Next Article in Journal
Soluble ST2 as a Potential Biomarker for Abdominal Aortic Aneurysms—A Single-Center Retrospective Cohort Study
Next Article in Special Issue
Dual Inhibition of EGFR and IGF-1R Signaling Leads to Enhanced Antitumor Efficacy against Esophageal Squamous Cancer
Previous Article in Journal
Mitochonic Acid 5 Improves Duchenne Muscular Dystrophy and Parkinson’s Disease Model of Caenorhabditis elegans
Previous Article in Special Issue
Antigen Peptide Transporter 1 (TAP1) Promotes Resistance to MEK Inhibitors in Pancreatic Cancers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Targeting HDAC6 to Overcome Autophagy-Promoted Anti-Cancer Drug Resistance

Department of Biochemistry, College of Natural Sciences, Kangwon National University, Chuncheon 24341, Korea
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(17), 9592; https://doi.org/10.3390/ijms23179592
Submission received: 28 July 2022 / Revised: 18 August 2022 / Accepted: 18 August 2022 / Published: 24 August 2022

Abstract

:
Histone deacetylases (HDACs) regulate gene expression through the epigenetic modification of chromatin structure. HDAC6, unlike many other HDACs, is present in the cytoplasm. Its deacetylates non-histone proteins and plays diverse roles in cancer cell initiation, proliferation, autophagy, and anti-cancer drug resistance. The development of HDAC6-specific inhibitors has been relatively successful. Mechanisms of HDAC6-promoted anti-cancer drug resistance, cancer cell proliferation, and autophagy are discussed. The relationship between autophagy and anti-cancer drug resistance is discussed. The effects of combination therapy, which includes HDAC6 inhibitors, on the sensitivity of cancer cells to chemotherapeutics and immune checkpoint blockade are presented. A summary of clinical trials involving HDAC6-specific inhibitors is also presented. This review presents HDAC6 as a valuable target for developing anti-cancer drugs.

1. Role of HDAC6 in Cancer Cell Proliferation

Histone deacetylase 6 (HDAC6) deacetylates non-histone proteins, including cortactin, peroxiredoxin, and α-tubulin [1,2,3]. For example, HDAC6 deacetylates α-tubulin and mediates the disassembly of primary cilia [4]. It mostly deacetylates cytoplasmic proteins [5]. HDAC6 displays unique structure and cellular localization. It has a wider range of biological functions than other isoforms [6,7]. Unlike many other HDACs, HDAC6 shows cytoplasmic localization. It contains two catalytic domains (DD1 and DD2) and nuclear export signal sequences (NES) (Figure 1). The NES might be responsible for the cytoplasmic localization of HDAC6. HDAC6 catalyzes the removal of acetyl groups from various cytoplasmic proteins [8] and contains a ubiquitin binding domain necessary for targeting misfolded and damaged proteins to autophagosomes for removal (Figure 1). HDAC6 also contains a dynein motor biding domain necessary for aggresome formation for autophagic clearance (Figure 1).
Glioblastoma is associated with high levels of the mesenchymal markers Snail, Slug, and N-cadherin. Glioblastoma cell lines show high levels of autophagic flux and activation of the sonic hedgehog (Shh) pathway [9]. HDAC6 is the most overexpressed isoform in glioblastoma [9]. The downregulation of HDAC6 was shown to inhibit the proliferation and migration of glioblastoma cell lines [9], and ACY-1215 (Ricolinostat, a specific HDAC6 inhibitor) inhibits glioblastoma growth [10]. HDAC6 promotes the invasion and metastasis of melanoma cells [11]. It also makes tumor cells resistant to anoikis, thereby facilitating tumor invasion and metastasis [12]. The downregulation of HDAC6 was reported to inhibit cancer stem cell (CSC) growth and autophagy and increase the apoptosis of breast cancer cells [13]. A high level of HDAC6 has been reported in patients with glioblastoma multiforme (GBM) and in a subset of human gastric cancer cells [14]. HDAC6 downregulation inhibited gastric cancer cell growth without affecting cell cycle transition or the processing of cell death [14]. Thus, HDAC6 can regulate cancer cell proliferation.

2. Mechanism of HDAC6-Promoted Cancer Cell Proliferation

HDAC6 promotes the growth of GBM cells by inhibiting the suppressor of mothers against decapentaplegic 2 (SMAD2) phosphorylation to downregulate [10] p21.The Ras oncogene increases HDAC6 expression. HDAC6 confers resistance to Ras-induced oncogenesis, implying that HDAC6 can promote the activation of Ras and its downstream phosphoinositide 3-kinase (PI3K) and mitogen-activated protein kinase (MAPK) pathways [15]. Extracellular regulated kinase (ERK) can bind to and phosphorylate HDAC6, which enhances cell migration by deacetylating α-tubulin [16]. HDAC6 was shown to promote ERK1 activity by deacetylating Lys-72 [17]. It can bind to tyrosine-protein phosphatase non-receptor type 1 (PTPN1), activate ERK1/2, inhibit apoptosis, and promote melanoma cell proliferation [18]. HDAC6 knockdown inhibited colon cancer cell growth and migration by suppressing the MAPK/ERK pathway [19]. HDAC6 was reported to promote glioblastoma proliferation by increasing the expression of mitogen activated protein kinase 7 (MKK7) and enhancing Jun N-terminal kinase (JNK) activity [20].
HDAC6 was reported to increase the level of the Ewing sarcoma breakpoint region 1 fusion gene (EWSR1-FLI1) [21]. A high level of HDAC6 can predict low survival of patients with Ewing Sarcoma [21]. HDAC6 was shown to promote the invasion and migration of rhabdomyosarcoma (RMS) cells via RAC1, a Rho family GTPase [22]. HDAC6 also promoted the proliferation of glioblastoma by increasing the levels of DNA damage response genes such as DNA repair protein 51 (RAD51) and checkpoint kinase 1 (CHEK1) [23].
HDAC6 was shown to decrease the expression of tumor suppressor mammalian STE20-like kinase 1 (MST1) by deacetylating MST1 in breast cancer cells [24]. HDAC6 increased by nuclear factor-κB (NF-κB) promoted hepatocellular carcinoma cell proliferation by inducing the degradation of [25] p53 and promoted the invasion of breast cancer cells by decreasing E-cadherin while increasing the signal transducer and activator of transcription 3 (STAT3) level [26].
Upon epidermal growth factor (EGF) stimulation, HDAC6 was shown to interact with β-catenin at the caveolae membrane to deacetylate β-catenin (lysine residue 49) and inhibit the phosphorylation of β-catenin (Serine 45) [27]. HDAC6 inhibition prevented the nuclear translocation of β-catenin, decreased c-myc expression, and inhibited tumor proliferation [27]. HDAC6 and wingless-related integration site 5a (Wnt5a) are highly expressed in human papillomavirus (HR-HPV)-positive cervical cancer tissues [28]. HDAC6 promoted the proliferation and inhibited the apoptosis of HPV-infected cervical carcinoma cells by decreasing the expression of miR-199a, an inhibitor of wnt signaling [28].
Pentaspan membrane glycoprotein CD133 (Prominin-1), a marker of cancer stemness, can be used to predict the poor prognosis of patients with many different types of tumors. HDAC6 interacts with CD133 to inhibit CD133 degradation [29]. CD133, HDAC6, and β-catenin can form a ternary complex [29], which stabilizes β-catenin via HDAC6 deacetylase activity. The downregulation of either CD133 or HDAC6 was shown to increase β-catenin acetylation and degradation, which leads to decreased proliferation both in vitro and tumor xenograft growth in vivo [29]. CD133, a stem cell marker implicated in tumor initiation, differentiation, and anti-cancer drug resistance, is known to be associated with extracellular vesicles (EVs) in various types of cancer. Tubacin, an inhibitor of HDAC6, promoted the extracellular release of CD133+ EVs from human FEMX-I metastatic melanoma and Caco-2 colorectal carcinoma cells, leading to the downregulation of intracellular CD133 [30]. Tubacin-induced EV release altered cellular lipid composition and decreased clonogenic capacity and the formation of multicellular aggregates [30].
A high level of transmembrane serine protease 4 (TMPRSS4) is associated with the poor prognosis of patients with non-small cell lung cancer (NSCLC), gastric cancer, colorectal cancer, prostate cancer, and other cancers. TMPRSS4 promoted tumor cell proliferation and metastasis by inducing specificity protein 1/3 (Sp1/3), activator protein-1 (AP-1), and NF-κB transcription factors [31]. Sp1 is a novel substrate of HDAC6 [23].
A high level of ubiquitin specific protease 10 (USP10) is associated with the poor overall survival of patients NSCLC with p53 mutations [32]. USP10 was shown to interact with, deubiquitinate, and stabilize HDAC6 [32]. The genetic deletion or inhibition of USP10 was reported to inhibit the growth of lung cancer xenografts lacking wild-type p53 and enhance cisplatin sensitivity [32].
In patients with colorectal cancer (CRC), expression levels of SET domain containing 7 (SET7), a histone lysine methyl transferase, in cancer tissues are lower than those in adjacent tissue. Decreases in the expression of SET7 can predict poor patient prognosis [33]. SET7 and HDAC6 displayed reciprocal interactions [33], where SET7-HDAC6 interaction decreased colorectal cancer cell viability and migration. SET7 was shown to act as a tumor suppressor by increasing the level of acetylated-α-tubulin [33]. SET7-HDAC6 interaction was shown to decrease the p-ERK/ERK ratio [33]. Table 1 and Figure 2 show the mechanism of HDAC6-promoted cancer cell proliferation.

3. Roles of HDAC6-Targeting miRNAs in Cancer Cell Proliferation

The increased expression of HDAC6 was observed in anti-cancer drug-resistant cancer cells [34,35]. Figure 3A shows the potential transcription factors that may regulate the expression of HDAC6. However, the roles of these transcription factors in regulating the expression of HDAC6 have not been studied extensively.
MicroRNAs (miRNAs) are small non-coding RNAs that can downregulate the expression of genes involved in carcinogenesis [36,37]. The expression of miR-601 is decreased in esophageal squamous cell carcinoma (ESCC) tissues and cells [37]. miR-601 is predicted to regulate the expression of HDAC6 and suppress the proliferation of ESCC cells [37]. miR-22 is down-regulated in cervical cancer and shows inverse correlation with its downstream target HDAC6 [38]. miR-22-overexpressing macrophages inhibited glioma formation by targeting HDAC6 and NF-κB signaling in tumor-associated macrophages (TAMs) [39].
HDAC6 shows aberrant expression in diffuse large B-cell lymphoma (DLBCL). HDAC6 inhibition exerts anti-tumor effects both in vitro and in vivo [40]. The decreased expression of miR-27b in DLBCL tissues predicts the poor overall survival of patients with DLBCL [40]. Rel A/p65 negatively regulates miR-27b expression. HDAC6 inhibition was shown to increase miR-27b expression by acetylation and block of nuclear translocation of RelA/p65 [40]. miR-27b represses mesenchymal-to-epithelial transition (MET) and MET/PI3K/AKT pathway [40].
miR-206 directly decreased the expression of glycolytic enhancer 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase 3 (PFKFB3) and suppressed the proliferation of ovarian cancer cells [41]. miR-206 was reported to suppress the proliferation of head and neck carcinoma cells by targeting HDAC6 [42]. HDAC6 is a direct target of miR-206 and can promote endometrial carcinoma (EC) cell proliferation and metastasis [43]. HDAC6 promotes carcinogenesis through the phosphatase and tensin homolog deleted on chromosome 10 (PTEN)/AKT/the mammalian target of rapamycin (mTOR) pathway [43].
TargetScan analysis predicted that miR-141 could target HDAC6 to protect against lung injury [44] (Figure 3B). miR-433-3p is predicted to target HDAC6 (Figure 3B). miR-141-3p inhibits Janus kinase (JAK2)/signal transducer and the activator of T cells 3 (STAT3) and targets pre-B-cell leukemia homeobox-1 (PBX1) to suppress oral squamous cell carcinoma (OSCC) proliferation [45]. miR-433, targeting HDAC6, decreased the malignant phenotype of cholangiocarcinoma [46]. The downregulation of exportin-5, a protein necessary for the nuclear export of miRNAs, restored the expression of HDAC6 [46], and exporitn-5 overexpression increased the level of mature miR-433 [46]. Table 2 shows the miRNAs that target HDAC6 and the roles of these miRNAs in cancer cell proliferation.

4. HDAC6-Selective Inhibitors

Unlike other HDACs, the development of HDAC6-specific inhibitors has been relatively successful [6,7,47]. NN-390, the first HDAC6-selective inhibitor, has shown therapeutic potential for Group 3 medulloblastoma (MB), an aggressive pediatric brain tumor associated with leptomeningeal metastases and therapy resistance [48]. NM-390 targets MB stem cells and demonstrates a 45-fold increased efficacy over HDAC6 inhibitor citarinostat (ACY-241) [48]. ACY-1215 was shown to inhibit the translocation of GRP78 to the plasma membrane by inhibiting the PI3K/AKT signaling [47]. It suppressed tumor growth by 50% in a xenograft model of cholangiocarcinoma cells [47]. ACY-1215 inhibited BCR-ABL signaling while increasing the expression of PTEN in chronic myeloid leukemia cells [49] and suppressed the proliferation of esophageal squamous cell carcinoma by inhibiting the PI3K/AKT/ERK pathway [50]. ACY-1215 suppressed the proliferation of esophageal squamous cells in mouse model of xenograft [50]. Zeta55, a selective inhibitor of HDAC6, was reported to degrade the androgen receptor (AR) and reduce the growth of AR-overexpressing prostate cancer cells both in vitro and in a castration-resistant prostate cancer (CRPC) xenograft model [51]. MPT0G612, an inhibitor of HDAC6, induced the apoptosis of colorectal cancer (CRC) cells and decreased the expression of programmed death ligand-1 (PD-L1) [52]. WT-161, an inhibitor of HDAC6, induced the apoptosis of osteosarcoma cells by increasing the expression of PTEN [53]. WT-161 also induced the apoptotic cell death of various breast cancer cells by decreasing the expression levels of epidermal growth factor receptor (EGFR), epidermal growth factor receptor 2 (HER2), and estrogen receptor α (ERα), as well as downstream signaling [54]. Azaindolylsulfonamide, an HDAC6 inhibitor, was shown to target the long non-coding RNA LINC00461, causing cell-cycle arrest and suppressing the proliferation of glioblastoma cells [55]. MPT0B451, a dual inhibitor of HDAC6 and tubulin inhibitor, suppressed tumor growth in HL-60 and PC-3 xenograft models [56].
LSD1 (lysine-specific demethylase 1) and HDAC6 selective dual inhibitors exerted synergistic effects in a xenograft model of multiple myeloma, exhibiting better effects than treatment with the single agent [57]. The combination of tubacin and temozolomide (TMZ), an alkylating agent used to treat glioblastoma, induces severe glioma cell death by blocking the fusion of autophagosome and lysosome [58]. HDAC6 inhibitors (ACY-1215, CAY10603, and tubastatin A) enhance the sensitivity of glioblastoma cells to TMZ by increasing the levels of DNA mismatch repair proteins such as mutS homolog 2 (MSH2) and MSH6 [59]. ACY-1215 was shown to enhance the sensitivity of gallbladder cancer cells to gemcitabine and oxaliplatin by decreasing the expression of Bcl-2 while increasing expression levels of caspase-3, caspase-7, and Bax [60]. The combination of ACY-1215 and gemcitabine enhanced the sensitivity of prostate cancer cells to gemcitabine in pancreatic ductal adenocarcinoma (PDAC) xenografts by inducing the apoptotic effects such as cleavage of caspase-3 [61]. ACY-1215 enhanced the sensitivity of triple-negative breast cancer cells (TNBC) to eribulin by increasing tubulin acetylation [62]. The combination of AC-1215 and the WEE1 G2 checkpoint kinase (WEE1) inhibitor adavosertib (Adv) suppressed checkpoint kinase 1 (Chk1) activity, which synergistically enhanced the apoptosis in head and neck squamous cell carcinoma (HNSCC) cells via mitotic catastrophe in a p53-dependent manner in [63]. In combination with 5-Fu, WT161 synergistically inhibited osteosarcoma cells both in vitro and in vivo by increasing the level of PTEN [53]. HDAC6-selective inhibitor A452 can enhance sensitivity of acquired bortezomib (BTZ)-resistant multiple myeloma cells to bortezomib by inhibiting the activation of ERK and NF-κB in [64]. The combination of imatinib, a tyrosine kinase inhibitor, and the HDAC6 inhibitor 7b was shown to synergistically induce caspase-dependent apoptotic cell death and decrease the proportion of leukemia stem cells [65]. HDAC6 inhibitor C1A, in combination with phosphatidylinositol 3′-kinase (PI3K) inhibitor, also had synergistic effects on caspase 3/7 activity in various cancer cells [66]. Table 3 shows HDAC6 inhibitors that enhance the sensitivity of cancer cells to anti-cancer drugs.

5. Role of HDAC6 in Anti-Cancer Drug Resistance

The overexpression of ATP binding cassette subfamily B member 1 (ABCB1) or ATP binding cassette subfamily G member 2 (ABCG2) decreased the sensitivity of human cancer cells to ACY-241 [67]. Thus, HDAC6 might play a role in anti-cancer drug resistance. Sahaquine, a selective HDAC6 inhibitor, reduced the level of p-glycoprotein and EGFR activity and enhanced the sensitivity of glioblastoma to anti-cancer cancer drugs such as TMZ, quercetin, and buthionine sulfoximine [68] (Figure 4A). A high level of HDAC6 was reported to be closely related to tamoxifen resistance [69].
HDAC6 confers resistance to gefitinib by stabilizing EGFR [70]. CAY10603, an inhibitor of HDAC6, acted synergistically with gefitinib to induce the apoptosis of lung adenocarcinoma cell lines by destabilizing EGFR [70] (Figure 4A), whereas sorafenib activated EGFR signaling by stabilizing HDAC6 [71] (Figure 4A). The inhibition of HDAC6 can synergize with sorafenib to induce the apoptotic cell death of NSCLCs by suppressing sorafenib-mediated EGFR pathway activation [71]. HDAC6 binds to tubulin β3 and confers resistance to microtubule-targeting anti-cancer drugs in melanoma cells [35] (Figure 4B). HDAC6 downregulation decreased the expression levels of MDR1 and tubulin β3 in anti-cancer drug-resistant melanoma cells [35] (Figure 4B).
Pan-HDAC inhibitors, including trichostatin A, suberoylanilide hydroxamic acid, and sodium butyrate, increased the PTEN expression by inducing the acetylation of transcription factor p53-related p63 protein (ΔNp63α). The overexpression of ΔNp63α downregulated membrane-bound PTEN but enhanced the nuclear translocation of PTEN, leading to cisplatin resistance in oral cancer cells [72]. Inhibiting either HDAC1 or HDAC6 prevented the nuclear translocation of PTEN and attenuated cisplatin resistance in oral cancer cells [72]. Figure 4 shows the mechanism of HDAC6-promoted anti-cancer drug resistance.

6. Correlations between HDAC6 and PD-L1

A high level of HDAC6 can predict the poor prognosis of glioblastoma patients [73,74]. Glioblastoma patients with poor prognoses display the activation of the transforming growth factor-β (TGF-β)/Smad pathway compared to long-term survivors [74]. The short overall survival (OS) group exhibited a decrease in Smad 7 expression and a low level of p21 [74]. A high level of HDAC6 and the activation of the TGF-β/Smad pathway promoted glioblastoma progression [74]. A high level of HDAC6 predicted low progression-free survival (p = 0.001) and overall survival (p = 0.008) in patients with serous carcinoma [75]. A low level of HDAC6 was reported to predict a high overall response rate (ORR) and prolonged low progression-free survival (PFS) of patients with NSCLC [76].
Cancer cells can activate immune checkpoint pathways to evade immune surveillance. Immune checkpoint molecules, such as programmed cell death-1 (PD-1) and PD-L1, allow cancer cells to proliferate by inhibiting cytolytic T cell (CTL) and natural killer (NK) cell activity [77,78]. EGFR-STAT signaling was shown to increase the expression of PD-L1 in ovarian cancer cells [79]. EGFR-P38MAPK signaling was critical for the increased expression of PD-L1 in hepatocellular carcinoma cells [80]. MYC can act as a transcriptional activator of PD-L1 [81]. It was shown to drive immune evasion by decreasing immune cell infiltration and HLA class I expression [82]. The downregulation of PTEN promoted the proliferation of NSCLCs by increasing the expression of PD-L1 [83]. Colorectal cancer patients with p53 mutation displayed a high level of PD-L1 [84]. Figure 5A shows the regulation of PD-L1 expression in various cancer cells.
High PD-L1 expression can predict the poor prognosis of patients with colorectal cancers [85]. A high PD-L1 level in endometrial cancer stem-like cells (ECSCs) was correlated with self-renewal capability [86]. PD-L1 downregulation decreased the expression levels of pluripotency-related genes (aldehyde dehydrogenase 1 (ALDH1), CD133, OCT4, SOX2, and NANOG), impaired the proliferation of ECSCs, and decreased the number of CD133 positive ECSCs and the number of stem-like spheres [86]. PD-L1 downregulation decreased the tumorigenicity of ECSCs. The self-renewal capability of ECSCs induced by PD-L1 was shown to depend on hypoxia-inducible factor-1α (HIF-1α) and HIF-2α activation [86]. These reports indicate the role of PD-L1 in cancer cell proliferation.
The elevated expression of PD-L1 conferred cisplatin resistance to NSCLCs [87]. PD-L1 promoted the resistance of breast cancer cells to HER2-targeting anti-cancer drugs such as trastuzumab [88]. PD-L1 amplification was shown to be responsible for the acquired resistance of NSCLCs to EGFR-tyrosine kinase inhibitors (EGFR-TKIs) [89].
The PD-L1 level showed a positive correlation with the HDAC6 level in serous carcinoma [75]. The selective HDAC6 inhibitor nexturastat A decreased PD-L1 expression in A549 cells [76]. HDAC6 inhibition decreased the expression level of PD-L1 in urothelial cancer cell lines [90]. The pharmacological or genetic abrogation of HDAC6 in osteosarcoma cell lines decreased PD-L1 expression, which activated the inhibitory regulatory pathway of PD-1 expression in T cells [91]. HDAC6 increased PD-L1 expression by activating STAT3 signaling [91]. Thus, HDAC6 inhibitors may enhance the sensitivity of cancer cells to the PD-L1 blockade. Figure 5B shows the correlation between HDAC6 and PD-L1 in various types of cancer cells.

7. Combination of HDAC6 Inhibition with Immune Check Point Blockade

Tumor immunotherapy can exert anti-tumor effects by blocking the aggresome and proteasome pathways and reducing the number of M2 macrophages while increasing sensitivity to PD-L1 blockade [75]. Since HDAC6 can increase the expression of PD-L1, inhibiting HDAC6 might enhance the efficacy of tumor immunotherapy. XP5, an inhibitor of HDAC6, in combination with a small-molecule PD-L1 inhibitor (NP19) synergistically enhanced anti-tumor immune responses by increasing tumor-infiltrating lymphocytes while decreasing PD-L1 expression in melanoma [92]. HDAC6 inhibition enhanced the sensitivity of cancer cells to anti-PD-L1 blockade by reducing the number of M2 macrophages [93]. AT-rich interactive domain-containing protein 1A (ARID1A), an SWI/SNF component, was shown to be mutated in more than 50% of ovarian clear cell carcinomas (OCCCs). ARID1A targets the inositol-requiring transmembrane kinase/endoribonuclease 1α (IRE1α)-X-box binding protein 1 (XBP1) axis of ER stress response [94]. B-109, an inhibitor of IRE1α, suppressed the growth of ARID1A-mutant OCCCs [94]. B-109, in combination with HDAC6 inhibition, synergistically inhibited the growth of ARID1A-inactivated OCCCs [94]. ARID1A-deficient EC cells were shown to require HDAC6 for progression [95]. ARID1A repressed the expression of PD-L1 [96]. The combination of ACY-1215 and an anti-PD-L1 antibody improved the anti-tumor response in ARIDA-inactivated OCCCs by increasing the number of interferon γ (IFNγ)-positive cytotoxic CD8+ T cells [96]. Combination treatment with ACY-738 can augment the anti-tumor efficacy of anti-PD-1 and anti-PD-L1 monoclonal antibodies in an Eμ-TCL1 adoptive transfer murine model by activating the cytotoxic CD8+ T-cell phenotype in chronic lymphocytic leukemia [97]. The combination of the HDAC6 inhibitor nexturastat and an anti-PD-1 antibody induced anti-tumor immune responses in syngeneic melanoma tumor models by increasing the infiltration of immune cells, increasing central and effector T cell memory, and decreasing pro-tumorigenic M2 macrophages [98]. Combining A452 (an HDAC6 selective inhibitor) with either lenalidomide or pomalidomide (immunomodulatory drugs (IMiDs)) synergistically increased the apoptosis in multiple myeloma cells by inactivating AKT and extracellular signal-regulated kinase (ERK)1/2 [99]. Figure 5C and Table 4 show the effects of HDAC6 inhibition in terms of enhancing the sensitivity of cancer cells to tumor immune therapy.

8. Role of HDAC6 in Autophagy

Autophagy either promotes or blocks tumorigenesis. Epigenetic regulation is involved in autophagy [9,100,101]. Increased autophagy activity was associated with the decreased expression of enhancer of zeste homolog 2 (EZH2), a histone methyltransferase [102], and EZH2 was shown to negatively regulate autophagy in NSCLC cell lines [102].
Autophagy promotes the degradation and recycling of damaged organelles regulating cellular homeostasis and energy metabolism [103,104]. HDAC6 is necessary for removal of misfolded proteins [105] and ubiquitin-mediated protein degradation (Figure 6). Among HDACs, HDAC6 is unique in that it has intrinsic ubiquitin-binding activity [106], which can modulate autophagy [103,107,108]. Proteins that are not degraded by proteasomes are sequestered into an insoluble aggresome in HDAC6- and dynein-dependent manners [109] (Figure 6). In other words, aggresome formation occurs when protein degradation by proteasome function is overwhelmed. Aggresome formation is a cytoprotective response to misfolded/damaged proteins and induces their clearance by autophagy. These proteasomes are co-localized with autophagic receptor sequestosome 1 (SQSTM1) and cleared through selective macroautophagy [109]. HDAC6 can interact with cortactin and promote the polymerization of F-actin. HDAC6 is necessary for autophagosome-lysosome fusion during an autophagic clearance process [110] (Figure 6). Since HDAC6 can transport ubiquitinated protein aggregates to the microtubule organizing center (MTOC) for aggresome formation and autophagosomal clearance [111], the dynein adapter HDAC6 is critical for microtubule transport and the assembly of inflammasomes [111].
ACY-1215 inhibited aggresome formation and autophagy, resulting in apoptotic cell death [112] (Figure 6). The inhibition of HDAC6 promoted the degradation of oncoprotein by acetylating heat shock protein 90 (HSP90) [113] (Figure 6).
Nutrient deprivation increased the expression of transactive response DNA binding protein-43 (TDP-43) in glioblastoma cell lines [114]. TDP-43 activated autophagy while suppressing stress-induced apoptosis in an HDAC6-dependent manner [114]. High levels of TDP-43 and HDAC6 were shown to predict low relapse-free survival in patients with glioblastoma [114]. P62, a selective receptor of autophagy, increased the expression of HDAC6, promoted epithelial-to-mesenchymal transition (EMT), and enhanced the proliferation of prostate cancer cells [115]. J22352, an inhibitor of HDAC6, promoted HDAC6 degradation and induced anti-cancer effects by inhibiting autophagy in glioblastoma [116]. The downregulation of HDAC6 inhibited autophagosome-lysosome fusion and decreased the expression of Myc [117]. The growth of Myc-positive neuroblastoma cells was inhibited in response to HDAC6 inhibitors [117]. Thus, HDAC6-promoted autophagy plays a role in cancer cell proliferation.
TGF-β was shown to induce autophagy in cancer-associated fibroblasts and promote EMT and metastasis [118]. Both the downregulation of HDAC6 and the overexpression of transmembrane protein 100 (TMEM100) were shown to inhibit TGF-β1-induced EMT and suppress the activation of the Wnt/β-catenin signaling pathway in NSCLCs [119]. A low level of TMEM100 predicted the poor prognosis of patients with lung cancers [119]. Snail2 interacted with HDAC6 and recruited HDAC6 to the promoter sequences of E-cadherin, which decreased E-cadherin expression in colorectal cancer cells [120]. Zeb1, a critical regulator of EMT, increased the expression of LC-3II and the resistance of TNBCs to anti-cancer drugs [121]. Butyrate enhanced the anti-cancer effects of HDAC6 inhibitors in cholangiocarcinoma (CCA) by decreasing Zeb1 expression [122]. Enhanced autophagy increased the expression of the multi-drug resistance gene (MDR) [123]. Thus, HDAC6-promoted autophagy is closely related to anti-cancer drug resistance.
Stem cell-like breast cancer cells are known to acquire resistance to metformin by increasing glycolysis [124]. The anti-Warbrug effect of R406 induced the apoptosis of glioma stem cells [125], and the inhibition of autophagy suppressed Ras-mediated cellular proliferation and glycolytic capacity [126]. Sorafenib induced autophagy by increasing the expression of HDAC6 and promoting glycolysis in hepatocellular carcinoma cells [127]. HDAC6 enhanced glycolytic enzyme pyruvate kinase M2 (PKM2) activity by deacetylating HSP90 [127]. The inhibition or knockout of HDAC6 reduced glycolytic metabolism in triple-negative breast cancers [128]. The inhibition of 6-Phosphofructo-2-Kinase/Fructose-2,6-Biphosphatase 3 (PFKFB3) enhanced the sensitivity of NSCLCs to erlotinib by decreasing autophagy [129], and inhibiting glycolysis was shown to overcome taxol-resistance in colorectal cancer cells [130]. These reports suggest that autophagy increases the expression of HDAC6, which in turn enhances glycolysis and anti-cancer drug resistance.

9. Targeting Autophagy for Overcoming Anti-Cancer Drug Resistance

Autophagy is considered a cellular adaptive response against hypoxia, nutrient deprivation, and energy deprivation, indicating its cytoprotective role [131]. Protective autophagy might be necessary for the survival of cancer cells in response to anti-cancer therapies [132,133,134]. Autophagy enables tumor cells to maintain functional mitochondria for survival [135]. ROS-mediated JNK activation increased the level of autophagic flux such as autophagy related 5 (ATG5) and ATG7 [136]. Lysosomal protein transmembrane 4 beta (LAPTM4B) is upregulated in many types of cancers, is necessary for cancer cell proliferation, and confers anti-cancer drug resistance [137]. LAPTM4B promoted cancer cell proliferation via the PI3K/AKT signaling pathway and mediated EGFR family-promoted autophagy [137]. Glioblastoma cells were shown to activate protective autophagy in response to hypoxia and displayed an increased expression of ATG9A [138].
A high level of ATG9A or ATG16L1 was reported to predict the poor overall survival and earlier relapse of patients with oral squamous cell carcinoma [139]. A high level of SQSTM1/p62 can predict a poor response to cetuximab [140]. Autophagy is induced by chemotherapy and is associated with chemo-resistance [141]. Increased autophagic flux was shown to play a role in the resistance to anti-cancer therapies, including radiation therapy and chemotherapy [142,143,144]. Autophagy enhanced cellular tolerance to various stresses [145]. Anti-cancer drugs such as erlotinib can induce autophagy [129,146]. Anti-cancer drug-resistant gastric cancer cells (AGSR) displayed an increased expression level of CAGE (a cancer-associated gene) and autophagic flux compared to anti-cancer drug-sensitive parental gastric cancer cells (AGS cells) [36]. CAGE was shown to bind to Beclin1 (a mediator of autophagy) and confers resistance to various anti-cancer drugs [36,147]. These reports imply that targeting autophagy may enhance sensitivity to anti-cancer drugs.
Protective autophagy was shown to be responsible for resistance to AG1478, an inhibitor of EGFR tyrosine kinase, in ovarian cancer cells [148]. Protective autophagy also conferred resistance to erlotinib in head and neck squamous cell carcinomas [149]. Microtubule-associated protein 1 light chain 3-alpha (LC3A)-mediated autophagy conferred resistance to EGFR-TKIs in carcinoma cells [150]. Jolkinolide B (JB) inhibited both AKT signaling and cytoprotective autophagy, potentiating the anti-proliferative efficacy of the mTOR inhibitor in both PTEN-deficient and cisplatin-resistant bladder cancer cells [151].
Treatment with low-intensity focused ultrasound and microbubble (LIFU+MB) combined with paclitaxel (PTX) increased the apoptosis of paclitaxel-resistant ovarian cells by decreasing autophagy [152]. Thus, targeting the autophagy process can improve the efficacy of chemotherapy. Autophagy inhibition by dichloroacetate sensitized breast cancer cells to paclitaxel [153]. The inhibition of autophagy by baffilomycin enhanced advanced glycation end products (AGEs) and induced apoptosis [154]. The inhibition of cytoprotective autophagy by JB enhanced the sensitivity of bladder cancer cells to mTOR inhibitors such as temsirolimus, rapamycin, and everolimus [151]. Ulinastatin (UTI) inhibited epirubicin (EPI)-induced protective autophagy, promoted apoptosis, and enhanced the sensitivity of the hepatic cancer cells to EPI [155]. The combination of the autophagy inhibitor chloroquine (CQ) with bevacizumab (Avastin) synergistically inhibited glioblastoma growth in vivo [138]. Blocking autophagy was shown to enhance the sensitivity of cancer cells to toosendanin, a vacuolar-type H(+)-translocating ATPase inhibitor [156]. Thus, the HDAC6 inhibitor can enhance the sensitivity of cancer cells to anti-cancer drugs by inhibiting autophagy.

10. Conclusions

Unlike other HDACs, HDAC6 knockout did not induce a toxic effect [157]. Thus, HDAC6 can serve as a target for developing anti-cancer drugs. Epigenetic modifications play critical roles in autophagy and anti-cancer drug resistance. Based on its role in autophagy and anti-cancer drug resistance, HDAC6 can serve as a target for developing single or combined treatments for cancer.
Investigations into the mechanism of the regulation of HDAC6 expression are necessary for a better understanding of HDAC6-promoted autophagy and anti-cancer drug resistance. However, extensive efforts have not been made to identify transcription factors that regulate HDAC6 expression. It will also be necessary to identify genes regulated by HDAC6. These genes might serve as targets for developing anti-cancer drugs.
Many reports showed that miRNAs targeting HDAC6 regulated cancer cell proliferation [42]. TargetScan analysis can predict miRNAs that target HDAC6 (Figure 3B), and these microRNAs can be anti-cancer drugs. miRNA-mimics or -inhibitors can treat cancer and overcome resistance to anti-cancer drugs [158,159]. Unlike siRNAs, most miRNA-based therapeutics (miR-mimics or miR-inhibitors) are in clinical phase I or phase II trials. miR-mimics can cause off-target effects, be degraded by RNase, and have difficulty penetrating cell membrane. Improved delivery systems for miR-mimics are needed to enhance the therapeutic values of miR-mimics.
The identification of novel HDAC6-binding proteins is necessary for developing anti-cancer drugs for cancer patients expressing high level of HDAC6. The identification of the domain of HDAC6 critical for binding to partner of HDAC6 is necessary for designing HDAC6-targeting anti-cancer drugs. Peptides that correspond to the binding domain of HDAC6 might regulate autophagy and overcome resistance to anti-cancer drugs.
The tumor microenvironment (TME) consists of cancer cells, endothelial cells, cancer associated fibroblasts, and various innate and adaptive immune cells. These immune cells include B and T cells, dendritic cells, myeloid-derived suppressor cells (MDSCs), and macrophages (M1/M2). The TME is critical for cancer initiation and promotion. Cellular interactions involving cancer cells and immune cells can promote cancer cell progression [160]. Exosomal molecules regulated by HDAC6 can serve as targets for developing anti-cancer drugs.
Currently, ACY-1215, ACY-241, KA2507, and JBI-802 are in clinical trials (Table 5). Clinical trials involving HDAC6 inhibitors are mostly in phase I or phase 2. Since tumor displays heterogeneity, it is advisable to employ combination therapy to treat cancer. The combination of ACY-241 with nivolumab (anti-PD-L1 antibody) was shown to cause side effects such as dyspenia (1/17) and pneumonia (2/17) in patients with NSCLCs who had not received an HDAC6 inhibitor or immune checkpoint inhibitor (n = 17) [161]. Among 17 patients, three showed partial or complete responses to combination therapy [161]. At the recommended daily dose of ACY-1215 (160 mg), the combination with BTZ was safe, tolerable, and active in relapsed or refractory patients with multiple myeloma [162]. The overall response rate to this combination therapy in bortezomib refractory patients was 14% [162]. ACY-241 showed synergistic anti-cancer activity with paclitaxel in multiple solid tumor models [163]. Three (3/20) patients showed partial responses [163]. A combination of ACY-241 with paclitaxel showed no appreciable toxicities [163].
The screening of small molecules based on a full-length HDAC6 protein structure is needed to develop HDAC6-specific inhibitors. Concerted effects are needed to minimize cytotoxic effects, improve pharmacokinetics and pharmacodynamics, and improve the chemotherapeutic efficacy of HDAC6-specicfic inhibitors.
Resistance to HDAC6 inhibitors will eventually develop. Thus, the identification of genes that confer resistance to HDAC6 inhibitors will be necessary. For this, the identification of genes regulated by HDAC6 inhibitors is necessary. Mechanisms of resistance to HDAC6 inhibitors may involve the overexpression of PD-L1 and PD-1, epigenetic regulation (acetylation and methylation), drug efflux, the activation of EGFR/PI3K/AKT signaling, the overexpression of oncogenes, and the downregulation of tumor suppressor genes.

Author Contributions

D.J. conceived the idea and wrote the manuscript. H.J. and K.S. critically reviewed the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Research Foundation Grants (2020R1A2C1006996, 2022R1F1A1060031, and 2017M3A9G7072417) and a grant from the BK21 Plus Program.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vergani, B.; Sandrone, G.; Marchini, M.; Ripamonti, C.; Cellupica, E.; Galbiati, E.; Caprini, G.; Pavich, G.; Porro, G.; Rocchio, I.; et al. Novel Benzohydroxamate-Based Potent and Selective Histone Deacetylase 6 (HDAC6) Inhibitors Bearing a Pentaheterocyclic Scaffold: Design, Synthesis, and Biological Evaluation. J. Med. Chem. 2019, 62, 10711–10739. [Google Scholar] [CrossRef] [PubMed]
  2. Leyk, J.; Daly, C.; Janssen-Bienhold, U.; Kennedy, B.N.; Richter-Landsberg, C. HDAC6 inhibition by tubastatin A is protective against oxidative stress in a photoreceptor cell line and restores visual function in a zebrafish model of inherited blindness. Cell Death Dis. 2017, 8, e3028. [Google Scholar] [CrossRef] [PubMed]
  3. Varga, J.K.; Diffley, K.; Welker Leng, K.R.; Fierke, C.A.; Schueler-Furman, O. Structure-based prediction of HDAC6 substrates validated by enzymatic assay reveals determinants of promiscuity and detects new potential substrates. Sci. Rep. 2022, 12, 1788. [Google Scholar] [CrossRef] [PubMed]
  4. Shi, P.; Hoang-Minh, L.B.; Tian, J.; Cheng, A.; Basrai, R.; Kalaria, N.; Lebowitz, J.J.; Khoshbouei, H.; Deleyrolle, L.P.; Sarkisian, M.R. HDAC6 Signaling at Primary Cilia Promotes Proliferation and Restricts Differentiation of Glioma Cells. Cancers 2021, 13, 1644. [Google Scholar] [CrossRef]
  5. Kim, J.Y.; Hwang, H.G.; Lee, J.Y.; Kim, M.; Kim, J.Y. Cortactin deacetylation by HDAC6 and SIRT2 regulates neuronal migration and dendrite morphogenesis during cerebral cortex development. Mol. Brain 2020, 13, 105. [Google Scholar] [CrossRef]
  6. Zhang, X.H.; Kang, H.Q.; Tao, Y.Y.; Li, Y.H.; Zhao, J.R.; Ya, G.; Ma, L.Y.; Liu, H.M. Identification of novel 1,3-diaryl-1,2,4-triazole-capped histone deacetylase 6 inhibitors with potential anti-gastric cancer activity. Eur. J. Med. Chem. 2021, 218, 113392. [Google Scholar] [CrossRef]
  7. Zhang, X.H.; Qin, M.; Wu, H.P.; Khamis, M.Y.; Li, Y.H.; Ma, L.Y.; Liu, H.M. A Review of Progress in Histone Deacetylase 6 Inhibitors Research: Structural Specificity and Functional Diversity. J. Med. Chem. 2021, 64, 1362–1391. [Google Scholar] [CrossRef]
  8. Xu, X.; Ding, P.; Shi, L.; Wu, G.; Ma, X. LukS-PV inhibits hepatocellular carcinoma cells migration by downregulating HDAC6 expression. BMC Cancer 2022, 22, 630. [Google Scholar] [CrossRef]
  9. Urdiciain, A.; Erausquin, E.; Zelaya, M.V.; Zazpe, I.; Lanciego, J.L.; Melendez, B.; Rey, J.A.; Idoate, M.A.; Riobo-Del Galdo, N.A.; Castresana, J.S. Silencing of Histone Deacetylase 6 Decreases Cellular Malignancy and Contributes to Primary Cilium Restoration, Epithelial-to-Mesenchymal Transition Reversion, and Autophagy Inhibition in Glioblastoma Cell Lines. Biology 2021, 10, 467. [Google Scholar] [CrossRef]
  10. Li, S.; Liu, X.; Chen, X.; Zhang, L.; Wang, X. Histone deacetylase 6 promotes growth of glioblastoma through inhibition of SMAD2 signaling. Tumour. Biol. 2015, 36, 9661–9665. [Google Scholar] [CrossRef]
  11. Liu, J.; Gu, J.; Feng, Z.; Yang, Y.; Zhu, N.; Lu, W.; Qi, F. Both HDAC5 and HDAC6 are required for the proliferation and metastasis of melanoma cells. J. Transl. Med. 2016, 14, 7. [Google Scholar] [CrossRef] [PubMed]
  12. Geiger, T.R.; Peeper, D.S. Critical role for TrkB kinase function in anoikis suppression, tumorigenesis, and metastasis. Cancer Res. 2007, 67, 6221–6229. [Google Scholar] [CrossRef] [PubMed]
  13. Sharif, T.; Martell, E.; Dai, C.; Ghassemi-Rad, M.S.; Hanes, M.R.; Murphy, P.J.; Margam, N.N.; Parmar, H.B.; Giacomantonio, C.A.; Duncan, R.; et al. HDAC6 differentially regulates autophagy in stem-like versus differentiated cancer cells. Autophagy 2019, 15, 686–706. [Google Scholar] [CrossRef] [PubMed]
  14. Park, S.J.; Kim, J.K.; Bae, H.J.; Eun, J.W.; Shen, Q.; Kim, H.S.; Shin, W.C.; Yang, H.D.; Lee, E.K.; You, J.S.; et al. HDAC6 sustains growth stimulation by prolonging the activation of EGF receptor through the inhibition of rabaptin-5-mediated early endosome fusion in gastric cancer. Cancer Lett. 2014, 354, 97–106. [Google Scholar] [CrossRef] [PubMed]
  15. Lee, Y.S.; Lim, K.H.; Guo, X.; Kawaguchi, Y.; Gao, Y.; Barrientos, T.; Ordentlich, P.; Wang, X.F.; Counter, C.M.; Yao, T.P. The cytoplasmic deacetylase HDAC6 is required for efficient oncogenic tumorigenesis. Cancer Res. 2008, 68, 7561–7569. [Google Scholar] [CrossRef]
  16. Williams, K.A.; Zhang, M.; Xiang, S.; Hu, C.; Wu, J.Y.; Zhang, S.; Ryan, M.; Cox, A.D.; Der, C.J.; Fang, B.; et al. Extracellular signal-regulated kinase (ERK) phosphorylates histone deacetylase 6 (HDAC6) at serine 1035 to stimulate cell migration. J. Biol. Chem. 2013, 288, 33156–33170. [Google Scholar] [CrossRef]
  17. Wu, J.Y.; Xiang, S.; Zhang, M.; Fang, B.; Huang, H.; Kwon, O.K.; Zhao, Y.; Yang, Z.; Bai, W.; Bepler, G.; et al. Histone deacetylase 6 (HDAC6) deacetylates extracellular signal-regulated kinase 1 (ERK1) and thereby stimulates ERK1 activity. J. Biol. Chem. 2018, 293, 1976–1993. [Google Scholar] [CrossRef]
  18. Liu, J.; Luan, W.; Zhang, Y.; Gu, J.; Shi, Y.; Yang, Y.; Feng, Z.; Qi, F. HDAC6 interacts with PTPN1 to enhance melanoma cells progression. Biochem. Biophys. Res. Commun. 2018, 495, 2630–2636. [Google Scholar] [CrossRef]
  19. Zhang, S.L.; Zhu, H.Y.; Zhou, B.Y.; Chu, Y.; Huo, J.R.; Tan, Y.Y.; Liu, D.L. Histone deacetylase 6 is overexpressed and promotes tumor growth of colon cancer through regulation of the MAPK/ERK signal pathway. Onco Targets Ther. 2019, 12, 2409–2419. [Google Scholar] [CrossRef]
  20. Huang, Z.; Xia, Y.; Hu, K.; Zeng, S.; Wu, L.; Liu, S.; Zhi, C.; Lai, M.; Chen, D.; Xie, L.; et al. Histone deacetylase 6 promotes growth of glioblastoma through the MKK7/JNK/c-Jun signaling pathway. J. Neurochem. 2020, 152, 221–234. [Google Scholar] [CrossRef]
  21. Garcia-Dominguez, D.J.; Hajji, N.; Sanchez-Molina, S.; Figuerola-Bou, E.; de Pablos, R.M.; Espinosa-Oliva, A.M.; Andres-Leon, E.; Terron-Camero, L.C.; Flores-Campos, R.; Pascual-Pasto, G.; et al. Selective inhibition of HDAC6 regulates expression of the oncogenic driver EWSR1-FLI1 through the EWSR1 promoter in Ewing sarcoma. Oncogene 2021, 40, 5843–5853. [Google Scholar] [CrossRef] [PubMed]
  22. Pham, T.Q.; Robinson, K.; Xu, L.; Pavlova, M.N.; Skapek, S.X.; Chen, E.Y. HDAC6 promotes growth, migration/invasion, and self-renewal of rhabdomyosarcoma. Oncogene 2021, 40, 578–591. [Google Scholar] [CrossRef]
  23. Yang, W.B.; Wu, A.C.; Hsu, T.I.; Liou, J.P.; Lo, W.L.; Chang, K.Y.; Chen, P.Y.; Kikkawa, U.; Yang, S.T.; Kao, T.J.; et al. Histone deacetylase 6 acts upstream of DNA damage response activation to support the survival of glioblastoma cells. Cell Death Dis. 2021, 12, 884. [Google Scholar] [CrossRef] [PubMed]
  24. Li, L.; Fang, R.; Liu, B.; Shi, H.; Wang, Y.; Zhang, W.; Zhang, X.; Ye, L. Deacetylation of tumor-suppressor MST1 in Hippo pathway induces its degradation through HBXIP-elevated HDAC6 in promotion of breast cancer growth. Oncogene 2016, 35, 4048–4057. [Google Scholar] [CrossRef] [PubMed]
  25. Ding, G.; Liu, H.D.; Huang, Q.; Liang, H.X.; Ding, Z.H.; Liao, Z.J.; Huang, G. HDAC6 promotes hepatocellular carcinoma progression by inhibiting P53 transcriptional activity. FEBS Lett. 2013, 587, 880–886. [Google Scholar] [CrossRef] [PubMed]
  26. Banik, D.; Noonepalle, S.; Hadley, M.; Palmer, E.; Gracia-Hernandez, M.; Zevallos-Delgado, C.; Manhas, N.; Simonyan, H.; Young, C.N.; Popratiloff, A.; et al. HDAC6 Plays a Noncanonical Role in the Regulation of Antitumor Immune Responses, Dissemination, and Invasiveness of Breast Cancer. Cancer Res. 2020, 80, 3649–3662. [Google Scholar] [CrossRef]
  27. Li, Y.; Zhang, X.; Polakiewicz, R.D.; Yao, T.P.; Comb, M.J. HDAC6 is required for epidermal growth factor-induced beta-catenin nuclear localization. J. Biol. Chem. 2008, 283, 12686–12690. [Google Scholar] [CrossRef]
  28. Shao, Y.; Zhu, F.; Zhu, S.; Bai, L. HDAC6 suppresses microRNA-199a transcription and augments HPV-positive cervical cancer progression through Wnt5a upregulation. Int. J. Biochem. Cell Biol. 2021, 136, 106000. [Google Scholar] [CrossRef]
  29. Mak, A.B.; Nixon, A.M.; Kittanakom, S.; Stewart, J.M.; Chen, G.I.; Curak, J.; Gingras, A.C.; Mazitschek, R.; Neel, B.G.; Stagljar, I.; et al. Regulation of CD133 by HDAC6 promotes beta-catenin signaling to suppress cancer cell differentiation. Cell Rep. 2012, 2, 951–963. [Google Scholar] [CrossRef]
  30. Chao, O.S.; Chang, T.C.; Di Bella, M.A.; Alessandro, R.; Anzanello, F.; Rappa, G.; Goodman, O.B.; Lorico, A. The HDAC6 Inhibitor Tubacin Induces Release of CD133(+) Extracellular Vesicles From Cancer Cells. J. Cell Biochem. 2017, 118, 4414–4424. [Google Scholar] [CrossRef]
  31. Kim, S.; Ko, D.; Lee, Y.; Jang, S.; Lee, Y.; Lee, I.Y.; Kim, S. Anti-cancer activity of the novel 2-hydroxydiarylamide derivatives IMD-0354 and KRT1853 through suppression of cancer cell invasion, proliferation, and survival mediated by TMPRSS4. Sci. Rep. 2019, 9, 10003. [Google Scholar] [CrossRef] [PubMed]
  32. Hu, C.; Zhang, M.; Moses, N.; Hu, C.L.; Polin, L.; Chen, W.; Jang, H.; Heyza, J.; Malysa, A.; Caruso, J.A.; et al. The USP10-HDAC6 axis confers cisplatin resistance in non-small cell lung cancer lacking wild-type p53. Cell Death Dis. 2020, 11, 328. [Google Scholar] [CrossRef] [PubMed]
  33. Zhang, S.L.; Du, X.; Tan, L.N.; Deng, F.H.; Zhou, B.Y.; Zhou, H.J.; Zhu, H.Y.; Chu, Y.; Liu, D.L.; Tan, Y.Y. SET7 interacts with HDAC6 and suppresses the development of colon cancer through inactivation of HDAC6. Am. J. Transl. Res. 2020, 12, 602–611. [Google Scholar]
  34. Gu, X.; Li, X.; Guan, M.; Jiang, C.; Song, Q.; Sun, N.; Zou, Y.; Zhou, Q.; Chen, J.; Qiu, J. Discovery of thiosemicarbazone-containing compounds with potent anti-proliferation activity against drug-resistant K562/A02 cells. Bioorganic Med. Chem. Lett. 2020, 30, 127638. [Google Scholar] [CrossRef]
  35. Kim, Y.; Kim, H.; Jeoung, D. Tubulin Beta3 Serves as a Target of HDAC3 and Mediates Resistance to Microtubule-Targeting Drugs. Mol. Cells 2015, 38, 705–714. [Google Scholar] [CrossRef] [PubMed]
  36. Yeon, M.; Kim, Y.; Pathak, D.; Kwon, E.; Kim, D.Y.; Jeong, M.S.; Jung, H.S.; Jeoung, D. The CAGE-MiR-181b-5p-S1PR1 Axis Regulates Anticancer Drug Resistance and Autophagy in Gastric Cancer Cells. Front. Cell Dev. Biol. 2021, 9, 666387. [Google Scholar] [CrossRef]
  37. Liu, C.; Tian, X.; Sun, H.B.; Wang, Z.F.; Jiang, L.F.; Li, Z.X. MiR-601 inhibits the proliferation and metastasis of esophageal squamous cell carcinoma (ESCC) by targeting HDAC6. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 1069–1076. [Google Scholar]
  38. Wongjampa, W.; Ekalaksananan, T.; Chopjitt, P.; Chuerduangphui, J.; Kleebkaow, P.; Patarapadungkit, N.; Pientong, C. Suppression of miR-22, a tumor suppressor in cervical cancer, by human papillomavirus 16 E6 via a p53/miR-22/HDAC6 pathway. PLoS ONE 2018, 13, e0206644. [Google Scholar] [CrossRef]
  39. Tu, J.; Fang, Y.; Han, D.; Tan, X.; Xu, Z.; Jiang, H.; Wang, X.; Hong, W.; Wei, W. MicroRNA-22 represses glioma development via activation of macrophage-mediated innate and adaptive immune responses. Oncogene 2022, 41, 2444–2457. [Google Scholar] [CrossRef]
  40. Jia, Y.J.; Liu, Z.B.; Wang, W.G.; Sun, C.B.; Wei, P.; Yang, Y.L.; You, M.J.; Yu, B.H.; Li, X.Q.; Zhou, X.Y. HDAC6 regulates microRNA-27b that suppresses proliferation, promotes apoptosis and target MET in diffuse large B-cell lymphoma. Leukemia 2018, 32, 703–711. [Google Scholar] [CrossRef]
  41. Boscaro, C.; Baggio, C.; Carotti, M.; Sandona, D.; Trevisi, L.; Cignarella, A.; Bolego, C. Targeting of PFKFB3 with miR-206 but not mir-26b inhibits ovarian cancer cell proliferation and migration involving FAK downregulation. FASEB J. 2022, 36, e22140. [Google Scholar] [CrossRef] [PubMed]
  42. Liu, F.; Zhao, X.; Qian, Y.; Zhang, J.; Zhang, Y.; Yin, R. MiR-206 inhibits Head and neck squamous cell carcinoma cell progression by targeting HDAC6 via PTEN/AKT/mTOR pathway. Biomed. Pharmacother. 2017, 96, 229–237. [Google Scholar] [CrossRef] [PubMed]
  43. Zheng, Y.; Yang, X.; Wang, C.; Zhang, S.; Wang, Z.; Li, M.; Wang, Y.; Wang, X.; Yang, X. HDAC6, modulated by miR-206, promotes endometrial cancer progression through the PTEN/AKT/mTOR pathway. Sci. Rep. 2020, 10, 3576. [Google Scholar] [CrossRef] [PubMed]
  44. Zheng, F.; Zhu, J.; Zhang, W.; Fu, Y.; Lin, Z. Thal protects against paraquat-induced lung injury through a microRNA-141/HDAC6/IkappaBalpha-NF-kappaB axis in rat and cell models. Basic Clin. Pharmacol. Toxicol. 2021, 128, 334–347. [Google Scholar] [CrossRef]
  45. Cao, M.; Tian, K.; Sun, W.; Xu, J.; Tang, Y.; Wu, S. MicroRNA-141-3p inhibits the progression of oral squamous cell carcinoma via targeting PBX1 through the JAK2/STAT3 pathway. Exp. Ther. Med. 2022, 23, 97. [Google Scholar] [CrossRef]
  46. Mansini, A.P.; Lorenzo Pisarello, M.J.; Thelen, K.M.; Cruz-Reyes, M.; Peixoto, E.; Jin, S.; Howard, B.N.; Trussoni, C.E.; Gajdos, G.B.; LaRusso, N.F.; et al. MicroRNA (miR)-433 and miR-22 dysregulations induce histone-deacetylase-6 overexpression and ciliary loss in cholangiocarcinoma. Hepatology 2018, 68, 561–573. [Google Scholar] [CrossRef]
  47. Kim, C.; Lee, S.; Kim, D.; Lee, D.S.; Lee, E.; Yoo, C.; Kim, K.P. Blockade of GRP78 Translocation to the Cell Surface by HDAC6 Inhibition Suppresses Proliferation of Cholangiocarcinoma Cells. Anticancer Res. 2022, 42, 471–482. [Google Scholar] [CrossRef]
  48. Nawar, N.; Bukhari, S.; Adile, A.A.; Suk, Y.; Manaswiyoungkul, P.; Toutah, K.; Olaoye, O.O.; Raouf, Y.S.; Sedighi, A.; Garcha, H.K.; et al. Discovery of HDAC6-Selective Inhibitor NN-390 with in Vitro Efficacy in Group 3 Medulloblastoma. J. Med. Chem. 2022, 65, 3193–3217. [Google Scholar] [CrossRef]
  49. Qin, Y.; Liang, Y.; Jiang, G.; Peng, Y.; Feng, W. ACY-1215 suppresses the proliferation and induces apoptosis of chronic myeloid leukemia cells via the ROS/PTEN/Akt pathway. Cell Stress Chaperones 2022, 27, 383–396. [Google Scholar] [CrossRef]
  50. Cao, J.; Lv, W.; Wang, L.; Xu, J.; Yuan, P.; Huang, S.; He, Z.; Hu, J. Ricolinostat (ACY-1215) suppresses proliferation and promotes apoptosis in esophageal squamous cell carcinoma via miR-30d/PI3K/AKT/mTOR and ERK pathways. Cell Death Dis. 2018, 9, 817. [Google Scholar] [CrossRef]
  51. Zhou, M.; Zheng, H.; Li, Y.; Huang, H.; Min, X.; Dai, S.; Zhou, W.; Chen, Z.; Xu, G.; Chen, Y. Discovery of a novel AR/HDAC6 dual inhibitor for prostate cancer treatment. Aging 2021, 13, 6982–6998. [Google Scholar] [CrossRef] [PubMed]
  52. Chen, M.C.; Lin, Y.C.; Liao, Y.H.; Liou, J.P.; Chen, C.H. MPT0G612, a Novel HDAC6 Inhibitor, Induces Apoptosis and Suppresses IFN-gamma-Induced Programmed Death-Ligand 1 in Human Colorectal Carcinoma Cells. Cancers 2019, 11, 1617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Sun, J.; Wu, W.; Tang, X.; Zhang, F.; Ju, C.; Liu, R.; Liang, Y.; Yu, B.; Lv, B.; Guo, Y.; et al. HDAC6 inhibitor WT161 performs anti-tumor effect on osteosarcoma and synergistically interacts with 5-FU. Biosci. Rep. 2021, 41, BSR20203905. [Google Scholar] [CrossRef] [PubMed]
  54. Hideshima, T.; Mazitschek, R.; Qi, J.; Mimura, N.; Tseng, J.C.; Kung, A.L.; Bradner, J.E.; Anderson, K.C. HDAC6 inhibitor WT161 downregulates growth factor receptors in breast cancer. Oncotarget 2017, 8, 80109–80123. [Google Scholar] [CrossRef] [PubMed]
  55. Wu, A.C.; Yang, W.B.; Chang, K.Y.; Lee, J.S.; Liou, J.P.; Su, R.Y.; Cheng, S.M.; Hwang, D.Y.; Kikkawa, U.; Hsu, T.I.; et al. HDAC6 involves in regulating the lncRNA-microRNA-mRNA network to promote the proliferation of glioblastoma cells. J. Exp. Clin. Cancer Res. 2022, 41, 47. [Google Scholar] [CrossRef] [PubMed]
  56. Wu, Y.W.; Hsu, K.C.; Lee, H.Y.; Huang, T.C.; Lin, T.E.; Chen, Y.L.; Sung, T.Y.; Liou, J.P.; Hwang-Verslues, W.W.; Pan, S.L.; et al. A Novel Dual HDAC6 and Tubulin Inhibitor, MPT0B451, Displays Anti-tumor Ability in Human Cancer Cells in Vitro and in Vivo. Front. Pharmacol. 2018, 9, 205. [Google Scholar] [CrossRef]
  57. Naveen Sadhu, M.; Sivanandhan, D.; Gajendran, C.; Tantry, S.; Dewang, P.; Murugan, K.; Chickamunivenkatappa, S.; Zainuddin, M.; Nair, S.; Vaithilingam, K.; et al. Novel dual LSD1/HDAC6 inhibitors for the treatment of multiple myeloma. Bioorganic Med. Chem. Lett. 2021, 34, 127763. [Google Scholar] [CrossRef]
  58. Yin, C.; Li, P. Growth Suppression of Glioma Cells Using HDAC6 Inhibitor, Tubacin. Open Med. 2018, 13, 221–226. [Google Scholar] [CrossRef]
  59. Kim, G.W.; Lee, D.H.; Yeon, S.K.; Jeon, Y.H.; Yoo, J.; Lee, S.W.; Kwon, S.H. Temozolomide-resistant Glioblastoma Depends on HDAC6 Activity Through Regulation of DNA Mismatch Repair. Anticancer Res. 2019, 39, 6731–6741. [Google Scholar] [CrossRef]
  60. Ruan, Y.; Wang, L.; Lu, Y. HDAC6 inhibitor, ACY1215 suppress the proliferation and induce apoptosis of gallbladder cancer cells and increased the chemotherapy effect of gemcitabine and oxaliplatin. Drug Dev. Res. 2021, 82, 598–604. [Google Scholar] [CrossRef]
  61. Laschanzky, R.S.; Humphrey, L.E.; Ma, J.; Smith, L.M.; Enke, T.J.; Shukla, S.K.; Dasgupta, A.; Singh, P.K.; Howell, G.M.; Brattain, M.G.; et al. Selective Inhibition of Histone Deacetylases 1/2/6 in Combination with Gemcitabine: A Promising Combination for Pancreatic Cancer Therapy. Cancers 2019, 11, 1327. [Google Scholar] [CrossRef] [PubMed]
  62. Oba, T.; Ono, M.; Matoba, H.; Uehara, T.; Hasegawa, Y.; Ito, K.I. HDAC6 inhibition enhances the anti-tumor effect of eribulin through tubulin acetylation in triple-negative breast cancer cells. Breast Cancer Res. Treat. 2021, 186, 37–51. [Google Scholar] [CrossRef] [PubMed]
  63. Miyake, K.; Takano, N.; Kazama, H.; Kikuchi, H.; Hiramoto, M.; Tsukahara, K.; Miyazawa, K. Ricolinostat enhances adavosertibinduced mitotic catastrophe in TP53mutated head and neck squamous cell carcinoma cells. Int. J. Oncol. 2022, 60, 54. [Google Scholar] [CrossRef] [PubMed]
  64. Lee, S.W.; Yeon, S.K.; Kim, G.W.; Lee, D.H.; Jeon, Y.H.; Yoo, J.; Kim, S.Y.; Kwon, S.H. HDAC6-Selective Inhibitor Overcomes Bortezomib Resistance in Multiple Myeloma. Int. J. Mol. Sci. 2021, 22, 1341. [Google Scholar] [CrossRef] [PubMed]
  65. Losson, H.; Gajulapalli, S.R.; Lernoux, M.; Lee, J.Y.; Mazumder, A.; Gerard, D.; Seidel, C.; Hahn, H.; Christov, C.; Dicato, M.; et al. The HDAC6 inhibitor 7b induces BCR-ABL ubiquitination and downregulation and synergizes with imatinib to trigger apoptosis in chronic myeloid leukemia. Pharmacol. Res. 2020, 160, 105058. [Google Scholar] [CrossRef]
  66. Kaliszczak, M.; Trousil, S.; Ali, T.; Aboagye, E.O. AKT activation controls cell survival in response to HDAC6 inhibition. Cell Death Dis. 2016, 7, e2286. [Google Scholar] [CrossRef]
  67. Wu, C.P.; Hung, C.Y.; Lusvarghi, S.; Chang, Y.F.; Hsiao, S.H.; Huang, Y.H.; Hung, T.H.; Yu, J.S.; Ambudkar, S.V. Overexpression of Human ABCB1 and ABCG2 Reduces the Susceptibility of Cancer Cells to the Histone Deacetylase 6-Specific Inhibitor Citarinostat. Int. J. Mol. Sci. 2021, 22, 2592. [Google Scholar] [CrossRef]
  68. Zhang, I.; Beus, M.; Stochaj, U.; Le, P.U.; Zorc, B.; Rajic, Z.; Petrecca, K.; Maysinger, D. Inhibition of glioblastoma cell proliferation, invasion, and mechanism of action of a novel hydroxamic acid hybrid molecule. Cell Death Discov. 2018, 4, 41. [Google Scholar] [CrossRef]
  69. Yu, S.; Cai, X.; Wu, C.; Liu, Y.; Zhang, J.; Gong, X.; Wang, X.; Wu, X.; Zhu, T.; Mo, L.; et al. Targeting HSP90-HDAC6 Regulating Network Implicates Precision Treatment of Breast Cancer. Int. J. Biol. Sci. 2017, 13, 505–517. [Google Scholar] [CrossRef]
  70. Wang, Z.; Tang, F.; Hu, P.; Wang, Y.; Gong, J.; Sun, S.; Xie, C. HDAC6 promotes cell proliferation and confers resistance to gefitinib in lung adenocarcinoma. Oncol. Rep. 2016, 36, 589–597. [Google Scholar] [CrossRef]
  71. Wang, Z.; Hu, P.; Tang, F.; Xie, C. HDAC6-mediated EGFR stabilization and activation restrict cell response to sorafenib in non-small cell lung cancer cells. Med. Oncol. 2016, 33, 50. [Google Scholar] [CrossRef] [PubMed]
  72. Hao, T.; Gan, Y.H. DeltaNp63alpha promotes the expression and nuclear translocation of PTEN, leading to cisplatin resistance in oral cancer cells. Am. J. Transl. Res. 2020, 12, 6187–6203. [Google Scholar] [PubMed]
  73. Yang, W.B.; Hsu, C.C.; Hsu, T.I.; Liou, J.P.; Chang, K.Y.; Chen, P.Y.; Liu, J.J.; Yang, S.T.; Wang, J.Y.; Yeh, S.H.; et al. Increased activation of HDAC1/2/6 and Sp1 underlies therapeutic resistance and tumor growth in glioblastoma. Neuro Oncol. 2020, 22, 1439–1451. [Google Scholar] [CrossRef]
  74. Sferra, R.; Pompili, S.; Festuccia, C.; Marampon, F.; Gravina, G.L.; Ventura, L.; Di Cesare, E.; Cicchinelli, S.; Gaudio, E.; Vetuschi, A. The possible prognostic role of histone deacetylase and transforming growth factor beta/Smad signaling in high grade gliomas treated by radio-chemotherapy: A preliminary immunohistochemical study. Eur. J. Histochem. 2017, 61, 2732. [Google Scholar] [CrossRef] [PubMed]
  75. Yano, M.; Miyazawa, M.; Ogane, N.; Ogasawara, A.; Hasegawa, K.; Narahara, H.; Yasuda, M. Up-regulation of HDAC6 Results in Poor Prognosis and Chemoresistance in Patients With Advanced Ovarian High-grade Serous Carcinoma. Anticancer Res. 2021, 41, 1647–1654. [Google Scholar] [CrossRef]
  76. Shin, H.S.; Choi, J.; Lee, J.; Lee, S.Y. Histone Deacetylase as a Valuable Predictive Biomarker and Therapeutic Target in Immunotherapy for Non-Small Cell Lung Cancer. Cancer Res. Treat. 2022, 54, 458–468. [Google Scholar] [CrossRef]
  77. Chakrabarti, J.; Koh, V.; Steele, N.; Hawkins, J.; Ito, Y.; Merchant, J.L.; Wang, J.; Helmrath, M.A.; Ahmad, S.A.; So, J.B.Y.; et al. Disruption of Her2-Induced PD-L1 Inhibits Tumor Cell Immune Evasion in Patient-Derived Gastric Cancer Organoids. Cancers 2021, 13, 6158. [Google Scholar] [CrossRef]
  78. Mimura, K.; Kua, L.F.; Xiao, J.F.; Asuncion, B.R.; Nakayama, Y.; Syn, N.; Fazreen, Z.; Soong, R.; Kono, K.; Yong, W.P. Combined inhibition of PD-1/PD-L1, Lag-3, and Tim-3 axes augments antitumor immunity in gastric cancer-T cell coculture models. Gastric. Cancer 2021, 24, 611–623. [Google Scholar] [CrossRef]
  79. Chen, Y.; Mei, J.; Zhang, P.; Liu, J.; Chen, L.; Wu, L.; Zhang, Y. IQGAP1 is positively correlated with PD-L1 and regulates its expression via mediating STAT proteins phosphorylation. Int. Immunopharmacol. 2022, 108, 108897. [Google Scholar] [CrossRef]
  80. Liu, Z.; Ning, F.; Cai, Y.; Sheng, H.; Zheng, R.; Yin, X.; Lu, Z.; Su, L.; Chen, X.; Zeng, C.; et al. The EGFR-P38 MAPK axis up-regulates PD-L1 through miR-675-5p and down-regulates HLA-ABC via hexokinase-2 in hepatocellular carcinoma cells. Cancer Commun. 2021, 41, 62–78. [Google Scholar] [CrossRef]
  81. Chen, M.L.; Fan, L.; Huang, G.R.; Sun, Z.F. lncRNA EGFR-AS1 facilitates leiomyosarcoma progression and immune escape via the EGFR-MYC-PD-L1 axis. Int. Immunol. 2022, 34, 365–377. [Google Scholar] [CrossRef] [PubMed]
  82. Lee, J.V.; Housley, F.; Yau, C.; Nakagawa, R.; Winkler, J.; Anttila, J.M.; Munne, P.M.; Savelius, M.; Houlahan, K.E.; Van de Mark, D.; et al. Combinatorial immunotherapies overcome MYC-driven immune evasion in triple negative breast cancer. Nat. Commun. 2022, 13, 3671. [Google Scholar] [CrossRef] [PubMed]
  83. Gong, J.; Shen, Y.; Jiang, F.; Wang, Y.; Chu, L.; Sun, J.; Shen, P.; Chen, M. MicroRNA-20a promotes non-small cell lung cancer proliferation by upregulating PD-L1 by targeting PTEN. Oncol. Lett. 2022, 23, 148. [Google Scholar] [CrossRef]
  84. Liu, N.; Jiang, X.; Guo, L.; Zhang, C.; Jiang, M.; Sun, Z.; Zhang, Y.; Mi, W.; Li, J.; Fu, Y.; et al. Mutant p53 achieved Gain-of-Function by promoting tumor growth and immune escape through PHLPP2/AKT/PD-L1 pathway. Int. J. Biol. Sci. 2022, 18, 2419–2438. [Google Scholar] [CrossRef]
  85. Tang, M.; Zheng, Z.; Shang, J.; Zhang, S. Risk Analysis of Positive PD-L1 Expression and Clinicopathological Features and Survival Prognosis in Patients with Colorectal Cancer: Systematic Review and Meta-Analysis. J. Healthc. Eng. 2022, 2022, 8212486. [Google Scholar] [CrossRef] [PubMed]
  86. Yin, S.; Guo, Y.; Wen, X.; Zeng, H.; Chen, G. Increased expression of PD-L1 in endometrial cancer stem-like cells is regulated by hypoxia. Front. Biosci. 2022, 27, 23. [Google Scholar] [CrossRef]
  87. Jiang, Z.; Yang, Y.; Yang, Y.; Zhang, Y.; Yue, Z.; Pan, Z.; Ren, X. Ginsenoside Rg3 attenuates cisplatin resistance in lung cancer by downregulating PD-L1 and resuming immune. Biomed. Pharmacother. 2017, 96, 378–383. [Google Scholar] [CrossRef]
  88. Martinez, V.G.; O’Neill, S.; Salimu, J.; Breslin, S.; Clayton, A.; Crown, J.; O’Driscoll, L. Resistance to HER2-targeted anti-cancer drugs is associated with immune evasion in cancer cells and their derived extracellular vesicles. Oncoimmunology 2017, 6, e1362530. [Google Scholar] [CrossRef]
  89. Aboubakar Nana, F.; Ocak, S. Targeting BRAF Activation as Acquired Resistance Mechanism to EGFR Tyrosine Kinase Inhibitors in EGFR-Mutant Non-Small-Cell Lung Cancer. Pharmaceutics 2021, 13, 1478. [Google Scholar] [CrossRef]
  90. Kuroki, H.; Anraku, T.; Kazama, A.; Shirono, Y.; Bilim, V.; Tomita, Y. Histone deacetylase 6 inhibition in urothelial cancer as a potential new strategy for cancer treatment. Oncol. Lett. 2021, 21, 64. [Google Scholar] [CrossRef]
  91. Keremu, A.; Aimaiti, A.; Liang, Z.; Zou, X. Role of the HDAC6/STAT3 pathway in regulating PD-L1 expression in osteosarcoma cell lines. Cancer Chemother. Pharmacol. 2019, 83, 255–264. [Google Scholar] [CrossRef] [PubMed]
  92. Peng, X.; Li, L.; Chen, J.; Ren, Y.; Liu, J.; Yu, Z.; Cao, H.; Chen, J. Discovery of Novel Histone Deacetylase 6 (HDAC6) Inhibitors with Enhanced Antitumor Immunity of Anti-PD-L1 Immunotherapy in Melanoma. J. Med. Chem. 2022, 65, 2434–2457. [Google Scholar] [CrossRef] [PubMed]
  93. Yussuf Khamis, M.; Wu, H.P.; Ma, Q.; Li, Y.H.; Ma, L.Y.; Zhang, X.H.; Liu, H.M. Overcome the tumor immunotherapy resistance by combination of the HDAC6 inhibitors with antitumor immunomodulatory agents. Bioorganic Chem. 2021, 109, 104754. [Google Scholar] [CrossRef]
  94. Zundell, J.A.; Fukumoto, T.; Lin, J.; Fatkhudinov, N.; Nacarelli, T.; Kossenkov, A.V.; Liu, Q.; Cassel, J.; Hu, C.A.; Wu, S.; et al. Targeting the IRE1alpha/XBP1 Endoplasmic Reticulum Stress Response Pathway in ARID1A-Mutant Ovarian Cancers. Cancer Res. 2021, 81, 5325–5335. [Google Scholar] [CrossRef] [PubMed]
  95. Megino-Luque, C.; Siso, P.; Mota-Martorell, N.; Navaridas, R.; de la Rosa, I.; Urdanibia, I.; Alberti-Valls, M.; Santacana, M.; Pinyol, M.; Bonifaci, N.; et al. ARID1A-deficient cells require HDAC6 for progression of endometrial carcinoma. Mol. Oncol. 2022, 16, 2235–2259. [Google Scholar] [CrossRef]
  96. Fukumoto, T.; Fatkhutdinov, N.; Zundell, J.A.; Tcyganov, E.N.; Nacarelli, T.; Karakashev, S.; Wu, S.; Liu, Q.; Gabrilovich, D.I.; Zhang, R. HDAC6 Inhibition Synergizes with Anti-PD-L1 Therapy in ARID1A-Inactivated Ovarian Cancer. Cancer Res. 2019, 79, 5482–5489. [Google Scholar] [CrossRef]
  97. Maharaj, K.; Powers, J.J.; Mediavilla-Varela, M.; Achille, A.; Gamal, W.; Quayle, S.; Jones, S.S.; Sahakian, E.; Pinilla-Ibarz, J. HDAC6 Inhibition Alleviates CLL-Induced T-Cell Dysfunction and Enhances Immune Checkpoint Blockade Efficacy in the Emu-TCL1 Model. Front. Immunol. 2020, 11, 590072. [Google Scholar] [CrossRef]
  98. Knox, T.; Sahakian, E.; Banik, D.; Hadley, M.; Palmer, E.; Noonepalle, S.; Kim, J.; Powers, J.; Gracia-Hernandez, M.; Oliveira, V.; et al. Author Correction: Selective HDAC6 inhibitors improve anti-PD-1 immune checkpoint blockade therapy by decreasing the anti-inflammatory phenotype of macrophages and down-regulation of immunosuppressive proteins in tumor cells. Sci. Rep. 2019, 9, 14824. [Google Scholar] [CrossRef]
  99. Won, H.R.; Lee, D.H.; Yeon, S.K.; Ryu, H.W.; Kim, G.W.; Kwon, S.H. HDAC6 selective inhibitor synergistically enhances the anticancer activity of immunomodulatory rugs in multiple myeloma. Int. J. Oncol. 2019, 55, 499–512. [Google Scholar]
  100. Son, S.M.; Park, S.J.; Lee, H.; Siddiqi, F.; Lee, J.E.; Menzies, F.M.; Rubinsztein, D.C. Leucine Signals to mTORC1 via Its Metabolite Acetyl-Coenzyme A. Cell. Metab. 2019, 29, 192–201.e197. [Google Scholar] [CrossRef]
  101. Son, S.M.; Park, S.J.; Stamatakou, E.; Vicinanza, M.; Menzies, F.M.; Rubinsztein, D.C. Leucine regulates autophagy via acetylation of the mTORC1 component raptor. Nat. Commun. 2020, 11, 3148. [Google Scholar] [CrossRef] [PubMed]
  102. Cao, P.; Li, Y.; Shi, R.; Yuan, Y.; Gong, H.; Zhu, G.; Zhang, Z.; Chen, C.; Zhang, H.; Liu, M.; et al. Combining EGFR-TKI With SAHA Overcomes EGFR-TKI-Acquired Resistance by Reducing the Protective Autophagy in Non-Small Cell Lung Cancer. Front. Chem. 2022, 10, 837987. [Google Scholar] [CrossRef]
  103. Passaro, E.; Papulino, C.; Chianese, U.; Toraldo, A.; Congi, R.; Del Gaudio, N.; Nicoletti, M.M.; Benedetti, R.; Altucci, L. HDAC6 Inhibition Extinguishes Autophagy in Cancer: Recent Insights. Cancers 2021, 13, 6280. [Google Scholar] [CrossRef] [PubMed]
  104. Viry, E.; Paggetti, J.; Baginska, J.; Mgrditchian, T.; Berchem, G.; Moussay, E.; Janji, B. Autophagy: An adaptive metabolic response to stress shaping the antitumor immunity. Biochem. Pharmacol. 2014, 92, 31–42. [Google Scholar] [CrossRef] [PubMed]
  105. Chan, A.M.; Fletcher, S. Shifting the paradigm in treating multi-factorial diseases: Polypharmacological co-inhibitors of HDAC6. RSC Med. Chem. 2021, 12, 178–196. [Google Scholar] [CrossRef] [PubMed]
  106. Park, S.Y.; Kim, J.S. A short guide to histone deacetylases including recent progress on class II enzymes. Exp. Mol. Med. 2020, 52, 204–212. [Google Scholar] [CrossRef]
  107. Zhang, Z.; Zhang, X.; Huang, A. Aggresome-Autophagy Associated Gene HDAC6 Is a Potential Biomarker in Pan-Cancer, Especially in Colon Adenocarcinoma. Front. Oncol. 2021, 11, 718589. [Google Scholar] [CrossRef]
  108. Chang, P.; Li, H.; Hu, H.; Li, Y.; Wang, T. The Role of HDAC6 in Autophagy and NLRP3 Inflammasome. Front. Immunol. 2021, 12, 763831. [Google Scholar] [CrossRef]
  109. Choi, W.H.; Yun, Y.; Park, S.; Jeon, J.H.; Lee, J.; Lee, J.H.; Yang, S.A.; Kim, N.K.; Jung, C.H.; Kwon, Y.T.; et al. Aggresomal sequestration and STUB1-mediated ubiquitylation during mammalian proteaphagy of inhibited proteasomes. Proc. Natl. Acad. Sci. USA 2020, 117, 19190–19200. [Google Scholar] [CrossRef]
  110. Wang, R.; Tan, J.; Chen, T.; Han, H.; Tian, R.; Tan, Y.; Wu, Y.; Cui, J.; Chen, F.; Li, J.; et al. ATP13A2 facilitates HDAC6 recruitment to lysosome to promote autophagosome-lysosome fusion. J. Cell Biol. 2019, 218, 267–284. [Google Scholar] [CrossRef]
  111. Magupalli, V.G.; Negro, R.; Tian, Y.; Hauenstein, A.V.; Di Caprio, G.; Skillern, W.; Deng, Q.; Orning, P.; Alam, H.B.; Maliga, Z.; et al. HDAC6 mediates an aggresome-like mechanism for NLRP3 and pyrin inflammasome activation. Science 2020, 369, eaas8995. [Google Scholar] [CrossRef] [PubMed]
  112. Mishima, Y.; Santo, L.; Eda, H.; Cirstea, D.; Nemani, N.; Yee, A.J.; O’Donnell, E.; Selig, M.K.; Quayle, S.N.; Arastu-Kapur, S.; et al. Ricolinostat (ACY-1215) induced inhibition of aggresome formation accelerates carfilzomib-induced multiple myeloma cell death. Br. J. Haematol. 2015, 169, 423–434. [Google Scholar] [CrossRef] [PubMed]
  113. Kekatpure, V.D.; Dannenberg, A.J.; Subbaramaiah, K. Withdrawal: HDAC6 modulates Hsp90 chaperone activity and regulates activation of aryl hydrocarbon receptor signaling. J. Biol. Chem. 2020, 295, 297. [Google Scholar] [CrossRef] [Green Version]
  114. Lin, T.W.; Chen, M.T.; Lin, L.T.; Huang, P.I.; Lo, W.L.; Yang, Y.P.; Lu, K.H.; Chen, Y.W.; Chiou, S.H.; Wu, C.W. TDP-43/HDAC6 axis promoted tumor progression and regulated nutrient deprivation-induced autophagy in glioblastoma. Oncotarget 2017, 8, 56612. [Google Scholar] [CrossRef] [PubMed]
  115. Jiang, X.; Huang, Y.; Liang, X.; Jiang, F.; He, Y.; Li, T.; Xu, G.; Zhao, H.; Yang, W.; Jiang, G.; et al. Metastatic prostate cancer-associated P62 inhibits autophagy flux and promotes epithelial to mesenchymal transition by sustaining the level of HDAC6. Prostate 2018, 78, 426–434. [Google Scholar] [CrossRef] [PubMed]
  116. Liu, J.R.; Yu, C.W.; Hung, P.Y.; Hsin, L.W.; Chern, J.W. High-selective HDAC6 inhibitor promotes HDAC6 degradation following autophagy modulation and enhanced antitumor immunity in glioblastoma. Biochem. Pharmacol. 2019, 163, 458–471. [Google Scholar] [CrossRef] [PubMed]
  117. Kaliszczak, M.; van Hechanova, E.; Li, Y.; Alsadah, H.; Parzych, K.; Auner, H.W.; Aboagye, E.O. The HDAC6 inhibitor C1A modulates autophagy substrates in diverse cancer cells and induces cell death. Br. J. Cancer 2018, 119, 1278–1287. [Google Scholar] [CrossRef]
  118. Jena, B.C.; Das, C.K.; Banerjee, I.; Bharadwaj, D.; Majumder, R.; Das, S.; Biswas, A.; Kundu, M.; Roy, P.K.; Kundu, C.N.; et al. TGF-beta1 induced autophagy in cancer associated fibroblasts during hypoxia contributes EMT and glycolysis via MCT4 upregulation. Exp. Cell Res. 2022, 417, 113195. [Google Scholar] [CrossRef]
  119. Wang, Y.; Ha, M.; Li, M.; Zhang, L.; Chen, Y. Histone deacetylase 6-mediated downregulation of TMEM100 expedites the development and progression of non-small cell lung cancer. Hum. Cell. 2022, 35, 271–285. [Google Scholar] [CrossRef]
  120. Hu, Y.; Dai, M.; Zheng, Y.; Wu, J.; Yu, B.; Zhang, H.; Kong, W.; Wu, H.; Yu, X. Epigenetic suppression of E-cadherin expression by Snail2 during the metastasis of colorectal cancer. Clin. Epigenetics 2018, 10, 154. [Google Scholar] [CrossRef]
  121. Fedorova, O.; Daks, A.; Parfenyev, S.; Shuvalov, O.; Netsvetay, S.; Vasileva, J.; Gudovich, A.; Golotin, V.; Semenov, O.; Petukhov, A.; et al. Zeb1-mediated autophagy enhances resistance of breast cancer cells to genotoxic drugs. Biochem. Biophys. Res. Commun. 2022, 589, 29–34. [Google Scholar] [CrossRef] [PubMed]
  122. Pant, K.; Richard, S.; Gradilone, S.A. Short-Chain Fatty Acid Butyrate Induces Cilia Formation and Potentiates the Effects of HDAC6 Inhibitors in Cholangiocarcinoma Cells. Front. Cell Dev. Biol. 2021, 9, 809382. [Google Scholar] [CrossRef] [PubMed]
  123. Zhang, L.H.; Yang, A.J.; Wang, M.; Liu, W.; Wang, C.Y.; Xie, X.F.; Chen, X.; Dong, J.F.; Li, M. Enhanced autophagy reveals vulnerability of P-gp mediated epirubicin resistance in triple negative breast cancer cells. Apoptosis 2016, 21, 473–488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Banerjee, A.; Birts, C.N.; Darley, M.; Parker, R.; Mirnezami, A.H.; West, J.; Cutress, R.I.; Beers, S.A.; Rose-Zerilli, M.J.J.; Blaydes, J.P. Stem cell-like breast cancer cells with acquired resistance to metformin are sensitive to inhibitors of NADH-dependent CtBP dimerization. Carcinogenesis 2019, 40, 871–882. [Google Scholar] [CrossRef]
  125. Sun, S.; Xue, D.; Chen, Z.; Ou-Yang, Y.; Zhang, J.; Mai, J.; Gu, J.; Lu, W.; Liu, X.; Liu, W.; et al. R406 elicits anti-Warburg effect via Syk-dependent and -independent mechanisms to trigger apoptosis in glioma stem cells. Cell Death Dis. 2019, 10, 358. [Google Scholar] [CrossRef]
  126. Lock, R.; Roy, S.; Kenific, C.M.; Su, J.S.; Salas, E.; Ronen, S.M.; Debnath, J. Autophagy facilitates glycolysis during Ras-mediated oncogenic transformation. Mol. Biol. Cell 2011, 22, 165–178. [Google Scholar] [CrossRef]
  127. Yan, X.; Tian, R.; Sun, J.; Zhao, Y.; Liu, B.; Su, J.; Li, M.; Sun, W.; Xu, X. Sorafenib-Induced Autophagy Promotes Glycolysis by Upregulating the p62/HDAC6/HSP90 Axis in Hepatocellular Carcinoma Cells. Front. Pharmacol. 2021, 12, 788667. [Google Scholar] [CrossRef]
  128. Dowling, C.M.; Hollinshead, K.E.R.; Di Grande, A.; Pritchard, J.; Zhang, H.; Dillon, E.T.; Haley, K.; Papadopoulos, E.; Mehta, A.K.; Bleach, R.; et al. Multiple screening approaches reveal HDAC6 as a novel regulator of glycolytic metabolism in triple-negative breast cancer. Sci. Adv. 2021, 7, eabc4897. [Google Scholar] [CrossRef]
  129. Lypova, N.; Dougherty, S.M.; Lanceta, L.; Chesney, J.; Imbert-Fernandez, Y. PFKFB3 Inhibition Impairs Erlotinib-Induced Autophagy in NSCLCs. Cells 2021, 10, 1679. [Google Scholar] [CrossRef]
  130. Wang, H.P.; Chen, W.J.; Shen, J.M.; Ye, T.; Xie, H.W. Attenuating glucose metabolism by Fbxw7 promotes Taxol sensitivity of colon cancer cells through downregulating NADPH oxidase 1 (Nox1). Ann. Transl. Med. 2021, 9, 886. [Google Scholar] [CrossRef]
  131. Dikic, I.; Elazar, Z. Mechanism and medical implications of mammalian autophagy. Nat. Rev. Mol. Cell Biol. 2018, 19, 349–364. [Google Scholar] [CrossRef] [PubMed]
  132. Shen, S.; Kepp, O.; Michaud, M.; Martins, I.; Minoux, H.; Metivier, D.; Maiuri, M.C.; Kroemer, R.T.; Kroemer, G. Association and dissociation of autophagy, apoptosis and necrosis by systematic chemical study. Oncogene 2011, 30, 4544–4556. [Google Scholar] [CrossRef] [PubMed]
  133. Sui, X.; Chen, R.; Wang, Z.; Huang, Z.; Kong, N.; Zhang, M.; Han, W.; Lou, F.; Yang, J.; Zhang, Q.; et al. Autophagy and chemotherapy resistance: A promising therapeutic target for cancer treatment. Cell Death Dis. 2013, 4, e838. [Google Scholar] [CrossRef] [PubMed]
  134. Thorburn, A.; Thamm, D.H.; Gustafson, D.L. Autophagy and cancer therapy. Mol. Pharmacol. 2014, 85, 830–838. [Google Scholar] [CrossRef] [PubMed]
  135. Guo, J.Y.; Chen, H.Y.; Mathew, R.; Fan, J.; Strohecker, A.M.; Karsli-Uzunbas, G.; Kamphorst, J.J.; Chen, G.; Lemons, J.M.; Karantza, V.; et al. Activated Ras requires autophagy to maintain oxidative metabolism and tumorigenesis. Genes Dev. 2011, 25, 460–470. [Google Scholar] [CrossRef]
  136. Kim, M.J.; Woo, S.J.; Yoon, C.H.; Lee, J.S.; An, S.; Choi, Y.H.; Hwang, S.G.; Yoon, G.; Lee, S.J. Involvement of autophagy in oncogenic K-Ras-induced malignant cell transformation. J. Biol. Chem. 2011, 286, 12924–12932. [Google Scholar] [CrossRef]
  137. Ji, X.; Ma, H.; Du, Y. Role and mechanism of action of LAPTM4B in EGFR-mediated autophagy. Oncol. Lett. 2022, 23, 109. [Google Scholar] [CrossRef]
  138. Abdul Rahim, S.A.; Dirkse, A.; Oudin, A.; Schuster, A.; Bohler, J.; Barthelemy, V.; Muller, A.; Vallar, L.; Janji, B.; Golebiewska, A.; et al. Regulation of hypoxia-induced autophagy in glioblastoma involves ATG9A. Br. J. Cancer 2017, 117, 813–825. [Google Scholar] [CrossRef]
  139. Tang, J.Y.; Hsi, E.; Huang, Y.C.; Hsu, N.C.; Chen, Y.K.; Chu, P.Y.; Chai, C.Y. ATG9A overexpression is associated with disease recurrence and poor survival in patients with oral squamous cell carcinoma. Virchows Arch. 2013, 463, 737–742. [Google Scholar] [CrossRef]
  140. Lei, Y.; Kansy, B.A.; Li, J.; Cong, L.; Liu, Y.; Trivedi, S.; Wen, H.; Ting, J.P.; Ouyang, H.; Ferris, R.L. EGFR-targeted mAb therapy modulates autophagy in head and neck squamous cell carcinoma through NLRX1-TUFM protein complex. Oncogene 2016, 35, 4698–4707. [Google Scholar] [CrossRef]
  141. Belounis, A.; Nyalendo, C.; Le Gall, R.; Imbriglio, T.V.; Mahma, M.; Teira, P.; Beaunoyer, M.; Cournoyer, S.; Haddad, E.; Vassal, G.; et al. Autophagy is associated with chemoresistance in neuroblastoma. BMC Cancer 2016, 16, 891. [Google Scholar] [CrossRef] [PubMed]
  142. Hu, Y.L.; Jahangiri, A.; Delay, M.; Aghi, M.K. Tumor cell autophagy as an adaptive response mediating resistance to treatments such as antiangiogenic therapy. Cancer Res. 2012, 72, 4294–4299. [Google Scholar] [CrossRef] [PubMed]
  143. Park, S.Y.; Kim, W.J.; Byun, J.H.; Lee, J.J.; Jeoung, D.; Park, S.T.; Kim, Y. Role of DDX53 in taxol-resistance of cervix cancer cells in vitro. Biochem. Biophys. Res. Commun. 2018, 506, 641–647. [Google Scholar] [CrossRef] [PubMed]
  144. Zou, Z.; Yuan, Z.; Zhang, Q.; Long, Z.; Chen, J.; Tang, Z.; Zhu, Y.; Chen, S.; Xu, J.; Yan, M.; et al. Aurora kinase A inhibition-induced autophagy triggers drug resistance in breast cancer cells. Autophagy 2012, 8, 1798–1810. [Google Scholar] [CrossRef]
  145. Mascia, F.; Mazo, I.; Alterovitz, W.L.; Karagiannis, K.; Wu, W.W.; Shen, R.F.; Beaver, J.A.; Rao, V.A. In search of autophagy biomarkers in breast cancer: Receptor status and drug agnostic transcriptional changes during autophagy flux in cell lines. PLoS ONE 2022, 17, e0262134. [Google Scholar] [CrossRef]
  146. Lee, H.K.; Noh, M.H.; Hong, S.W.; Kim, S.M.; Kim, S.H.; Kim, Y.S.; Broaddus, V.C.; Hur, D.Y. Erlotinib Activates Different Cell Death Pathways in EGFR-mutant Lung Cancer Cells Grown in 3D Versus 2D Culture Systems. Anticancer Res. 2021, 41, 1261–1269. [Google Scholar] [CrossRef]
  147. Yeon, M.; Byun, J.; Kim, H.; Kim, M.; Jung, H.S.; Jeon, D.; Kim, Y.; Jeoung, D. CAGE Binds to Beclin1, Regulates Autophagic Flux and CAGE-Derived Peptide Confers Sensitivity to Anti-cancer Drugs in Non-small Cell Lung Cancer Cells. Front. Oncol. 2018, 8, 599. [Google Scholar] [CrossRef]
  148. Yang, Z.; Liu, Y.; Wei, X.; Zhou, X.; Gong, C.; Zhang, T.; Jin, P.; Xu, S.; Ma, D.; Gao, Q. Co-targeting EGFR and Autophagy Impairs Ovarian Cancer Cell Survival during Detachment from the ECM. Curr. Cancer Drug Targets 2015, 15, 215–226. [Google Scholar] [CrossRef]
  149. Sobhakumari, A.; Schickling, B.M.; Love-Homan, L.; Raeburn, A.; Fletcher, E.V.; Case, A.J.; Domann, F.E.; Miller, F.J., Jr.; Simons, A.L. NOX4 mediates cytoprotective autophagy induced by the EGFR inhibitor erlotinib in head and neck cancer cells. Toxicol. Appl. Pharmacol. 2013, 272, 736–745. [Google Scholar] [CrossRef]
  150. Nihira, K.; Miki, Y.; Iida, S.; Narumi, S.; Ono, K.; Iwabuchi, E.; Ise, K.; Mori, K.; Saito, M.; Ebina, M.; et al. An activation of LC3A-mediated autophagy contributes to de novo and acquired resistance to EGFR tyrosine kinase inhibitors in lung adenocarcinoma. J. Pathol. 2014, 234, 277–288. [Google Scholar]
  151. Sang, J.; Gan, L.; Zou, M.F.; Lin, Z.J.; Fan, R.Z.; Huang, J.L.; Li, W.; Tang, G.H.; Yin, S. Jolkinolide B sensitizes bladder cancer to mTOR inhibitors via dual inhibition of Akt signaling and autophagy. Cancer Lett. 2022, 526, 352–362. [Google Scholar] [CrossRef] [PubMed]
  152. Fan, G.; Qin, J.; Fu, X.; Si, X.; Li, L.; Yang, K.; Wang, B.; Lou, H.; Zhu, J. Low-Intensity Focused Ultrasound Targeted Microbubble Destruction Enhanced Paclitaxel Sensitivity by Decreasing Autophagy in Paclitaxel-Resistant Ovarian Cancer. Front. Oncol. 2022, 12, 823956. [Google Scholar] [CrossRef] [PubMed]
  153. Wang, M.; Liao, C.; Hu, Y.; Qinwen, P.; Jiang, J. Sensitization of breast cancer cells to paclitaxel by dichloroacetate through inhibiting autophagy. Biochem. Biophys. Res. Commun. 2017, 489, 103–108. [Google Scholar] [CrossRef] [PubMed]
  154. Xu, L.; Fan, Q.; Wang, X.; Zhao, X.; Wang, L. Inhibition of autophagy increased AGE/ROS-mediated apoptosis in mesangial cells. Cell. Death Dis 2016, 7, e2445. [Google Scholar] [CrossRef] [PubMed]
  155. Song, B.; Bian, Q.; Shao, C.H.; Li, G.; Liu, A.A.; Jing, W.; Liu, R.; Zhang, Y.J.; Zhou, Y.Q.; Hu, X.G.; et al. Ulinastatin reduces the resistance of liver cancer cells to epirubicin by inhibiting autophagy. PLoS ONE 2015, 10, e0120694. [Google Scholar]
  156. Dong, Y.; Zhu, G.; Wang, S.F.; Keon, K.A.; Rubinstein, J.L.; Zeng, S.X.; Zhang, S.; Chen, Q.L.; Fu, J.; Li, M.; et al. Toosendanin, a novel potent vacuolar-type H(+)-translocating ATPase inhibitor, sensitizes cancer cells to chemotherapy by blocking protective autophagy. Int. J. Biol. Sci. 2022, 18, 2684–2702. [Google Scholar] [CrossRef]
  157. Zhang, Y.; Kwon, S.; Yamaguchi, T.; Cubizolles, F.; Rousseaux, S.; Kneissel, M.; Cao, C.; Li, N.; Cheng, H.L.; Chua, K.; et al. Mice lacking histone deacetylase 6 have hyperacetylated tubulin but are viable and develop normally. Mol. Cell Biol. 2008, 28, 1688–1701. [Google Scholar] [CrossRef]
  158. Slaby, O.; Laga, R.; Sedlacek, O. Therapeutic targeting of non-coding RNAs in cancer. Biochem. J. 2017, 474, 4219–4251. [Google Scholar] [CrossRef]
  159. Sun, D.; Wu, Y.; Zhang, S.; Han, Y.; Shen, J.; Zheng, W.; Wei, L.; Liu, Y.; Ren, L.; Gu, Z.; et al. Distinct roles of miR-34 family members on suppression of lung squamous cell carcinoma. Biomed. Pharmacother. 2021, 142, 111967. [Google Scholar] [CrossRef]
  160. Xu, W.; Liu, W.; Anwaier, A.; Tian, X.; Su, J.; Shi, G.; Wei, S.; Qu, Y.; Zhang, H.; Ye, D. Deciphering the role of miR-187-3p/LRFN1 axis in modulating progression, aerobic glycolysis and immune microenvironment of clear cell renal cell carcinoma. Discov. Oncol. 2022, 13, 59. [Google Scholar] [CrossRef]
  161. Awad, M.M.; Le Bruchec, Y.; Lu, B.; Ye, J.; Miller, J.; Lizotte, P.H.; Cavanaugh, M.E.; Rode, A.J.; Dumitru, C.D.; Spira, A. Selective Histone Deacetylase Inhibitor ACY-241 (Citarinostat) Plus Nivolumab in Advanced Non-Small Cell Lung Cancer: Results from a Phase Ib Study. Front. Oncol. 2021, 11, 696512. [Google Scholar] [CrossRef] [PubMed]
  162. Vogl, D.T.; Raje, N.; Jagannath, S.; Richardson, P.; Hari, P.; Orlowski, R.; Supko, J.G.; Tamang, D.; Yang, M.; Jones, S.S.; et al. Ricolinostat, the First Selective Histone Deacetylase 6 Inhibitor, in Combination with Bortezomib and Dexamethasone for Relapsed or Refractory Multiple Myeloma. Clin. Cancer Res. 2017, 23, 3307–3315. [Google Scholar] [CrossRef] [PubMed]
  163. Gordon, M.S.; Shapiro, G.I.; Sarantopoulos, J.; Juric, D.; Lu, B.; Zarotiadou, A.; Connarn, J.N.; Le Bruchec, Y.; Dumitru, C.D.; Harvey, R.D. Phase Ib Study of the Histone Deacetylase 6 Inhibitor Citarinostat in Combination With Paclitaxel in Patients with Advanced Solid Tumors. Front. Oncol. 2021, 11, 786120. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Functional domains of HDAC6. NLS, nuclear localization signal sequences; NES, nuclear export signal sequences; DD1/DD2, catalytic domains; DMB, dynein motor binding; SE14, Ser-Glu tetra-decapeptide repeat; ZnF, zinc finger domain; and UBP, ubiquitin binding domain.
Figure 1. Functional domains of HDAC6. NLS, nuclear localization signal sequences; NES, nuclear export signal sequences; DD1/DD2, catalytic domains; DMB, dynein motor binding; SE14, Ser-Glu tetra-decapeptide repeat; ZnF, zinc finger domain; and UBP, ubiquitin binding domain.
Ijms 23 09592 g001
Figure 2. Mechanism of HDAC6-promoted cancer cell proliferation. (A) Genes and signaling pathways regulated by HDAC6. (B) HDAC6 decreases the expression of p21 to promote cancer cell proliferation. USP10 increases HDAC6 stability, which in turn promotes cancer cell proliferation. HDAC6 enhances β-catenin signaling to promote cancer cell proliferation. HDAC6 binds to PTPN1, which activates ERK signaling and promotes cancer cell proliferation. ↓ denotes negative regulation; ↑ denotes positive regulation; T bar arrows denote negative regulation; and hollow arrows denote direction of reaction.
Figure 2. Mechanism of HDAC6-promoted cancer cell proliferation. (A) Genes and signaling pathways regulated by HDAC6. (B) HDAC6 decreases the expression of p21 to promote cancer cell proliferation. USP10 increases HDAC6 stability, which in turn promotes cancer cell proliferation. HDAC6 enhances β-catenin signaling to promote cancer cell proliferation. HDAC6 binds to PTPN1, which activates ERK signaling and promotes cancer cell proliferation. ↓ denotes negative regulation; ↑ denotes positive regulation; T bar arrows denote negative regulation; and hollow arrows denote direction of reaction.
Ijms 23 09592 g002
Figure 3. Expression regulation of HDAC6 and roles of HDAC6-targeting miRNAs in cancer cell proliferation. (A) Shows potential binding sites for various transcription factors in promoter sequences of HDAC6. (B) miRNAs that potentially target HDAC6. The potential binding of the miRNAs to the 3′ UTR of HDAC6 is shown. Seed sequences of miRNAs are shown. UTR denotes untranslated region.
Figure 3. Expression regulation of HDAC6 and roles of HDAC6-targeting miRNAs in cancer cell proliferation. (A) Shows potential binding sites for various transcription factors in promoter sequences of HDAC6. (B) miRNAs that potentially target HDAC6. The potential binding of the miRNAs to the 3′ UTR of HDAC6 is shown. Seed sequences of miRNAs are shown. UTR denotes untranslated region.
Ijms 23 09592 g003
Figure 4. Mechanism of HDAC6-promoted anti-cancer drug resistance. (A) HDAC6 prevents EGFR degradation, activates EGFR signaling, and confers resistance to anti-cancer drugs. HDAC6 inhibition decreases the expression of EGFR, which enhances sensitivity of cancer cells to anti-cancer drugs. (B) HDAC6 enhances resistance to anti-cancer drugs by deacetylation of hsp90. HDAC6 binding to β-tubulin is necessary for conferring resistance to anti-cancer drugs. Downregulation of HDAC6 decreases the expression of MDR1. ↓ denotes negative regulation; ↑ denotes positive regulation; T bar arrows denote negative regulation; and hollow arrows denote direction of reaction.
Figure 4. Mechanism of HDAC6-promoted anti-cancer drug resistance. (A) HDAC6 prevents EGFR degradation, activates EGFR signaling, and confers resistance to anti-cancer drugs. HDAC6 inhibition decreases the expression of EGFR, which enhances sensitivity of cancer cells to anti-cancer drugs. (B) HDAC6 enhances resistance to anti-cancer drugs by deacetylation of hsp90. HDAC6 binding to β-tubulin is necessary for conferring resistance to anti-cancer drugs. Downregulation of HDAC6 decreases the expression of MDR1. ↓ denotes negative regulation; ↑ denotes positive regulation; T bar arrows denote negative regulation; and hollow arrows denote direction of reaction.
Ijms 23 09592 g004
Figure 5. Expression regulation of PD-L1 and effects of HDAC6 inhibitors on sensitivity of cancer cells to immune checkpoint inhibitors. (A) EGFR signaling, MYC, and HDAC6 increase the expression of PD-L1. PTEN and P53 negatively regulate PD-L1 expression. (B) Downregulation of HDAC6 decreases the expression of PD-L1. (C) HDAC6 inhibitors enhance sensitivity of cancer cells to immune checkpoint inhibitors by increasing activated CD8+ T cells while decreasing tumor promoting M2 macrophages. ↓ denotes negative regulation; ↑denotes positive regulation; → denotes increases in transcription; T bar arrows denote negative regulation; and hollow arrows denote direction of reaction.
Figure 5. Expression regulation of PD-L1 and effects of HDAC6 inhibitors on sensitivity of cancer cells to immune checkpoint inhibitors. (A) EGFR signaling, MYC, and HDAC6 increase the expression of PD-L1. PTEN and P53 negatively regulate PD-L1 expression. (B) Downregulation of HDAC6 decreases the expression of PD-L1. (C) HDAC6 inhibitors enhance sensitivity of cancer cells to immune checkpoint inhibitors by increasing activated CD8+ T cells while decreasing tumor promoting M2 macrophages. ↓ denotes negative regulation; ↑denotes positive regulation; → denotes increases in transcription; T bar arrows denote negative regulation; and hollow arrows denote direction of reaction.
Ijms 23 09592 g005
Figure 6. Role of HDAC6 in autophagic clearance of protein aggregates. Unfolded and/or misfolded proteins are ubiquitinated and taken into proteasome for degradation. Inhibition of proteasome activity by bortezomib (BTZ) leads to formation of abnormal proteins. Accumulation of abnormal proteins leads to apoptotic cell death. Aggresome formation occurs when proteasome cannot further process unfolded and/or misfolded proteins. HDAC6 loads ubiquitinated protein aggregates (misfolded and/or unfolded) onto dynein motor protein for transport to aggresome for autophagic clearance. HDAC6 inhibition prevents fusion of autophagosome with lysosome, leading to accumulation of protein aggregates and resulting in apoptotic cell death. Autophagic clearance of protein aggregates leads to cancer cell survival and confers anti-cancer drug resistance. DD1 and DD2 are catalytic domains of HDAC6; Ub denotes ubiquitin; UBD denotes ubiquitin binding domain; DMB denotes dynein motor binding domain; MT denotes microtubule; F-actin denotes actin filaments; and ↓ denotes negative regulation.
Figure 6. Role of HDAC6 in autophagic clearance of protein aggregates. Unfolded and/or misfolded proteins are ubiquitinated and taken into proteasome for degradation. Inhibition of proteasome activity by bortezomib (BTZ) leads to formation of abnormal proteins. Accumulation of abnormal proteins leads to apoptotic cell death. Aggresome formation occurs when proteasome cannot further process unfolded and/or misfolded proteins. HDAC6 loads ubiquitinated protein aggregates (misfolded and/or unfolded) onto dynein motor protein for transport to aggresome for autophagic clearance. HDAC6 inhibition prevents fusion of autophagosome with lysosome, leading to accumulation of protein aggregates and resulting in apoptotic cell death. Autophagic clearance of protein aggregates leads to cancer cell survival and confers anti-cancer drug resistance. DD1 and DD2 are catalytic domains of HDAC6; Ub denotes ubiquitin; UBD denotes ubiquitin binding domain; DMB denotes dynein motor binding domain; MT denotes microtubule; F-actin denotes actin filaments; and ↓ denotes negative regulation.
Ijms 23 09592 g006
Table 1. Mechanism of HDAC6-promoted cancer cell proliferation.
Table 1. Mechanism of HDAC6-promoted cancer cell proliferation.
Targets/MechanismCancer TypesReferences
SMAD 2 phosphorylation ↓
P21 ↓
Glioblastoma[10]
Binds to PTPN1
ERK1/2 activity ↑
Apoptosis ↓
Melanoma[18]
HDAC6 ↓
MAPK/ERK ↓
Colon Cancer[19]
MKK7 activity ↑Glioblastoma[20]
Oncogene EWSR1-FLI1 ↑Ewing Sarcoma[21]
RAC1 (Rho GTPase) activity ↑Rhabdomyosarcoma[22]
RAD51 ↑ CHEK activity ↑Glioblastoma[23]
Tumor suppressor MST1 ↓Breast Cancer[24]
P53 ↓Hepatocellular Carcinoma[25]
E-cadherin ↓ STAT activity ↑Breast Cancer[26]
C-Myc ↑ β-catenin Signaling ↑Colon Cancer[27]
miR-199 ↓ (negative regulator of Wnt signaling)Cervical Carcinoma [28]
↓ denotes negative regulation; ↑ denotes positive regulation.
Table 2. Role of HDAC6-targeting microRNAs in cancer cell proliferation.
Table 2. Role of HDAC6-targeting microRNAs in cancer cell proliferation.
miRNAsMechanism Cancer TypesReferences
miR-601
  • HDAC6 ↓
  • Suppresses proliferation, invasion, and migration
Esophageal Squamous Cell Carcinoma[37]
miR-22
  • HDAC6 ↓
  • Inhibits NF-kB signaling
Glioma [39]
miR-27b
  • HDAC6 ↓
  • MET/PI3K/AKT signaling ↓
Diffuse Large B-cell Lymphoma[40]
miR-206
  • HDAC6 ↓
  • PTEN ↑
  • AKT ↓ mTOR ↓
Endometrial Carcinoma[43]
mIR-433
  • HDAC6 ↓
  • Exportin 5 ↓
Cholangiocarcinoma [46]
↓ denotes negative regulation; ↑ denotes positive regulation.
Table 3. HDAC6-selective inhibitors overcome resistance to anti-cancer drugs.
Table 3. HDAC6-selective inhibitors overcome resistance to anti-cancer drugs.
HDAC6-Selective
Inhibitor
Target/MechanismEnhances
Sensitivity to
Cancer TypesReferences
WT161HDAC6 ↓
PTEN ↑ Apoptosis ↑
5-FUOsteosarcoma[53]
WT-161HDAC6 ↓
EGFR/HER2/ERα ↓
BortezomibBreast Cancer[54]
ACY-1215 (Ricolinostat)HDAC6 ↓
Apoptosis ↑
Gemcitabine and OxaliplatinGallbladder Cancer[60]
ACY-1215HDAC6 ↓
Apoptosis ↑
GemcitabinePancreatic Ductal Adenocarcinoma[61]
ACY-1215HDAC6 ↓
Acetylation of α-tubulin ↑
EribulinBreast Cancer[62]
ACY-1215HDAC6 ↓
Chk ↓
Mitotic catastrophe
AdavosertibHead and Neck Squamous Cell Carcinoma[63]
A452HDAC6 ↓
ERK ↓
NF-κB ↓
Bortezomib (BTZ)Multiple Myeloma[64]
7bHDAC6 ↓
BCR-ABL ↓
Leukemic stem cells ↓
ImatinibChronic Myeloid Leukemia[65]
ACY-241HDAC6 ↓PaclitaxelAdvanced Solid Tumors [66]
↓ denotes negative regulation; ↑ denotes positive regulation.
Table 4. Effects of HDAC6-selective inhibitors on sensitivity of cancer cells to immune checkpoint inhibitors.
Table 4. Effects of HDAC6-selective inhibitors on sensitivity of cancer cells to immune checkpoint inhibitors.
HDAC6
Inhibitor
Enhances Sensitivity toMechanismCancer TypesReferences
NexturastatAnti-PD-1 antibodyCTL ↑
IL-1β/IL-6 ↓
PD-L1 ↓
Non-Small Cell Lung Cancer [76]
XP-5Small molecule PD-L1 inhibitor PD-L1 ↓
Tumor-Infiltrating Lymphocytes ↓
Melanoma[92]
ACY1215Anti-PD-L1 antibodyIFN-γ positive CTL ↑Ovarian Cancer[96]
ACY738Anti-PD1
Anti-PD-L1
CTL ↑
JAK/STAT ↓
IL-10 ↓
Acetylation of HSP90 ↑
Chronic Lymphocytic Leukemia[97]
NexturastatAnti-PD-1 blockadePro-Tumorigenic M2 Macrophages ↓
Central/Memory T cells ↑
Melanoma[98]
A542Immunomodulatory drugs (lenalidomide or pomalidomide)AKT/ERK signaling ↓ Multiple Myeloma[99]
↓ denotes negative regulation; ↑ denotes positive regulation.
Table 5. Clinical trials of HDAC6-specific inhibitors registered in https://clinicaltrials.gov (accessed on 14 July 2022).
Table 5. Clinical trials of HDAC6-specific inhibitors registered in https://clinicaltrials.gov (accessed on 14 July 2022).
TitleInhibitorsStudy DesignTypes of CancersPhaseStudy DatesNCT Number
Safety, Tolerability, and MTD of KA2507 (HDAC6 inhibitor) KA2507
  • Enrollment: Twenty participants with solid tumors
  • Administration: twice-daily oral dosing using a 3 + 3 dose-escalation design
  • Adverse events: well-tolerated
  • Outcome: stable disease
Solid Tumor, AdultPhase 1
  • Start:
7 August 2017
  • Completion:
10 June 2020
NCT03008018
HDAC6 Inhibitor ACY-241 in Combination with Ipilimumab and NivolumabACY-241, nivolumab, ipilimumab
  • Enrollment: One participant with unresectable melanoma
  • Administration: Phase 2 dose (RP2D) of ACY-241 (oral) in combination with ipilimumab (infusion) at 1 mg/kg and nivolumab (infusion) at 3 mg/kg every 3 weeks for 4 doses each during a 12-week induction period, then administered with nivolumab at a flat dose of 240 mg every 2 weeks in maintenance for up to 1 year
  • Adverse events: not provided
  • Outcome: not provided
Malignant MelanomaPhase 1
(Stage III/Stage IV melanoma)
  • Start:
30 September 2016
  • Completion:
7 April 2017
NCT02935790
HDAC6 Inhibitor KA2507 in Advanced Biliary Tract CancerKA2507
  • Enrollment: No participant with standard of care chemotherapy (ABC-11)
  • Adverse events: not provided
  • Outcome: not provided
Biliary Tract CancerPhase 2
  • Start:
5 March 2020
  • Completion:
October 2023
NCT04186156
ACY-241 in Combination with Nivolumab in Patients with Unresectable Non-Small Cell Lung CancerACY-241, nivolumab
  • Enrollment: Eighteen participants with unresectable non-small cell lung cancer
  • Administration: The orally administered ACY-241 dose was escalated (180, 360, or 480 mg once daily). Nivolumab was administered at 240 mg (day 15 of cycle 1, then every 2 weeks thereafter)
  • Adverse events: Dyspenia (n = 3; 18%), pneumonia (n = 3; 18%)
  • Outcome: ① At the 180-mg dose, 1 complete response and 2 partial responses (PRs) were observed ② At the 360-mg dose, 3 PRs were observed; 1 patient achieved stable disease (SD) and 1 experienced progressive disease (PD)
Non-Small Cell Lung CancerPhase 1
  • Start:
25 August 2016
  • Completion:
30 June 2022
NCT02635061
Orally Administered JBI-802, an LSD1/HDAC6 InhibitorJBI-802
  • Enrollment: One hundred twenty-six participants with advanced solid tumors
  • Administration: 10 mg JBI-802 once daily as the starting dose with 4 days on/3 days off cycle
  • Adverse events: unknown
  • Outcome: unknown
Locally Advanced Solid Tumor, Metastatic Solid TumorPhase 1
Phase 2
  • Start:
8 April 2022
  • Completion:
August 2025
NCT05268666
ACY-1215 Alone and in Combination With Bortezomib and DexamethasoneACY-1215
  • Enrollment: One hundred twenty participants with relapsed/refractory multiple myeloma
  • Administration: Liquid oral dose on Days 1–5 and 8–12 of 21-day treatment cycle
  • Adverse events: Combination therapy with bortezomib and dexamethasone was well-tolerated during dose escalation but led to dose-limiting diarrhea in an expansion cohort at a ricolinostat dose of 160 mg twice daily
  • Outcome: The overall response rate in combination with daily ricolinostat at ≥160 mg was 37%
  • The response rate to combination therapy among bortezomib-refractory patients was 14%
Multiple MyelomaPhase 1: To evaluate the side effects and determine the best dose of oral ACY-1215 as monotherapy, and in combination with bortezomib and dexamethasone in patients with relapsed or relapsed/refractory multiple myeloma.
Phase 2: To determine the objective response rate of oral ACY-1215 in combination with bortezomib and dexamethasone in patients with relapsed or relapsed/refractory multiple myeloma.
  • Start:
July 2011
  • Completion:
3 December 2016
NCT01323751
ACY-1215 for Relapsed/Refractory Lymphoid MalignanciesACY-1215
  • Enrollment: Twenty-four participants with relapsed or refractory lymphoid malignancies
  • Administration: Oral administration
  • Adverse events: Not provided
  • Outcome: Not provided
Lymphoma, Lymphoid MalignanciesPhase 1: All patients will take the oral ACY-1215, 160 mg for 28 consecutive days on a 28-day treatment cycle
Phase 2: All patients will take the oral ACY-1215, 160 mg for 28 consecutive days on a 28-day treatment cycle.
  • Start:
2 April 2014
  • Completion:
5 May 2019
NCT02091063
ACY-1215+Nab-paclitaxel in Metastatic Breast CancerACY-1215, Nab-paclitaxel
  • Enrollment: Seventeen participants with Nab-paclitaxel in unresectable or metastatic breast cancer
  • Administration:
Drug: ACY-1215
An orally active, selective HDAC6 inhibitor
-
Assigned dosing 80 mg, 120 mg, 180 mg, 240 mg PO, and once daily Days 1–21 in a 28-day cycle
-
Drug: Nab-paclitaxel
-
100 mg/m2 30 min IV infusion Days 1, 8, and 15 in a 28-day cycle
  • Adverse events: not provided
  • Outcome: not provided
Metastatic Breast Cancer, Breast CarcinomaPhase 1
  • Start:
1 March 2016
  • Completion:
30 September 2020
NCT02632071
Safety, Pharmacokinetics, and Preliminary Antitumor Activity of ACY 241 in Combination with Paclitaxel in Patients With Advanced Solid TumorsACY-241 Paclitaxel
  • Enrollment: Twenty participants; fifteen had received prior taxane therapy
  • Administration:
① ACY-241: oral administration once daily (QD) or, if supported by PK and safety data, twice daily on 21 consecutive days of a 28-day treatment cycle
② Paclitaxel: administered to patients at 80 mg/m2 IV over 1 h on Days 1, 8, and 15 of the 28-day treatment cycle
  • Adverse events: The combination of ACY-241 plus paclitaxel showed an acceptable safety profile, with no unexpected or dose-limiting toxicities, and potential evidence of antitumor activity in patients with heavily pretreated advanced solid tumors
  • Outcome: Three patients showed partial response. Thirteen patients showed stable disease
Advanced solid tumors Phase 1b
  • Start:
22 December 2015
  • Completion:
4 October 2019
NCT02551185
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jo, H.; Shim, K.; Jeoung, D. Targeting HDAC6 to Overcome Autophagy-Promoted Anti-Cancer Drug Resistance. Int. J. Mol. Sci. 2022, 23, 9592. https://doi.org/10.3390/ijms23179592

AMA Style

Jo H, Shim K, Jeoung D. Targeting HDAC6 to Overcome Autophagy-Promoted Anti-Cancer Drug Resistance. International Journal of Molecular Sciences. 2022; 23(17):9592. https://doi.org/10.3390/ijms23179592

Chicago/Turabian Style

Jo, Hyein, Kyeonghee Shim, and Dooil Jeoung. 2022. "Targeting HDAC6 to Overcome Autophagy-Promoted Anti-Cancer Drug Resistance" International Journal of Molecular Sciences 23, no. 17: 9592. https://doi.org/10.3390/ijms23179592

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

Article Metrics

Back to TopTop