Next Article in Journal
Unleashing Ferroptosis in Human Cancers: Targeting Ferroptosis Suppressor Protein 1 for Overcoming Therapy Resistance
Next Article in Special Issue
Evaluation of Glutathione in Spike Protein of SARS-CoV-2 Induced Immunothrombosis and Cytokine Dysregulation
Previous Article in Journal
Oxidative Stress in a Mother Consuming Alcohol during Pregnancy and in Her Newborn: A Case Report
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cytoprotective Role of Heme Oxygenase-1 in Cancer Chemoresistance: Focus on Antioxidant, Antiapoptotic, and Pro-Autophagy Properties

1
Department of Gynecology, Hospital of Obstetrics and Gynecology, Fudan University, Shanghai 200011, China
2
Shanghai Key Laboratory of Female Reproductive Endocrine Related Diseases, Shanghai 200011, China
*
Authors to whom correspondence should be addressed.
Antioxidants 2023, 12(6), 1217; https://doi.org/10.3390/antiox12061217
Submission received: 19 April 2023 / Revised: 23 May 2023 / Accepted: 25 May 2023 / Published: 5 June 2023
(This article belongs to the Special Issue Antioxidant Enzymes in Cancer Biology)

Abstract

:
Chemoresistance remains the foremost challenge in cancer therapy. Targeting reactive oxygen species (ROS) manipulation is a promising strategy in cancer treatment since tumor cells present high levels of intracellular ROS, which makes them more vulnerable to further ROS elevation than normal cells. Nevertheless, dynamic redox evolution and adaptation of tumor cells are capable of counteracting therapy-induced oxidative stress, which leads to chemoresistance. Hence, exploring the cytoprotective mechanisms of tumor cells is urgently needed to overcome chemoresistance. Heme oxygenase-1 (HO-1), a rate-limiting enzyme of heme degradation, acts as a crucial antioxidant defense and cytoprotective molecule in response to cellular stress. Recently, emerging evidence indicated that ROS detoxification and oxidative stress tolerance owing to the antioxidant function of HO-1 contribute to chemoresistance in various cancers. Enhanced HO-1 expression or enzymatic activity was revealed to promote apoptosis resistance and activate protective autophagy, which also involved in the development of chemoresistance. Moreover, inhibition of HO-1 in multiple cancers was identified to reversing chemoresistance or improving chemosensitivity. Here, we summarize the most recent advances regarding the antioxidant, antiapoptotic, and pro-autophagy properties of HO-1 in mediating chemoresistance, highlighting HO-1 as a novel target for overcoming chemoresistance and improving the prognosis of cancer patients.

1. Introduction

The development of resistance to chemotherapy is one of the main obstacles for tumor management. Although many in-depth studies have been conducted, the specific molecular mechanisms underlying this inevitable resistance remain to be fully elucidated. Currently, therapeutic strategies targeting reactive oxygen species (ROS) production to initiate cellular damage and programmed cell death in cancer treatment are an area of emerging interest [1,2,3]. ROS refers to highly reactive molecules produced by incomplete reduction of oxygen in normal metabolic processes [4], which are mainly categorized into two types, i.e., free radicals and nonradical ROS. Free radicals contain one or more unpaired electron(s) in their outer orbital, including superoxide radical anion (O2●−), hydroxyl radical (HO), alkoxyl radical (RO), and peroxyl radical (ROO), while nonradical ROS, such as hydrogen peroxide (H2O2), organic peroxides (ROOH), and singlet oxygen (1O2), do not possess unpaired electrons but are the downstream products of incomplete reduction of molecular oxygen and have oxidizing potential [5,6]. The generation and elimination of ROS are balanced under physiological conditions, and excessive ROS can be scavenged by a tightly regulated endogenous antioxidant defense system consisting of enzymatic antioxidants such as superoxide dismutase (SOD), catalase (CAT), glutathione peroxidase (GPx), peroxiredoxins (Prx), thioredoxins(Trx), and nonenzymatic antioxidants including glutathione, uric acid, coenzyme Q, and lipoic acid [7,8]. ROS play an essential role in maintaining cell homeostasis by regulating intracellular signaling and the immune response at low concentrations. However, high levels of ROS are deleterious to nucleic acids, proteins, and cell membranes, and once an imbalance tends toward ROS overproduction, oxidative stress occurs, leading to cytotoxicity, cell dysfunction, and cell death [9,10]. Cancer cells are hypersensitive to fluctuations in ROS levels, which makes inducing ROS production over cytotoxic threshold levels an effective anticancer strategy [11].
Beyond causing cellular damage directly, ROS have also emerged as important signaling molecules in cancer cell fate decisions. The impact of ROS on the regulation of programmed cell death, such as apoptosis and autophagy, has been well established recently [12,13,14]. Remarkably, autophagy and apoptosis have been shown to interact with each other [15], while ROS may act as a molecular switch between autophagy and apoptosis [16]. Apoptosis, also defined as type I cell death, can be triggered by ROS via both the death receptor or extrinsic pathway and the mitochondrial or intrinsic pathway [17]. Several standard chemotherapeutic agents, including paclitaxel, cisplatin, bortezomib, and etoposide, have been revealed to achieve tumor therapeutic efficacy partially through initiating ROS-mediated apoptosis [18]. Adaptive autophagy is essential for cell survival, proliferation, differentiation, and homeostasis, while hyperactivation of autophagy results in type II cell death—autophagic cell death [19]. Interestingly, paradoxical effects have been described for ROS in autophagy. It was reported that excessive ROS levels can activate autophagic flux via the ROS-FOXO3-LC3/BNIP3, ROS-NRF2-P62, ROS-HIF1-BNIP3/NIX, and ROS-TIGAR pathways [20]. ROS accumulation was also shown to inhibit autophagy either by directly increasing AKT/mTOR or by indirectly transactivating antioxidant gene responses to oxidative stress, which serves as a feedback loop to repress autophagy [21]. Additionally, ROS are also relevant to other types of cell death, including ferroptosis [22], pyroptosis [23], and necroptosis [24]. ROS-based chemotherapy strategies have been widely adopted for clinical applications, yet exposure to chemotherapeutic drugs significantly stimulates the overactivation of the antioxidant defense system in cancer cells, which induces a high level of cellular redox balance, triggering adaptive responses to counteract the lethal effects of ROS [25]. Thus, the bright side and dark side of ROS rely on each other in cancer chemotherapy, and chemoresistance emerges when the cellular antioxidant capacity exceeds the cell killing capacity of ROS. In parallel, accumulating evidence suggests that antioxidative factor-targeted therapy can recover a low level of redox status in cancer cells and thus remodel the cytotoxic effect of chemotherapy agents [26,27]. Therefore, proteins engaged in antioxidant activities may play a crucial role in the occurrence of tumor chemoresistance.
Heme oxygenase 1 (HO-1), a rate-limiting enzyme that promotes the catalytic breakdown of heme into carbon monoxide (CO), free iron, and biliverdin, is an essential antioxidant enzyme that induces the host defense response to oxidative stress [28]. HO-1 is normally expressed at a low level under physiological conditions, while it can be activated in stressful situations, which is important for maintaining cellular homeostasis against exogenous stress and injury [29]. Nevertheless, HO-1 is a double-edged sword that plays a cytoprotective role not only in normal cells but also in tumor cells. HO-1 was indicated to be highly induced in multiple cancers, such as breast cancer [30], liver cancer [31], gastric cancer [32], colorectal cancer [33], esophageal squamous cell carcinoma [34], prostate cancer [35], pancreatic cancer [36], and neuroblastoma [37], and was closely related to poor patient prognosis. Upregulated expression and improved enzymatic activity of HO-1 not only represents a potent risk factor for tumorigenesis and tumor progression, but it also relates to therapy resistance [38]. HO-1 inhibition significantly improves chemosensitivity, which motivated an increasing number of preclinical studies committed to exploring and optimizing the ideal strategy for HO-1 targeting to resolve chemoresistance [39,40].
As a stress-related biosensor, the mechanisms of action of HO-1 in chemotherapy resistance are complicated. In addition to its antioxidant function, HO-1 has been revealed to be associated with apoptosis avoidance and autophagy induction, which both result in resistance to cell death [41,42,43]. Thus, approaches targeting HO-1 in chemoresistance have attracted great attention. In this review, we comprehensively summarized the impact of HO-1 on tumor chemoresistance, with a special focus on HO-1-mediated antioxidant, anti-apoptosis, and pro-autophagy activity, as well as proposed potential prospects of HO-1 inhibition in promoting therapeutic benefits in cancers.

2. Biological Functions of HO-1

HO-1, a member of the heme oxygenase (HO) family, also known as heat shock protein 32 (Hsp32), is encoded by the ubiquitous stress-responsive gene HMOX-1, which maps to chromosome 22q12.3 [44]. HO-1 is widely expressed as a type II membrane protein throughout mammals, which predominantly anchors to the endoplasmic reticulum membrane with its hydrophobic carboxy-termini. In addition, the distinct subcellular localization of HO-1 has also been reported, including the caveolae, the nucleus, and the mitochondria [45]. Under physiological conditions, normal cells express low levels of HO-1, which maintains cellular redox homeostasis [46]. Notably, HO-1 is highly inducible by various stress stimuli, including free heme, pathogen infection, tissue injury, altered oxygen tension (hypoxia/hyperoxia), inflammatory cytokines, UV radiation, and oxidant stress, which also serves as a protective feedback mechanism for cellular injury. As a stress response antioxidant enzyme, HO-1 was identified to play a critical role in heme metabolism by catabolizing toxic heme into free iron (Fe2+), carbon monoxide (CO), and the linear tetrapyrrole biliverdin, of which Fe2+ is sequestered by ferritin, and biliverdin is further reduced to bilirubin [47,48,49,50]. HO-1 is widely considered a remarkable cytoprotective factor and represents the cellular defense mechanism due to its enzymatic activity as well as byproducts, which will be further discussed below.

2.1. Degradation of Heme

Heme is synthesized in mitochondria by adding iron to protoporphyrin IX with ferrochelatase and transported to different cellular locations by membrane trafficking proteins as well as heme chaperones, which represent an indispensable cofactor for multiple biological processes, including oxygen binding and transport, electron transfer, oxidative metabolism, gas sensing, signal transduction related to apoptosis, and cell proliferation [51,52]. Intracellular heme homeostasis is tightly controlled because excess free heme is highly toxic by catalyzing the generation of free radicals, together with exacerbating lipid peroxidation, mitochondrial dysfunction, and inflammatory reactions [53,54]. Free heme triggers programmed cell death through various mechanisms. For instance, Fortes et al. indicated that heme induces macrophage necrosis through TLR4/Myd88-dependent expression of TNF and TLR4-independent generation of ROS [55]. Petrillo et al. found that heme accumulation in endothelial cells activates paraptosis by augmenting endoplasmic reticulum (ER) stress [56]. In addition, heme interacts with distinct immune cells to function as a proinflammatory molecule [57,58,59]. Heme itself is an activator of HO-1; in turn, the enzymatic activity of HO-1 eliminates excessive free heme, thereby alleviating cellular oxidative stress and cell injury. Hemin, an artificially synthesized form of heme, was demonstrated to cause DNA strand breaks and oxidative DNA damage in human colonic epithelial (HCEC) and colorectal cancer (CRC) cells, which subsequently promotes Nrf2 stabilization and translocation to the nucleus, along with robust expression of cytosolic HO-1, thereby conferring protection against the deleterious effects of hemin [60]. Nevertheless, a recent study described the opposite role of heme in HO-1 regulation: excessive concentrations of heme obviously compromised the inducibility of HO-1 in normal colonic epithelial cells [61]. Overall, degradation of free heme by HO-1 counteracts its cytotoxic effect, improving cellular homeostasis.

2.2. Metabolites of HO-1

2.2.1. CO

CO is widely known as a poisonous gas due to its ability to bind to hemoglobin with great affinity and thereby block oxygen transport to vital organs, which can cause respiratory depression [62]. However, emerging evidence has suggested that CO, as a gasotransmitter, exerts antioxidant, anti-inflammatory, anti-apoptotic, anti-proliferative, anti-fibrotic, and anti-thrombotic properties at low to moderate doses [63,64,65,66]. Mitochondria are considered the central target organelle in the pro-survival action of CO [62,67,68,69]. The dominant function of mitochondria is ATP generation. Soluble guanylyl cyclase (sGC) is one of the most widely discerned molecular targets of CO [70]. CO elevates ATP production through the induction of sGC and inhibits TNF-α-mediated hepatocyte apoptosis [71,72]. In parallel, low concentrations of CO were found to protect astrocytes against oxidative stress-induced apoptosis by improving mitochondrial oxidative phosphorylation and ATP production, which relies on Bcl-2 expression and its interaction with cytochrome c oxidase (COX) [73]. CO was described to promote ATP production by activating AMPK phosphorylation, which coordinates with the activation of HIF-1α and Nrf2 to suppress hypoxia-induced tubular cell damage [74]. Additionally, CO was shown to increase the nuclear translocation of transcription factor EB (TFEB) via PERK-dependent Ca2+ signaling and calcineurin activation, which subsequently enhanced mitophagy and mitochondrial biogenesis [75]. Due to the great potential therapeutic value of CO, the development of carbon monoxide-releasing molecules (CORMs), which allow CO to be safely and selectively delivered, has recently attracted great attention [76].

2.2.2. Fe2+

Iron is a vital element for numerous biological processes, including mitochondrial respiration, DNA synthesis, signal transduction, and oxygen transport. Iron is a transition element that can exist as reduced ferrous (Fe2+) or oxidized ferric (Fe3+). Fe2+ is implicated in the formation of cellular oxidative stress via the Fenton and Haber–Weiss reactions [77]. Concurrent with HO-1 activation, excessively produced Fe2+ is rapidly sequestered by ferritin in a redox-inactive form. Ferritin is a ubiquitously expressed cytosolic iron storage protein composed of two subunits, ferritin heavy chain (FTH) with ferroxidase activity and ferritin light chain (FTL), which facilitates iron nucleation [78].
Practically, ferritin can be regarded not only as an iron regulatory protein but also as a crucial cellular defense molecule against oxidative stress. In an in vitro oxidative stress model evoked by H2O2, robust and prolonged activation of JNK and p38 MAPKs was observed in a labile iron-dependent manner to trigger cell apoptosis. Along with this, ferritin showed a rapid elevation, which was considered a protective mechanism against external stimuli. Regrettably, the regulatory mechanism of iron in redox signaling was not fully elucidated in this study [79]. Another study showed that Nrf2, the master transcription factor involved in oxidative stress signaling modulation, may directly induce the transcription of the FTL and FTH genes [80]. Recently, ferritin was implied to interact with ferroptosis through selective autophagic degradation of ferritin via nuclear receptor coactivator 4 (NCOA4), namely, ferritinophagy [81,82], which sheds light on the treatment of various diseases, including cancers [83,84], neurodegenerative disorders [85], and immune dysfunction [86,87,88]. Ferroptosis exerts opposite therapeutic effects in a context-dependent manner; for instance, activating ferroptosis is beneficial for tumor therapy but has been found to promote the progression of neurodegenerative diseases. Thus, boosting or inhibiting ferroptosis reasonably in variable contexts to achieve satisfactory therapeutic effects might be possible.

2.2.3. Biliverdin

Biliverdin (BV) is a water-soluble bile pigment that is rapidly converted to bilirubin (BR) by biliverdin reductase (BVR). The BV/BR redox cycle elicits robust defenses against oxidative stress and the inflammatory response. On the one hand, BV and BVR are well-characterized signaling cascades, and BVR exhibits cytoprotective effects by converting BV into the potent antioxidant BR. A previous study revealed that biliverdin reductase mRNA appeared to be increased by exogenous supplementation with biliverdin, which in turn reduced the infiltration of neutrophils and the expression of proinflammatory proteins in an HO-deficient corneal epithelial injury mouse model [89]. Similarly, recent studies have declared that the BVR-dependent BV/BR redox process protects lens epithelial cells (LECs) from oxidative stress-induced apoptosis by enhancing intracellular redox homeostasis, suppressing the NF-κB/iNOS pathway, and activating the Nrf2/HO-1 pathway [90]. On the other hand, BV itself evokes beneficial biological effects. BV could effectively diminish the upregulation of proinflammatory mediators such as IL-6, CCL2, and iNOS and relieve lipid peroxidation, giving rise to a reduction in intestinal mucosal injury in an in vivo intestinal transplantation model [91]. BV also inhibits cerebral ischemia reperfusion-induced cell apoptosis and ameliorates cerebral ischemia reperfusion injury (CIR) in the rat cortex by downregulating lncRNA H19 [92]. Another investigation proposed that elevated miRNA204-5p expression and its direct interaction with Ets1 was the underlying mechanism of BV-related neural improvement [93].
In addition, it is widely accepted that BR replaces BV to function as the final product of HO-1 metabolism in humans. Studies have provided evidence for the protective effects of BR via antioxidation, anti-inflammation, and immune modulation [94] in numerous diseases, including inflammatory bowel disease [95], diabetes [96], ischemia-reperfusion injury [97], and immunological diseases [98].

2.3. Subcellular Localization of HO-1

Apart from functions exerted by the enzyme activity, distinct subcellular compartments of HO-1 also affect its biological activity and function. The C-terminus of endoplasmic reticulum-located HO-1 can be proteolyzed under hypoxic or oxidative stress conditions, leading to different subcellular compartments of HO-1. However, it is somewhat surprising since HO-1 is inactive when translocated into the nucleus, whereas its enzyme activity is preserved when migrated to mitochondria or caveolae [45,99]. Nuclear HO-1 without enzymatic activity displayed the noncanonical function of HO-1, which may mediate oxidative stress protection and cell death prevention through the transcriptional regulation of antioxidant genes or signaling pathways, e.g., G6PDH, NQO1, PI3K/Akt/GSK3 [100], and MAPK pathways [101]. Furthermore, mitochondrial HO-1 was shown to protect lung epithelial cells from mitochondria-mediated cell death [102]. Moreover, upon mitochondrial translocation, HO-1 detoxified accumulated intramitochondrial free heme and reduced mitochondrial oxidative stress and dysfunction, which ultimately repaired apoptotic tissue injury in a gastric mucosal injury rat model [103].
Taken together, these findings imply that both increased expression or enhanced activity of HO-1 and the different subcellular locations of HO-1 variants constitute an important cellular protective mechanism against stressful stimuli.

3. Mechanisms of HO-1 Regulation

3.1. Transcription Factors

The human HO-1 gene (HMOX-1) contains specific DNA-binding elements in the promoter, namely, the proximal promoter (PP), distal enhancer E1 (DE1), and distal enhancer E2 (DE2), located at approximately −0.3 kb, −4 kb, and −10 kb, respectively [104,105,106]. Accumulating evidence has revealed that multiple transcription factors may directly bind to enhancer sequences termed “stress-related response elements” (StREs) in the promoter region of HMOX1, which regulates the expression of HO-1 [106].
Nuclear factor erythroid derived 2-like 2 (Nrf2) and BTB and CNC homology 1 (BACH1) are considered the key transcription factors that conversely modulate HO-1 expression (Figure 1). Nrf2 and BACH1 are both pivotal defense molecules against oxidative stress that bind to similar DNA sequences by forming respective heterodimers with one or more members of the small Maf (musculoaponeurotic fibrosarcoma) protein family. Under basal conditions, Nrf2 is sequestered in the cytoplasm by Kelch-like ECH-associated protein 1 (Keap1), leading to the ubiquitination and subsequent proteolysis of Nrf2, which retains the low activity of Nrf2 [107]. In response to oxidative stress, Nrf2 dissociates from Keap1 and translocates to the nucleus. Then, through the basic leucine zipper (bZip) domain, Nrf2 heterodimerizes with Maf family proteins to bind StRE or ARE (antioxidant response elements) of the target genes [108,109]. It is through this mechanism that Nrf2 positively regulates HO-1 expression under stimulus conditions. In contrast to Nrf2, under normal conditions, BACH1 is stably expressed in the cytosol and transferred into the nucleus to bind to Maf proteins as a transcriptional repressor, which blocks the nuclear translocation of Nrf2 and consequently represses the expression of HO-1 [110]. Under high heme concentrations, BACH1 dissociates from Maf proteins, is exported from the nucleus, and is degraded after tyrosine phosphorylation. Then, Nrf2 directly binds to the HO-1 promoter, leading to the activation of HO-1 expression [111]. Moreover, existing studies also suggest that HMOX1 induction can be BACH1 dependent but NRF2 independent [112,113].
In addition to Nrf2 and BACH1, transcription factors such as HIF1α, NF-κB, ATF4, AP-1, and KLF7 also regulate HO-1 at the transcriptional level by converging on the HO-1 promoter, among which HIF1α [114,115], NF-κB [116], ATF4 [117], and AP-1 [118] positively regulate HO-1 expression, while KLF7 [119] is known as a repressor of HO-1. Of particular interest, HIF1α, a key regulator of hypoxia-stimulated metabolic adaptation, has been considered to be an activator of HO-1 by multiple in vitro and in vivo studies, which has attracted research attention. It was implied in a recent study that HO-1 acts both downstream and upstream of HIF-1α, and HO-1-stabilized HIF-1α may partly be due to its enzymatic activity [120]. The interaction between the two stress adaptation molecules is sophisticated, and its impact on physiological and pathological states is worthy of further study.

3.2. Promoter Polymorphisms

Gene polymorphisms located in the promoter region are thought to affect gene transcription. In the 5’-noncoding region of the HMOX-1 gene, three polymorphic sites were identified to be functional and regulate HO-1 expression, including the (GT)n repeat dinucleotide length polymorphism and T(−413)A and G(−1135)A single nucleotide polymorphism (SNP) sites [121,122]. Among these, the (GT)n polymorphism was the most studied. Current evidence proved that the (GT)n repeat varies from 12–45 repeats [123], and a GT repeat sequence < 25 was defined as a short (S) allele, relating to higher HO-1 inducibility and enhanced HO-1 enzymatic activity compared to those with ≥25 repeat sequences, which were considered long (L) alleles [124]. The other two single-nucleotide polymorphisms, T(−413)A and G(−1135)A, were detected in research aimed at exploring the correlation between HO-1 promoter variants and human essential hypertension. It was shown that the A(−413) allele was associated with significantly improved HMOX-1 promoter activity in comparison with the T(−413) allele [125]. A previous study comprehensively considered both T(−413)A and (GT)n polymorphisms to investigate the promoter activity of HMOX-1, and the results revealed that the transcriptional activity of the A(−413)-(GT)30 allele was approximately six times higher than that of the T(−413)-(GT)23 allele [126]. However, the effect of the G(−1135)A polymorphism on HO-1 transcriptional activity remains unclear.

3.3. MicroRNAs

Emerging evidence indicates that microRNAs are involved in the transcriptional and posttranscriptional modulation of the HO-1 gene, which opens up a new horizon for research on HO-1 regulation. For instance, it was demonstrated that miR-155 efficiently upregulated HMOX1 mRNA and protein expression by inhibiting BACH1 protein translation in human umbilical vein endothelial cells [127]. Gu et al. also found that miR-155 positively regulated HO-1, which favors lung cancer resistance to arsenic trioxide [128]. In contrast to the aforementioned results, Zhang et al. reported that miR-155 could directly target the 3′UTR of HO-1, which repressed its expression in tolerant CD4+ T cells [129]. In addition, Li and colleagues found that miR-155 and miR-181a synergistically engaged in cadmium-induced kidney immunotoxicological effects by reducing HO-1 expression [130]. miR-494 was suggested to upregulate HO-1 expression through a BACH1-independent mechanism in neuroblastoma cells under oxidative stress [131]. miR-1254 and miR-193a-5p were shown to be HO-1 suppressors and inducers in prostate and non-small cell lung cancer cells, respectively [132,133]. Moreover, microRNAs may modulate HO-1 expression at a posttranslational level, as it was proposed that overexpression of miR-217 combined with miR-377 decreased HO-1 protein expression but did not change HMOX1 mRNA levels [134].

4. Cytoprotective Role of HO-1 in Cancer

The primary roles of HO-1 in antioxidant, anti-inflammatory, antiapoptotic, and proangiogenic effects have been highlighted [135]. Nevertheless, HO-1 displays equivalent cytoprotective effects on tumor cells against oxidative stress caused by the accumulation of ROS [136]. Thus, HO-1 has been implicated in a broad spectrum of protumorigenic effects and various cancer hallmarks. Survival analysis in the PrognoScan database (www.prognoscan.org/, accessed on 10 March 2023) provided strong evidence that high HO-1 mRNA expression predicts significantly worse prognosis in lung cancer, breast cancer, blood cancer, brain cancer, colorectal cancer, and ovarian cancer (Figure 2). Meanwhile, in the Cancer Treatment Response Gene Signature Database (CTR-DB, http://ctrdb.cloudna.cn/) [137], the HMOX-1 expression level was shown to be upregulated in treatment-nonresponsive patients with breast and brain cancer compared with responsive patients (Figure 3).
HO-1 overexpression drives tumor growth, metastasis, angiogenesis, therapy resistance, and immune evasion. Wang et al. found that GRIM-19 deficiency resulted in aberrant HO-1 activation in a ROS-Nrf2 axis-dependent manner in gastric cancer cells, along with significantly increased HO-1 expression in metastatic lung and liver tissues, while HO-1 inhibition limited GC cell migration and invasion and directly abrogated GC metastasis in vivo [138]. Overexpressed HO-1 also augmented the bone metastasis of prostate cancer by modulating bone turnover and remodeling. More importantly, HO-1 enhanced the cell-cell interactions of osteoblasts and cancer cells [139]. A more recent study illustrated that exosomes may serve as the basis for HO-1-involved intracellular communication because androgen-independent prostate cancer (AIPC) cell-derived exosomes activated HMOX1 expression in androgen-dependent prostate cancer (ADPC) cells in vivo and in vitro, which promoted the transformation of ADPC cells to AIPC cells and the development of castration-resistant prostate cancer [140]. In addition, HO-1-related angiogenesis plays an essential role in the progression of solid tumors, and the underlying mechanism is likely due to its upregulation or activation of proangiogenic factors such as VEGF and stroma cell-derived factor-1 (SDF-1) [141,142,143].
Notably, an increasing amount of evidence suggests that HO-1 acts as an immunosuppressive agent in immune responses [144,145]. HO-1 expression was induced during the differentiation of monocytic cells into macrophages in the tumor microenvironment (TME) and exerted a strong immunosuppressive effect by limiting antigen-specific CD8+ T-cell effector function against tumor cells, while myeloid-restricted HO-1 ablation boosted the effectiveness of therapeutic immunization [146]. Consistently, Khojandi et al. revealed that HO-1 endowed tumors with the ability to resist immune-mediated apoptosis, and combined treatment with HO-1 inhibition and anti-PD1 significantly decreased tumor volume in a mouse model of breast cancer and melanoma [147]. Indeed, HO-1 may affect the immune response in many distinct ways, among which two main approaches should be highlighted. On the one hand, HO-1 expressed by antigen-presenting cells (APCs) interferes with the development of regulatory T (Treg) cells, along with restricted infiltration and activity of effector T (Teff) cells, which was considered as the fundamental mechanism of the HO-1-related immunosuppressive phenotype [148,149,150]. On the other hand, the promotion of HO-1 expression is associated with biased M2 macrophage polarization, thus facilitating suppressed immune responses and enhancing tumor progression [151,152]. Furthermore, the inhibition of HO-1 augments tumor cell recognition and elimination by NK cells [153,154]. Therefore, HO-1 represents a promising target to reprogram the TME and improve the efficacy of cancer immunotherapy.
Based on their promising therapeutic potential, the development of HO-1 inhibitors has attracted increasing attention. The most reported HO-1 inhibitors for experimental use are metalloporphyrines (Mps), known as heme derivatives, including SnPP, SnMP, and ZnPP, which are structurally similar to protoporphyrin but have different metal ions in their center. Mps can competitively inhibit heme binding to HO-1 to decrease the enzyme activity. However, the clinical application of Mps is limited due to the inhibition of other heme-containing enzymes with poor selectivity, which largely limits its clinical translation [155,156]. Azole-based derivatives such as ketoconazole, terconazole, sulconazole nitrate, and imidazole-based compounds used for HO-1 inhibitors are also reported (Figure 4) [157,158]. Ketoconazole, as a widely used antifungal agent, was indicated to inhibit both HO-1 and HO-2 activity at typical therapeutic concentrations, and the potential mechanism is that the KTZ imidazole moiety directly interacts with heme iron and forms a complex blocking heme from binding to the HO catalytic site [158]. Aside from its antifungal activity, ketoconazole has been applied off-label as a second-line hormonal therapy agent for castration-resistant prostate cancer (CRPC) since the 1980s, due to its additional antiandrogenic function via inhibition of CYP17A1 to block androgen synthesis. Numerous studies have demonstrated that ketoconazole significantly decreases PSA levels, alleviates clinical symptoms, and delays disease progression in CRPC patients, regardless of whether it is used in combination with docetaxel or applied alone before and after docetaxel [159,160]. Furthermore, artesunate, a powerful anti-malaria drug, was declared to have a potential impact on HO-1 modulation [161], iron metabolism [162], and ferroptosis [163]. Concurrently, another approved drug, artesunate, was announced to be active against cancers, including ovarian cancer [164], hepatocellular carcinoma [165], renal cell carcinoma [166], breast cancer [167], and colorectal cancer [168]. Although these drugs showed both HO-1 modulation and cancer therapy effects, whether the HO inhibitory effect directly contributes to the effectiveness of tumor therapy remains unclear.
Conspicuously, aberrant activation of HO-1 was universal in chemoresistant cancer cells compared to that of chemosensitive cancer cells, while the inhibition of HO-1 obviously mitigated resistance to anticancer therapies. These results suggest that HO-1 is implicated in the chemoresistance of various human cancers and could be a potential therapeutic target to overcome chemotherapy failure. This aspect will be discussed in the next section.

5. Mechanisms of HO-1-Targeted Chemoresistance in Cancer

Chemotherapeutic agents related ROS accumulation can eliminate cancer cells by inducing oxidative stress, causing DNA damage, disrupting mitochondrial function, and facilitating synergistic effects [9]. However, its efficacy and long-term use are seriously limited by drug resistance. Recently, a pivotal role of HO-1, a cytoprotective enzyme with known antioxidant defense functions, in chemoresistance has been highlighted. The upregulation of HO-1 after chemotherapy in cancer cells can mitigate ROS-mediated oxidative damage and counteract chemotherapeutic agent-induced cytotoxicity, which contributes to acquired chemoresistance. It has also been well documented that the inhibition of HO-1 is capable of restoring the chemosensitivity of resistant cells in multiple cancer types, including ovarian cancer [169], acute myeloid leukemia [170], and melanoma [154].
To date, the underlying mechanisms of how HO-1 promotes chemoresistance have not been fully elucidated. An increasing number of studies have suggested that antioxidative, anti-apoptosis, and pro-autophagy activities might be involved in the development of HO-1-dependent chemoresistance. In this section, we will comprehensively summarize the role of these mechanisms in tumor chemoresistance associated with HO-1.

5.1. Antioxidative Activity of HO-1

As described above, induction of intracellular ROS by conventional chemotherapeutic agents can be beneficial in terms of their cytotoxic effects on cancer cell. Yet, enhanced HO-1 expression and enzymatic activity protect cancer cells from ROS during stress response, leading to acquired drug resistance. Sun et al. reported that the Nrf2/HO-1 pathway positively regulated by SIRT5 contributed to cisplatin resistance in ovarian cancer cells by suppressing cisplatin-induced DNA damage in an ROS-dependent manner [169]. The upregulation of Nrf2 and its downstream target HO-1 has also been reported to be correlated with 5-Fu resistance in gastric adenocarcinoma by inhibiting ROS production [171]. These observations broadly coincide with the results of our previous work. We found that Nrf2 and its target genes, including HO-1 and NQO-1, were upregulated in cisplatin-resistant ovarian cancer cells, and Nrf2, the main regulator of HO-1, was of great importance for the development of cisplatin resistance in ovarian cancer [172,173,174]. In addition, while induction of HO-1 expression by hemin was shown to protect laryngeal squamous cancer cells from cisplatin-induced oxidative damage and apoptosis in vitro, the suppression of HO-1 expression and enzyme activity by the HO-1 inhibitor ZnPPIX significantly promoted ROS downstream signaling pathway activation, such as P38 and JNK phosphorylation, which conclusively enhanced the cisplatin sensitivity of cancer cells [175].
The mechanisms underlying chemoresistance are also involved in the antioxidant properties of heme metabolites catalyzed by HO-1. Recently, Rios-Arrabal et al. reported that CO produced after HO-1 overexpression induced ECE-1 expression through the activation of pNF-kβ and pc-Jun in p53 wild-type colorectal cancer (CRC) cells, leading to 5-FU resistance. However, CO released by HO-1 was unable to modify 5-FU sensitivity in P53 null CRC cells, suggesting that CO-involved chemoresistance may partly rely on P53 [176]. Although HO-1-produced free ferrous iron itself is considered a mediator of ferroptosis [177], a new type of cell death mediated by iron-dependent ROS and lipid peroxidation, free ferrous iron can be rapidly sequestered and stored by ferritin. The latter has been well documented to have a significant impact on chemoresistance. For instance, ferritin was found to prevent doxorubicin-mediated cell death by inhibiting intracellular ROS formation and then reduce doxorubicin sensitivity in a dose- and time-dependent manner in breast cancer cells [178]. Salatino A et al. observed that chemoresistant ovarian cancer patients may be characterized by higher ferritin heavy chain levels than chemosensitive patients, and overexpression of ferritin heavy chain significantly eliminated cisplatin-mediated ROS, subsequently leading to reduced responsiveness in ovarian cancer cells [179]. Indeed, studies on the association between the cytoprotective role of HO-1 and newly identified forms of cell death, such as ferroptosis [180], necroptosis [181], and pyroptosis [182], are gradually emerging; however, the role of this linkage in chemoresistance remains poorly defined.
In conclusion, these data implied that the antioxidative and cytoprotective features of HO-1 and its byproducts’ attributes to the development of chemoresistance and poor patient prognosis. However, evidence has not been provided that HO-1 possess direct ROS-scavenging activity like “classical” antioxidant enzymes (SOD, catalase, peroxiredoxins, etc.). Instead, HO-1 exert antioxidant effects indirectly, and the possible mechanism includes its removal of “prooxidant” heme and production of antioxidant biliverdin/bilirubin and/or CO products. Moreover, co-expression of HO-1 with other protective enzymes may obscure the real driving force of the antioxidant effect.
Meanwhile, emerging studies have revealed that HO-1 is a double-edged sword in ROS regulation; when excessively activated, HO-1 can induce ROS overload and cancer cell death [183,184]. The probable mechanism may involve uncoordinated stress response mechanisms mainly caused by imbalanced iron metabolism. Some scholars have concluded that the iron deprivation status of the body impels the antioxidant function of HO-1 [185]. Since HO-1 regulates iron metabolism via the degradation of heme, once excessively produced iron cannot be utilized for biosynthesis or sequestered by ferritin, oxidative stress occurs, which also initiates ferroptotic cell death [186,187]. Hence, the dual role of HO-1 in ROS modulation potentially substantiates its paradoxical action in ferroptosis.

5.2. Anti-Apoptosis Activity of HO-1

Apoptotic cell death, a form of programmed cell death provoked in tumor cells following exposure to chemotherapeutic agents and excessive oxidative stress [188,189], is integral to the success of cancer chemotherapy. Since enforced expression and activation of HO-1 in response to chemotherapy-mediated oxidative stress could provide cytoprotective effects against chemotherapy-induced apoptosis in tumor cells, this antiapoptotic mechanism might contribute to acquired chemoresistance. A protein-protein interaction (PPI) network of HMOX1-related genes was constructed using the STRING database (Figure 5A). To further explore the function of HMOX1, these interacting genes from the PPI network were used for Gene Ontology (GO) enrichment analysis. The GO analysis results showed that the HMOX1-related genes were mainly enriched in antioxidative stress and anti-apoptosis biological processes (Figure 5B). Based on this, we will fully address the influence of HO-1 on apoptosis, apoptosis-related proteins, and signaling pathways in resistant tumor cells.
The transcriptional and protein levels of pro-apoptosis-related genes, such as Bax, Smac, Survivin, Fas/FasL, caspase-3 or caspase-9, and anti-apoptosis genes, such as Bcl-2, BFAR, Bcl-xL, A1 or Mcl-1, are often measured to estimate the degree of cellular apoptosis [190,191]. Barbagallo et al. described that increased HO-1 gene expression and enzymatic activity drive resistance to carfilzomib in neuroblastoma cells, while cotreatment with carfilzomib and LS1/71, a noncompetitive inhibitor of HO-1, triggered cell apoptosis with significantly activated proapoptotic BAX gene expression and reduced antiapoptotic BFAR gene expression [192]. Chakraborty et al. determined that the cytoprotective effect of the c-Met-HGF-Nrf2-HO-1 pathway in sorafenib-treated renal cancer cells was mainly implicated in the reduction in ROS generation, promotion of Bcl-2 and Bcl-xL expression, and downregulation of cleaved caspase-3 expression, which ultimately led to apoptosis resistance [193]. These results are consistent with the observation that HO-1 inhibition sensitized pancreatic ductal adenocarcinoma cells to gemcitabine by increasing the production of reactive oxygen species, disrupting the glutathione cycle, and enhancing apoptosis [194]. Recently, microRNAs involved in the regulation of HO-1 were suggested to participate in the development of chemoresistance by decreasing apoptosis. Gu et al. confirmed that HO-1 expression modulated by miR-155 was partly responsible for the downregulation of arsenic trioxide-induced apoptosis, leading to resistance to arsenic trioxide in lung cancer cells [128]. miR-193a-5p-induced HO-1 expression was shown to counteract docetaxel-induced apoptosis in PC3 cells [133].
In addition, considerable research efforts have been devoted to investigating the impact of HO-1 end products generated by heme metabolism on apoptosis evasion. Huang et al. proposed that activation of the HO-1/CO axis protects lens epithelial cells (LECs) from H2O2-induced apoptosis by inhibiting oxidative stress and activating NF-κB signaling, along with decreasing apoptotic molecules (Bax, Bcl-2, and caspase-3) [195]. A recent in vitro study revealed that CO produced by HO-1 may contribute to the resistance of myeloma cells to bortezomib. Specifically, the treatment of myeloma cells with bortezomib induced cross-regulation between TLR4 and the HO-1/CO signaling pathway, which subsequently increased the unfolded protein response and protected the function of mitochondria, leading to decreased apoptosis resulting from the cytotoxic effects of bortezomib [196]. In fact, it should be noted that existing studies have mainly focused on the apoptosis evasion role of HO-1 itself in chemoresistance, and further effort is still required to understand the antiapoptotic mechanisms regulated by HO-1 byproducts in chemoresistance.

5.3. Pro-Autophagy Activity of HO-1

Autophagy, an evolutionarily conserved catabolic process, plays an essential role in maintaining cellular homeostasis under various metabolic stresses, such as nutritional deficiency, hypoxia, pathogen infection, ER stress, oxidative stress, and even chemotherapy intervention, by capturing and transporting aggregated or misfolded proteins and damaged organelles to lysosomes for degradation [197,198]. To date, three major types have been identified, including macroautophagy, chaperone-mediated autophagy, and microautophagy. Of these, macroautophagy (hereafter autophagy) is the most studied.
In cancer cells, autophagy is a double-edged sword for cell fate determination. On the one hand, excessive or persistent autophagy promotes autophagic cell death, limits tumorigenesis and enhances therapeutic efficacy in some specific contexts [199]. On the other hand, the prosurvival function of autophagy is thought to be involved in oncogenesis, tumor progression, and therapy resistance [200].
HO-1 and autophagy are both inducible by stress conditions. Indeed, HO-1 was revealed to be coregulated with autophagy in the development of tumor chemoresistance, and manipulating HO-1-related autophagy to improve therapeutic responses has attracted attention. A previous study showed that HO-1 expression was upregulated by rapamycin and sorafenib treatment in renal cancer cells, which subsequently promoted cancer cell survival by suppressing both apoptotic and autophagic cell death. The underlying mechanism of HO-1-downregulated autophagy was associated with increased linkage between Beclin-1 and Rubicon [201]. Nevertheless, the positive role of HO-1 in autophagy activation to promote cell survival has been highlighted. HO-1 activated autophagic flux by increasing the LC3BI/LC3BII ratio and upregulating ERK and JNK, which was indicated to drive resistance against HER2-targeted therapies in breast cancer [202]. HO-1 boosted autophagy modulation by the Src-STAT3 and PI3K/Akt signaling pathways and was also discovered to contribute to doxorubicin and pharmorubicin resistance, respectively, in breast cancer cells [42,43]. In chronic myeloid leukemia cells, silencing HO-1 was shown to reverse imatinib resistance by inhibiting autophagy through the activation of the mTOR pathway [203]. In addition, increased HO-1 induced by silencing BACH2 was demonstrated to facilitate bortezomib resistance in mantle cell lymphoma cells by triggering cytoprotective autophagy formation and maintaining ROS at a minimal tumor-promoting level [204].
The underlying mechanism of HO-1-regulated autophagy is complicated and not fully elucidated. Autophagy and apoptosis are tightly intertwined, and the regulation of autophagy by HO-1 is often accompanied by changes in apoptosis, either by inducing autophagy to inhibit apoptosis and play a protective role [205,206,207] or activating excessive autophagy to blow up apoptotic cell death [208]. Thus, appropriate regulation of HO-1-associated autophagy might become a potential therapeutic strategy to ameliorate chemoresistance. It is noteworthy that mitochondria are critical to the manipulation of apoptosis, energy production and cellular homeostasis [209]. Mitophagy, a selective form of autophagy that facilitates the elimination of damaged mitochondria before they cause cellular function impairments, can be amplified by overactivation of HO-1 and then undergo cell demise in glioma cells [210]. In accordance with this, HO-1-dependent NRF-1 induction was validated to arrest cardiomyocyte cell death and prevent fibrosis after oxidative stress through the transcriptional regulation of key mitophagy proteins that preserve mitochondrial homeostasis [211]. In addition, AMPK, a crucial regulator of energy metabolic homeostasis, was recently found to antagonize mTOR and facilitate HO-1-associated autophagy [212]. HO-1 was implicated in maintaining mitochondrial quality by inhibiting mitochondrial fission [213,214]. Based on the evidence reviewed above, the crosstalk between autophagy and apoptosis as well as the participation of mitochondria in biological processes might be relevant to the regulation of autophagy by HO-1, which still remains to be thoroughly investigated.

6. Conclusions

HO-1 is an established component of the antioxidant defense system that exerts cytoprotective functions by scavenging ROS and recovering redox homeostasis to counteract oxidative stress. However, recent developments have suggested that aberrant activation of HO-1 may be the decisive factor for chemoresistance in various cancers. In this review, we highlighted that increased antioxidative, antiapoptotic, and proautophagic activities are crucial mechanisms underlying HO-1-mediated chemoresistance (Figure 6).
Research to date has revealed that inhibition of HO-1 may be a promising strategy to overcome chemoresistance and improve the efficacy of chemotherapeutic regimens. Both pharmacological suppression of HO-1 activity by zinc protoporphyrin IX (ZnPP-IX) or tin protoporphyrin IX (SnPP-IX) and targeted knockdown of HO-1 expression by siRNA significantly enhanced chemosensitivity in cancer cells and suppressed chemoresistant xenograft tumor growth in vivo. Advances have been made in understanding the mechanisms underlying HO-1-mediated chemoresistance, and further development of HO-1-targeted therapies provides encouraging prospects for clinical application to optimize patient prognosis in different cancers. However, translating preclinical results to the clinic has been challenging due to the lack of safe and effective HO-1 inhibitors. It is noteworthy that suppression of HO-1 might break its protective effect on normal cells when exerting a killing effect on tumor cells; thus, the side effects of HO-1 inhibitors on normal cells and the incidence of adverse outcomes remain to be properly measured in the development of new inhibitors. Furthermore, whether HO-1-dependent therapy resistance is universal or specific for certain tumor types remains largely unexplored. As discussed above, overexpressed HO-1 leads to therapy resistance in both solid and hematological tumors. As a future perspective, the detailed molecular mechanisms of how HO-1 meditates chemoresistance in a specific cancer type and the development of safe, effective HO-1 inhibitors warrant further investigation.

Author Contributions

H.W., Q.C. and L.B. collected and reviewed the literature. X.Y. and K.C. conceived the review and designed the review outline. H.W. performed the bioinformatic analyses and drafted the manuscript. M.L., X.Y. and K.C. revised the manuscript. All authors contributed to the article and approved the submitted version. X.Y. and K.C. contributed equally to this work. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Natural Science Foundation of China (81872432).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data generated or analyzed during this study are included in this published article.

Acknowledgments

The authors acknowledgment support from the Natural Science Foundation of China (No. 81872432). The authors gratefully acknowledge the editors and reviewers for their insightful comments and constructive suggestions that helped us improve the manuscript.

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Hayes, J.D.; Dinkova-Kostova, A.T.; Tew, K.D. Oxidative Stress in Cancer. Cancer Cell 2020, 38, 167–197. [Google Scholar] [CrossRef] [PubMed]
  2. Costa, R.G.A.; Silva, S.L.R.; Dias, I.R.S.B.; Oliveira, M.S.; Rodrigues, A.C.B.D.C.; Dias, R.B.; Bezerra, D.P. Emerging drugs targeting cellular redox homeostasis to eliminate acute myeloid leukemia stem cells. Redox Biol. 2023, 62, 102692. [Google Scholar] [CrossRef]
  3. He, M.; Wang, M.; Xu, T.; Zhang, M.; Dai, H.; Wang, C.; Ding, D.; Zhong, Z. Reactive oxygen species-powered cancer immunotherapy: Current status and challenges. J. Control. Release 2023, 356, 623–648. [Google Scholar] [CrossRef] [PubMed]
  4. Hämäläinen, R.H.; Ahlqvist, K.J.; Ellonen, P.; Lepistö, M.; Logan, A.; Otonkoski, T.; Murphy, M.P.; Suomalainen, A. mtDNA mutagenesis disrupts pluripotent stem cell function by altering redox signaling. Cell Rep. 2015, 11, 1614–1624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Trachootham, D.; Alexandre, J.; Huang, P. Targeting cancer cells by ROS-mediated mechanisms: A radical therapeutic approach? Nat. Rev. Drug Discov. 2009, 8, 579–591. [Google Scholar] [CrossRef] [PubMed]
  6. Lü, J.M.; Jiang, J.; Jamaluddin, M.S.; Liang, Z.; Yao, Q.; Chen, C. Ginsenoside Rb1 blocks ritonavir-induced oxidative stress and eNOS downregulation through activation of estrogen receptor-beta and upregulation of SOD in human endothelial cells. Int. J. Mol. Sci. 2019, 20, 294. [Google Scholar] [CrossRef] [Green Version]
  7. Xu, L.; Yang, J.; Xu, M.; Shan, D.; Wu, Z.; Yuan, D. Speciation and adaptive evolution reshape antioxidant enzymatic system diversity across the phylum Nematoda. BMC Biol. 2020, 18, 181. [Google Scholar] [CrossRef]
  8. Matschke, V.; Theiss, C.; Matschke, J. Oxidative stress: The lowest common denominator of multiple diseases. Neural Regen. Res. 2019, 14, 238–241. [Google Scholar] [CrossRef]
  9. Nakamura, H.; Takada, K. Reactive oxygen species in cancer: Current findings and future directions. Cancer Sci. 2021, 112, 3945–3952. [Google Scholar] [CrossRef]
  10. Checa, J.; Aran, J.M. Reactive oxygen species: Drivers of physiological and pathological Processes. J. Inflamm. Res. 2020, 13, 1057–1073. [Google Scholar] [CrossRef] [PubMed]
  11. Gorrini, C.; Harris, I.S.; Mak, T.W. Modulation of oxidative stress as an anticancer strategy. Nat. Rev. Drug Discov. 2013, 12, 931–947. [Google Scholar] [CrossRef] [PubMed]
  12. Sun, J.; Feng, Y.; Wang, Y.; Ji, Q.; Cai, G.; Shi, L.; Wang, Y.; Huang, Y.; Zhang, J.; Li, Q. α-hederin induces autophagic cell death in colorectal cancer cells through reactive oxygen species dependent AMPK/mTOR signaling pathway activation. Int. J. Oncol. 2019, 54, 1601–1612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Liu, K.; Guo, C.; Lao, Y.; Yang, J.; Chen, F.; Zhao, Y.; Yang, Y.; Yang, J.; Yi, J. A fine-tuning mechanism underlying self-control for autophagy: deSUMOylation of BECN1 by SENP3. Autophagy 2020, 16, 975–990. [Google Scholar] [CrossRef] [PubMed]
  14. Kou, J.Y.; Li, Y.; Zhong, Z.Y.; Jiang, Y.Q.; Li, X.S.; Han, X.B.; Liu, Z.N.; Tian, Y.; Yang, L.M. Berberine-sonodynamic therapy induces autophagy and lipid unloading in macrophage. Cell Death Dis. 2017, 8, e2558. [Google Scholar] [CrossRef] [Green Version]
  15. Yousefi, S.; Perozzo, R.; Schmid, I.; Ziemiecki, A.; Schaffner, T.; Scapozza, L.; Brunner, T.; Simon, H.U. Calpain-mediated cleavage of Atg5 switches autophagy to apoptosis. Nat. Cell Biol. 2006, 8, 1124–1132. [Google Scholar] [CrossRef] [PubMed]
  16. Shen, Y.; Yang, J.; Zhao, J.; Xiao, C.; Xu, C.; Xiang, Y. The switch from ER stress-induced apoptosis to autophagy via ROS-mediated JNK/p62 signals: A survival mechanism in methotrexate-resistant choriocarcinoma cells. Exp. Cell Res. 2015, 334, 207–218. [Google Scholar] [CrossRef]
  17. Redza-Dutordoir, M.; Averill-Bates, D.A. Activation of apoptosis signaling pathways by reactive oxygen species. Biochim. Biophys. Acta 2016, 1863, 2977–2992. [Google Scholar] [CrossRef]
  18. Wondrak, G.T. Redox-directed cancer therapeutics: Molecular mechanisms and opportunities. Antioxid. Redox Signal. 2009, 11, 3013–3069. [Google Scholar] [CrossRef] [Green Version]
  19. Levine, B.; Kroemer, G. Autophagy in the pathogenesis of disease. Cell 2008, 132, 27–42. [Google Scholar] [CrossRef] [Green Version]
  20. Li, L.; Tan, J.; Miao, Y.; Lei, P.; Zhang, Q. ROS and autophagy: Interactions and molecular regulatory mechanisms. Cell Mol. Neurobiol. 2015, 35, 615–621. [Google Scholar] [CrossRef]
  21. Kaminskyy, V.O.; Zhivotovsky, B. Free radicals in cross talk between autophagy and apoptosis. Antioxid. Redox Signal. 2014, 21, 86–102. [Google Scholar] [CrossRef] [PubMed]
  22. Liu, J.; Kang, R.; Tang, D. Signaling pathways and defense mechanisms of ferroptosis. FEBS J. 2022, 289, 7038–7050. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, Y.; Shi, P.; Chen, Q.; Huang, Z.; Zou, D.; Zhang, J.; Gao, X.; Lin, Z. Mitochondrial ROS promote macrophage pyroptosis by inducing GSDMD oxidation. J. Mol. Cell Biol. 2019, 11, 1069–1082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Deragon, M.A.; McCaig, W.D.; Patel, P.S.; Haluska, R.J.; Hodges, A.L.; Sosunov, S.A.; Murphy, M.P.; Ten, V.S.; LaRocca, T.J. Mitochondrial ROS prime the hyperglycemic shift from apoptosis to necroptosis. Cell Death Discov. 2020, 6, 132. [Google Scholar] [CrossRef] [PubMed]
  25. Kim, E.K.; Jang, M.; Song, M.J.; Kim, D.; Kim, Y.; Jang, H.H. Redox-mediated mechanism of chemoresistance in cancer cells. Antioxidants 2019, 8, 471. [Google Scholar] [CrossRef] [Green Version]
  26. Wang, J.; Lu, Q.; Cai, J.; Wang, Y.; Lai, X.; Qiu, Y.; Huang, Y.; Ke, Q.; Zhang, Y.; Guan, Y.; et al. Nestin regulates cellular redox homeostasis in lung cancer through the Keap1-Nrf2 feedback loop. Nat. Commun. 2019, 10, 5043. [Google Scholar] [CrossRef] [Green Version]
  27. Wang, K.; Jiang, J.; Lei, Y.; Zhou, S.; Wei, Y.; Huang, C. Targeting metabolic-redox circuits for cancer therapy. Trends Biochem. Sci. 2019, 44, 401–414. [Google Scholar] [CrossRef]
  28. Ko, W.; Yoon, C.S.; Kim, K.W.; Lee, H.; Kim, N.; Woo, E.R.; Kim, Y.C.; Kang, D.G.; Lee, H.S.; Oh, H.; et al. Neuroprotective and anti-inflammatory effects of kuwanon C from Cudrania tricuspidata are mediated by heme oxygenase-1 in HT22 hippocampal cells, RAW264.7 macrophage, and BV2 microglia. Int. J. Mol. Sci. 2020, 21, 4839. [Google Scholar] [CrossRef]
  29. Ortiz, E.; Sanchis, P.; Bizzotto, J.; Lage-Vickers, S.; Labanca, E.; Navone, N.; Cotignola, J.; Vazquez, E.; Gueron, G. Myxovirus resistance protein 1 (MX1), a novel HO-1 interactor, tilts the balance of endoplasmic reticulum stress towards pro-death events in prostate cancer. Biomolecules 2020, 10, 1005. [Google Scholar] [CrossRef]
  30. Noh, S.J.; Bae, J.S.; Jamiyandorj, U.; Park, H.S.; Kwon, K.S.; Jung, S.H.; Youn, H.J.; Lee, H.; Park, B.H.; Chung, M.J.; et al. Expression of nerve growth factor and heme oxygenase-1 predict poor survival of breast carcinoma patients. BMC Cancer 2013, 13, 516. [Google Scholar] [CrossRef] [Green Version]
  31. Park, C.S.; Eom, D.W.; Ahn, Y.; Jang, H.J.; Hwang, S.; Lee, S.G. Can heme oxygenase-1 be a prognostic factor in patients with hepatocellular carcinoma? Medicine 2019, 98, e16084. [Google Scholar] [CrossRef]
  32. Noh, S.J.; Kim, K.M.; Jang, K.Y. Individual and co-expression patterns of nerve growth factor and heme oxygenase-1 predict shorter survival of gastric carcinoma patients. Diagn. Pathol. 2017, 12, 48. [Google Scholar] [CrossRef] [PubMed]
  33. Kimura, S.; Aung, N.Y.; Ohe, R.; Yano, M.; Hashimoto, T.; Fujishima, T.; Kimura, W.; Yamakawa, M. Increasing heme oxygenase-1-expressing macrophages indicates a tendency of poor prognosis in advanced colorectal cancer. Digestion 2020, 101, 401–410. [Google Scholar] [CrossRef] [PubMed]
  34. Akaishi, R.; Fujishima, F.; Ishida, H.; Tsunokake, J.; Yamauchi, T.; Gokon, Y.; Ueki, S.; Fukutomi, T.; Okamoto, H.; Takaya, K.; et al. Correlation between TXNRD1/HO-1 expression and response to neoadjuvant chemoradiation therapy in patients with esophageal squamous cell carcinoma. Esophagus 2022, 19, 436–443. [Google Scholar] [CrossRef] [PubMed]
  35. Alaoui-Jamali, M.A.; Bismar, T.A.; Gupta, A.; Szarek, W.A.; Su, J.; Song, W.; Xu, Y.; Xu, B.; Liu, G.; Vlahakis, J.Z.; et al. A novel experimental heme oxygenase-1-targeted therapy for hormone-refractory prostate cancer. Cancer Res. 2009, 69, 8017–8024. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Berberat, P.O.; Dambrauskas, Z.; Gulbinas, A.; Giese, T.; Giese, N.; Künzli, B.; Autschbach, F.; Meuer, S.; Büchler, M.W.; Friess, H. Inhibition of heme oxygenase-1 increases responsiveness of pancreatic cancer cells to anticancer treatment. Clin. Cancer Res. 2005, 11, 3790–3798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Fest, S.; Soldati, R.; Christiansen, N.M.; Zenclussen, M.L.; Kilz, J.; Berger, E.; Starke, S.; Lode, H.N.; Engel, C.; Zenclussen, A.C.; et al. Targeting of heme oxygenase-1 as a novel immune regulator of neuroblastoma. Int. J. Cancer 2016, 138, 2030–2042. [Google Scholar] [CrossRef] [Green Version]
  38. Chau, L.Y. Heme oxygenase-1: Emerging target of cancer therapy. J. Biomed. Sci. 2015, 22, 22. [Google Scholar] [CrossRef] [Green Version]
  39. Kim, E.J.; Kim, Y.J.; Lee, H.I.; Jeong, S.H.; Nam, H.J.; Cho, J.H. Nrf2 knockdown resensitizes 5-Fluorouracil-resistant pancreatic cancer cells by suppressing HO-1 and ABCG2 expression. Int. J. Mol. Sci. 2020, 21, 4646. [Google Scholar] [CrossRef] [PubMed]
  40. Mucha, O.; Podkalicka, P.; Mikulski, M.; Barwacz, S.; Andrysiak, K.; Biela, A.; Mieczkowski, M.; Kachamakova-Trojanowska, N.; Ryszawy, D.; Białas, A.; et al. Development and characterization of a new inhibitor of heme oxygenase activity for cancer treatment. Arch. Biochem. Biophys. 2019, 671, 130–142. [Google Scholar] [CrossRef]
  41. Zhang, L.; Zhang, X.; Wu, T.; Pan, X.; Wang, Z. Isoflurane reduces septic neuron injury by HO-1-mediated abatement of inflammation and apoptosis. Mol. Med. Rep. 2021, 23, 115. [Google Scholar] [CrossRef]
  42. Tan, Q.; Wang, H.; Hu, Y.; Hu, M.; Li, X.; Aodengqimuge, M.Y.; Wei, C.; Song, L. Src/STAT3-dependent heme oxygenase-1 induction mediates chemoresistance of breast cancer cells to doxorubicin by promoting autophagy. Cancer Sci. 2015, 106, 1023–1032. [Google Scholar] [CrossRef] [PubMed]
  43. Pei, L.; Kong, Y.; Shao, C.; Yue, X.; Wang, Z.; Zhang, N. Heme oxygenase-1 induction mediates chemoresistance of breast cancer cells to pharmorubicin by promoting autophagy via PI3K/Akt pathway. J. Cell Mol. Med. 2018, 22, 5311–5321. [Google Scholar] [CrossRef] [PubMed]
  44. Hill-Kapturczak, N.; Sikorski, E.; Voakes, C.; Garcia, J.; Nick, H.S.; Agarwal, A. An internal enhancer regulates heme- and cadmium-mediated induction of human heme oxygenase-1. Am. J. Physiol. Ren. Physiol. 2003, 285, F515–F523. [Google Scholar] [CrossRef] [Green Version]
  45. Dunn, L.L.; Midwinter, R.G.; Ni, J.; Hamid, H.A.; Parish, C.R.; Stocker, R. New insights into intracellular locations and functions of heme oxygenase-1. Antioxid. Redox Signal. 2014, 20, 1723–1742. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Ryter, S.W.; Alam, J.; Choi, A.M. Heme oxygenase-1/carbon monoxide: From basic science to therapeutic applications. Physiol. Rev. 2006, 86, 583–650. [Google Scholar] [CrossRef]
  47. Salerno, L.; Vanella, L.; Sorrenti, V.; Consoli, V.; Ciaffaglione, V.; Fallica, A.N.; Canale, V.; Zajdel, P.; Pignatello, R.; Intagliata, S. Novel mutual prodrug of 5-fluorouracil and heme oxygenase-1 inhibitor (5-FU/HO-1 hybrid): Design and preliminary in vitro evaluation. J. Enzym. Inhib. Med. Chem. 2021, 36, 1378–1386. [Google Scholar] [CrossRef]
  48. Alves de Souza, R.W.; Gallo, D.; Lee, G.R.; Katsuyama, E.; Schaufler, A.; Weber, J.; Csizmadia, E.; Tsokos, G.C.; Koch, L.G.; Britton, S.L.; et al. Skeletal muscle heme oxygenase-1 activity regulates aerobic capacity. Cell Rep. 2021, 35, 109018. [Google Scholar] [CrossRef]
  49. Lim, S.; Kwon, M.; Joung, E.J.; Shin, T.; Oh, C.W.; Choi, J.S.; Kim, H.R. Meroterpenoid-rich fraction of the ethanolic extract from Sargassum serratifolium suppressed oxidative stress induced by tert-butyl hydroperoxide in HepG2 cells. Mar. Drugs 2018, 16, 374. [Google Scholar] [CrossRef] [Green Version]
  50. Youssef, L.A.; Rebbaa, A.; Pampou, S.; Weisberg, S.P.; Stockwell, B.R.; Hod, E.A.; Spitalnik, S.L. Increased erythrophagocytosis induces ferroptosis in red pulp macrophages in a mouse model of transfusion. Blood 2018, 131, 2581–2593. [Google Scholar] [CrossRef]
  51. Dutt, S.; Hamza, I.; Bartnikas, T.B. Molecular mechanisms of iron and heme metabolism. Annu. Rev. Nutr. 2022, 42, 311–335. [Google Scholar] [CrossRef] [PubMed]
  52. Mendez, D.L.; Lowder, E.P.; Tillman, D.E.; Sutherland, M.C.; Collier, A.L.; Rau, M.J.; Fitzpatrick, J.A.J.; Kranz, R.G. Cryo-EM of CcsBA reveals the basis for cytochrome c biogenesis and heme transport. Nat. Chem. Biol. 2022, 18, 101–108. [Google Scholar] [CrossRef] [PubMed]
  53. Sparkenbaugh, E.M.; Chantrathammachart, P.; Wang, S.; Jonas, W.; Kirchhofer, D.; Gailani, D.; Gruber, A.; Kasthuri, R.; Key, N.S.; Mackman, N.; et al. Excess of heme induces tissue factor-dependent activation of coagulation in mice. Haematologica 2015, 100, 308–314. [Google Scholar] [CrossRef] [Green Version]
  54. Merle, N.S.; Grunenwald, A.; Figueres, M.L.; Chauvet, S.; Daugan, M.; Knockaert, S.; Robe-Rybkine, T.; Noe, R.; May, O.; Frimat, M.; et al. Characterization of renal injury and inflammation in an experimental model of intravascular hemolysis. Front. Immunol. 2018, 9, 179. [Google Scholar] [CrossRef] [PubMed]
  55. Fortes, G.B.; Alves, L.S.; de Oliveira, R.; Dutra, F.F.; Rodrigues, D.; Fernandez, P.L.; Souto-Padron, T.; De Rosa, M.J.; Kelliher, M.; Golenbock, D.; et al. Heme induces programmed necrosis on macrophages through autocrine TNF and ROS production. Blood 2012, 119, 2368–2375. [Google Scholar] [CrossRef] [Green Version]
  56. Petrillo, S.; Chiabrando, D.; Genova, T.; Fiorito, V.; Ingoglia, G.; Vinchi, F.; Mussano, F.; Carossa, S.; Silengo, L.; Altruda, F.; et al. Heme accumulation in endothelial cells impairs angiogenesis by triggering paraptosis. Cell Death Differ. 2018, 25, 573–588. [Google Scholar] [CrossRef] [Green Version]
  57. Liu, X.; Spolarics, Z. Methemoglobin is a potent activator of endothelial cells by stimulating IL-6 and IL-8 production and E-selectin membrane expression. Am. J. Physiol. Cell Physiol. 2003, 285, C1036–C1046. [Google Scholar] [CrossRef] [Green Version]
  58. Nyakundi, B.B.; Tóth, A.; Balogh, E.; Nagy, B.; Erdei, J.; Ryffel, B.; Paragh, G.; Cordero, M.D.; Jeney, V. Oxidized hemoglobin forms contribute to NLRP3 inflammasome-driven IL-1beta production upon intravascular hemolysis. Biochim. Biophys. Acta Mol. Basis Dis. 2019, 1865, 464–475. [Google Scholar] [CrossRef]
  59. Erdei, J.; Tóth, A.; Balogh, E.; Nyakundi, B.B.; Bányai, E.; Ryffel, B.; Paragh, G.; Cordero, M.D.; Jeney, V. Induction of NLRP3 inflammasome activation by heme in human endothelial cells. Oxid. Med. Cell Longev. 2018, 2018, 4310816. [Google Scholar] [CrossRef] [Green Version]
  60. Seiwert, N.; Wecklein, S.; Demuth, P.; Hasselwander, S.; Kemper, T.A.; Schwerdtle, T.; Brunner, T.; Fahrer, J. Heme oxygenase 1 protects human colonocytes against ROS formation, oxidative DNA damage and cytotoxicity induced by heme iron, but not inorganic iron. Cell Death Dis. 2020, 11, 787. [Google Scholar] [CrossRef]
  61. Gamage, S.M.K.; Nanayakkara, S.; Macfarlane, L.; Hewage, D.; Cheng, T.; Aktar, S.; Lu, C.T.; Dissabandara, L.; Islam, F.; Lam, A.K.; et al. Heme oxygenase-1 & 2 and their potential contribution in heme induced colorectal carcinogenesis. Pathol. Res. Pract. 2022, 233, 153885. [Google Scholar]
  62. Kapetanaki, S.M.; Burton, M.J.; Basran, J.; Uragami, C.; Moody, P.C.E.; Mitcheson, J.S.; Schmid, R.; Davies, N.W.; Dorlet, P.; Vos, M.H.; et al. A mechanism for CO regulation of ion channels. Nat. Commun. 2018, 9, 907. [Google Scholar] [CrossRef] [Green Version]
  63. Parfenova, H.; Leffler, C.W.; Basuroy, S.; Liu, J.; Fedinec, A.L. Antioxidant roles of heme oxygenase, carbon monoxide, and bilirubin in cerebral circulation during seizures. J. Cereb. Blood Flow Metab. 2012, 32, 1024–1034. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Knauert, M.; Vangala, S.; Haslip, M.; Lee, P.J. Therapeutic applications of carbon monoxide. Oxid. Med. Cell Longev. 2013, 2013, 360815. [Google Scholar] [CrossRef]
  65. Ryter, S.W. Heme oxygenase-1/carbon monoxide as modulators of autophagy and inflammation. Arch. Biochem. Biophys. 2019, 678, 108186. [Google Scholar] [CrossRef] [PubMed]
  66. Di Pietro, C.; Öz, H.H.; Murray, T.S.; Bruscia, E.M. Targeting the heme oxygenase 1/carbon monoxide pathway to resolve lung hyper-inflammation and restore a regulated immune response in cystic fibrosis. Front. Pharmacol. 2020, 11, 1059. [Google Scholar] [CrossRef] [PubMed]
  67. Zhou, H.; Liu, J.; Pan, P.; Jin, D.; Ding, W.; Li, W. Carbon monoxide inhalation decreased lung injury via anti-inflammatory and anti-apoptotic effects in brain death rats. Exp. Biol. Med. 2010, 235, 1236–1243. [Google Scholar] [CrossRef]
  68. Averilla, J.N.; Oh, J.; Kim, J.S. Carbon monoxide partially mediates protective effect of resveratrol against UVB-induced oxidative stress in human keratinocytes. Antioxidants 2019, 8, 432. [Google Scholar] [CrossRef] [Green Version]
  69. Tiwari, R.; Shinde, P.S.; Sreedharan, S.; Dey, A.K.; Vallis, K.A.; Mhaske, S.B.; Pramanik, S.K.; Das, A. Photoactivatable prodrug for simultaneous release of mertansine and CO along with a BODIPY derivative as a luminescent marker in mitochondria: A proof of concept for NIR image-guided cancer therapy. Chem. Sci. 2020, 12, 2667–2673. [Google Scholar] [CrossRef]
  70. Lu, W.; Yang, X.; Wang, B. Carbon monoxide signaling and soluble guanylyl cyclase: Facts, myths, and intriguing possibilities. Biochem. Pharmacol. 2022, 200, 115041. [Google Scholar] [CrossRef]
  71. Tsui, T.Y.; Siu, Y.T.; Schlitt, H.J.; Fan, S.T. Heme oxygenase-1-derived carbon monoxide stimulates adenosine triphosphate generation in human hepatocyte. Biochem. Biophys. Res. Commun. 2005, 336, 898–902. [Google Scholar] [CrossRef]
  72. Tsui, T.Y.; Obed, A.; Siu, Y.T.; Yet, S.F.; Prantl, L.; Schlitt, H.J.; Fan, S.T. Carbon monoxide inhalation rescues mice from fulminant hepatitis through improving hepatic energy metabolism. Shock 2007, 27, 165–171. [Google Scholar] [CrossRef] [PubMed]
  73. Almeida, A.S.; Queiroga, C.S.; Sousa, M.F.; Alves, P.M.; Vieira, H.L. Carbon monoxide modulates apoptosis by reinforcing oxidative metabolism in astrocytes: Role of Bcl-2. J. Biol. Chem. 2012, 287, 10761–10770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Nagasaki, T.; Maeda, H.; Taguchi, K.; Yanagisawa, H.; Nishida, K.; Kobayashi, K.; Wada, N.; Noguchi, I.; Murata, R.; Sakai, H.; et al. A bioinspired carbon monoxide delivery system prevents acute kidney injury and the progression to chronic kidney disease. Redox Biol. 2022, 54, 102371. [Google Scholar] [CrossRef] [PubMed]
  75. Kim, H.J.; Joe, Y.; Rah, S.Y.; Kim, S.K.; Park, S.U.; Park, J.; Kim, J.; Ryu, J.; Cho, G.J.; Surh, Y.J.; et al. Carbon monoxide-induced TFEB nuclear translocation enhances mitophagy/mitochondrial biogenesis in hepatocytes and ameliorates inflammatory liver injury. Cell Death Dis. 2018, 9, 1060. [Google Scholar] [CrossRef] [Green Version]
  76. Li, Y.; Dang, J.; Liang, Q.; Yin, L. Thermal-responsive carbon monoxide (CO) delivery expedites metabolic exhaustion of cancer cells toward reversal of chemotherapy resistance. ACS Cent. Sci. 2019, 5, 1044–1058. [Google Scholar] [CrossRef] [Green Version]
  77. Chen, C.; Garcia-Santos, D.; Ishikawa, Y.; Seguin, A.; Li, L.; Fegan, K.H.; Hildick-Smith, G.J.; Shah, D.I.; Cooney, J.D.; Chen, W.; et al. Snx3 regulates recycling of the transferrin receptor and iron assimilation. Cell. Metab. 2013, 17, 343–352. [Google Scholar] [CrossRef] [Green Version]
  78. Chakraborty, T.; Tóth, Z.; Tóth, R.; Vágvölgyi, C.; Gácser, A. Iron metabolism, pseudohypha production, and biofilm formation through a multicopper oxidase in the human-pathogenic fungus candida parapsilosis. mSphere 2020, 3, e00227-20. [Google Scholar] [CrossRef]
  79. Kitsati, N.; Mantzaris, M.I.D.; Galaris, D. Hydroxytyrosol inhibits hydrogen peroxide-induced apoptotic signaling via labile iron chelation. Redox Biol. 2016, 10, 233–242. [Google Scholar] [CrossRef] [Green Version]
  80. Kerins, M.J.; Ooi, A. The roles of NRF2 in modulating cellular iron homeostasis. Antioxid. Redox Signal. 2018, 29, 1756–1773. [Google Scholar] [CrossRef] [Green Version]
  81. Mancias, J.D.; Wang, X.; Gygi, S.P.; Harper, J.W.; Kimmelman, A.C. Quantitative proteomics identifies NCOA4 as the cargo receptor mediating ferritinophagy. Nature 2014, 509, 105–109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Santana-Codina, N.; Gikandi, A.; Mancias, J.D. The role of NCOA4-mediated ferritinophagy in ferroptosis. Adv. Exp. Med. Biol. 2021, 1301, 41–57. [Google Scholar] [PubMed]
  83. Li, K.; Chen, B.; Xu, A.; Shen, J.; Li, K.; Hao, K.; Hao, R.; Yang, W.; Jiang, W.; Zheng, Y.; et al. TRIM7 modulates NCOA4-mediated ferritinophagy and ferroptosis in glioblastoma cells. Redox Biol. 2022, 56, 102451. [Google Scholar] [CrossRef]
  84. Lee, J.; You, J.; Roh, J.L. Poly(rC)-binding protein 1 represses ferritinophagy-mediated ferroptosis in head and neck cancer. Redox Biol. 2022, 51, 102276. [Google Scholar] [CrossRef] [PubMed]
  85. Tian, Y.; Lu, J.; Hao, X.; Li, H.; Zhang, G.; Liu, X.; Li, X.; Zhao, C.; Kuang, W.; Chen, D.; et al. FTH1 inhibits ferroptosis through ferritinophagy in the 6-OHDA model of Parkinson’s disease. Neurotherapeutics 2020, 17, 1796–1812. [Google Scholar] [CrossRef] [PubMed]
  86. Mesquita, G.; Silva, T.; Gomes, A.C.; Oliveira, P.F.; Alves, M.G.; Fernandes, R.; Almeida, A.A.; Moreira, A.C.; Gomes, M.S. H-Ferritin is essential for macrophages’ capacity to store or detoxify exogenously added iron. Sci. Rep. 2020, 10, 3061. [Google Scholar] [CrossRef] [Green Version]
  87. Jia, F.; Liu, H.; Kang, S. NCOA4-mediated ferritinophagy: A vicious culprit in COVID-19 pathogenesis? Front. Mol. Biosci. 2021, 8, 761793. [Google Scholar] [CrossRef]
  88. Moreira, A.C.; Mesquita, G.; Gomes, M.S. Ferritin: An inflammatory player keeping iron at the core of pathogen-host interactions. Microorganisms 2020, 8, 589. [Google Scholar] [CrossRef] [Green Version]
  89. Bellner, L.; Wolstein, J.; Patil, K.A.; Dunn, M.W.; Laniado-Schwartzman, M. Biliverdin rescues the HO-2 null mouse phenotype of unresolved chronic inflammation following corneal epithelial injury. Invest. Ophthalmol. Vis. Sci. 2011, 52, 3246–3253. [Google Scholar] [CrossRef]
  90. Huang, Y.; Li, J.; Li, W.; Ai, N.; Jin, H. Biliverdin/bilirubin redox pair protects lens epithelial cells against oxidative stress in age-related cataract by regulating NF-B/iNOS and Nrf2/HO-1 pathways. Oxid. Med. Cell Longev. 2022, 2022, 7299182. [Google Scholar] [CrossRef]
  91. Nojima, T.; Obara, T.; Yamamoto, H.; Yumoto, T.; Igawa, T.; Aokage, T.; Seya, M.; Nakao, A.; Naito, H. Luminal administration of biliverdin ameliorates ischemia-reperfusion injury following intestinal transplant in rats. Surgery 2022, 172, 1522–1528. [Google Scholar] [CrossRef] [PubMed]
  92. Li, J.; Jiang, H.; Peng, P.; Zhang, Q.; Bai, W.; Yang, Y.; Huo, S.; Zhou, G.; Shao, J. Biliverdin modulates the long non-coding RNA H19/microRNA-181b-5p/endothelial cell specific molecule 1 axis to alleviate cerebral ischemia reperfusion injury. Biomed. Pharmacother. 2022, 153, 113455. [Google Scholar] [CrossRef]
  93. Zou, Z.Y.; Liu, J.; Chang, C.; Li, J.J.; Luo, J.; Jin, Y.; Ma, Z.; Wang, T.H.; Shao, J.L. Biliverdin administration regulates the microRNA-mRNA expressional network associated with neuroprotection in cerebral ischemia reperfusion injury in rats. Int. J. Mol. Med. 2019, 43, 1356–1372. [Google Scholar] [CrossRef] [Green Version]
  94. Vítek, L.; Tiribelli, C. Bilirubin, intestinal integrity, the microbiome, and inflammation. N. Engl. J. Med. 2020, 383, 684–686. [Google Scholar] [CrossRef]
  95. Lee, Y.; Sugihara, K.; Gillilland, M.G., 3rd; Jon, S.; Kamada, N.; Moon, J.J. Hyaluronic acid-bilirubin nanomedicine for targeted modulation of dysregulated intestinal barrier, microbiome and immune responses in colitis. Nat. Mater. 2020, 19, 118–126. [Google Scholar] [CrossRef]
  96. Hull, T.D.; Agarwal, A. Bilirubin: A potential biomarker and therapeutic target for diabetic nephropathy. Diabetes 2014, 63, 2613–2616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Kim, J.Y.; Lee, D.Y.; Kang, S.; Miao, W.; Kim, H.; Lee, Y.; Jon, S. Bilirubin nanoparticle preconditioning protects against hepatic ischemia-reperfusion injury. Biomaterials 2017, 133, 1–10. [Google Scholar] [CrossRef] [PubMed]
  98. Nitti, M.; Furfaro, A.L.; Mann, G.E. Heme oxygenase dependent bilirubin generation in vascular cells: A role in preventing endothelial dysfunction in local tissue microenvironment? Front. Physiol. 2020, 11, 23. [Google Scholar] [CrossRef]
  99. Mascaró, M.; Alonso, E.N.; Alonso, E.G.; Lacunza, E.; Curino, A.C.; Facchinetti, M.M. Nuclear localization of heme oxygenase-1 in pathophysiological conditions: Does it explain the dual role in cancer? Antioxidants 2021, 10, 87. [Google Scholar] [CrossRef]
  100. Biswas, C.; Shah, N.; Muthu, M.; La, P.; Fernando, A.P.; Sengupta, S.; Yang, G.; Dennery, P.A. Nuclear heme oxygenase-1 (HO-1) modulates subcellular distribution and activation of Nrf2, impacting metabolic and anti-oxidant defenses. J. Biol. Chem. 2014, 289, 26882–26894. [Google Scholar] [CrossRef] [Green Version]
  101. Wang, P.; Zhao, M.; Chen, Z.; Wu, G.; Fujino, M.; Zhang, C.; Zhou, W.; Zhao, M.; Hirano, S.I.; Li, X.K.; et al. Hydrogen gas attenuates hypoxic-ischemic brain injury via regulation of the MAPK/HO-1/PGC-1a pathway in neonatal rats. Oxid. Med. Cell Longev. 2020, 2020, 6978784. [Google Scholar] [CrossRef] [PubMed]
  102. Slebos, D.J.; Ryter, S.W.; van der Toorn, M.; Liu, F.; Guo, F.; Baty, C.J.; Karlsson, J.M.; Watkins, S.C.; Kim, H.P.; Wang, X.; et al. Mitochondrial localization and function of heme oxygenase-1 in cigarette smoke-induced cell death. Am. J. Respir. Cell Mol. Biol. 2007, 36, 409–417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Bindu, S.; Pal, C.; Dey, S.; Goyal, M.; Alam, A.; Iqbal, M.S.; Dutta, S.; Sarkar, S.; Kumar, R.; Maity, P.; et al. Translocation of heme oxygenase-1 to mitochondria is a novel cytoprotective mechanism against non-steroidal anti-inflammatory drug-induced mitochondrial oxidative stress, apoptosis, and gastric mucosal injury. J. Biol. Chem. 2011, 286, 39387–39402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Alam, J.; Cai, J.; Smith, A. Isolation and characterization of the mouse heme oxygenase-1 gene. Distal 5′ sequences are required for induction by heme or heavy metals. J. Biol. Chem. 1994, 269, 1001–1009. [Google Scholar] [CrossRef] [PubMed]
  105. Alam, J.; Camhi, S.; Choi, A.M. Identification of a second region upstream of the mouse heme oxygenase-1 gene that functions as a basal level and inducer-dependent transcription enhancer. J. Biol. Chem. 1995, 270, 11977–11984. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Kwok, S.C.; Samuel, S.P.; Handal, J. Atorvastatin activates heme oxygenase-1 at the stress response elements. J. Cell Mol. Med. 2012, 16, 394–400. [Google Scholar] [CrossRef] [PubMed]
  107. LeBoeuf, S.E.; Wu, W.L.; Karakousi, T.R.; Karadal, B.; Jackson, S.R.; Davidson, S.M.; Wong, K.K.; Koralov, S.B.; Sayin, V.I.; Papagiannakopoulos, T. Activation of oxidative stress response in cancer generates a druggable dependency on exogenous non-essential amino acids. Cell Metab. 2020, 31, 339–350.e4. [Google Scholar] [CrossRef]
  108. Ashino, T.; Yamamoto, M.; Numazawa, S. Nrf2 antioxidative system is involved in cytochrome P450 gene expression and activity: A delay in pentobarbital metabolism in Nrf2-deficient mice. Drug Metab. Dispos. 2020, 48, 673–680. [Google Scholar] [CrossRef]
  109. Bose, C.; Alves, I.; Singh, P.; Palade, P.T.; Carvalho, E.; Børsheim, E.; Jun, S.R.; Cheema, A.; Boerma, M.; Awasthi, S.; et al. Sulforaphane prevents age-associated cardiac and muscular dysfunction through Nrf2 signaling. Aging Cell 2020, 19, e13261. [Google Scholar] [CrossRef] [PubMed]
  110. Huang, X.; Zheng, J.; Li, J.; Che, X.; Tan, W.; Tan, W.; Shao, M.; Cheng, X.; Du, Z.; Zhao, Y.; et al. Functional role of BTB and CNC Homology 1 gene in pancreatic cancer and its association with survival in patients treated with gemcitabine. Theranostics 2018, 8, 3366–3379. [Google Scholar] [CrossRef]
  111. Jiao, Y.; Watts, T.; Xue, J.; Hannafon, B.; Ding, W.Q. Sorafenib and docosahexaenoic acid act in synergy to suppress cancer cell viability: A role of heme oxygenase 1. BMC Cancer 2018, 18, 1042. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Sun, J.; Hoshino, H.; Takaku, K.; Nakajima, O.; Muto, A.; Suzuki, H.; Tashiro, S.; Takahashi, S.; Shibahara, S.; Alam, J.; et al. Hemoprotein Bach1 regulates enhancer availability of heme oxygenase-1 gene. EMBO J. 2002, 21, 5216–5224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Casares, L.; García, V.; Garrido-Rodríguez, M.; Millán, E.; Collado, J.A.; García-Martín, A.; Peñarando, J.; Calzado, M.A.; de la Vega, L.; Muñoz, E. Cannabidiol induces antioxidant pathways in keratinocytes by targeting BACH1. Redox Biol. 2020, 28, 101321. [Google Scholar] [CrossRef] [PubMed]
  114. Dawn, B.; Bolli, R. HO-1 induction by HIF-1: A new mechanism for delayed cardioprotection? Am. J. Physiol. Heart Circ. Physiol. 2005, 289, H522–H524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Shen, H.H.; Wang, C.J.; Zhang, X.Y.; Sheng, Y.R.; Yang, S.L.; Zheng, Z.M.; Shi, J.L.; Qiu, X.M.; Xie, F.; Li, M.Q. HIF1α-induced heme oxygenase 1 promotes the survival of decidual stromal cells against excess heme-mediated oxidative stress. Reproduction 2021, 163, 33–43. [Google Scholar] [CrossRef] [PubMed]
  116. Huang, J.; Guo, P.; Ma, D.; Lin, X.; Fang, Q.; Wang, J. Overexpression of heme oxygenase-1 induced by constitutively activated NF-κB as a potential therapeutic target for activated B-cell-like diffuse large B-cell lymphoma. Int. J. Oncol. 2016, 49, 253–264. [Google Scholar] [CrossRef] [Green Version]
  117. Dey, S.; Sayers, C.M.; Verginadis, I.I.; Lehman, S.L.; Cheng, Y.; Cerniglia, G.J.; Tuttle, S.W.; Feldman, M.D.; Zhang, P.J.; Fuchs, S.Y.; et al. ATF4-dependent induction of heme oxygenase 1 prevents anoikis and promotes metastasis. J. Clin. Invest. 2015, 125, 2592–2608. [Google Scholar] [CrossRef] [Green Version]
  118. Lee, Y.J.; Beak, S.Y.; Choi, I.; Sung, J.S. Quercetin and its metabolites protect hepatocytes against ethanol-induced oxidative stress by activation of Nrf2 and AP-1. Food Sci. Biotechnol. 2018, 27, 809–817. [Google Scholar] [CrossRef]
  119. Chen, C.; Hu, F.; Miao, S.; Sun, L.; Jiao, Y.; Xu, M.; Huang, X.; Yang, Y.; Zhou, R. Transcription factor KLF7 promotes osteoclast differentiation by suppressing HO-1. Front. Genet. 2022, 13, 798433. [Google Scholar] [CrossRef] [PubMed]
  120. Dunn, L.L.; Kong, S.M.Y.; Tumanov, S.; Chen, W.; Cantley, J.; Ayer, A.; Maghzal, G.J.; Midwinter, R.G.; Chan, K.H.; Ng, M.K.C.; et al. Hmox1 (Heme oxygenase-1) protects against ischemia-mediated injury via stabilization of HIF-1α (Hypoxia-inducible factor-1α). Arterioscler. Thromb. Vasc. Biol. 2021, 41, 317–330. [Google Scholar] [CrossRef]
  121. Nakasone, R.; Ashina, M.; Abe, S.; Tanimura, K.; Van Rostenberghe, H.; Fujioka, K. The role of heme oxygenase-1 promoter polymorphisms in perinatal disease. Int. J. Environ. Res. Public Health 2021, 18, 3520. [Google Scholar] [CrossRef] [PubMed]
  122. Exner, M.; Minar, E.; Wagner, O.; Schillinger, M. The role of heme oxygenase-1 promoter polymorphisms in human disease. Free Radic. Biol. Med. 2004, 37, 1097–1104. [Google Scholar] [CrossRef] [PubMed]
  123. Gill, A.J.; Garza, R.; Ambegaokar, S.S.; Gelman, B.B.; Kolson, D.L. Heme oxygenase-1 promoter region (GT)n polymorphism associates with increased neuroimmune activation and risk for encephalitis in HIV infection. J. Neuroinflamm. 2018, 15, 70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Kaplan, M.; Wong, R.J.; Stevenson, D.K. Heme oxygenase-1 promoter polymorphisms: Do they modulate neonatal hyperbilirubinemia? J. Perinatol. 2017, 37, 901–905. [Google Scholar] [CrossRef]
  125. Bao, W.; Song, F.; Li, X.; Rong, S.; Yang, W.; Wang, D.; Xu, J.; Fu, J.; Zhao, Y.; Liu, L. Association between heme oxygenase-1 gene promoter polymorphisms and type 2 diabetes mellitus: A HuGE review and meta-analysis. Am. J. Epidemiol. 2010, 172, 631–636. [Google Scholar] [CrossRef] [Green Version]
  126. Ono, K.; Mannami, T.; Iwai, N. Association of a promoter variant of the haeme oxygenase-1 gene with hypertension in women. J. Hypertens. 2003, 21, 1497–1503. [Google Scholar] [CrossRef]
  127. Pulkkinen, K.H.; Ylä-Herttuala, S.; Levonen, A.-L. Heme oxygenase 1 is induced by miR-155 via reduced BACH1 translation in endothelial cells. Free Radic. Biol. Med. 2011, 51, 2124–2131. [Google Scholar] [CrossRef]
  128. Gu, S.; Lai, Y.; Chen, H.; Liu, Y.; Zhang, Z. miR-155 mediates arsenic trioxide resistance by activating Nrf2 and suppressing apoptosis in lung cancer cells. Sci. Rep. 2017, 7, 12155. [Google Scholar] [CrossRef] [Green Version]
  129. Zhang, J.; Vandevenne, P.; Hamdi, H.; Van Puyvelde, M.; Zucchi, A.; Bettonville, M.; Weatherly, K.; Braun, M.Y. Micro-RNA-155-mediated control of heme oxygenase 1 (HO-1) is required for restoring adaptively tolerant CD4+ T-cell function in rodents. Eur. J. Immunol. 2015, 45, 829–842. [Google Scholar] [CrossRef]
  130. Li, H.; Di, G.; Zhang, Y.; Xue, R.; Zhang, J.; Liang, J. MicroRNA-155 and microRNA-181a, via HO-1, participate in regulating the immunotoxicity of cadmium in the kidneys of exposed Cyprinus carpio. Fish Shellfish Immunol. 2019, 95, 473–480. [Google Scholar] [CrossRef]
  131. Piras, S.; Furfaro, A.L.; Caggiano, R.; Brondolo, L.; Garibaldi, S.; Ivaldo, C.; Marinari, U.M.; Pronzato, M.A.; Faraonio, R.; Nitti, M. microRNA-494 favors HO-1 expression in neuroblastoma cells exposed to oxidative stress in a Bach1-independent way. Front. Oncol. 2018, 8, 199. [Google Scholar] [CrossRef] [Green Version]
  132. Pu, M.; Li, C.; Qi, X.; Chen, J.; Wang, Y.; Gao, L.; Miao, L.; Ren, J. MiR-1254 suppresses HO-1 expression through seed region-dependent silencing and non-seed interaction with TFAP2A transcript to attenuate NSCLC growth. PLoS Genet. 2017, 13, e1006896. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Yang, Z.; Chen, J.S.; Wen, J.K.; Gao, H.T.; Zheng, B.; Qu, C.B.; Liu, K.L.; Zhang, M.L.; Gu, J.F.; Li, J.D.; et al. Silencing of miR-193a-5p increases the chemosensitivity of prostate cancer cells to docetaxel. J. Exp. Clin. Cancer Res. 2017, 36, 178. [Google Scholar] [CrossRef] [Green Version]
  134. Beckman, J.D.; Chen, C.; Nguyen, J.; Thayanithy, V.; Subramanian, S.; Steer, C.J.; Vercellotti, G.M. Regulation of heme oxygenase-1 protein expression by miR-377 in combination with miR-217. J. Biol. Chem. 2011, 286, 3194–3202. [Google Scholar] [CrossRef] [Green Version]
  135. Nowis, D.; Legat, M.; Grzela, T.; Niderla, J.; Wilczek, E.; Wilczynski, G.M.; Głodkowska, E.; Mrówka, P.; Issat, T.; Dulak, J.; et al. Heme oxygenase-1 protects tumor cells against photodynamic therapy-mediated cytotoxicity. Oncogene 2006, 25, 3365–3374. [Google Scholar] [CrossRef] [Green Version]
  136. Neuzil, J.; Stocker, R. Bilirubin attenuates radical-mediated damage to serum albumin. FEBS Lett. 1993, 331, 281–284. [Google Scholar] [CrossRef] [Green Version]
  137. Liu, Z.; Liu, J.; Liu, X.; Wang, X.; Xie, Q.; Zhang, X.; Kong, X.; He, M.; Yang, Y.; Deng, X.; et al. CTR-DB, an omnibus for patient-derived gene expression signatures correlated with cancer drug response. Nucleic Acids Res. 2022, 50, D1184–D1199. [Google Scholar] [CrossRef]
  138. Wang, X.; Ye, T.; Xue, B.; Yang, M.; Li, R.; Xu, X.; Zeng, X.; Tian, N.; Bao, L.; Huang, Y. Mitochondrial GRIM-19 deficiency facilitates gastric cancer metastasis through oncogenic ROS-Nrf2-HO-1 axis via a Nrf2-HO-1 loop. Gastric Cancer 2021, 24, 117–132. [Google Scholar] [CrossRef] [PubMed]
  139. Anselmino, N.; Starbuck, M.; Labanca, E.; Cotignola, J.; Navone, N.; Gueron, G.; Zenclussen, A.C.; Vazquez, E. Heme oxygenase-1 Is a pivotal modulator of bone turnover and remodeling: Molecular implications for prostate cancer bone metastasis. Antioxid. Redox Signal. 2020, 32, 1243–1258. [Google Scholar] [CrossRef]
  140. Zhang, Y.; Chen, B.; Xu, N.; Xu, P.; Lin, W.; Liu, C.; Huang, P. Exosomes promote the transition of androgen-dependent prostate cancer cells into androgen-independent manner through up-regulating the heme oxygenase-1. Int. J. Nanomed. 2021, 16, 315–327. [Google Scholar] [CrossRef] [PubMed]
  141. Lin, S.; Zhang, Q.; Li, S.; Zhang, T.; Wang, L.; Qin, X.; Zhang, M.; Shi, S.; Cai, X. Antioxidative and angiogenesis-promoting effects of tetrahedral framework nucleic acids in diabetic wound healing with activation of the Akt/Nrf2/HO-1 pathway. ACS Appl. Mater. Interfaces 2020, 12, 11397–11408. [Google Scholar] [CrossRef]
  142. Cheng, C.C.; Guan, S.S.; Yang, H.J.; Chang, C.C.; Luo, T.Y.; Chang, J.; Ho, A.S. Blocking heme oxygenase-1 by zinc protoporphyrin reduces tumor hypoxia-mediated VEGF release and inhibits tumor angiogenesis as a potential therapeutic agent against colorectal cancer. J. Biomed. Sci. 2016, 23, 18. [Google Scholar] [CrossRef] [Green Version]
  143. Deshane, J.; Chen, S.; Caballero, S.; Grochot-Przeczek, A.; Was, H.; Li Calzi, S.; Lach, R.; Hock, T.D.; Chen, B.; Hill-Kapturczak, N.; et al. Stromal cell-derived factor 1 promotes angiogenesis via a heme oxygenase 1-dependent mechanism. J. Exp. Med. 2007, 204, 605–618. [Google Scholar] [CrossRef] [PubMed]
  144. Silva, R.C.M.C.; Vasconcelos, L.R.; Travassos, L.H. The different facets of heme-oxygenase 1 in innate and adaptive immunity. Cell Biochem. Biophys. 2022, 80, 609–631. [Google Scholar] [CrossRef]
  145. De Wilde, V.; Van Rompaey, N.; Hill, M.; Lebrun, J.F.; Lemaître, P.; Lhommé, F.; Kubjak, C.; Vokaer, B.; Oldenhove, G.; Charbonnier, L.M.; et al. Endotoxin-induced myeloid-derived suppressor cells inhibit alloimmune responses via heme oxygenase-1. Am. J. Transplant. 2009, 9, 2034–2047. [Google Scholar] [CrossRef] [PubMed]
  146. Alaluf, E.; Vokaer, B.; Detavernier, A.; Azouz, A.; Splittgerber, M.; Carrette, A.; Boon, L.; Libert, F.; Soares, M.; Le Moine, A.; et al. Heme oxygenase-1 orchestrates the immunosuppressive program of tumor-associated macrophages. JCI Insight 2020, 5, e133929. [Google Scholar] [CrossRef] [PubMed]
  147. Khojandi, N.; Kuehm, L.M.; Piening, A.; Donlin, M.J.; Hsueh, E.C.; Schwartz, T.L.; Farrell, K.; Richart, J.M.; Geerling, E.; Pinto, A.K.; et al. Oxidized lipoproteins promote resistance to cancer immunotherapy independent of patient obesity. Cancer Immunol. Res. 2021, 9, 214–226. [Google Scholar] [CrossRef] [PubMed]
  148. George, J.F.; Braun, A.; Brusko, T.M.; Joseph, R.; Bolisetty, S.; Wasserfall, C.H.; Atkinson, M.A.; Agarwal, A.; Kapturczak, M.H. Suppression by CD4+CD25+ regulatory T cells is dependent on expression of heme oxygenase-1 in antigen-presenting cells. Am. J. Pathol. 2008, 173, 154–160. [Google Scholar] [CrossRef] [Green Version]
  149. Chabannes, D.; Hill, M.; Merieau, E.; Rossignol, J.; Brion, R.; Soulillou, J.P.; Anegon, I.; Cuturi, M.C. A role for heme oxygenase-1 in the immunosuppressive effect of adult rat and human mesenchymal stem cells. Blood 2007, 110, 3691–3694. [Google Scholar] [CrossRef] [Green Version]
  150. Yoon, S.J.; Kim, S.J.; Lee, S.M. Overexpression of HO-1 contributes to sepsis-induced immunosuppression by modulating the Th1/Th2 balance and regulatory T-cell function. J. Infect. Dis. 2017, 215, 1608–1618. [Google Scholar] [CrossRef]
  151. Deng, R.; Wang, S.M.; Yin, T.; Ye, T.H.; Shen, G.B.; Li, L.; Zhao, J.Y.; Sang, Y.X.; Duan, X.G.; Wei, Y.Q. Inhibition of tumor growth and alteration of associated macrophage cell type by an HO-1 inhibitor in breast carcinoma-bearing mice. Oncol. Res. 2013, 20, 473–482. [Google Scholar] [CrossRef] [PubMed]
  152. Magri, S.; Musca, B.; Pinton, L.; Orecchini, E.; Belladonna, M.L.; Orabona, C.; Bonaudo, C.; Volpin, F.; Ciccarino, P.; Baro, V.; et al. The immunosuppression pathway of tumor-associated macrophages is controlled by heme oxygenase-1 in glioblastoma patients. Int. J. Cancer 2022, 151, 2265–2277. [Google Scholar] [CrossRef] [PubMed]
  153. Gómez-Lomelí, P.; Bravo-Cuellar, A.; Hernández-Flores, G.; Jave-Suárez, L.F.; Aguilar-Lemarroy, A.; Lerma-Díaz, J.M.; Domínguez-Rodríguez, J.R.; Sánchez-Reyes, K.; Ortiz-Lazareno, P.C. Increase of IFN-γ and TNF-α production in CD107a + NK-92 cells co-cultured with cervical cancer cell lines pre-treated with the HO-1 inhibitor. Cancer Cell Int. 2014, 14, 100. [Google Scholar] [CrossRef] [Green Version]
  154. Furfaro, A.L.; Ottonello, S.; Loi, G.; Cossu, I.; Piras, S.; Spagnolo, F.; Queirolo, P.; Marinari, U.M.; Moretta, L.; Pronzato, M.A.; et al. HO-1 downregulation favors BRAF melanoma cell death induced by Vemurafenib/PLX4032 and increases NK recognition. Int. J. Cancer 2020, 146, 1950–1962. [Google Scholar] [CrossRef] [PubMed]
  155. Fang, J.; Sawa, T.; Akaike, T.; Akuta, T.; Sahoo, S.K.; Khaled, G.; Hamada, A.; Maeda, H. In vivo antitumor activity of pegylated zinc protoporphyrin: Targeted inhibition of heme oxygenase in solid tumor. Cancer Res. 2003, 63, 3567–3574. [Google Scholar]
  156. Pittalà, V.; Salerno, L.; Romeo, G.; Modica, M.N.; Siracusa, M.A. A focus on heme oxygenase-1 (HO-1) inhibitors. Curr. Med. Chem. 2013, 20, 3711–3732. [Google Scholar] [CrossRef]
  157. Floresta, G.; Amata, E.; Gentile, D.; Romeo, G.; Marrazzo, A.; Pittalà, V.; Salerno, L.; Rescifina, A. Fourfold filtered statistical/computational approach for the identification of imidazole compounds as HO-1 inhibitors from natural products. Mar. Drugs 2019, 17, 113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Kinobe, R.T.; Dercho, R.A.; Vlahakis, J.Z.; Brien, J.F.; Szarek, W.A.; Nakatsu, K. Inhibition of the enzymatic activity of heme oxygenases by azole-based antifungal drugs. J. Pharmacol. Exp. Ther. 2006, 319, 277–284. [Google Scholar] [CrossRef] [Green Version]
  159. Trump, D.L.; Havlin, K.H.; Messing, E.M.; Cummings, K.B.; Lange, P.H.; Jordan, V.C. High-dose ketoconazole in advanced hormone-refractory prostate cancer: Endocrinologic and clinical effects. J. Clin. Oncol. 1989, 7, 1093–1098. [Google Scholar] [CrossRef]
  160. Patel, V.; Liaw, B.; Oh, W. The role of ketoconazole in current prostate cancer care. Nat. Rev. Urol. 2018, 15, 643–651. [Google Scholar] [CrossRef]
  161. Cao, T.H.; Jin, S.G.; Fei, D.S.; Kang, K.; Jiang, L.; Lian, Z.Y.; Pan, S.H.; Zhao, M.R.; Zhao, M.Y. Artesunate protects against sepsis-induced lung injury via heme oxygenase-1 modulation. Inflammation 2016, 39, 651–662. [Google Scholar] [CrossRef]
  162. Jiang, Z.; Wang, Z.; Chen, L.; Zhang, C.; Liao, F.; Wang, Y.; Wang, Y.; Luo, P.; Luo, M.; Shi, C. Artesunate induces ER-derived-ROS-mediated cell death by disrupting labile iron pool and iron redistribution in hepatocellular carcinoma cells. Am. J. Cancer Res. 2021, 11, 691–711. [Google Scholar] [PubMed]
  163. Su, Y.; Zhao, B.; Zhou, L.; Zhang, Z.; Shen, Y.; Lv, H.; AlQudsy, L.H.H.; Shang, P. Ferroptosis, a novel pharmacological mechanism of anti-cancer drugs. Cancer Lett. 2020, 483, 127–136. [Google Scholar] [CrossRef] [PubMed]
  164. Li, G.; Ling, M.; Yu, K.; Yang, W.; Liu, Q.; He, L.; Cai, X.; Zhong, M.; Mai, Z.; Sun, R.; et al. Synergetic delivery of artesunate and isosorbide 5-mononitrate with reduction-sensitive polymer nanoparticles for ovarian cancer chemotherapy. J. Nanobiotechnol. 2022, 20, 471. [Google Scholar] [CrossRef]
  165. Chen, W.; Ma, Z.; Yu, L.; Mao, X.; Ma, N.; Guo, X.; Yin, X.; Jiang, F.; Wang, Q.; Wang, J.; et al. Preclinical investigation of artesunate as a therapeutic agent for hepatocellular carcinoma via impairment of glucosylceramidase-mediated autophagic degradation. Exp. Mol. Med. 2022, 54, 1536–1548. [Google Scholar] [CrossRef]
  166. Markowitsch, S.D.; Schupp, P.; Lauckner, J.; Vakhrusheva, O.; Slade, K.S.; Mager, R.; Efferth, T.; Haferkamp, A.; Juengel, E. Artesunate inhibits growth of sunitinib-resistant renal cell carcinoma cells through cell cycle arrest and induction of ferroptosis. Cancers 2020, 12, 3150. [Google Scholar] [CrossRef] [PubMed]
  167. Pirali, M.; Taheri, M.; Zarei, S.; Majidi, M.; Ghafouri, H. Artesunate, as a HSP70 ATPase activity inhibitor, induces apoptosis in breast cancer cells. Int. J. Biol. Macromol. 2020, 164, 3369–3375. [Google Scholar] [CrossRef] [PubMed]
  168. Krishna, S.; Ganapathi, S.; Ster, I.C.; Saeed, M.E.; Cowan, M.; Finlayson, C.; Kovacsevics, H.; Jansen, H.; Kremsner, P.G.; Efferth, T.; et al. A randomised, double blind, placebo-controlled pilot study of oral artesunate therapy for colorectal cancer. EBioMedicine 2014, 2, 82–90. [Google Scholar] [CrossRef] [Green Version]
  169. Sun, X.; Wang, S.; Gai, J.; Guan, J.; Li, J.; Li, Y.; Zhao, J.; Zhao, C.; Fu, L.; Li, Q. SIRT5 promotes cisplatin resistance in ovarian cancer by suppressing dna damage in a ROS-dependent manner via regulation of the Nrf2/HO-1 pathway. Front. Oncol. 2019, 9, 754. [Google Scholar] [CrossRef] [Green Version]
  170. Kannan, S.; Irwin, M.E.; Herbrich, S.M.; Cheng, T.; Patterson, L.L.; Aitken, M.J.L.; Bhalla, K.; You, M.J.; Konopleva, M.; Zweidler-McKay, P.A.; et al. Targeting the NRF2/HO-1 antioxidant pathway in FLT3-ITD-positive AML enhances therapy efficacy. Antioxidants 2022, 11, 717. [Google Scholar] [CrossRef]
  171. Wen, L.; Yang, S.; Li, P.; Chen, R.; Wang, Q.; Kaspo, B.; Fan, H.; Hu, J. iASSP-mediated ROS inhibition drives 5-Fu resistance dependent on Nrf2 antioxidative signaling pathway in gastric adenocarcinoma. Dig. Dis. Sci. 2020, 65, 2873–2883. [Google Scholar] [CrossRef] [PubMed]
  172. Bao, L.J.; Jaramillo, M.C.; Zhang, Z.B.; Zheng, Y.X.; Yao, M.; Zhang, D.D.; Yi, X.F. Nrf2 induces cisplatin resistance through activation of autophagy in ovarian carcinoma. Int. J. Clin. Exp. Pathol. 2014, 7, 1502–1513. [Google Scholar] [PubMed]
  173. Bao, L.; Wu, J.; Dodson, M.; Rojo de la Vega, E.M.; Ning, Y.; Zhang, Z.; Yao, M.; Zhang, D.D.; Xu, C.; Yi, X. ABCF2, an Nrf2 target gene, contributes to cisplatin resistance in ovarian cancer cells. Mol. Carcinog. 2017, 56, 1543–1553. [Google Scholar] [CrossRef]
  174. Wu, J.; Bao, L.; Zhang, Z.; Yi, X. Nrf2 induces cisplatin resistance via suppressing the iron export related gene SLC40A1 in ovarian cancer cells. Oncotarget 2017, 8, 93502–93515. [Google Scholar] [CrossRef] [Green Version]
  175. Lv, X.; Song, D.M.; Niu, Y.H.; Wang, B.S. Inhibition of heme oxygenase-1 enhances the chemosensitivity of laryngeal squamous cell cancer Hep-2 cells to cisplatin. Apoptosis 2016, 21, 489–501. [Google Scholar] [CrossRef]
  176. Ríos-Arrabal, S.; Puentes-Pardo, J.D.; Moreno-SanJuan, S.; Szuba, Á.; Casado, J.; García-Costela, M.; Escudero-Feliu, J.; Verbeni, M.; Cano, C.; González-Puga, C.; et al. Endothelin-1 as a mediator of heme oxygenase-1-induced stemness in colorectal cancer: Influence of p53. J. Pers. Med. 2021, 11, 509. [Google Scholar] [CrossRef]
  177. Lippmann, J.; Petri, K.; Fulda, S.; Liese, J. Redox modulation and induction of ferroptosis as a new therapeutic strategy in hepatocellular carcinoma. Transl. Oncol. 2020, 13, 100785. [Google Scholar] [CrossRef] [PubMed]
  178. Buranrat, B.; Connor, J.R. Cytoprotective effects of ferritin on doxorubicin-induced breast cancer cell death. Oncol. Rep. 2015, 34, 2790–2796. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Salatino, A.; Aversa, I.; Battaglia, A.M.; Sacco, A.; Di Vito, A.; Santamaria, G.; Chirillo, R.; Veltri, P.; Tradigo, G.; Di Cello, A.; et al. H-Ferritin affects cisplatin-induced cytotoxicity in ovarian cancer cells through the modulation of ROS. Oxid. Med. Cell Longev. 2019, 2019, 3461251. [Google Scholar] [CrossRef]
  180. Tang, Z.; Ju, Y.; Dai, X.; Ni, N.; Liu, Y.; Zhang, D.; Gao, H.; Sun, H.; Zhang, J.; Gu, P. HO-1-mediated ferroptosis as a target for protection against retinal pigment epithelium degeneration. Redox Biol. 2021, 43, 101971. [Google Scholar] [CrossRef]
  181. Yin, W.; Wang, C.; Peng, Y.; Yuan, W.; Zhang, Z.; Liu, H.; Xia, Z.; Ren, C.; Qian, J. Dexmedetomidine alleviates H2O2-induced oxidative stress and cell necroptosis through activating of α2-adrenoceptor in H9C2 cells. Mol. Biol. Rep. 2020, 47, 3629–3639. [Google Scholar] [CrossRef] [PubMed]
  182. Li, H.B.; Zhang, X.Z.; Sun, Y.; Zhou, Q.; Song, J.N.; Hu, Z.F.; Li, Y.; Wu, J.N.; Guo, Y.; Zhang, Y.; et al. HO-1/PINK1 regulated mitochondrial fusion/fission to inhibit pyroptosis and attenuate septic acute kidney injury. Biomed. Res. Int. 2020, 2020, 2148706. [Google Scholar] [CrossRef] [PubMed]
  183. Song, W.; Cressatti, M.; Zukor, H.; Liberman, A.; Galindez, C.; Schipper, H.M. Parkinsonian features in aging GFAP.HMOX1 transgenic mice overexpressing human HO-1 in the astroglial compartment. Neurobiol. Aging 2017, 58, 163–179. [Google Scholar] [CrossRef] [PubMed]
  184. Chiang, S.K.; Chen, S.E.; Chang, L.C. The role of HO-1 and its crosstalk with oxidative stress in cancer cell survival. Cells 2021, 10, 2401. [Google Scholar] [CrossRef]
  185. Yang, S.; Ouyang, J.; Lu, Y.; Harypursat, V.; Chen, Y. A dual role of heme oxygenase-1 in tuberculosis. Front. Immunol. 2022, 13, 842858. [Google Scholar] [CrossRef]
  186. Chang, L.C.; Chiang, S.K.; Chen, S.E.; Yu, Y.L.; Chou, R.H.; Chang, W.C. Heme oxygenase-1 mediates BAY 11-7085 induced ferroptosis. Cancer Lett. 2018, 416, 124–137. [Google Scholar] [CrossRef]
  187. NaveenKumar, S.K.; SharathBabu, B.N.; Hemshekhar, M.; Kemparaju, K.; Girish, K.S.; Mugesh, G. The role of reactive oxygen species and ferroptosis in heme-mediated activation of human platelets. ACS Chem. Biol. 2018, 13, 1996–2002. [Google Scholar] [CrossRef]
  188. Perillo, B.; Di Donato, M.; Pezone, A.; Di Zazzo, E.; Giovannelli, P.; Galasso, G.; Castoria, G.; Migliaccio, A. ROS in cancer therapy: The bright side of the moon. Exp. Mol. Med. 2020, 52, 192–203. [Google Scholar] [CrossRef]
  189. Mohammad, R.M.; Muqbil, I.; Lowe, L.; Yedjou, C.; Hsu, H.Y.; Lin, L.T.; Siegelin, M.D.; Fimognari, C.; Kumar, N.B.; Dou, Q.P.; et al. Broad targeting of resistance to apoptosis in cancer. Semin. Cancer Biol. 2015, 35 (Suppl. S0), S78–S103. [Google Scholar] [CrossRef] [PubMed]
  190. Dietrich, J.B. Apoptosis and anti-apoptosis genes in the Bcl-2 family. Arch. Physiol. Biochem. 1997, 105, 125–135. [Google Scholar] [CrossRef]
  191. Zhang, L.; Zhang, S. Modulating Bcl-2 family proteins and caspase-3 in induction of apoptosis by paeoniflorin in human cervical cancer cells. Phytother. Res. 2011, 25, 1551–1557. [Google Scholar] [CrossRef] [PubMed]
  192. Barbagallo, I.; Giallongo, C.; Volti, G.L.; Distefano, A.; Camiolo, G.; Raffaele, M.; Salerno, L.; Pittalà, V.; Sorrenti, V.; Avola, R.; et al. Heme oxygenase inhibition sensitizes neuroblastoma cells to carfilzomib. Mol. Neurobiol. 2019, 56, 1451–1460. [Google Scholar] [CrossRef] [PubMed]
  193. Chakraborty, S.; Balan, M.; Flynn, E.; Zurakowski, D.; Choueiri, T.K.; Pal, S. Activation of c-Met in cancer cells mediates growth-promoting signals against oxidative stress through Nrf2-HO-1. Oncogenesis 2019, 8, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Abdalla, M.Y.; Ahmad, I.M.; Rachagani, S.; Banerjee, K.; Thompson, C.M.; Maurer, H.C.; Olive, K.P.; Bailey, K.L.; Britigan, B.E.; Kumar, S. Enhancing responsiveness of pancreatic cancer cells to gemcitabine treatment under hypoxia by heme oxygenase-1 inhibition. Transl. Res. 2019, 207, 56–69. [Google Scholar] [CrossRef] [PubMed]
  195. Huang, Y.; Ma, T.; Ye, Z.; Li, H.; Zhao, Y.; Chen, W.; Fu, Y.; Ye, Z.; Sun, A.; Li, Z. Carbon monoxide (CO) inhibits hydrogen peroxide (H2O2)-induced oxidative stress and the activation of NF-κB signaling in lens epithelial cells. Exp. Eye Res. 2018, 166, 29–39. [Google Scholar] [CrossRef]
  196. Scandura, G.; Giallongo, C.; Puglisi, F.; Romano, A.; Parrinello, N.L.; Zuppelli, T.; Longhitano, L.; Giallongo, S.; Di Rosa, M.; Musumeci, G.; et al. TLR4 signaling and heme oxygenase-1/carbon monoxide pathway crosstalk induces resiliency of myeloma plasma cells to bortezomib treatment. Antioxidants 2022, 11, 767. [Google Scholar] [CrossRef]
  197. Wang, J.; Zhang, J.; Lee, Y.M.; Koh, P.L.; Ng, S.; Bao, F.; Lin, Q.; Shen, H.M. Quantitative chemical proteomics profiling of de novo protein synthesis during starvation-mediated autophagy. Autophagy 2016, 12, 1931–1944. [Google Scholar] [CrossRef] [Green Version]
  198. Musial, C.; Siedlecka-Kroplewska, K.; Kmiec, Z.; Gorska-Ponikowska, M. Modulation of autophagy in cancer cells by dietary polyphenols. Antioxidants 2021, 10, 123. [Google Scholar] [CrossRef]
  199. Chen, M.H.; Zheng, Y.; Cai, X.J.; Zhang, H.; Wang, F.X.; Tan, C.P.; Chen, W.H.; Ji, L.N.; Mao, Z.W. Inhibition of autophagic flux by cyclometalated iridium(iii) complexes through anion transportation. Chem. Sci. 2019, 10, 3315–3323. [Google Scholar] [CrossRef] [Green Version]
  200. Zhang, P.; Zheng, Z.; Ling, L.; Yang, X.; Zhang, N.; Wang, X.; Hu, M.; Xia, Y.; Ma, Y.; Yang, H.; et al. w09, a novel autophagy enhancer, induces autophagy-dependent cell apoptosis via activation of the EGFR-mediated RAS-RAF1-MAP2K-MAPK1/3 pathway. Autophagy 2017, 13, 1093–1112. [Google Scholar] [CrossRef]
  201. Banerjee, P.; Basu, A.; Wegiel, B.; Otterbein, L.E.; Mizumura, K.; Gasser, M.; Waaga-Gasser, A.M.; Choi, A.M.; Pal, S. Heme oxygenase-1 promotes survival of renal cancer cells through modulation of apoptosis- and autophagy-regulating molecules. J. Biol. Chem. 2012, 287, 32113–32123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Tracey, N.; Creedon, H.; Kemp, A.J.; Culley, J.; Muir, M.; Klinowska, T.; Brunton, V.G. HO-1 drives autophagy as a mechanism of resistance against HER2-targeted therapies. Breast Cancer Res. Treat. 2020, 179, 543–555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Cao, L.; Wang, J.; Ma, D.; Wang, P.; Zhang, Y.; Fang, Q. Heme oxygenase-1 contributes to imatinib resistance by promoting autophagy in chronic myeloid leukemia through disrupting the mTOR signaling pathway. Biomed. Pharmacother. 2016, 78, 30–38. [Google Scholar] [CrossRef] [PubMed]
  204. Feng, M.; Wang, J.; Sun, M.; Li, G.; Li, B.; Zhang, H. 3-Methyladenine but not antioxidants to overcome BACH2-mediated bortezomib resistance in mantle cell lymphoma. Cancer Cell Int. 2021, 21, 279. [Google Scholar] [CrossRef]
  205. Zou, L.; Lei, H.; Shen, J.; Liu, X.; Zhang, X.; Wu, L.; Hao, J.; Jiang, W.; Hu, Z. HO-1 induced autophagy protects against IL-1 β-mediated apoptosis in human nucleus pulposus cells by inhibiting NF-κB. Aging 2020, 12, 2440–2452. [Google Scholar] [CrossRef]
  206. Fan, W.X.; Wen, X.L.; Xiao, H.; Yang, Q.P.; Liang, Z. MicroRNA-29a enhances autophagy in podocytes as a protective mechanism against high glucose-induced apoptosis by targeting heme oxygenase-1. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 8909–8917. [Google Scholar]
  207. Peng, Z.; Liao, Y.; Wang, X.; Chen, L.; Wang, L.; Qin, C.; Wang, Z.; Cai, M.; Hu, J.; Li, D.; et al. Heme oxygenase-1 regulates autophagy through carbon-oxygen to alleviate deoxynivalenol-induced hepatic damage. Arch. Toxicol. 2020, 94, 573–588. [Google Scholar] [CrossRef]
  208. Yang, Y.; Wang, H.X.; Zhang, L.; Huo, W.; Li, X.D.; Qi, R.Q.; Song, X.Y.; Wei, S.; Gao, X.H.; Han, S.; et al. Inhibition of heme oxygenase-1 enhances hyperthermia-induced autophagy and antiviral effect. Int. J. Biol. Sci. 2019, 15, 568–578. [Google Scholar] [CrossRef] [Green Version]
  209. Jacoby, E.; Ben Yakir-Blumkin, M.; Blumenfeld-Kan, S.; Brody, Y.; Meir, A.; Melamed-Book, N.; Napso, T.; Pozner, G.; Saadi, E.; Shabtay-Orbach, A.; et al. Mitochondrial augmentation of CD34 cells from healthy donors and patients with mitochondrial DNA disorders confers functional benefit. NPJ Regen. Med. 2021, 6, 58. [Google Scholar] [CrossRef]
  210. Meyer, N.; Zielke, S.; Michaelis, J.B.; Linder, B.; Warnsmann, V.; Rakel, S.; Osiewacz, H.D.; Fulda, S.; Mittelbronn, M.; Münch, C.; et al. AT 101 induces early mitochondrial dysfunction and HMOX1 (heme oxygenase 1) to trigger mitophagic cell death in glioma cells. Autophagy 2018, 14, 1693–1709. [Google Scholar] [CrossRef]
  211. Suliman, H.B.; Keenan, J.E.; Piantadosi, C.A. Mitochondrial quality-control dysregulation in conditional HO-1 mice. JCI Insight 2017, 2, e89676. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Guan, X.; Shen, J.; Xu, Y.; Feng, X.; Zhou, R. Heme oxygenase-1 enhances autophagy by modulating the AMPK/mTORC1 signaling pathway as a renoprotective mechanism to mitigate lead-induced nephrotoxicity. Am. J. Transl. Res. 2020, 12, 4807–4818. [Google Scholar]
  213. Han, Q.; Li, G.; Ip, M.S.; Zhang, Y.; Zhen, Z.; Mak, J.C.; Zhang, N. Haemin attenuates intermittent hypoxia-induced cardiac injury via inhibiting mitochondrial fission. J. Cell Mol. Med. 2018, 22, 2717–2726. [Google Scholar] [CrossRef] [PubMed]
  214. Hull, T.D.; Boddu, R.; Guo, L.; Tisher, C.C.; Traylor, A.M.; Patel, B.; Joseph, R.; Prabhu, S.D.; Suliman, H.B.; Piantadosi, C.A.; et al. Heme oxygenase-1 regulates mitochondrial quality control in the heart. JCI Insight 2016, 1, e85817. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The regulatory mechanism of HO-1. Upon stressful stimuli, Nrf2 dissociates form the Keap1–Nrf2 complex and translocates into nucleus, and then Nrf2 binds with small Maf proteins to stress response elements (StRE) in the regulatory regions of HMOX1, leading to the activation of HO-1. BACH1 acts as a negative regulator of Nrf2 through competing with Nrf2 for binding to Maf proteins, which inhibits HO-1 expression. Several other important transcription factors are also implicated in the modulation of HO-1 expression including HIF1α [114,115], NF-κB [116], ATF4 [117], AP-1 [118], and KLF7 [119]. Among these transcription factors, HIF1α demonstrate a complex crosstalk with HO-1 [120]. Genetic polymorphisms in the promoter region of HMOX1 including (GT)n repeat and T(−413)A SNP are also involved in the transcriptional regulation HO-1 [121,122,123,124,125,126]. MiR-155 displayed opposite roles in HO-1 regulation. On the one hand, miR-155 activates HO-1 expression partially via repressing BACH1 [127,128], while on the other hand, miR-155 alone or in cooperation with miR-181a inhibits HO-1 expression [129,130]. Furthermore, it has been reported that miR-494 and miR-193a-5p enhanced HO-1 expression, wihle miR-1254 and miR-217 combined with miR-377 suppress the activity of HO-1 [131,132,133,134].
Figure 1. The regulatory mechanism of HO-1. Upon stressful stimuli, Nrf2 dissociates form the Keap1–Nrf2 complex and translocates into nucleus, and then Nrf2 binds with small Maf proteins to stress response elements (StRE) in the regulatory regions of HMOX1, leading to the activation of HO-1. BACH1 acts as a negative regulator of Nrf2 through competing with Nrf2 for binding to Maf proteins, which inhibits HO-1 expression. Several other important transcription factors are also implicated in the modulation of HO-1 expression including HIF1α [114,115], NF-κB [116], ATF4 [117], AP-1 [118], and KLF7 [119]. Among these transcription factors, HIF1α demonstrate a complex crosstalk with HO-1 [120]. Genetic polymorphisms in the promoter region of HMOX1 including (GT)n repeat and T(−413)A SNP are also involved in the transcriptional regulation HO-1 [121,122,123,124,125,126]. MiR-155 displayed opposite roles in HO-1 regulation. On the one hand, miR-155 activates HO-1 expression partially via repressing BACH1 [127,128], while on the other hand, miR-155 alone or in cooperation with miR-181a inhibits HO-1 expression [129,130]. Furthermore, it has been reported that miR-494 and miR-193a-5p enhanced HO-1 expression, wihle miR-1254 and miR-217 combined with miR-377 suppress the activity of HO-1 [131,132,133,134].
Antioxidants 12 01217 g001
Figure 2. Prognostic significance of HMOX1 expression in various cancer types. Kaplan-Meier survival curves obtained from PrognoScan database for patients with high (red) and low (blue) HO-1 mRNA expression in lung cancer (AC), breast cancer (DF), blood cancer (G), brain cancer (H,I), colorectal cancer (J,K), and ovarian cancer (L). OS, overall survival; RFS, relapse free survival; DMFS, distant metastasis free survival; DFS, disease free survival; HR, hazard ratio.
Figure 2. Prognostic significance of HMOX1 expression in various cancer types. Kaplan-Meier survival curves obtained from PrognoScan database for patients with high (red) and low (blue) HO-1 mRNA expression in lung cancer (AC), breast cancer (DF), blood cancer (G), brain cancer (H,I), colorectal cancer (J,K), and ovarian cancer (L). OS, overall survival; RFS, relapse free survival; DMFS, distant metastasis free survival; DFS, disease free survival; HR, hazard ratio.
Antioxidants 12 01217 g002
Figure 3. HMOX-1 expression differences between non-responsive and responsive groups in breast (A) and brain cancer (B) during chemotherapy. * p < 0.05.
Figure 3. HMOX-1 expression differences between non-responsive and responsive groups in breast (A) and brain cancer (B) during chemotherapy. * p < 0.05.
Antioxidants 12 01217 g003
Figure 4. Chemical structure of mainstream HO-1 inhibitors. (AC). Protoporphyrin class (cited from PubChem). (DF). Azole-based derivatives [158].
Figure 4. Chemical structure of mainstream HO-1 inhibitors. (AC). Protoporphyrin class (cited from PubChem). (DF). Azole-based derivatives [158].
Antioxidants 12 01217 g004
Figure 5. Protein-protein interaction network of HMOX1-related genes associated with antioxidative stress and anti-apoptosis factors. (A) The protein-protein interaction network of HMOX1-related genes based on STRING database (https://www.string-db.org/). (B) Gene Ontology (GO) enrichment analysis of HMOX1-related genes in protein-protein interaction network categorized by biological process.
Figure 5. Protein-protein interaction network of HMOX1-related genes associated with antioxidative stress and anti-apoptosis factors. (A) The protein-protein interaction network of HMOX1-related genes based on STRING database (https://www.string-db.org/). (B) Gene Ontology (GO) enrichment analysis of HMOX1-related genes in protein-protein interaction network categorized by biological process.
Antioxidants 12 01217 g005
Figure 6. Schematic representation of heme oxygenase 1 (HO-1) regulation and the potential mechanism of HO-1-mediated chemoresistance.
Figure 6. Schematic representation of heme oxygenase 1 (HO-1) regulation and the potential mechanism of HO-1-mediated chemoresistance.
Antioxidants 12 01217 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, H.; Cheng, Q.; Bao, L.; Li, M.; Chang, K.; Yi, X. Cytoprotective Role of Heme Oxygenase-1 in Cancer Chemoresistance: Focus on Antioxidant, Antiapoptotic, and Pro-Autophagy Properties. Antioxidants 2023, 12, 1217. https://doi.org/10.3390/antiox12061217

AMA Style

Wang H, Cheng Q, Bao L, Li M, Chang K, Yi X. Cytoprotective Role of Heme Oxygenase-1 in Cancer Chemoresistance: Focus on Antioxidant, Antiapoptotic, and Pro-Autophagy Properties. Antioxidants. 2023; 12(6):1217. https://doi.org/10.3390/antiox12061217

Chicago/Turabian Style

Wang, Huan, Qi Cheng, Lingjie Bao, Mingqing Li, Kaikai Chang, and Xiaofang Yi. 2023. "Cytoprotective Role of Heme Oxygenase-1 in Cancer Chemoresistance: Focus on Antioxidant, Antiapoptotic, and Pro-Autophagy Properties" Antioxidants 12, no. 6: 1217. https://doi.org/10.3390/antiox12061217

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