Next Article in Journal
Chromosome Transplantation: Opportunities and Limitations
Previous Article in Journal
Focal Adhesion’s Role in Cardiomyocytes Function: From Cardiomyogenesis to Mechanotransduction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Therapeutic Potential of Clusterin Inhibition in Human Cancer

by
Desirée Martín-García
1,2,3,4,†,
Marilina García-Aranda
2,3,4,† and
Maximino Redondo
1,2,3,4,*
1
Surgical Specialties, Biochemistry and Immunology Department, Faculty of Medicine, University of Málaga, 29010 Málaga, Spain
2
Red de Investigación en Servicios de Salud en Enfermedades Crónicas (REDISSEC), Red de Investigación en Cronicidad, Atención Primaria y Promoción de la Salud (RICAPPS), Instituto de Investigación Biomédica de Málaga (IBIMA), 29590 Málaga, Spain
3
Instituto de Investigación Biomédica de Málaga y Plataforma en Nanomedicina—IBIMA Plataforma BIONAND, 29590 Málaga, Spain
4
Research and Innovation Unit, Hospital Costa del Sol, 29602 Marbella, Spain
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Cells 2024, 13(8), 665; https://doi.org/10.3390/cells13080665
Submission received: 5 February 2024 / Revised: 11 March 2024 / Accepted: 9 April 2024 / Published: 10 April 2024

Abstract

:
Clusterin (CLU) protein is involved in various pathophysiological processes including carcinogenesis and tumor progression. In recent years, the role of the secretory isoform has been demonstrated in tumor cells, where it inhibits apoptosis and favors the acquisition of resistance to conventional treatments used to treat cancer. To determine the possible therapeutic potential of inhibiting this protein, numerous studies have been carried out in this field. In this article, we present the existing knowledge to date on the inhibition of this protein in different types of cancer and analyze the importance it could have in the development of new therapies targeted against this disease.

Graphical Abstract

1. Introduction

Clusterin (CLU), also known by various names such as Apoliprotein J (APOJ), Complement Lysis Inhibitor (CLI), Complimented associated protein SP-40, 40 (SP-40), Sulfated Glycoprotein 2 (SGP-2), Testosterone-Repressed Prostate Message 2 (TRPM2), and Ku70-Binding Protein 1 or KUB1, is a multifunctional protein encoded by the CLU gene [1], widely expressed in various human tissues [2,3]. Over the years, CLU has been identified as a key molecule in a variety of physiological processes [3], including cell differentiation, morphogenesis [4], sperm maturation, lipid transport, complement inhibition, tissue remodeling, membrane recycling, cell–cell and cell–substrate interactions, the stabilization of stressed proteins, and cell proliferation, survival, and apoptosis [5,6].
Alterations in CLU expression have been associated with serious physiological diseases, such as spongiform encephalopathies, hippocampal and heart ischemic injuries, atherosclerosis [7], schizophrenia [8], cardiovascular diseases [9], cancer, vascular damage, diabetes, and osteoarthritis [10], and degenerative conditions such as age-related macular degeneration, retinitis pigmentosa [11], Parkinson’s disease [12], and Alzheimer’s [13,14]. These conditions are often more prominent in advanced aging [13,14], where CLU overexpression has been observed in phenomena such as enhanced cell migration [15], chemotherapy resistance [16,17], and more aggressive biological behaviors [18] in malignant cells. CLU has been proposed as a cancer biomarker [19] and cellular senescence [20], generating interest in research to better understand its implications and explore its utility in the diagnosis, prevention, and treatment of these pathologies.

2. Isoforms and Regulation of CLU Expression Gene

The human CLU gene, located on chromosome 8p21-p12, is a highly conserved gene with at least 63 orthologs in different species [21], and is structured with 11 exons. This gene produces at least 17 splice variants, both coding and non-coding [21,22]. However, Variant 1, Variant 2, and Variant 3 (NM-001831, NR_038335, NR_045494, respectively) are known as three transcription products that share some parts, such as exons 2-11, but differ in the first exon (1a, 1b, 1c) and untranslated terminal regions [23,24,25]. This suggests that there are different starting points for the transcription of each variant, but it is still unclear what biological function these different mRNAs have.
The synthesis of sCLU begins with the initiation codon located in exon 2 of the mRNA, leading to the formation of a preprotein comprising 449 amino acids. Initially, the sequence encodes a translocation signal that guides the preprotein to the endoplasmic reticulum (ER). In the ER, this signal is removed, and N-glycosylation takes place at six asparagine residues. This process converts the preprotein into an almost presecreted mature isoform of 53 kDa (psCLU) [26]. Subsequently, psCLU moves to the Golgi apparatus, where it undergoes further intricate glycosylation, resulting in a weight increase to between 70 and 80 kDa. The secreted isoform (sCLU), which represents the mature form of the protein, is created through the process of cleavage and the formation of disulfide bonds between residues 227 and 228 (Figure 1). Thus, sCLU is finally secreted as a heterodimeric glycoprotein complex composed of two polypeptide chains, an alpha chain (34–36 kDa) and a beta chain (36–39 kDa), connected by five disulfide bonds. It acts as an extracellular chaperone, inhibiting the aggregation of partially unfolded proteins, enhancing phagocytosis and cellular degradation, and exhibiting antiapoptotic activity in cancer cells [27,28,29]. Additionally, this variant has the capability to enhance the production of the p53 protein, which plays a role in triggering genes responsible for pausing the cell cycle, consequently inhibiting cell proliferation [3,30].
In situations of alternative splicing between exons 1 and 3, exon 2 is deleted, generating a non-functional prenuclear isoform of 49 kDa (pnCLU). In response to stress, the isoform undergoes a transformation into a functional nuclear variant known as nCLU, weighing 55 kDa. After translocating to the nucleus and binding with the Ku70 protein, it facilitates cellular apoptosis [31,32,33,34] (Figure 2).
In stressful situations, the nearly mature presecreted isoform of 53 kDa (psCLU) first interacts with the chaperone GRP78 (Bip) in the endoplasmic reticulum (ER). Subsequently, it translocates to the mitochondria, where it associates with the activated form of the Bax protein. This interaction influences the Bax protein’s ability to form homodimers and impedes the formation of Bax-Bak complexes. Within the mitochondrial, it associates with the activated form of the Bax protein, influencing its capacity to create homodimers and hindering the formation of Bax-Bak complexes [35]. This antiapoptotic influence extends to the cytoplasm, where psCLU can stabilize the Ku70-Bax complex. By doing so, it hinders Bax from reaching the mitochondria and fosters the proteasomal breakdown of cytotoxic substances [3,36].
The expression of CLU is meticulously regulated in both healthy and pathological contexts, where various pathways exert influence on its basal expression in healthy tissues as well as its upregulation in conditions such as apoptosis, pathologies, and aging. In this scenario, two forms of CLU with antagonistic functions are distinguished: sCLU, playing a cytoprotective role, and nCLU, capable of promoting cell death [37,38]. Each CLU variant is subject to different signaling pathways, depending not only on its molecular configuration but also on the cell type in which it manifests. The sophisticated regulation of both CLU isoforms encompasses a broad spectrum of factors, including growth factors such as EGF, NFGF, TGF-β, and IGF1, cytokines like IL-1, IL-6, TNF-α, and IL-2, transcription factors such as TCF1 and Jak/STAT1, as well as the Wnt signaling pathway [39,40,41,42] and epigenetic modifications. Furthermore, various factors like oxidate stress, chemotherapeutic agents, ionizing and ultraviolet radiation, estrogen, and androgen deprivation can induce their expression in hormone-sensitive tumors [43,44].
Gene expression control in vertebrates is an intricate process orchestrated by numerous regulatory proteins that collaborate to influence the expression of a specific gene or transcript within a cell. Typically, vertebrate gene expression is regulated by the basal promoter, a DNA region spanning from -200bp to the transcription start site (TSS) (+1), facilitating the binding of regulatory proteins, including transcription factors [45]. The variation in CLU expression across different tissues suggests its finely tuned and tissue-specific regulation at all levels [46]. The human CLU promoter, initially identified as a unique element, contains conserved DNA motifs, including a palindrome (-73 to -87) and a TATAA box (-26), crucial for decoding the genetic sequence. Furthermore, a GCATT box (-93) plays a key role in regulating transcription frequency [47]. Additional studies have uncovered a second promoter region (P2) within the first CLU intron, potentially controlling CLU2 expression [23].
In the regulation of CLU expression during aging and cancer progression, epigenetics emerges as a pivotal player, investigating heritable changes in gene expression without altering the DNA sequence [48]. The intricate interplay of DNA methylation and histone acetylation [11,49], influenced by aging and cancer, suggests epigenetic regulation’s significant role in controlling CLU [11,23,48,50].
The Histone Code Hypothesis outlines the diverse histone modifications governing chromatin structure and transcriptional status [51]. Histone modifications impact gene silencing or activation by modulating DNA accessibility. Notably, histone 3 lysine 9 trimethylation (H3K9me3) and histone 3 lysine 27 trimethylation (H3K27me) contribute to nCLU downregulation [48] and cell survival. Conversely, histone 3 lysine 4 trimethylation (H3K4me3) or histone 3 lysine 9 acetylation and serine 10 phosphorylation (H3K9AcS10P) lead to nCLU activation [48] and cell death. Epigenetic drug treatments alter histone modifications, impacting CLU1 and CLU2 transcription [23].
Consistent findings across diverse cell types, including cancer cells [11,23], retinal pigment epithelial cells [11], hepatocellular carcinoma cell lines [52], and endothelial tumor cells [49], emphasize the role of histone hyperacetylation induced by histone deacetylase inhibition in promoting CLU expression.
In differentiated mammalian cells, gene repression often involves DNA methylation, co-coordinated with histone modifications. While colon cancer studies indicate the predominant regulation of CLU by histone modifications [48], the presence of a G+C-rich region and a CpG mini-island within the CLU promoter suggests epigenetic control via methylation [47]. This promoter methylation is associated with decreased protein expression, observed in breast [50], ovarian epithelial cancer [18], and hormone-refractory prostate carcinoma samples [53]. Induced demethylation of CLU promoters significantly increases CLU1 and CLU2 transcripts, enhancing CLU expression across various colon cell lines [48] and tissues [54].
Furthermore, micro-RNAs (mi-RNAs) contribute to the post-transcriptional regulation of CLU expression [55]. In head and neck squamous cell carcinoma, oncogenic miRNA-21 specifically targets the CLU1 isoform, downregulating its growth-suppressive variant [56]. Elevated levels of miRNA-378 inhibit lung adenocarcinoma cell growth and sCLU expression [57].
While this field’s studies are recent, the emerging evidence underscores the intricate epigenetic control mechanisms orchestrating CLU expression in diverse cellular contexts. These insights provide a foundation for understanding the molecular intricacies of CLU regulation, offering potential avenues for therapeutic interventions in aging and cancer-related conditions.

3. Clusterin and Its Involvement in Cancer

Despite initially being believed to be linked to biological processes such as scar formation, membrane remodeling, or sperm maturation [47], early studies revealed a correlation between the overexpression of mRNA-CLU and cell death after different toxic signals [58,59,60]. However, later research presented contradictory results, establishing a relationship between CLU overexpression and cell protection [61,62].
The identification of two sets of CLU isoforms, with opposing roles and cellular locations in terms of cell survival, has contributed to clarifying these contradictions [34,63]. It has been suggested that the ratio between the expression of sCLU and nCLU is crucial in determining cancer aggressiveness. In tumor cells, a survival-favorable environment is observed with a higher sCLU/nCLU ratio, indicating elevated sCLU levels with concomitantly reduced levels of nCLU protein in the nucleus of neoplastic cells [64,65].
Post-translational modifications appear to regulate the distinctive roles of CLU, determining the expression of specific isoforms based on the presence or absence of a leader sequence in the CLU peptide [46]. Inhibition of exon 2 of mRNA-CLU, leading to the silencing of the sCLU isoform, has been crucial for the design of cancer-targeted therapies that are currently under study [66].
Early studies suggest that CLU overexpression persists after tissue degeneration, suggesting that this gene is probably not induced as a part of apoptosis that may lead to a degenerative disorder but as a secondary consequence of the disease phenotype [47]. nCLU-mediated apoptosis is associated with the accumulation of mature nCLU protein, while the rapid degradation of CLU and the regulation of its nuclear export/import can explain why highly malignant tumors avoid accumulating nCLU levels [66]. The lethality of nCLU is linked to members of the BCL-2 protein family, with nCLU-mediated apoptosis depending on Bax [64]. Additionally, nCLU can sequester the antiapoptotic BCL-XL, promoting apoptosis by forming pores in mitochondrial membranes [67].
The proapoptotic role of nCLU is also related to Ku70, which binds to CLU in response to DNA damage. The formation of a trimeric protein complex (nCLU/Ku70/Ku80) with reduced DNA end-binding activity and interaction with other proteins like Ku70 in the C-terminal coiled-coil domain of nCLU is essential for inducing apoptosis [68].
Although the constant expression of nCLU is related to cell maintenance and apoptosis induction, cells can also express cytoplasmic clusterin (cCLU) or sCLU isoforms in response to different cellular triggers. This results in increased cell survival associated with tumorigenesis, with CLU promoting apoptosis under low or moderate stress conditions but favoring the overexpression of the antiapoptotic isoform [19] under extreme cellular conditions, resulting in increased cell proliferation, viability, and invasiveness [7].
Furthermore, CLU has garnered significant attention from the scientific society because of its crucial involvement in a multitude of biological processes, including resistance to chemotherapy, cell proliferation, or apoptosis (Table 1) [69].

3.1. Tumorigenesis

Tumorigenesis, or tumor formation, is a complex process where normal cells transform into cancerous cells, proliferating uncontrollably. In this context, the protein CLU plays a significant role. Under normal conditions, CLU expression is low, but it significantly increases in response to stress, especially during tumorigenesis (Table 1) [72,83].
The sCLU isoform of CLU acts as an extracellular chaperone during stressful situations. This protects cells by preventing apoptosis (programmed cell death) and conferring resistance to cytotoxic agents. sCLU also plays a crucial role by interacting with protein complexes like Ku70-bax, acting as a Bax retention factor in the cytosol, inhibiting its proapoptotic function. Under normal conditions, inhibition of CLU weakens this complex, allowing Bax to translocate to the mitochondria, triggering cytochrome c release, and activating caspase 9, initiating apoptosis [70].
On the other hand, the proapoptotic isoform nCLU of CLU induces apoptosis in certain types of cancer, such as breast and prostate, through specific interactions with proteins like Ku70 and Bcl-XL [37,71].
Additionally, CLU is involved in various signaling pathways, such as B-MYB, Akt, and the PI3K/Akt pathway, regulating both apoptotic and antiapoptotic functions [84,85]. The ratio between CLU isoforms is critical for the regulation of these functions. CLU overexpression has been associated with the promotion of tumorigenesis and resistance to chemotherapy in various types of cancer, including breast and prostate [86].

3.2. Cell Proliferation

Cell proliferation, a cornerstone of biology, involves the proliferation of cells through repeated divisions, ensuring a dynamic equilibrium between cell death and metabolic promotion. Conversely, cell growth pertains to the physical enlargement of cell volume or size as the cell matures [87]. These processes are intricately regulated by various growth factors and cytokines in normal cells [88]. However, in cancer cells, cell proliferation becomes uncontrolled and exhibits a sustained proliferation property due to the hyperactivation of proliferative signaling pathways and evasion of growth suppressors [88].
Recent evidence highlights the involvement of the CLU protein in promoting cell proliferation and growth across different cancer types [69]. This is characterized by the acquisition of sustained proliferative signaling pathways by cancer cells (Table 1). Specifically, research indicates that the transcription factor c-Myc, encoded by the oncogene MYC and implicated in tumorigenesis, suppresses the expression of nCLU. It does so by elevating the levels of the microRNA cluster miRNA-17 ~ 92 and reducing the activity of the TGF-β axis, thus facilitating angiogenesis and tumor progression in colon cancer [73].
Both sCLU and nCLU exert notable influence on the canonical NF-κB, ERK, and AKT pathways, thereby impacting cell proliferation and growth in diverse cancer types. In prostate cancer, inhibition of sCLU with apocynin, an inhibitor of NADPH oxidase, arrests the MEK-ERK1/2 pathway, leading to suppressed cell proliferation [89]. Conversely, in osteosarcoma, induction of sCLU by DDP triggers phosphorylation of ERK1/2, stimulating cell growth and bolstering resistance to DDP [74].
The activation of the NF-κB pathway is pivotal for cell proliferation, as it prompts the expression of Bcl-2 [75]. In pancreatic cancer, melittin disrupts sCLU, thereby impeding both the cholesterol/NF-κB/Bcl-2 axis and the cholesterol/p-ERK axis, resulting in the suppression of tumor growth [75]. Interestingly, sCLU exhibits antiproliferative effects by deactivating the TAK1/NF-κB axis, preventing the recruitment of the TNF receptor-activating factor 6 (TRAF6)/TAK1-binding protein 2 (TAB2)/TGF-β-activated kinase 1 (TAK1) complex by the transforming growth factor beta receptor 1 (TGFBR1). This inhibition effectively hampers tumor proliferation and growth in human non-small cell lung cancer (NSCLC) [90].
In addition to the ERK and NF-κB pathways, the PI3K/AKT axis is also involved in regulating cell proliferation and growth. Notably, sCLU activates AKT by downregulating the expression of the protein phosphatase 2A catalytic subunit C (PP2AC) to promote the proliferation of PC-3 prostate cancer cells [91]. This aligns with the finding that silencing CLU gene transcription decreases the phosphorylation level of AKT and GSK-3β, subsequently inhibiting the proliferation of HCCLM3 cells in hepatocellular carcinoma (HCC) [92].
CLU expression and cell proliferation are influenced by various external factors. Insulin-like growth factor 1 (IGF-1), an essential component of insulin receptor signaling known for sustaining cell proliferation [81], has been found to upregulate sCLU expression. For instance, IGF-1 increases sCLU expression, activating the PI3K/AKT axis and promoting A549 cell proliferation in non-small cell lung cancer [85]. Moreover, in prostate cancer, IGF-1 enhances the transcriptional activity of the CLU gene by activating the STAT-3/Twist-1 axis. Furthermore, pituitary tumor-transforming gene (PTTG) and forkhead box protein L2 (FOXL2) influence sCLU expression and CLU gene transcription [93]. PTTG activates the ATM/IGF-1/p38MAPK/CLU axis, while FOXL2 directly binds to the CLU promoter. Interestingly, sCLU exhibits antiproliferative effects by suppressing PTTG expression, thereby restraining cell proliferation in pituitary carcinoma [76].
Metformin, a well-known medication used to treat type II diabetes, demonstrates anticancer properties. In bladder cancer, metformin works by reducing the activity of sCLU, thus inhibiting tumor growth through the deactivation of the SREBP-1c/fatty acid synthase (FASN) axis [77]. Furthermore, in renal cancer, sCLU enhances cell growth and proliferation by increasing the expression of the calcium-binding protein S100A4 [94]. In addition to metformin and melittin, compounds like epigallocatechin-3-gallate (EGCG) and green tea extract (GTE) also boost sCLU expression. They accomplish this by suppressing the activity of β-catenin, thereby promoting the proliferation of COLO 205 cells in colon cancer [95].

3.3. Epithelial–Mesenchymal Transition and Metastasis

Epithelial–mesenchymal transition (EMT) is a biological process involving significant changes in cell structure and function, such as loss of polarity, cytoskeletal remodeling, and acquisition of invasive properties. EMT facilitates invasion, metastasis, and tumor progression. CLU has emerged as a key regulator in these events, influencing matrix metalloproteinase (MMP) activity and the ERK1/2 and PI3K/Akt signaling pathways [96]. NF-κB activation by CLU also plays a crucial role in increasing MMP-9 and MMP-2 expression, thus promoting metastasis [79].
In different cancer types, CLU overexpression has been associated with metastasis (Table 1). In nasopharyngeal carcinoma, CLU is positively regulated by N, N′-dinitrosopiperazine (DNP), inducing MMP-9 and VEGF expression, contributing to metastasis [80]. In breast cancer, CLU collaborates with eHsp90α to activate key signaling pathways, promoting EMT, migration, and tumor metastasis [82]. In colon cancer, CLU interaction with platelets activates the p38MAPK pathway and positively regulates MMP-9, facilitating invasion [85]. Additionally, studies on prostate cancer have demonstrated that miRNA-217-5p exerts control over the processes of invasion and migration by specifically targeting CLU [97].
In patients with pancreatic ductal adenocarcinoma (PDAC), hepatocyte nuclear factor 1 b (HNF1B) has been shown to positively regulate CLU, and lower expression of both was associated with poor patient survival. Studies with pancreatic cancer cell lines have revealed that CLU inhibits cell proliferation, invasiveness, or EMT. These findings suggest that the HNF1B/CLU pathway is crucial in slowing the progression of pancreatic cancer [98].

3.4. Chemoresistance and Chemosensitivity with Clusterin

Chemoresistance, which involves cancer cells’ ability to withstand chemotherapy, poses a significant obstacle in cancer treatment. This phenomenon can occur through different mechanisms, including genetic alterations, changes in drug metabolism, and the suppression of apoptosis [99]. Chaperone proteins like CLU are pivotal in conferring resistance to cancer therapy, promoting the growth of malignant tumors and shielding drug-resistant cells (Table 1) [78].
The stress response triggered by treatments like radiotherapy and chemotherapy leads to increased expression of sCLU, a protective chaperone. sCLU interacts with activated Bax, inhibiting the release of cytochrome c and thus preventing apoptosis [100]. Conversely, reducing CLU levels, as observed in certain cancers like testicular seminoma, enhances sensitivity to radiotherapy and chemotherapy [101].
In hepatocellular carcinoma (HCC), sCLU is markedly overexpressed, contributing to resistance against oxaliplatin by downregulating Gadd45a expression and activating the PI3K/Akt pathway. Inhibiting CLU in HepG2/ADM HCC cells replenishes sensitivity to various drugs [102].

4. Clusterin as a Biomarker and Therapeutic Target in Cancer

In the field of medicine, biomarkers have become essential tools to provide valuable information about patient health and the progression of various diseases. From measuring simple protein levels to identifying genetic mutations, these markers play a crucial role in the diagnosis, prognosis, monitoring, and treatment of diseases. In this context, CLU has been studied as a potential biomarker in various types of cancer due to its tendency to be overexpressed in stressful situations (Table 2).
However, further research is needed in this field to determine the exact value of CLU levels and their association with each type of cancer. Currently, it can only be used as a tool to assess risks [140].
As a therapeutic target, the inhibition of CLU has demonstrated excellent therapeutic effects in various cancers both in vitro and in vivo, prolonging the survival of patients. Custirsen (OGX-011), a second-generation antisense oligonucleotide, has proven effective by interacting with the ATG sequence in exon 2 of the secretory isoform of CLU (sCLU), inhibiting its translation and suppressing cancer progression such as prostate cancer [15], renal cell carcinoma [141], bladder [142], liver [143], lung [144], prostate [145], breast [146], lung adenocarcinoma [104], melanoma [131], osteosarcoma [135], and ovary [147].
When combined with other cancer therapies in clinical trials, OGX 011 enhances antitumor activity. In the case of renal cell carcinoma, CLU inhibition by custirsen improves sorafenib cytotoxicity [148]. In ovarian cancer, it has been shown to improve the survival of patients treated with paclitaxel by enhancing its response [149], similar to what happens in castration-resistant prostate cancer with mitoxantrone and docetaxel by reducing sCLU expression [150,151]. Clinical studies have also explored the efficacy of custirsen in metastatic castration-resistant prostate cancer, showing promising results in early-phase trials, although the standard treatment remains prednisone and cabazitaxel [152].
In metastatic breast cancer, similar effects have been shown with combinations of OGX 011 and docetaxel [153], while in lung cancer, combinations with gemcitabine or cisplatin have extended survival [154].
Therapeutic approaches beyond OGX-011, such as RNA interference strategies including microRNA (miRNA), short hairpin RNA (shRNA), and small interfering RNA (siRNA), have demonstrated efficacy in suppressing CLU expression across various cancers. For instance, miRNA-217-5p and miRNA-195 have been found to downregulate CLU expression, impeding cell migration and invasion while fostering apoptosis in prostate cancer cell cultures [155,156]. Furthermore, several other therapeutic agents, including melittin, green tea extract (GTE), apocynin, vitamin D, metformin, and verteporfin, have exhibited antitumor properties by inhibiting CLU protein expression. Consequently, these interventions have shown promise in impeding cancer progression [75,77,89,157,158,159].
In the realm of emerging therapies, the drug AB-16B5, a monoclonal antibody targeting sCLU, is being evaluated in a phase II clinical trial in combination with docetaxel in patients with metastatic non-small cell lung cancer [160].

5. Conclusions

CLU plays a pivotal role in cancer progression by modulating key processes such as programmed cell death, epithelial–mesenchymal transition (EMT), metastasis, and cell proliferation and growth through intricate signaling pathways. The complexity of its biological function stems from the existence of two alternatively spliced isoforms and the variable localization of their protein products both intra- and extracellularly. Despite the intricate nature of these isoforms, the coexistence of nCLU and sCLU within cells, coupled with the precise regulation of their balance, confers either pro- or antiapoptotic properties.
Various potent inhibitors of CLU, including OGX-011 and RNA interference (RNAi) agents, have been developed for use in cancer therapies, yielding encouraging therapeutic outcomes in patients. Given the diverse regulatory pathways governed by CLU in cancer progression and the demonstrated survival benefits associated with its inhibition across different cancer types, CLU emerges as a promising therapeutic target for the development of more efficacious agents and targeted therapies. Moreover, considering its oncogenic properties, the sCLU isoform holds potential as a blood biomarker for cancer diagnosis.
However, despite extensive research on the functions of CLU isoforms, it remains unclear which isoform contributes to specific biological effects. Therefore, it is imperative to delve deeper into the distinct functional role of each CLU isoform, particularly sCLU, to refine CLU targeting as a therapeutic strategy more effectively.

Author Contributions

D.M.-G., conception and design and wrote the paper with input from all authors; M.G.-A., review and editing and contribution to the final version of the manuscript; M.R., conception, critical revision, and contribution to the final version of the manuscript. All authors have read and agreed to the published version of the manuscripts.

Funding

The researcher Marilina García-Aranda is the benefactor of a postdoctoral contract financed by the European Social Fund—Operational Program of Andalusia 2014–2020 for the “Incorporation of Research Personnel with a PhD degree in the field of Health Sciences and Technologies in R&D and Innovation Centers of the Public Health System of Andalusia” (RH-0055-2020). This work was partially supported by grants from the University of Malaga—Consejería de Transformación Económica, Industria, Conocimiento y Universidades—Junta de Andalucia (UMA20-FEDERJA-161) and from Instituto de Salud Carlos III (PI18/01181, PI21/00252) and were cofunded by the European Regional Development fund. The funders had no role in the study design, data collection and analysis, decision to publish, or preparation of the manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hogg, S.D.; Embery, G. The isolation and partial characterization of a sulphated glycoprotein from human whole saliva which aggregates strains of Streptococcus sanguis but not Streptococcus mutans. Arch. Oral Biol. 1979, 24, 791–797. [Google Scholar] [CrossRef] [PubMed]
  2. Jones, S.E.; Jomary, C. Clusterin. Int. J. Biochem. Cell Biol. 2002, 34, 427–431. [Google Scholar] [CrossRef] [PubMed]
  3. Trougakos, I.P.; Gonos, E.S. Regulation of clusterin/apolipoprotein J, a functional homologue to the small heat shock proteins, by oxidative stress in aging and age-related diseases. Free Radic. Res. 2006, 40, 1324–1334. [Google Scholar] [CrossRef] [PubMed]
  4. Itahana, Y.; Piens, M.; Sumida, T.; Fong, S.; Muschler, J.; Desprez, P.Y. Regulation of clusterin expression in mammary epithelial cells. Exp. Cell Res. 2007, 313, 943–951. [Google Scholar] [CrossRef] [PubMed]
  5. Trougakos, I.P.; Gonos, E.S. Clusterin/apolipoprotein J in human aging and cancer. Int. J. Biochem. Cell Biol. 2002, 34, 1430–1448. [Google Scholar] [CrossRef]
  6. Bettuzzi, S. Chapter 1: Introduction. Adv. Cancer Res. 2009, 104, 1–8. [Google Scholar] [CrossRef]
  7. Prochnow, H.; Gollan, R.; Rohne, P.; Hassemer, M.; Koch-Brandt, C.; Baiersdörfer, M. Non-secreted clusterin isoforms are translated in rare amounts from distinct human mRNA variants and do not affect Bax-mediated apoptosis or the NF-kappaB signaling pathway. PLoS ONE 2013, 8, e75303. [Google Scholar] [CrossRef]
  8. Athanas, K.M.; Mauney, S.L.; Woo, T.U. Increased extracellular clusterin in the prefrontal cortex in schizophrenia. Schizophr. Res. 2015, 169, 381–385. [Google Scholar] [CrossRef]
  9. Yu, B.; Yang, Y.; Liu, H.; Gong, M.; Millard, R.W.; Wang, Y.G.; Ashraf, M.; Xu, M. Clusterin/Akt Up-Regulation Is Critical for GATA-4 Mediated Cytoprotection of Mesenchymal Stem Cells against Ischemia Injury. PLoS ONE 2016, 11, e0151542. [Google Scholar] [CrossRef]
  10. Fandridis, E.; Apergis, G.; Korres, D.S.; Nikolopoulos, K.; Zoubos, A.B.; Papassideri, I.; Trougakos, I.P. Increased expression levels of apolipoprotein J/clusterin during primary osteoarthritis. In Vivo 2011, 25, 745–749. [Google Scholar]
  11. Suuronen, T.; Nuutinen, T.; Ryhänen, T.; Kaarniranta, K.; Salminen, A. Epigenetic regulation of clusterin/apolipoprotein J expression in retinal pigment epithelial cells. Biochem. Biophys. Res. Commun. 2007, 357, 397–401. [Google Scholar] [CrossRef] [PubMed]
  12. Prikrylova Vranova, H.; Mareš, J.; Nevrlý, M.; Stejskal, D.; Zapletalová, J.; Hluštík, P.; Kaňovský, P. CSF markers of neurodegeneration in Parkinson’s disease. J. Neural Transm. 2010, 117, 1177–1181. [Google Scholar] [CrossRef] [PubMed]
  13. Weinstein, G.; Beiser, A.S.; Preis, S.R.; Courchesne, P.; Chouraki, V.; Levy, D.; Seshadri, S. Plasma clusterin levels and risk of dementia, Alzheimer’s disease, and stroke. Alzheimers Dement. 2016, 3, 103–109. [Google Scholar] [CrossRef] [PubMed]
  14. Schrijvers, E.M.; Koudstaal, P.J.; Hofman, A.; Breteler, M.M. Plasma clusterin and the risk of Alzheimer disease. JAMA 2011, 305, 1322–1326. [Google Scholar] [CrossRef]
  15. Chen, D.; Wang, Y.; Zhang, K.; Jiao, X.; Yan, B.; Liang, J. Antisense oligonucleotide against clusterin regulates human hepatocellular carcinoma invasion through transcriptional regulation of matrix metalloproteinase-2 and E-cadherin. Int. J. Mol. Sci. 2012, 13, 10594–10607. [Google Scholar] [CrossRef] [PubMed]
  16. Cheng, C.Y.; Wang, Y.; Zhang, K.; Jiao, X.; Yan, B.; Liang, J. Regulation of chemosensitivity and migration by clusterin in non-small cell lung cancer cells. Cancer Chemother. Pharmacol. 2012, 69, 145–154. [Google Scholar] [CrossRef] [PubMed]
  17. Lourda, M.; Trougakos, I.P.; Gonos, E.S. Development of resistance to chemotherapeutic drugs in human osteosarcoma cell lines largely depends on up-regulation of Clusterin/Apolipoprotein J. Int. J. Cancer 2007, 120, 611–622. [Google Scholar] [CrossRef] [PubMed]
  18. Yang, G.; Zhang, H.; Liu, Y.; Zhou, J.; He, W.; Quick, C.M.; Xie, D.; Smoller, B.R.; Fan, C.Y. Epigenetic and immunohistochemical characterization of the Clusterin gene in ovarian tumors. Arch. Gynecol. Obstet. 2013, 287, 989–995. [Google Scholar] [CrossRef]
  19. Tellez, T.; García-Aranda, M.; Redondo, M. The role of clusterin in carcinogenesis and its potential utility as a therapeutic target. Curr. Med. Chem. 2016, 23, 4297–4308. [Google Scholar] [CrossRef]
  20. Trougakos, I.P.; Pawelec, G.; Tzavelas, C.; Ntouroupi, T.; Gonos, E.S. Clusterin/Apolipoprotein J up-regulation after zinc exposure, replicative senescence, or differentiation of human hematopoietic cells. Biogerontology 2006, 7, 375–382. [Google Scholar] [CrossRef]
  21. Available online: http://www.ensembl.org/Homo_sapiens/Gene/Summary?g=ENSG00000120885,r=8:27596917-27615031 (accessed on 4 January 2024).
  22. Available online: http://vega.sanger.ac.uk/Homo_sapiens/Gene/Summary?g=OTTHUMG00000102114,r=8:27596917-27615031 (accessed on 4 January 2024).
  23. Bonacini, M.; Coletta, M.; Ramazzina, I.; Naponelli, V.; Modernelli, A.; Davalli, P.; Bettuzzi, S.; Rizzi, F. Distinct promoters, subjected to epigenetic regulation, drive the expression of two clusterin mRNAs in prostate cancer cells. Biochim. Biophys. Acta 2015, 1849, 44–54. [Google Scholar] [CrossRef] [PubMed]
  24. Rizzi, F.; Coletta, M.; Bettuzzi, S. Chapter 2: Clusterin (CLU): From one gene and two transcripts to many proteins. Adv. Cancer Res. 2009, 104, 9–23. [Google Scholar] [PubMed]
  25. Yu, J.T.; Tan, L. The role of clusterin in Alzheimer’s disease: Pathways, pathogenesis, and therapy. Mol. Neurobiol. 2012, 45, 314–326. [Google Scholar] [CrossRef] [PubMed]
  26. Foster, E.M.; Dangla-Valls, A.; Lovestone, S.; Ribe, E.M.; Buckley, N.J. Clusterin in Alzheimer’s Disease: Mechanisms, Genetics, and Lessons from Other Pathologies. Front. Neurosci. 2019, 13, 164. [Google Scholar] [CrossRef] [PubMed]
  27. Rohne, P.; Prochnow, H.; Koch-Brandt, C. The CLU-files: Disentanglement of a mystery. Biomol. Concepts 2016, 7, 1–15. [Google Scholar] [CrossRef]
  28. Zhang, Q.; Zhou, W.; Kundu, S.; Jang, T.L.; Yang, X.; Pins, M.; Smith, N.; Jovanovic, B.; Xin, D.; Liang, L.; et al. The leader sequence triggers and enhances several functions of clusterin and is instrumental in the progression of human prostate cancer in vivo and in vitro. BJU Int. 2006, 98, 452–460. [Google Scholar] [CrossRef] [PubMed]
  29. Poon, S.; Treweek, T.M.; Wilson, M.R.; Easterbrook-Smith, S.B.; Carver, J.A. Clusterin is an extracellular chaperone that specifically interacts with slowly aggregating proteins on their off-folding pathway. FEBS Lett. 2002, 513, 259–266. [Google Scholar] [CrossRef]
  30. Bettuzzi, S.; Scorcioni, F.; Astancolle, S.; Davalli, P.; Scaltriti, M.; Corti, A. Clusterin (SGP-2) transient overexpression decreases proliferation rate of SV40-immortalized human prostate epithelial cells by slowing down cell cycle progression. Oncogene 2002, 21, 4328–4334. [Google Scholar] [CrossRef]
  31. Pucci, S.; Bonanno, E.; Pichiorri, F.; Angeloni, C.; Spagnoli, L.G. Modulation of different clusterin isoforms in human colon tumorigenesis. Oncogene 2004, 23, 2298–2304. [Google Scholar] [CrossRef]
  32. Caccamo, A.E.; Scaltriti, M.; Caporali, A.; D’Arca, D.; Corti, A.; Corvetta, D.; Sala, A.; Bettuzzi, S. Ca2+ depletion induces nuclear clusterin, a novel effector of apoptosis in immortalized human prostate cells. Cell Death Differ. 2005, 12, 101–104. [Google Scholar] [CrossRef]
  33. Caccamo, A.E.; Scaltriti, M.; Caporali, A.; D’Arca, D.; Scorcioni, F.; Astancolle, S.; Mangiola, M.; Bettuzzi, S. Cell detachment and apoptosis induction of immortalized human prostate epithelial cells are associated with early accumulation of a 45 kDa nuclear isoform of clusterin. Biochem. J. 2004, 382, 157–168. [Google Scholar] [CrossRef] [PubMed]
  34. Leskov, K.S.; Klokov, D.Y.; Li, J.; Kinsella, T.J.; Boothman, D.A. Synthesis and functional analyses of nuclear clusterin, a cell death protein. J. Biol. Chem. 2003, 278, 11590–11600. [Google Scholar] [CrossRef]
  35. Li, N.; Zoubeidi, A.; Beraldi, E.; Gleave, M.E. GRP78 regulates clusterin stability, retrotranslocation and mitochondrial localization under ER stress in prostate cancer. Oncogene 2013, 32, 1933–1942. [Google Scholar] [CrossRef]
  36. Nizard, P.; Tetley, S.; Le Dréan, Y.; Watrin, T.; Le Goff, P.; Wilson, M.R.; Michel, D. Stress-Induced Retrotranslocation of Clusterin/ApoJ into the Cytosol. Traffic 2007, 8, 554–565. [Google Scholar] [CrossRef]
  37. Scaltriti, M.; Bettuzzi, S.; Sharrard, R.M.; Caporali, A.; Caccamo, A.E.; Maitland, N.J. Clusterin overexpression in both malignant and nonmalignant prostate epithelial cells induces cell cycle arrest and apoptosis. Br. J. Cancer 2004, 91, 1842–1850. [Google Scholar] [CrossRef]
  38. Scaltriti, M.; Santamaria, A.; Paciucci, R.; Bettuzzi, S. Intracellular clusterin induces G2-M phase arrest and cell death in PC-3 prostate cancer cells1. Cancer Res. 2004, 64, 6174–6182. [Google Scholar] [CrossRef] [PubMed]
  39. Park, S.; Mathis, K.W.; Lee, I.K. The physiological roles of apolipoprotein J/clusterin in metabolic and cardiovascular diseases. Rev. Endocr. Metab. Disord. 2014, 15, 45–53. [Google Scholar] [CrossRef]
  40. Zoubeidi, A.; Chi, K.; Gleave, M. Targeting the cytoprotective chaperone, clusterin, for treatment of advanced cancer. Clin. Cancer Res. 2010, 16, 1088–1093. [Google Scholar] [CrossRef]
  41. Goetz, E.M.; Shankar, B.; Zou, Y.; Morales, J.C.; Luo, X.; Araki, S.; Bachoo, R.; Mayo, L.D.; Boothman, D.A. ATM dependent IGF-1 induction regulates secretory clusterin expression after DNA damage and in genetic instability. Oncogene Nat. Publ. Group 2011, 30, 3745–3754. [Google Scholar] [CrossRef]
  42. Michel, D.; Chatelain, G.; North, S.; Brun, G. Stress-induced transcription of the clusterin/apoJ gene. Biochem. J. 1997, 50, 45–50. [Google Scholar] [CrossRef]
  43. Trougakos, I.P. The Molecular Chaperone Apolipoprotein J/Clusterin as a Sensor of Oxidative Stress: Implications in Therapeutic Approaches—A Mini-Review. Gerontology 2013, 59, 514–523. [Google Scholar] [CrossRef] [PubMed]
  44. July, L.V.; Akbari, M.; Zellweger, T.; Jones, E.C.; Goldenberg, S.L.; Gleave, M.E. Clusterin expression is significantly enhanced in prostate cancer cells following androgen withdrawal therapy. Prostate 2002, 50, 179–188. [Google Scholar] [CrossRef] [PubMed]
  45. FitzGerald, P.C.; Shlyakhtenko, A.; Mir, A.A.; Vinson, C. Clustering of DNA sequences in human promoters. Genome Res. 2004, 14, 1562–1574. [Google Scholar] [CrossRef] [PubMed]
  46. Kim, N.; Choi, W.S. Proapoptotic role of nuclear clusterin in brain. Anat. Cell Biol. 2011, 44, 169–175. [Google Scholar] [CrossRef] [PubMed]
  47. Wong, P.; Taillefer, D.; Lakins, J.; Pineault, J.; Chader, G.; Tenniswood, M. Molecular characterization of human TRPM-2/clusterin, a gene associated with sperm maturation, apoptosis and neurodegeneration. Eur. J. Biochem. 1994, 221, 917–925. [Google Scholar] [CrossRef] [PubMed]
  48. Deb, M.; Sengupta, D.; Rath, S.K.; Kar, S.; Parbin, S.; Shilpi, A.; Pradhan, N.; Bhutia, S.K.; Roy, S.; Patra, S.K. Clusterin gene is predominantly regulated by histone modifications in human colon cancer and ectopic expression of the nuclear isoform induces cell death. Biochim. Biophys. Acta 2015, 1852, 1630–1645. [Google Scholar] [CrossRef]
  49. Hellebrekers, D.M.; Melotte, V.; Viré, E.; Langenkamp, E.; Molema, G.; Fuks, F.; Herman, J.G.; Van Criekinge, W.; Griffioen, A.W.; van Engeland, M. Identification of epigenetically silenced genes in tumor endothelial cells. Cancer Res. 2007, 67, 4138–4148. [Google Scholar] [CrossRef]
  50. Serrano, A.; Redondo, M.; Tellez, T.; Castro-Vega, I.; Roldan, M.J.; Mendez, R.; Rueda, A.; Jiménez, E. Regulation of clusterin expression in human cancer via DNA methylation. Tumour Biol. 2009, 30, 286–291. [Google Scholar] [CrossRef]
  51. Handy, D.E.; Castro, R.; Loscalzo, J. Epigenetic modifications: Basic mechanisms and role in cardiovascular disease. Circulation 2011, 123, 2145–2156. [Google Scholar] [CrossRef]
  52. Liao, F.T.; Lee, Y.J.; Ko, J.L.; Tsai, C.C.; Tseng, C.J.; Sheu, G.T. Hepatitis delta virus epigenetically enhances clusterin expression via histone acetylation in human hepatocellular carcinoma cells. J. Gen. Virol. 2009, 90, 1124–1134. [Google Scholar] [CrossRef]
  53. Rauhala, H.E.; Porkka, K.P.; Saramäki, O.R.; Tammela, T.L.; Visakorpi, T. Clusterin is epigenetically regulated in prostate cancer. Int. J. Cancer 2008, 123, 1601–1609. [Google Scholar] [CrossRef] [PubMed]
  54. Rosemblit, N.; Chen, C.L. Regulators for the rat clusterin gene: DNA methylation and cis-acting regulatory elements. J. Mol. Endocrinol. 1994, 13, 69–76. [Google Scholar] [CrossRef] [PubMed]
  55. Ambros, V. The functions of animal microRNAs. Nature 2004, 431, 350–355. [Google Scholar] [CrossRef]
  56. Mydlarz, W.; Uemura, M.; Ahn, S.; Hennessey, P.; Chang, S.; Demokan, S.; Sun, W.; Shao, C.; Bishop, J.; Krosting, J.; et al. Clusterin is a gene-specific target of microRNA-21 in head and neck squamous cell carcinoma. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2014, 20, 868–877. [Google Scholar] [CrossRef]
  57. Chen, X.; Jiang, Y.; Huang, Z.; Li, D.; Chen, X.; Cao, M.; Meng, Q.; Pang, H.; Sun, L.; Zhao, Y.; et al. miRNA-378 reverses chemoresistance to cisplatin in lung adenocarcinoma cells by targeting secreted clusterin. Sci. Rep. 2016, 6, 19455. [Google Scholar] [CrossRef]
  58. Buttyan, R.; Olsson, C.A.; Pintar, J.; Chang, C.; Bandyk, M.; Ng, P.Y.; Sawczuk, I.S. Induction of the TRPM-2 gene in cells undergoing programmed death. Mol. Cell Biol. 1989, 9, 3473–3481. [Google Scholar] [CrossRef]
  59. Leger, J.G.; Montpetit, M.L.; Tenniswood, M.P. Characterization and cloning of androgen-repressed mRNAs from rat ventral prostate. Biochem. Biophys. Res. Commun. 1987, 147, 196–203. [Google Scholar] [CrossRef]
  60. Leger, J.G.; Le Guellec, R.; Tenniswood, M.P. Treatment with antiandrogens induces an androgen-repressed gene in the rat ventral prostate. Prostate 1988, 13, 131–142. [Google Scholar] [CrossRef]
  61. Sensibar, J.A.; Sutkowski, D.M.; Raffo, A.; Buttyan, R.; Griswold, M.D.; Sylvester, S.R.; Kozlowski, J.M.; Lee, C. Prevention of cell death induced by tumor necrosis factor alpha in LNCaP cells by overexpression of sulfated glycoprotein-2 (clusterin). Cancer Res. 1995, 55, 2431–2437. [Google Scholar]
  62. Schwochau, G.B.; Nath, K.A.; Rosenberg, M.E. Clusterin protects against oxidative stress in vitro through aggregative and nonaggregative properties. Kidney Int. 1998, 53, 1647–1653. [Google Scholar] [CrossRef]
  63. Yang, C.R.; Leskov, K.; Hosley-Eberlein, K.; Criswell, T.; Pink, J.J.; Kinsella, T.J.; Boothman, D.A. Nuclear Clusterin/XIP8, an x-ray-induced Ku70-binding protein that signals cell death. Proc. Natl. Acad. Sci. USA 2000, 97, 5907–5912. [Google Scholar] [CrossRef] [PubMed]
  64. Leskov, K.S.; Araki, S.; Lavik, J.P.; Gomez, J.A.; Gama, V.; Gonos, E.S.; Trougakos, I.P.; Matsuyama, S.; Boothman, D.A. CRM1 protein-mediated regulation of nuclear clusterin (nCLU), an ionizing radiation-stimulated, Bax-dependent pro-death factor. J. Biol. Chem. 2011, 286, 40083–40090. [Google Scholar] [CrossRef]
  65. Essabbani, A.; Garcia, L.; Zonetti, M.J.; Fisco, T.; Pucci, S.; Chiocchia, G. Exon-skipping strategy by ratio modulation between cytoprotective versus pro-apoptotic clusterin forms increased sensitivity of LNCaP to cell death. PLoS ONE 2013, 8, e54920. [Google Scholar] [CrossRef] [PubMed]
  66. García-Aranda, M.; Serrano, A.; Redondo, M. Regulation of Clusterin Gene Expression. Curr. Protein Pept. Sci. 2017, 19, 612–622. [Google Scholar] [CrossRef] [PubMed]
  67. Kim, N.; Yoo, J.C.; Han, J.Y.; Hwang, E.M.; Kim, Y.S.; Jeong, E.Y.; Sun, C.H.; Yi, G.S.; Roh, G.S.; Kim, H.J.; et al. Human nuclear clusterin mediates apoptosis by interacting with Bcl-XL through C-terminal coiled coil domain. J. Cell Physiol. 2012, 227, 1157–1167. [Google Scholar] [CrossRef] [PubMed]
  68. Shannan, B.; Seifert, M.; Leskov, K.; Willis, J.; Boothman, D.; Tilgen, W.; Reichrath, J. Challenge and promise: Roles for clusterin in pathogenesis, progression and therapy of cancer. Cell Death Differ. 2006, 13, 12–19. [Google Scholar] [CrossRef] [PubMed]
  69. Zhang, Y.; Lv, X.; Chen, L.; Liu, Y. The role and function of CLU in cancer biology and therapy. Clin. Exp. Med. 2023, 23, 1375–1391. [Google Scholar] [CrossRef]
  70. Trougakos, I.P.; Lourda, M.; Antonelou, M.H.; Kletsas, D.; Gorgoulis, V.G.; Papassideri, I.S.; Zou, Y.; Margaritis, L.H.; Boothman, D.A.; Gonos, E.S. Intracellular clusterin inhibits mitochondrial apoptosis by suppressing stress signals activating p53 and stabilizing the cytosolic Ku70-Bax protein complex. Clin. Cancer Res. 2009, 15, 48–59. [Google Scholar] [CrossRef]
  71. Blume, A.J.; Foster, C.J. Adenylate cyclase of mouse neuroblastoma cells: Regulation by 2-chloroadenosine, prostaglandin E1, and the cations Mg2+, Ca2+, and Mn2+. J. Neurochem. 1976, 26, 305–311. [Google Scholar] [CrossRef]
  72. Nicholson, G.; Lawrence, A.; Ala, F.A.; Bird, G.W. Semi-quantitative assay of antigen site density by flow cytometry analysis. Transfus. Med. 1991, 1, 87–90. [Google Scholar] [CrossRef]
  73. Dews, M.; Fox, J.L.; Hultine, S.; Sundaram, P.; Wang, W.; Liu, Y.Y.; Furth, E.; Enders, G.H.; El-Deiry, W.; Schelter, J.M.; et al. The myc-miR-17~92 axis blunts TGF{beta} signaling and production of multiple TGF{beta}-dependent antiangiogenic factors. Cancer Res. 2010, 70, 8233–8246. [Google Scholar] [CrossRef] [PubMed]
  74. Huang, H.; Wang, L.; Li, M.; Wang, X.; Zhang, L. Secreted clusterin (sCLU) regulates cell proliferation and chemosensitivity to cisplatin by modulating ERK1/2 signals in human osteosarcoma cells. World J. Surg. Oncol. 2014, 12, 255. [Google Scholar] [CrossRef]
  75. Wang, X.; Xie, J.; Lu, X.; Li, H.; Wen, C.; Huo, Z.; Xie, J.; Shi, M.; Tang, X.; Chen, H.; et al. Melittin inhibits tumor growth and decreases resistance to gemcitabine by downregulating cholesterol pathway gene CLU in pancreatic ductal adenocarcinoma. Cancer Lett. 2017, 399, 1–9. [Google Scholar] [CrossRef] [PubMed]
  76. Chesnokova, V.; Zonis, S.; Wawrowsky, K.; Tani, Y.; Ben-Shlomo, A.; Ljubimov, V.; Mamelak, A.; Bannykh, S.; Melmed, S. Clusterin and FOXL2 act concordantly to regulate pituitary gonadotroph adenoma growth. Mol. Endocrinol. 2012, 26, 2092–2103. [Google Scholar] [CrossRef] [PubMed]
  77. Deng, J.; Peng, M.; Zhou, S.; Xiao, D.; Hu, X.; Xu, S.; Wu, J.; Yang, X. Metformin targets Clusterin to control lipogenesis and inhibit the growth of bladder cancer cells through SREBP-1c/FASN axis. Signal Transduct. Target. Ther. 2021, 6, 98. [Google Scholar] [CrossRef] [PubMed]
  78. Whitesell, L.; Lindquist, S.L. HSP90 and the chaperoning of cancer. Nat. Rev. Cancer 2005, 5, 761–772. [Google Scholar] [CrossRef] [PubMed]
  79. Shim, Y.-J.; Kang, B.-H.; Jeon, H.-S.; Park, I.-S.; Lee, K.-U.; Lee, I.-K.; Park, G.-H.; Lee, K.-M.; Schedin, P.; Min, B.-H. Clusterin induces matrix metalloproteinase-9 expression via ERK1/2 and PI3K/Akt/NF-κB pathways in monocytes/macrophages. J. Leukoc. Biol. 2011, 90, 761–769. [Google Scholar] [CrossRef] [PubMed]
  80. Li, Y.; Lu, J.; Zhou, S.; Wang, W.; Tan, G.; Zhang, Z.; Dong, Z.; Kang, T.; Tang, F. N, N′-dinitrosopiperazine-induced clusterin participates in nasopharyngeal carcinoma metastasis. Oncotarget 2016, 7, 5548–5563. [Google Scholar] [CrossRef] [PubMed]
  81. Bailes, J.; Soloviev, M. Insulin-Like Growth Factor-1 (IGF-1) and Its Monitoring in Medical Diagnostic and in Sports. Biomolecules 2021, 11, 217. [Google Scholar] [CrossRef]
  82. Tian, Y.; Wang, C.; Chen, S.; Liu, J.; Fu, Y.; Luo, Y. Extracellular Hsp90alpha and clusterin synergistically promote breast cancer epithelial-to-mesenchymal transition and metastasis through LRP1. J. Cell Sci. 2019, 132, jcs228213. [Google Scholar] [CrossRef]
  83. Viard, I.; Wehrli, P.; Jornot, L.; Bullani, R.; Vechietti, J.-L.; French, L.E.; Schifferli, J.A.; Tschopp, J. Clusterin gene expression mediates resistance to apoptotic cell death induced by heat shock and oxidative stress. J. Investg. Dermatol. 1999, 112, 290–296. [Google Scholar] [CrossRef] [PubMed]
  84. Cervellera, M.; Raschella, G.; Santilli, G.; Tanno, B.; Ventura, A.; Mancini, C.; Sevignani, C.; Calabretta, B.; Sala, A. Direct transactivation of the antiapoptotic gene apolipoprotein J (Clusterin) by B-MYB. J. Biol. Chem. 2000, 275, 21055–21060. [Google Scholar] [CrossRef] [PubMed]
  85. Ma, X.; Bai, Y. IGF-1 activates the P13K/AKT signaling pathway by positively regulating secretory clusterin. Mol. Med. Rep. 2012, 6, 1433–1437. [Google Scholar] [CrossRef] [PubMed]
  86. Flanagan, L.; Whyte, L.; Chatterjee, N.; Tenniswood, M. Effects of clusterin overexpression on metastatic progression and therapy in breast cancer. BMC Cancer 2010, 10, 107. [Google Scholar] [CrossRef] [PubMed]
  87. Meunier, A.; Cornet, F.; Campos, M. Bacterial cell proliferation: From molecules to cells. FEMS Microbiol. Rev. 2020, 45, fuaa046. [Google Scholar] [CrossRef]
  88. Wang, Z. Regulation of cell cycle progression by growth factor-induced cell signaling. Cells 2021, 10, 33270. [Google Scholar] [CrossRef] [PubMed]
  89. Suzuki, S.; Shiraga, K.; Sato, S.; Punfa, W.; Naiki-Ito, A.; Yamashita, Y.; Shirai, T.; Takahashi, S. Apocynin, an NADPH oxidase inhibitor, suppresses rat prostate carcinogenesis. Cancer Sci. 2013, 104, 1711–1717. [Google Scholar] [CrossRef]
  90. Chen, Z.; Fan, Z.; Dou, X.; Zhou, Q.; Zeng, G.; Liu, L.; Chen, W.; Lan, R.; Liu, W.; Ru, G.; et al. Inactivation of tumor suppressor gene Clusterin leads to hyperactivation of TAK1-NF-kappaB signaling axis in lung cancer cells and denotes a therapeutic opportunity. Theranostics 2020, 10, 11520–11534. [Google Scholar] [CrossRef]
  91. Chun, Y.J. Knockdown of clusterin expression increases the in vitro sensitivity of human prostate cancer cells to paclitaxel. J. Toxicol. Environ. Health A 2014, 77, 1443–1450. [Google Scholar] [CrossRef]
  92. Zheng, W.; Yao, M.; Qian, Q.; Sai, W.; Qiu, L.; Yang, J.; Wu, W.; Dong, Z.; Yao, D. Oncogenic secretory clusterin in hepatocellular carcinoma: Expression at early staging and emerging molecular target. Oncotarget 2016, 8, 52321–52332. [Google Scholar] [CrossRef]
  93. Takeuchi, A.; Shiota, M.; Beraldi, E.; Thaper, D.; Takahara, K.; Ibuki, N.; Pollak, M.; Cox, M.E.; Naito, S.; Gleave, M.E.; et al. Insulin-like growth factor-I induces CLU expression through Twist1 to promote prostate cancer growth. Mol Cell Endocrinol. 2014, 384, 117–125. [Google Scholar] [CrossRef] [PubMed]
  94. Liu, Y.; Men, C.; Xu, Y.; Zhao, K.; Luo, L.; Dong, D.; Yu, Q. Clusterin promotes growth and invasion of clear cell renal carcinoma cell by upregulation of S100A4 expression. Cancer Biomark. 2018, 21, 915–923. [Google Scholar] [CrossRef]
  95. Pajak, B.; Kania, E.; Gajkowska, B.; Orzechowski, A. Lipid rafts mediate epigallocatechin-3-gallate- and green tea extract-dependent viability of human colon adenocarcinoma COLO 205 cells; clusterin affects lipid rafts-associated signaling pathways. J. Physiol. Pharmacol. Off. J. Pol. Physiol. Soc. 2011, 62, 449–459. [Google Scholar]
  96. Tang, Z.; Wang, W.; Liu, Z.; Sun, X.; Liao, Z.; Chen, F.; Jiang, G.; Huo, G. Blocking ERK signaling pathway reduces MMP-9 expression to alleviate cerebral edema after traumatic brain injury in rats. J. South. Med. Univ. 2020, 40, 1018–1022. [Google Scholar]
  97. Yang, P.; Yang, Z.; Dong, Y.; Yang, L.; Peng, S.; Yuan, L.; Hu, X.; Chen, S.; Tang, H.; Yang, X.; et al. Clusterin is a prognostic biomarker of breast cancer and correlates with the immune microenvironment. Transl. Cancer Res. 2023, 12, 31–45. [Google Scholar] [CrossRef] [PubMed]
  98. Yang, S.; Tang, W.; Azizian, A.; Gaedcke, J.; Strobel, P.; Wang, L.; Cawley, H.; Ohara, Y.; Valenzuela, P.; Zhang, L.; et al. Dysregulation of the HNF1B/Clusterin axis enhances disease progression in a highly aggressive subset of pancreatic cancer patients. Carcinogenesis 2022, 43, 1198–1210. [Google Scholar] [CrossRef]
  99. Mansoori, B.; Mohammadi, A.; Davudian, S.; Shirjang, S.; Baradaran, B. Different mechanisms of resistance to cancer drugs: A brief review. Adv. Pharm. Bull. 2017, 7, 339–348. [Google Scholar] [CrossRef]
  100. Zhang, H.; Kim, J.K.; Edwards, C.A.; Xu, Z.; Taichman, R.; Wang, C.-Y. Clusterin inhibits apoptosis by interacting with activated Bax. Nat. Cell Biol. 2005, 7, 909–915. [Google Scholar] [CrossRef]
  101. Tang, M.; Li, J.; Liu, B.; Song, N.; Wang, Z.; Yin, C. Clusterin expression in human testicular seminoma. Med. Hypotheses 2013, 81, 635–637. [Google Scholar] [CrossRef]
  102. Wang, X.; Zou, F.; Zhong, J.; Yue, L.; Wang, F.; Wei, H.; Yang, G.; Jin, T.; Dong, X.; Li, J.; et al. Secretory clusterin mediates resistance to oxaliplatin through the Gadd45a/PI3K/Akt signaling pathway in hepatocellular carcinoma. J. Cancer 2018, 9, 1403–1413. [Google Scholar] [CrossRef]
  103. Panico, F.; Rizzi, F.; Fabbri, L.M.; Bettuzzi, S.; Luppi, F. Clusterin (CLU) and lung cancer. Adv. Cancer Res. 2009, 105, 63–76. [Google Scholar]
  104. July, L.V.; Beraldi, E.; So, A.; Fazli, L.; Evans, K.; English, J.C.; Gleave, M.E. Nucleotide-based therapies targeting clusterin chemosensitize human lung adenocarcinoma cells both in vitro and in vivo. Mol. Cancer Ther. 2004, 3, 223–232. [Google Scholar] [CrossRef]
  105. Bi, J.; Guo, A.L.; Lai, Y.R.; Li, B.; Zhong, J.M.; Wu, H.Q.; Xie, Z.; He, Y.L.; Lv, Z.L.; Lau, S.H.; et al. Overexpression of clusterin correlates with tumor progression, metastasis in gastric cancer: A study on tissue microarrays. Neoplasm 2010, 57, 191. [Google Scholar] [CrossRef]
  106. Xie, D.; Lau, S.H.; Sham, J.S.; Wu, Q.L.; Fang, Y.; Liang, L.Z.; Che, L.H.; Zeng, Y.X.; Guan, X.Y. Up-regulated expression of cytoplasmic clusterin in human ovarian carcinoma. Cancer 2005, 103, 277–283. [Google Scholar] [CrossRef]
  107. Yang, G.F.; Li, X.M.; Xie, D. Overexpression of clusterin in ovarian cancer is correlated with impaired survival. Int. J. Gynecol. Cancer Off. J. Int. Gynecol. Cancer Soc. 2009, 19, 1342–1346. [Google Scholar] [CrossRef] [PubMed]
  108. Al-Maghrabi, J.A.; Butt, N.S.; Anfinan, N.; Sait, K.; Sait, H.; Bajouh, O.; Khabaz, M.N. Clusterin immunoexpression is associated with early stage endometrial carcinomas. Acta Histochemical. 2016, 118, 430–434. [Google Scholar] [CrossRef]
  109. Fuzio, P.; Valletti, A.; Napoli, A.; Napoli, G.; Cormio, G.; Selvaggi, L.; Liuni, S.; Pesole, G.; Maiorano, E.; Perlino, E. Regulation of the expression of CLU isoforms in endometrial proliferative diseases. Int. J. Oncol. 2013, 42, 1929–1944. [Google Scholar] [CrossRef]
  110. Won, Y.S.; Lee, S.J.; Yeo, S.G.; Park, D.C. Effects of female sex hormones on clusterin expression and paclitaxel resistance in endometrial cancer cell lines. Int. J. Med. Sci. 2012, 9, 86–92. [Google Scholar] [CrossRef] [PubMed]
  111. Park, D.C.; Yeo, S.G.; Shin, E.Y.; Mok, S.C.; Kim, D.H. Clusterin confers paclitaxel resistance in cervical cancer. Gynecol. Oncol. 2006, 103, 996–1000. [Google Scholar] [CrossRef]
  112. Niu, Z.; Li, X.; Hu, B.; Li, R.; Wang, L.; Wu, L.; Wang, X. Small interfering RNA targeted to secretory clusterin blocks tumor growth, motility, and invasion in breast cancer. Acta Biochim. Biophys. Sin. 2012, 44, 991–998. [Google Scholar] [CrossRef]
  113. Li, J.; Jia, L.; Zhao, P.; Jiang, Y.; Zhong, S.; Chen, D. Stable knockdown of clusterin by vectorbased RNA interference in a human breast cancer cell line inhibits tumour cell invasion and metastasis. J. Int. Med. Res. 2012, 40, 545–555. [Google Scholar] [CrossRef] [PubMed]
  114. Redondo, M.; Villar, E.; Torres-Muñoz, J.; Tellez, T.; Morell, M.; Petito, C.K. Overexpression of clusterin in human breast carcinoma. Am. J. Pathol. 2000, 157, 393–399. [Google Scholar] [CrossRef] [PubMed]
  115. Zhang, D.; Sun, B.; Zhao, X.; Cui, Y.; Xu, S.; Dong, X.; Zhao, J.; Meng, J.; Jia, X.; Chi, J. Secreted CLU is associated with the initiation of triple-negative breast cancer. Cancer Biol. Ther. 2012, 13, 321–329. [Google Scholar] [CrossRef] [PubMed]
  116. Niu, Z.H.; Wang, Y.; Chun, B.; Li, C.X.; Wu, L. Secretory clusterin (sCLU) overexpression is associated with resistance to preoperative neoadjuvant chemotherapy in primary breast cancer. Eur. Rev. Med. Pharmacol. Sci. 2013, 17, 1337–1344. [Google Scholar] [PubMed]
  117. Redondo, M.; Tellez, T.; Roldan, M.J. The role of clusterin (CLU) in malignant transformation and drug resistance in breast carcinomas. Adv. Cancer Res. 2009, 105, 21–43. [Google Scholar] [PubMed]
  118. Mazzarelli, P.; Pucci, S.; Spagnoli, L.G. CLU and colon cancer. The dual face of CLU: From the normal to the malignant phenotype. Adv. Cancer Res. 2009, 105, 45–61. [Google Scholar] [PubMed]
  119. Redondo, M.; Rodrigo, I.; Alcaide, J.; Tellez, T.; Roldan, M.J.; Funez, R.; Diaz-Martin, A.; Rueda, A.; Jiménez, E. Clusterin expression is associated with decreased disease-free survival of patients with colorectal carcinomas. Histopathology 2010, 56, 932–936. [Google Scholar] [CrossRef]
  120. Sun, B.; Moibi, J.A.; Mak, A.; Xiao, Z.; Roa, W.; Moore, R.B. Response of bladder carcinoma cells to TRAIL and antisense oligonucleotide, Bcl-2 or clusterin treatments. J. Urol. 2009, 181, 1361–1371. [Google Scholar] [CrossRef]
  121. Miyake, H.; Gleave, M.; Kamidono, S.; Hara, I. Overexpression of clusterin in transitional cell carcinoma of the bladder is related to disease progression and recurrence. Urology 2002, 59, 150–154. [Google Scholar] [CrossRef]
  122. Hazzaa, S.M.; Elashry, O.M.; Afifi, I.K. Clusterin as a diagnostic and prognostic marker for transitional cell carcinoma of the bladder. Pathol. Oncol. Res. POR 2010, 16, 101–109. [Google Scholar] [CrossRef]
  123. Lau, S.H.; Sham, J.S.; Xie, D.; Tzang, C.H.; Tang, D.; Ma, N.; Hu, L.; Wang, Y.; Wen, J.M.; Xiao, G.; et al. Clusterin plays an important role in hepatocellular carcinoma metastasis. Oncogene 2006, 25, 1242–1250. [Google Scholar] [CrossRef]
  124. Xiu, P.; Dong, X.; Dong, X.; Xu, Z.; Zhu, H.; Liu, F.; Wei, Z.; Zhai, B.; Kanwar, J.R.; Jiang, H.; et al. Secretory clusterin contributes to oxaliplatin resistance by activating Akt pathway in hepatocellular carcinoma. Cancer Sci. 2013, 104, 375–382. [Google Scholar] [CrossRef]
  125. Patarat, R.; Riku, S.; Kunadirek, P.; Chuaypen, N.; Tangkijvanich, P.; Mutirangura, A.; Puttipanyalears, C. The expression of FLNA and CLU in PBMCs as a novel screening marker for hepatocellular carcinoma. Sci. Rep. 2021, 11, 14838. [Google Scholar] [CrossRef]
  126. Gao, G.; Luan, X. Diagnostic performance of clusterin in hepatocellular carcinoma: A meta-analysis. Int. J. Biol. Markers 2022, 37, 404–411. [Google Scholar] [CrossRef]
  127. Rasmy, H.S.; Mohammed, H.A.; Mohammed, E.S.; Ahmed, A.S.M.; Isaac, A. Serum clusterin as a promising diagnostic and prognostic marker for hepatocellular carcinoma after locoregional treatment. Egypt. J. Immunol. 2022, 29, 26–40. [Google Scholar] [CrossRef]
  128. Li, Y.; Wang, X.; Chen, Y.H.; Tan, Q.Q.; Liu, X.B.; Tan, C. Clusterin is upregulated by erastin, a ferroptosis inducer and exerts cytoprotective effects in pancreatic adenocarcinoma cells. Anti-Cancer Drugs 2023, 35, 227–236. [Google Scholar] [CrossRef] [PubMed]
  129. Xie, M.J.; Motoo, Y.; Su, S.B.; Mouri, H.; Ohtsubo, K.; Matsubara, F.; Sawabu, N. Expression of clusterin in human pancreatic cancer. Pancreas 2002, 25, 234–238. [Google Scholar] [CrossRef] [PubMed]
  130. Hoeller, C.; Pratscher, B.; Thallinger, C.; Winter, D.; Fink, D.; Kovacic, B.; Sexl, V.; Wacheck, V.; Gleave, M.E.; Pehamberger, H.; et al. Clusterin regulates drug-resistance in melanoma cells. J. Investig. Dermatol. 2005, 124, 1300–1307. [Google Scholar] [CrossRef] [PubMed]
  131. He, L.R.; Liu, M.Z.; Li, B.K.; Rao, H.L.; Liao, Y.J.; Zhang, L.J.; Guan, X.Y.; Zeng, Y.X.; Xie, D. Clusterin as a predictor for chemoradiotherapy sensitivity and patient survival in esophageal squamous cell carcinoma. Cancer Sci. 2009, 100, 2354–2360. [Google Scholar] [CrossRef]
  132. Bijian, K.; Mlynarek, A.M.; Balys, R.L.; Jie, S.; Xu, Y.; Hier, M.P.; Black, M.J.; Di Falco, M.R.; LaBoissiere, S.; Alaoui-Jamali, M.A. Serum proteomic approach for the identification of serum biomarkers contributed by oral squamous cell carcinoma and host tissue microenvironment. J. Proteome Res. 2009, 8, 2173–2185. [Google Scholar] [CrossRef] [PubMed]
  133. Redondo, M.; Fùnez, R.; Esteban, F. Apoptosis in the Development and Treatment of Laryngeal Cancer: Role of p53, Bcl-2 and Clusterin. In Apoptosis in Carcinogenesis and Chemotherapy; Chen, G.G., Lai, P.B., Eds.; Springer: Dordrecht, The Netherlands, 2009. [Google Scholar]
  134. Wellmann, A.; Thieblemont, C.; Pittaluga, S.; Sakai, A.; Jaffe, E.S.; Siebert, P.; Raffeld, M. Detection of differentially expressed genes in lymphomas using cDNA arrays: Identification of clusterin as a new diagnostic marker for anaplastic large-cell lymphomas. Blood 2000, 96, 398–404. [Google Scholar] [CrossRef]
  135. Ma, J.; Gao, W.; Gao, J. sCLU as prognostic biomarker and therapeutic target in osteosarcoma. Bioengineered 2019, 10, 229–239. [Google Scholar] [CrossRef] [PubMed]
  136. Rizzi, F.; Bettuzzi, S. Clusterin (CLU) and prostate cancer. Adv. Cancer Res. 2009, 105, 1–19. [Google Scholar] [PubMed]
  137. Téllez, T.; Martin-García, D.; Redondo, M.; García-Aranda, M. Clusterin Expression in Colorectal Carcinomas. Int. J. Mol. Sci. 2023, 24, 14641. [Google Scholar] [CrossRef] [PubMed]
  138. Miyake, H.; Gleave, M.E.; Arakawa, S.; Kamidono, S.; Hara, I. Introducing the clusterin gene into human renal cell carcinoma cells enhances their metastatic potential. J. Urol. 2002, 167, 2203–2208. [Google Scholar] [CrossRef]
  139. Shi, H.; Deng, J.H.; Wang, Z.; Cao, K.Y.; Zhou, L.; Wan, H. Knockdown of clusterin inhibits the growth and migration of renal carcinoma cells and leads to differential gene expression. Mol. Med. Rep. 2013, 8, 35–40. [Google Scholar] [CrossRef] [PubMed]
  140. Beheshti Namdar, A.; Kabiri, M.; Mosanan Mozaffari, H.; Aminifar, E.; Mehrad-Majd, H. Circulating clusterin levels and cancer risk: A systematic review and meta-analysis. Cancer Control 2022, 29. [Google Scholar] [CrossRef] [PubMed]
  141. Nishikawa, M.; Miyake, H.; Gleave, M.; Fujisawa, M. Effect of targeting clusterin using OGX-011 on antitumor activity of temsirolimus in a human renal cell carcinoma model. Target Oncol. 2016, 12, 69–79. [Google Scholar] [CrossRef]
  142. Yamanaka, K.; Gleave, M.; Muramaki, M.; Hara, I.; Miyake, H. Enhanced radiosensitivity by inhibition of the anti-apoptotic gene clusterin using antisense oligodeoxynucleotide in a human bladder cancer model. Oncol. Rep. 2005, 13, 885. [Google Scholar] [CrossRef]
  143. Chen, Q.; Wang, Z.; Zhang, K.; Liu, X.; Cao, W.; Zhang, L.; Zhang, S.; Yan, B.; Wang, Y.; Xia, C. Clusterin confers gemcitabine resistance in pancreatic cancer. World J. Surgical. Oncol. 2011, 9, 59. [Google Scholar] [CrossRef]
  144. Cao, C.; Shinohara, E.T.; Li, H.; Niermann, K.J.; Kim, K.W.; Sekhar, K.R.; Gleave, M.; Freeman, M.; Lu, B. Clusterin as a therapeutic target for radiation sensitization in a lung cancer model. Int. J. Radiat. Oncol. Biol. Phys. 2005, 63, 1228–1236. [Google Scholar] [CrossRef]
  145. Zellweger, T.; Chi, K.; Miyake, H.; Adomat, H.; Kiyama, S.; Skov, K.; Gleave, M.E. Enhanced radiation sensitivity in prostate cancer by inhibition of the cell survival protein clusterin. Clin Cancer Res. 2002, 8, 3276–3284. [Google Scholar] [PubMed]
  146. So, A.; Chi, K.; Miyake, H.; Adomat, H.; Kiyama, S.; Skov, K.; Gleave, M.E. Knockdown of the cytoprotective chaperone, clusterin, chemosensitizes human breast cancer cells both in vitro and in vivo. Mol. Cancer Ther. 2005, 4, 1837–1849. [Google Scholar] [CrossRef] [PubMed]
  147. Zwain, I.; Amato, P. Clusterin protects granulosa cells from apoptotic cell death during follicular atresia. Exp. Cell Res. 2000, 257, 101–110. [Google Scholar] [CrossRef] [PubMed]
  148. Kususda, Y.; Miyake, H.; Gleave, M.E.; Fujisawa, M. Clusterin inhibition using OGX-011 synergistically enhances antitumour activity of sorafenib in a human renal cell carcinoma model. Br. J. Cancer 2012, 106, 1945–1952. [Google Scholar] [CrossRef]
  149. Hassan, M.K.; Watari, H.; Han, Y.; Mitamura, T.; Hosaka, M.; Wang, L.; Tanaka, S.; Sakuragi, N. Clusterin is a potential molecular predictor for ovarian cancer patient’s survival: Targeting clusterin improves response to paclitaxel. J. Exp. Clin. Cancer Res. CR 2011, 30, 113. [Google Scholar] [CrossRef]
  150. Blumenstein, B.; Saad, F.; Hotte, S.; Chi, K.N.; Eigl, B.; Gleave, M.; Jacobs, C. Reduction in serum clusterin is a potential therapeutic biomarker in patients with castration-resistant prostate cancer treated with custirsen. Cancer Med. 2013, 2, 468–477. [Google Scholar] [CrossRef] [PubMed]
  151. Higano, C.S. Potential use of custirsen to treat prostate cancer. OncoTargets Ther. 2013, 6, 785–797. [Google Scholar] [CrossRef]
  152. Chi, K.N.; Hotte, S.J.; Yu, E.Y.; Tu, D.; Eigl, B.J.; Tannock, I.; Saad, F.; North, S.; Powers, J.; Gleave, M.E.; et al. Randomized phase II study of docetaxel and prednisone with or without OGX-011 in patients with metastatic castration-resistant prostate cancer. J. Clin. Oncol. 2010, 28, 4247–4254. [Google Scholar] [CrossRef]
  153. Chia, S.; Dent, S.; Ellard, S.; Ellis, P.M.; Vandenberg, T.; Gelmon, K.; Powers, J.; Walsh, W.; Seymour, L.; Eisenhauer, E.A. Phase II trial of OGX-011 in combination with docetaxel in metastatic breast cancer. Clin. Cancer Res. 2009, 15, 708–713. [Google Scholar] [CrossRef]
  154. Laskin, J.J.; Nicholas, G.; Lee, C.; Gitlitz, B.; Vincent, M.; Cormier, Y.; Stephenson, J.; Ung, Y.; Sanborn, R.; Pressnail, B.; et al. Phase I/II trial of custirsen (OGX-011), an inhibitor of clusterin, in combination with a gemcitabine and platinum regimen in patients with previously untreated advanced non-small cell lung cancer. J. Thorac. Oncol. 2012, 7, 579–586. [Google Scholar] [CrossRef]
  155. Ma, X.; Zou, L.; Li, X.; Chen, Z.; Lin, Q.; Wu, X. MicroRNA-195 regulates docetaxel resistance by targeting clusterin in prostate cancer. Biomed. Pharmacother. 2018, 99, 445–450. [Google Scholar] [CrossRef] [PubMed]
  156. Zhao, W.; Wang, X.; Jiang, Y.; Jia, X.; Guo, Y. miR-217-5p inhibits invasion and metastasis of prostate cancer by targeting clusterin. Mamm Genome. 2021, 32, 371–380. [Google Scholar] [CrossRef] [PubMed]
  157. Davalli, P.; Rizzi, F.; Caldara, G.F.; Davoli, S.; Corti, A.; Silva, A.; Astancolle, S.; Vitale, M.; Bettuzzi, S.; Arcari, M.; et al. Chronic administration of green tea extract to TRAMP mice induces the collapse of Golgi apparatus in prostate secretory cells and results in alterations of protein post-translational processing. Int. J. Oncol. 2011, 39, 1521–1527. [Google Scholar] [PubMed]
  158. Zhu, Y.; Chen, P.; Gao, Y.; Ta, N.; Zhang, Y.; Cai, J.; Zhao, Y.; Liu, S.; Zheng, J. MEG3 activated by Vitamin D inhibits colorectal cancer cells proliferation and migration via regulating clusterin. EBioMedicine 2018, 30, 148–157. [Google Scholar] [CrossRef]
  159. Xiong, J.; Wang, S.; Chen, T.; Shu, X.; Mo, X.; Chang, G.; Chen, J.J.; Li, C.; Luo, H.; Lee, J.D. Verteporfin blocks Clusterin which is required for survival of gastric cancer stem cell by modulating HSP90 function. Int. J. Biol. Sci. 2019, 15, 312–324. [Google Scholar] [CrossRef]
  160. Available online: https://clinicaltrials.gov/study/NCT02412462?intr=clusterin&page=2&rank=11 (accessed on 10 January 2024).
Figure 1. Generation of sCLU. Initially, psCLU is conveyed to the endoplasmic reticulum (ER) through the leader signaling peptide. During transit to the Golgi apparatus, it undergoes cleavage and glycosylation processes. This results in an 80 kDa protein composed of alpha and beta subunits joined by disulfide bonds, ultimately being secreted from the cell. Images were created using Biorender.com (accessed on 9 March 2024).
Figure 1. Generation of sCLU. Initially, psCLU is conveyed to the endoplasmic reticulum (ER) through the leader signaling peptide. During transit to the Golgi apparatus, it undergoes cleavage and glycosylation processes. This results in an 80 kDa protein composed of alpha and beta subunits joined by disulfide bonds, ultimately being secreted from the cell. Images were created using Biorender.com (accessed on 9 March 2024).
Cells 13 00665 g001
Figure 2. Generation of sCLU pnCLU does not undergo any excision or glycosylation process; it is localized in the cytoplasm of unstressed cells. pnCLU is converted into a mature form inside the nucleus, nCLU. Images were created using Biorender.com (accessed on 9 March 2024).
Figure 2. Generation of sCLU pnCLU does not undergo any excision or glycosylation process; it is localized in the cytoplasm of unstressed cells. pnCLU is converted into a mature form inside the nucleus, nCLU. Images were created using Biorender.com (accessed on 9 March 2024).
Cells 13 00665 g002
Table 1. Biological processes involving CLU isoforms.
Table 1. Biological processes involving CLU isoforms.
Biological ProcessesnCLUsCLU
TumorigenesisThe proapoptotic isoform nCLU of CLU induces apoptosis in breast and prostate cancer, through specific interactions with proteins like Ku70 and Bcl-XL [70,71].sCLU also plays a crucial role by interacting with protein complexes like Ku70-bax, acting as a Bax retention factor in the cytosol, inhibiting its proapoptotic function. Under normal conditions, inhibition of CLU weakens this complex, allowing Bax to translocate to the mitochondria, triggering cytochrome c release, and activating caspase 9, initiating apoptosis [72].
Cell Proliferationc-Myc, a transcription factor encoded by the oncogene MYC involved in tumorigenesis, inhibits the expression of nCLU by upregulating the microRNA cluster miRNA-17 ~ 92 and attenuating the TGF-β axis, thus promoting angiogenesis and tumor growth in colon cancer [69].Blocking sCLU using apocynin, a substance that inhibits NADPH oxidase, halts the MEK-ERK1/2 pathway, resulting in reduced cell proliferation [73].
Melittin inhibits sCLU, inactivating both the cholesterol/NF-κB/Bcl-2 axis and the cholesterol/p-ERK axis to suppress tumor growth in pancreatic cancer [74].
sCLU exhibits an antiproliferative property by inactivating the TAK1/NF-κB axis, preventing the transforming growth factor beta receptor 1 (TGFBR1) from recruiting the TNF receptor-activating factor 6 (TRAF6)/TAK1-binding protein 2 (TAB2)/TGF-β-activated kinase 1 (TAK1) complex to inhibit tumor proliferation and growth in human non-small cell lung cancer (NSCLC) [75].
Metformin, a conventional medication for type II diabetes, exhibits antitumor effects. Metformin suppresses sCLU, thereby hindering tumor growth through the inactivation of the SREBP-1c/fatty acid synthase (FASN) axis [76].
sCLU promotes cell growth and proliferation by upregulating the expression of the calcium-binding protein S100A4 in renal cancer [77].
Chemoresistance and Chemosensitivity The stress response induced by treatments such as radiotherapy and chemotherapy lead to the overexpression of sCLU, a cytoprotective chaperone, which, by binding to activated Bax, it impedes the discharge of cytochrome c and prevents apoptosis [78].
Epithelial–Mesenchymal Transition and Metastasis In nasopharyngeal carcinoma, CLU undergoes positive regulation by N, N′-dinitrosopiperazine (DNP), triggering MMP-9 and VEGF expression, thus facilitating to metastasis [79].
In breast cancer, CLU collaborates with eHsp90α to activate key signaling pathways, promoting EMT, migration, and tumor metastasis [80].
In colon cancer, CLU interaction with platelets activates the p38MAPK pathway and positively regulates MMP-9, facilitating invasion [81].
Studies on prostate cancer it has been demonstrated that miRNA-217-5p exerts control over the processes of invasion and migration by specifically targeting CLU [82].
Table 2. Clusterin as a potential biomarker in various types of cancer.
Table 2. Clusterin as a potential biomarker in various types of cancer.
Types of CancerExpression of CLU In VitroExpression of CLU In Vivo
Non-small cell lungNon-small cell lung cancer cell lines show overexpression upon treatment with chemotherapy or radiotherapy. ASO therapy sensitizes cells to these treatments and decreases their metastatic potential [103]Patients exhibiting positive CLU expression tend to experience improved overall disease-free survival compared to those with negative CLU expression [104].
More than 80% of the tumors are immunoreactive for CLU [104].
Gastric Overexpression of sCLU correlates significantly with metastasis, tumor invasion, and TNM stage. In addition, it correlates with unfavorable survival for advanced-stage gastric cancers [105].
Ovarian Elevated sCLU levels show an inverse relationship with the tumor apoptotic index and are detected more frequently in metastatic lesions than in primary tumors [106].
Increased sCLU expression is associated with increased biological aggressiveness and decreased survival [107].
Endometrial When CLU is expressed in endometrial tumors, it is associated with a lower stage, supporting its role in the diagnosis of endometrial carcinoma [108].
There has been detected higher mRNA expression in both neoplastic and hyperplastic tissues compared to a normal endometrium. In this regard, an increase in mRNA expression of the specific sCLU isoform has been observed in neoplastic and hyperplastic endometrial diseases, but an increase in CLU protein has not been detected. Furthermore, specific CLU immunoreactivity has been observed in all glandular cells of the endometrium compared to other cellular compartments where CLU immunoreactivity was lower or absent [109].
Increased CLU expression enhances paclitaxel resistance in endometrial cancer [110,111].
BreastStudies with the MDA-MB-231 cell line show how sCLU silencing significantly inhibits cell proliferation and drastically reduces cell invasion, cell progression and metastatic potential [112,113].Unlike benign lesions, atypical hyperplasias, intraductal carcinomas, and invasive carcinomas are characterized by CLU overexpression [114].
Overexpression of sCLU is observed in a higher percentage of triple-negative breast cancer [115] and is associated with a negative estrogen and progesterone receptor status [114].
Likewise, overexpression in the stroma tends to directly correlate with resistance to preoperative neoadjuvant chemotherapy in the primary tumor and inversely with the apoptosis rate, indicating that gene expression may not be necessary for apoptotic cell death [114,116,117].
Colon sCLU is overexpressed, while nCLU is downregulated [118]. Likewise, increased sCLU expression was predominantly observed in the cytoplasm of highly invasive tumors and metastatic lymph nodes [31], indicating that CLU expression might serve as a marker to identify patients with more aggressive tumors who could potentially benefit from targeted treatments [119].
BladderTreatment with the antisense oligonucleotide (ASO) targeting negative regulation of Bcl-2 and CLU increases the sensitivity of partially resistant bladder carcinoma cells to the tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) [120].The recurrence-free survival time of patients with overexpression of CLU was shorter than that of patients with normal CLU expression [121,122].
Hepatocellular High levels of sCLU are associated with migration, invasion, and metastasis [15] due to increased MMP-2 expression and decreased E-cadherin expression [123].
Furthermore, sCLU overexpression contributes to oxaliplatin resistance [15].
In peripheral blood mononuclear cells (PBMC) from hepatocellular carcinoma patients, CLU has been proposed as a prospective detection biomarker along with other genes for its sensitivity and specificity [124]. The combination of CLU and AFP further improves diagnostic performance [125].
The initial levels of CLU are higher for patients with progressive disease than for those with partial or complete response, respectively [126,127].
PancreaticThe inducer of ferroptosis, a type of cell death characterized by the accumulation of reactive oxygen species (ROS), interferes with apoptotic cell death by regulating CLU [128].CLU expression in stages I and II is not significantly associated with apoptosis. Moreover, patients with positive CLU expression present better survival rates [129].
MelanomasIncreased expression is linked to heightened drug resistance and extended survival of tumor cells, whereas suppression diminishes resistance and lowers the survival rate of melanoma cells, both in laboratory settings and within living organisms [130].
Esophageal squamous cells Elevated CLU expression is associated with unfavorable outcomes in locoregional, overall, and distant progression-free survival. Additionally, individuals exhibiting CLU overexpression in both epithelium and stroma tend to have shorter survival times [131,132].
Head and neckOverexpression of CLU has been observed, but its implications have not yet been determined [133].Although CLU is detected in a low proportion of laryngeal carcinomas, it seems to exert a significant role in local invasiveness [133].
Anaplastic large cell lymphomas The role of CLU is unknown, but its expression within this lymphoma type provides an additional marker for diagnosis [134].
CLU expression does not correlate with the expression of anaplastic lymphoma kinase-1 (ALK-1). In reactive lymphoid tissues, only fibroblastic reticular cells and follicular dendritic cells exhibit positive expression [134].
Osteosarcoma sCLU overexpression is associated with metastasis and chemotherapy resistance [135].
ProstateThe expression of CLU increases in advanced stages of cancer, and its suppression sensitizes cells to chemotherapeutic drugs [136].It has been observed that CLU expression decreases considerably compared to benign tissues [137].
RenalThe introduction of the CLU gene enhances the metastatic potential of renal cell cancer [138], while the removal of the CLU gene inhibits growth and migration [139].
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

Martín-García, D.; García-Aranda, M.; Redondo, M. Therapeutic Potential of Clusterin Inhibition in Human Cancer. Cells 2024, 13, 665. https://doi.org/10.3390/cells13080665

AMA Style

Martín-García D, García-Aranda M, Redondo M. Therapeutic Potential of Clusterin Inhibition in Human Cancer. Cells. 2024; 13(8):665. https://doi.org/10.3390/cells13080665

Chicago/Turabian Style

Martín-García, Desirée, Marilina García-Aranda, and Maximino Redondo. 2024. "Therapeutic Potential of Clusterin Inhibition in Human Cancer" Cells 13, no. 8: 665. https://doi.org/10.3390/cells13080665

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