Takahashi Y, Tamamori-Adachi M, Kajitani T, Okinaga H, Fukusato T, Okazaki T. Histone acetyltransferase Hbo1 destabilizes estrogen receptor α by ubiquitination and modulates proliferation of breast cancers. Cancer Sci. 2013;104:1647–55. https://doiorg.publicaciones.saludcastillayleon.es/10.1111/cas.12303.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Iizuka M, Susa T, Tamamori-Adachi M, Okinaga H, Okazaki T. Intrinsic ubiquitin E3 ligase activity of histone acetyltransferase Hbo1 for estrogen receptor α. Proc Jpn Acad Ser B Phys Biol Sci. 2017;93:498–510. https://doiorg.publicaciones.saludcastillayleon.es/10.2183/pjab.93.030.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Hogquist KA, Baldwin TA, Jameson SC. Central tolerance: learning self-control in the thymus. Nat Rev Immunol. 2005;5:772–82. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/nri1707.

    Article  CAS  PubMed  Google Scholar 

  66. Abramson J, Anderson G. Thymic epithelial cells. Annu Rev Immunol. 2017;35:85–118. https://doiorg.publicaciones.saludcastillayleon.es/10.1146/annurev-immunol-051116-052320.

    Article  CAS  PubMed  Google Scholar 

  67. Derbinski J, Gäbler J, Brors B, Tierling S, Jonnakuty S, Hergenhahn M, Peltonen L, Walter J, Kyewski B. Promiscuous gene expression in thymic epithelial cells is regulated at multiple levels. J Exp Med. 2005;202:33–45. https://doiorg.publicaciones.saludcastillayleon.es/10.1084/jem.20050471.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Anderson MS, Venanzi ES, Klein L, Chen Z, Berzins SP, Turley SJ, von Boehmer H, Bronson R, Dierich A, Benoist C, Mathis D. Projection of an immunological self shadow within the thymus by the aire protein. Science. 2002;298:1395–401. https://doiorg.publicaciones.saludcastillayleon.es/10.1126/science.1075958.

    Article  CAS  PubMed  Google Scholar 

  69. Kisand K, Peterson P. Autoimmune polyendocrinopathy candidiasis ectodermal dystrophy. J Clin Immunol. 2015;35:463–78. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s10875-015-0176-y.

    Article  CAS  PubMed  Google Scholar 

  70. Mathis D, Benoist C. Aire. Annu Rev Immunol. 2009;27:287–312. https://doiorg.publicaciones.saludcastillayleon.es/10.1146/annurev.immunol.25.022106.141532.

    Article  CAS  PubMed  Google Scholar 

  71. Shum AK, Alimohammadi M, Tan CL, Cheng MH, Metzger TC, Law CS, Lwin W, Perheentupa J, Bour-Jordan H, Carel JC, et al. BPIFB1 is a lung-specific autoantigen associated with interstitial lung disease. Sci Transl Med. 2013;5:206ra139. https://doiorg.publicaciones.saludcastillayleon.es/10.1126/scitranslmed.3006998.

    Article  CAS  PubMed  Google Scholar 

  72. Shum AK, DeVoss J, Tan CL, Hou Y, Johannes K, O’Gorman CS, Jones KD, Sochett EB, Fong L, Anderson MS. Identification of an autoantigen demonstrates a link between interstitial lung disease and a defect in central tolerance. Sci Transl Med. 2009;1:9ra20. https://doiorg.publicaciones.saludcastillayleon.es/10.1126/scitranslmed.3000284.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Heinlein M, Gandolfo LC, Zhao K, Teh CE, Nguyen N, Baell JB, Goldfarb Y, Abramson J, Wichmann J, Voss AK, et al. The acetyltransferase KAT7 is required for thymic epithelial cell expansion, expression of AIRE target genes, and thymic tolerance. Sci Immunol. 2022;7:eabb6032. https://doiorg.publicaciones.saludcastillayleon.es/10.1126/sciimmunol.abb6032.

    Article  CAS  PubMed  Google Scholar 

  74. Taniuchi I, Ellmeier W. Transcriptional and epigenetic regulation of CD4/CD8 lineage choice. Adv Immunol. 2011;110:71–110. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/b978-0-12-387663-8.00003-x.

    Article  CAS  PubMed  Google Scholar 

  75. Collins A, Littman DR, Taniuchi I. RUNX proteins in transcription factor networks that regulate T-cell lineage choice. Nat Rev Immunol. 2009;9:106–15. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/nri2489.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Harker N, Naito T, Cortes M, Hostert A, Hirschberg S, Tolaini M, Roderick K, Georgopoulos K, Kioussis D. The CD8alpha gene locus is regulated by the Ikaros family of proteins. Mol Cell. 2002;10:1403–15. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/s1097-2765(02)00711-6.

    Article  CAS  PubMed  Google Scholar 

  77. Mishima Y, Wang C, Miyagi S, Saraya A, Hosokawa H, Mochizuki-Kashio M, Nakajima-Takagi Y, Koide S, Negishi M, Sashida G, et al. Histone acetylation mediated by Brd1 is crucial for Cd8 gene activation during early thymocyte development. Nat Commun. 2014;5:5872. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/ncomms6872.

    Article  CAS  PubMed  Google Scholar 

  78. Feng M, Liu X, Hao X, Ren Y, Dong G, Tian J, Wang Y, Du L, Wang Y, Wang C. Fatty acids support the fitness and functionality of tumor-resident CD8+ T cells by maintaining SCML4 expression. Cancer Res. 2023;83:3368–84. https://doiorg.publicaciones.saludcastillayleon.es/10.1158/0008-5472.Can-23-0287.

    Article  CAS  PubMed  Google Scholar 

  79. Boulais PE, Frenette PS. Making sense of hematopoietic stem cell niches. Blood. 2015;125:2621–9. https://doiorg.publicaciones.saludcastillayleon.es/10.1182/blood-2014-09-570192.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Wu M, Kwon HY, Rattis F, Blum J, Zhao C, Ashkenazi R, Jackson TL, Gaiano N, Oliver T, Reya T. Imaging hematopoietic precursor division in real time. Cell Stem Cell. 2007;1:541–54. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.stem.2007.08.009.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Will B, Vogler TO, Bartholdy B, Garrett-Bakelman F, Mayer J, Barreyro L, Pandolfi A, Todorova TI, Okoye-Okafor UC, Stanley RF, et al. Satb1 regulates the self-renewal of hematopoietic stem cells by promoting quiescence and repressing differentiation commitment. Nat Immunol. 2013;14:437–45. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/ni.2572.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Qian H, Buza-Vidas N, Hyland CD, Jensen CT, Antonchuk J, Månsson R, Thoren LA, Ekblom M, Alexander WS, Jacobsen SE. Critical role of thrombopoietin in maintaining adult quiescent hematopoietic stem cells. Cell Stem Cell. 2007;1:671–84. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.stem.2007.10.008.

    Article  CAS  PubMed  Google Scholar 

  83. Yoshihara H, Arai F, Hosokawa K, Hagiwara T, Takubo K, Nakamura Y, Gomei Y, Iwasaki H, Matsuoka S, Miyamoto K, et al. Thrombopoietin/MPL signaling regulates hematopoietic stem cell quiescence and interaction with the osteoblastic niche. Cell Stem Cell. 2007;1:685–97. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.stem.2007.10.020.

    Article  CAS  PubMed  Google Scholar 

  84. Tsai FY, Keller G, Kuo FC, Weiss M, Chen J, Rosenblatt M, Alt FW, Orkin SH. An early haematopoietic defect in mice lacking the transcription factor GATA-2. Nature. 1994;371:221–6. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/371221a0.

    Article  CAS  PubMed  Google Scholar 

  85. Rodrigues NP, Tipping AJ, Wang Z, Enver T. GATA-2 mediated regulation of normal hematopoietic stem/progenitor cell function, myelodysplasia and myeloid leukemia. Int J Biochem Cell Biol. 2012;44:457–60. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.biocel.2011.12.004.

    Article  CAS  PubMed  Google Scholar 

  86. Alharbi RA, Pettengell R, Pandha HS, Morgan R. The role of HOX genes in normal hematopoiesis and acute leukemia. Leukemia. 2013;27:1000–8. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/leu.2012.356.

    Article  CAS  PubMed  Google Scholar 

  87. Lawrence HJ, Christensen J, Fong S, Hu YL, Weissman I, Sauvageau G, Humphries RK, Largman C. Loss of expression of the Hoxa-9 homeobox gene impairs the proliferation and repopulating ability of hematopoietic stem cells. Blood. 2005;106:3988–94. https://doiorg.publicaciones.saludcastillayleon.es/10.1182/blood-2005-05-2003.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Karmodiya K, Krebs AR, Oulad-Abdelghani M, Kimura H, Tora L. H3K9 and H3K14 acetylation co-occur at many gene regulatory elements, while H3K14ac marks a subset of inactive inducible promoters in mouse embryonic stem cells. BMC Genomics. 2012;13:424. https://doiorg.publicaciones.saludcastillayleon.es/10.1186/1471-2164-13-424.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Yang Y, Kueh AJ, Grant ZL, Abeysekera W, Garnham AL, Wilcox S, Hyland CD, Di Rago L, Metcalf D, Alexander WS, et al. The histone lysine acetyltransferase HBO1 (KAT7) regulates hematopoietic stem cell quiescence and self-renewal. Blood. 2022;139:845–58. https://doiorg.publicaciones.saludcastillayleon.es/10.1182/blood.2021013954.

    Article  CAS  PubMed  Google Scholar 

  90. Reynolds BA, Weiss S. Generation of neurons and astrocytes from isolated cells of the adult mammalian central nervous system. Science. 1992;255:1707–10. https://doiorg.publicaciones.saludcastillayleon.es/10.1126/science.1553558.

    Article  CAS  PubMed  Google Scholar 

  91. Davis AA, Temple S. A self-renewing multipotential stem cell in embryonic rat cerebral cortex. Nature. 1994;372:263–6. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/372263a0.

    Article  CAS  PubMed  Google Scholar 

  92. Liu Z, Kraus WL. Catalytic-independent functions of PARP-1 determine Sox2 pioneer activity at intractable genomic loci. Mol Cell. 2017;65:589-603.e589. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.molcel.2017.01.017.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Kueh AJ, Bergamasco MI, Quaglieri A, Phipson B, Li-Wai-Suen CSN, Lönnstedt IM, Hu Y, Feng ZP, Woodruff C, May RE, et al. Stem cell plasticity, acetylation of H3K14, and de novo gene activation rely on KAT7. Cell Rep. 2023;42: 111980. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.celrep.2022.111980.

    Article  CAS  PubMed  Google Scholar 

  94. Chen TF, Hao HF, Zhang Y, Chen XY, Zhao HS, Yang R, Li P, Qiu LX, Sang YH, Xu C, Liu SX. HBO1 induces histone acetylation and is important for non-small cell lung cancer cell growth. Int J Biol Sci. 2022;18:3313–23. https://doiorg.publicaciones.saludcastillayleon.es/10.7150/ijbs.72526.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Gao YY, Ling ZY, Zhu YR, Shi C, Wang Y, Zhang XY, Zhang ZQ, Jiang Q, Chen MB, Yang S, Cao C. The histone acetyltransferase HBO1 functions as a novel oncogenic gene in osteosarcoma. Theranostics. 2021;11:4599–615. https://doiorg.publicaciones.saludcastillayleon.es/10.7150/thno.55655.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Zhong W, Liu H, Deng L, Chen G, Liu Y. HBO1 overexpression is important for hepatocellular carcinoma cell growth. Cell Death Dis. 2021;12:549. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41419-021-03818-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Xing B, Lan H, Li H. HBO1 as an important target for the treatment of CCL4-induced liver fibrosis and aged-related liver aging and fibrosis. Oxid Med Cell Longev. 2022;2022:1881519. https://doiorg.publicaciones.saludcastillayleon.es/10.1155/2022/1881519.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Chen Z, Zhou L, Wang L, Kazobinka G, Zhang X, Han X, Li B, Hou T. HBO1 promotes cell proliferation in bladder cancer via activation of Wnt/β-catenin signaling. Mol Carcinog. 2018;57:12–21. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/mc.22715.

    Article  CAS  PubMed  Google Scholar 

  99. Wu J, Li L, Jiang G, Zhan H, Zhu X, Yang W. NCAPG2 facilitates glioblastoma cells’ malignancy and xenograft tumor growth via HBO1 activation by phosphorylation. Cell Tissue Res. 2021;383:693–706. https://doiorg.publicaciones.saludcastillayleon.es/10.1007/s00441-020-03281-y.

    Article  CAS  PubMed  Google Scholar 

  100. Chen L, Luo L, Kang N, He X, Li T, Chen Y. The protective effect of HBO1 on cigarette smoke extract-induced apoptosis in airway epithelial cells. Int J Chron Obstruct Pulmon Dis. 2020;15:15–24. https://doiorg.publicaciones.saludcastillayleon.es/10.2147/copd.S234634.

    Article  PubMed  PubMed Central  Google Scholar 

  101. Mi YY, Ji Y, Zhang L, Sun CY, Wei BB, Yang DJ, Wan HY, Qi XW, Wu S, Zhu LJ. A first-in-class HBO1 inhibitor WM-3835 inhibits castration-resistant prostate cancer cell growth in vitro and in vivo. Cell Death Dis. 2023;14:67. https://doiorg.publicaciones.saludcastillayleon.es/10.1038/s41419-023-05606-5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

The present study was supported by the National Natural Sciences Foundation of China (82273219).

Author information

Authors and Affiliations

Authors

Contributions

Z.S. wrote the manuscript and drew the figures. J.T. and Y.Z. collected the related papers and helped to revise the manuscript. C.L. and Y.Z. designed and revised the manuscript. All the authors read and approved the final version of the review.

Corresponding authors

Correspondence to Yanhong Zhou or Chen Long.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

All authors approved the final manuscript and the submission to this journal.

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/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Su, Z., Zhang, Y., Tang, J. et al. Multifunctional acyltransferase HBO1: a key regulatory factor for cellular functions. Cell Mol Biol Lett 29, 141 (2024). https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s11658-024-00661-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s11658-024-00661-y

Keywords