Next Article in Journal
Structural, Mechanical, and Dielectric Properties of Polydimethylsiloxane and Silicone Elastomer for the Fabrication of Clinical-Grade Kidney Phantom
Next Article in Special Issue
The Triad Hsp60-miRNAs-Extracellular Vesicles in Brain Tumors: Assessing Its Components for Understanding Tumorigenesis and Monitoring Patients
Previous Article in Journal
Investigating Health-Related Features and Their Impact on the Prediction of Diabetes Using Machine Learning
Previous Article in Special Issue
The Neurochaperonopathies: Anomalies of the Chaperone System with Pathogenic Effects in Neurodegenerative and Neuromuscular Disorders
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Challenging Riddle about the Janus-Type Role of Hsp60 and Related Extracellular Vesicles and miRNAs in Carcinogenesis and the Promises of Its Solution

by
Sabrina David
1,†,
Alessandra Maria Vitale
1,2,†,
Alberto Fucarino
1,2,†,
Federica Scalia
1,2,
Giuseppe Vergilio
1,
Everly Conway de Macario
3,
Alberto J. L. Macario
2,3,*,
Celeste Caruso Bavisotto
1,2,* and
Alessandro Pitruzzella
1,2,4
1
Department Biomedicine, Neurosciences and Advanced Diagnostics, Section of Human Anatomy, University of Palermo, 90127 Palermo, Italy
2
Euro-Mediterranean Institute of Science and Technology (IEMEST), 90139 Palermo, Italy
3
Department of Microbiology and Immunology, School of Medicine, University of Maryland at Baltimore-Institute of Marine and Environmental Technology (IMET), Baltimore, MD 21202, USA
4
Consorzio Universitario Caltanissetta, University of Palermo, 93100 Caltanissetta, Italy
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Appl. Sci. 2021, 11(3), 1175; https://doi.org/10.3390/app11031175
Submission received: 21 December 2020 / Revised: 15 January 2021 / Accepted: 22 January 2021 / Published: 27 January 2021

Abstract

:
Hsp60 is one of the most ancient and evolutionarily conserved members of the chaperoning system. It typically resides within mitochondria, in which it contributes to maintaining the organelle’s proteome integrity and homeostasis. In the last few years, it has been shown that Hsp60 also occurs in other locations, intracellularly and extracellularly, including cytosol, plasma-cell membrane, and extracellular vesicles (EVs). Consequently, non-canonical functions and interacting partners of Hsp60 have been identified and it has been realized that it is a hub molecule in diverse networks and pathways and that it is implicated, directly or indirectly, in the development of various pathological conditions, the Hsp60 chaperonopathies. In this review, we will focus on the multi-faceted role of this chaperonin in human cancers, showing the contribution of intra- and extracellular Hsp60 in cancer development and progression, as well as the impact of miRNA-mediated regulation of Hsp60 in carcinogenesis. There are still various aspects of this intricate biological scenario that are poorly understood but ongoing research is steadily providing new insights and we will direct attention to them. For instance, we will highlight the possible applications of the Hsp60 involvement in carcinogenesis not only in diagnosis, but also in the development of specific anti-cancer therapies centered on the use of the chaperonin as therapeutic target or agent and depending on its role, pro- or anti-tumor.

1. Stress Responses and the Chaperoning System

Since their first appearance on Earth, living organisms have been oppressed by physical and chemical stressors, such as radiations, extreme temperatures and pH, and hypoxia. These challenges required adaptation, including the development of anti-stress mechanisms. Today, it is possible to observe the success of these protection strategies, since living beings are observable in a wide range of ecosystems, even in those seemingly incompatible with life, such as sulfurous lakes, deep depth of the oceans, and permafrost [1,2,3,4,5,6,7,8].
One of the most important cellular anti-stress machineries is the chaperoning (or chaperone) system, which is highly conserved in the three phylogenetic domains, Bacteria, Archaea, and Eucarya [9]. It is composed of molecular chaperones and co-chaperones, and their co-factors, interactors, and receptors, which form functional networks working together to ensure protein homeostasis, under normal and stressful conditions [10,11]. The genes of many molecular chaperones are constitutively expressed, but others, named Heat shock protein (Hsp), are transcriptionally upregulated by exposure to stressors (e.g., heat, hyperthermia, hypoxia, heavy metals, ethanol, infections, radiation, and UV) and, thereby, protect against protein misfolding and aggregation and maintain proteins in their functional native state [12,13,14,15]. Hsps are commonly classified according to their molecular weight, and, even if not all the chaperones can be considered Hsps, the two terms are commonly used as synonymous [16,17,18].

Hsp60, One of the Most Ancient Anti-Stress Molecules

Hsp60 belongs to one of the oldest and evolutionarily most conserved protein families of the chaperoning system [19,20,21]. These proteins are present in all living species, including plants, where they were first discovered [22,23,24], and, considering their unique molecular characteristics, they were named “chaperonins” to distinguish them from other chaperones [25].
The canonical classification divides chaperonins into two main groups. Group I chaperonins are found in bacteria, as well as inside eukaryotic organelles of endosymbiotic origin (mitochondria in animal cells, and chloroplasts in plant cells), and work together with a co-chaperonin which helps the closing of the folding cage. Group II chaperonins are found in the eukaryotic cytosol and in Archaea, and do not require a co-chaperonin since they have a built-in lid [26]. More recently, a third group (Group III chaperonins) has been discovered and it is now under characterization [27,28].
In humans, the Group I chaperonin is Hsp60 (or Cpn60, or HSPD1), and the Group II chaperonin is CCT (chaperonin-containing TCP-1) or TRiC (T-complex protein Ring Complex). The former typically resides inside mitochondria but also occurs in various other locations intra- and extracellularly, while CCT is in the cytosol.
Hsp60 and CCT form macromolecular double-ring complexes with a central internal cavity in which polypeptides in need of assistance for folding or refolding are encapsulated and assisted to achieve their functional final conformation (native state) via an ATP-dependent mechanism [29,30,31].
A considerable part of our knowledge about how human chaperonins assemble and work derives from the study of their bacterial homologues. The bacterial counterpart of human Hsp60 is the chaperonin GroEL [32], which, together with the co-chaperonin GroES (the bacterial counterpart of human the Hsp10 or Cpn10), forms a tubular, double-ring complex with a central cavity similar to that described above for the eukaryotic Hsp60 and CCT complexes, inside which protein folding occurs [33,34]. Human Hsp60 has been found in various conformations (monomer, single heptameric ring, double-ring tetradecamer), depending on their concentration, levels of ATP, and presence of Hsp10 and substrate [21,35,36,37]. Early studies showed that human Hsp60 can assist in productive protein folding without forming a macro-double ring-complex [38,39]. However, more recent studies, based on transmission electron microscopy and X-ray crystallographic investigations, have provided strong evidence that it likely uses both double- and single-ring intermediates during its ATPase cycle [40,41,42].
Hsp60 has been found in extramitochondrial sites, such as cytoplasm and plasma-cell membrane, as well as in extracellular sites, inside extracellular vesicles (EVs), in circulation, and in body fluids [43,44,45,46,47,48,49]. Consequently, in addition to its canonical chaperoning function, Hsp60 also performs various other non-canonical activities, “moonlighting functions” unrelated to protein quality control. (Figure 1) [21,49,50,51]. For instance, it has been observed that increased amounts of Hsp60 on the surface of cancer cells act as a signal to stimulate the immune system, leading to the activation and maturation of dendritic cells and the generation of an antitumor T-cell response [52,53,54]. A non-canonical function of human Hsp60 that is still under scrutiny is the regulation of cell apoptosis [55]. Some studies suggested a pro-apoptotic role, involving pro-caspase 3 proteolytic activation [56,57], whereas other investigations support an anti-apoptotic role, involving various mechanisms: the sequestration of Bax-containing complexes [58,59], the maintenance of mitochondria integrity and ATP generation [60], and the triggering of the IKK/NF-κB survival pathway [61]. This dual role in cell apoptosis/survival regulation has been observed both in normal and tumor cells [59,62], thus it may affect cancer progression either positively or negatively.

2. Hsp60 Chaperonopathies

Hsp60 is multifaceted and plays diverse physiological roles but when abnormal in structure and/or function it can become pathogenic and cause diseases, the Hsp60 chaperonopathies [63,64].
In genetic chaperonopathies, there are pathogenic variants, e.g., missense mutations, in the Hsp60 gene, whereas in the acquired chaperonopathies the gene is normal but the Hsp60 protein is altered structurally and/or functionally. Genetic chaperonopathies are infrequent and typically have an early clinical onset, while the acquired ones are more common, mostly occur in adults, and are often associated with other pathological conditions, especially age-related diseases [65,66]. Another classification of practical value sorts the chaperonopathies according to quantitative parameters pertaining to concentration and functionality into by defect, by excess, and by mistake or collaborationism (Table 1) [65,66].
Acquired chaperonopathies can be caused by aberrant post-translational modifications (PTMs) that have an impact on the structure/function of the chaperone molecule. The Hsp60 amino acid sequence contains various critical sites that can be affected by PTMs [67]. Possible PTMs are phosphorylation, O-GlcNAcylation, nitration, acetylation, S-nitrosylation, citrullination, methylation, oxidation, biotinylation, and ubiquitination [51]. Hsp60 PTMs can have beneficial or deleterious effects. For instance, Hsp60 hyperacetylation in the course of anti-osteosarcoma treatment lead to death of the malignant cells [68]. Hyperacetylation disrupted the Hsp60/p53 complex, restored replicative senescence, and diminished or stopped tumor growth [69]. Phosphorylated Hsp60 on the surface of breast cancer cells induced α3β1 integrin activation, resulting in enhanced motility and adhesion of these cells [70]. Tyrosine phosphorylation of Hsp60 helps malignant cells to escape immune surveillance by NK and CD8 T cells [71].
Chaperonopathies by mistake include all those pathological conditions in which a chaperone is normal as far it can be determined by current methodologies but contributes to the initiation and/or progression of disease, as it has been observed in various types of cancer, and autoimmune, inflammatory, and neurologic disorders [66].
Other examples of chaperonopathies by mistake are those autoimmune conditions in which human Hsp60 acts as auto-antigen. This situation has been described for various autoimmune conditions such as Behçet’s disease [72,73], diabetes mellitus [74], systemic lupus erythematosus and vasculitis-associated systemic autoimmune disorders [75,76,77,78], atherosclerosis [79,80,81,82], and rheumatoid arthritis [83]. The pathogenic autoimmune mechanism for some of these disorders is triggered by the presence of Hsp60 on the plasma-cell membrane where it becomes a target accessible to autoantibodies and, thereby, leads to apoptosis [75,76]. However, in other situations Hsp60 has shown cytoprotective activity, which reiterates the concept that this chaperonin can play apparently opposing roles (Table 2).
In view of the above considerations, it becomes evident that (1) the kind of activity the Hsp60 chaperonin undertakes depends on its context, i.e., it is determined and/or modulated by the composition of its surroundings, namely, by the receptors and interactors within its reach; and (2) Hsp60 has not only canonical functions, which pertain to protein quality control, but also displays other functions unrelated to protein homeostasis that are also important in health and disease. Therefore, advances in the treatment of Hsp60 chaperonopathies ought to include the development of modulators of the chaperonin in situ. Two types of chaperonin inhibitors are currently being investigated: type I that block the binding and hydrolysis of ATP, and type II that bind cysteine covalently [84]. Natural and synthetic compounds potentially useful have already been identified with various degrees of anti-cancer, anti-inflammatory, or anti-autoimmune potency [63,84].

3. Hsp60 in Carcinogenesis

An indication of Hsp60 involvement in carcinogenesis is its altered expression and localization observed in certain human cancers [47,49,85]. However, the exact role of Hsp60 in cancer remains undefined and seems to change depending on the molecular and cytological context (Table 2). Hsp60 was found increased in various malignancies: cervical and ovarian [86,87,88,89,90], breast [91,92], colorectal [93], lung [94,95], prostate [96], gastric [97,98], and thyroid cancers [99,100]; and in leukemias [101], and glioblastoma multiforme [102,103,104]. The data in general suggest that Hsp60 is actively involved in carcinogenesis as a pro-tumorigenic factor, because its increase and location changes positively correlated with tumor development and malignancy. Moreover, in some cancers, the Hsp60 increase was also associated with a heightened resistance to anti-cancer drugs and other treatments [89,105], and with metastasization and angiogenesis [106,107]. It has to be emphasized that the quantitative variations of Hsp60 in cancer cells, particularly its increase, may reflect the heightened need of the chaperone system by the malignant cells with their rapid and intensive metabolism and proliferation. Thus, the Hsp60 quantitative patterns observed would be the consequence of the disease but not a distinct etiological factor. We argue that even so, Hsp60 aids the tumor in what we call a chaperonopathy by mistake or collaborationism because the tumor depends on its help. What is the value of this concept? It puts Hsp60 in the stage’s center and presents it as a target for developing treatment strategies aimed at inhibiting-blocking this collaborator with the enemy from the inside. For example, Hsp60 downregulation by chemical compounds suppressed cancer-cell proliferation and tumor progression and enhanced the beneficial effects of anti-cancer treatments [62,69,90,108,109,110,111,112]. It was observed that Hsp60 knockdown inhibited tumor progression by altering mitochondrial homeostasis and inactivating the mTOR pathway, in ovarian cancer and glioblastoma [90,109]. In colorectal cancer, Hsp60 inhibition promoted the tumor-suppressive activity of insulin-like growth factor binding protein 7 (IGFBP7) [108]. Treatment of a neuroblastoma cell line with curcumin caused cell death by diminishing the cellular level of Hsp60 [111]. It was suggested that this cytotoxic effect was induced through the downregulation of survivin, whose expression was shown to be positively correlated with the expression of CCAR2 and Hsp60 in neuroblastoma tissues and cell lines [113]. Thus, it is becoming clear that the positive correlation between Hsp60 overexpression and increase in cancer cell proliferation and survival depends on the interaction of the chaperonin with proteins involved in cell cycle and apoptosis. In cancer cells lines, Hsp60 played a cytoprotective and pro-survival role by stabilizing the mitochondrial level of survivin and blocking p53-mediated apoptosis [62,113], or by inhibiting the intracellular isoform of clusterin [114]. In other cases, the pro-tumorigenic role of Hsp60 involved blocking the caspase-dependent apoptosis through the negative regulation of mitochondrial permeability transition [102], or the inhibition of pro-caspase 3 proteolytic activation [115,116].
However, others have reported that in some types of cancer, Hsp60 was decreased, downregulated not increased or, if increased, caused tumor suppression not enhancement [117,118]. For instance, Hsp60 was decreased in hepatocellular carcinoma (HCC) tissue compared to peritumor tissue, and this pattern positively correlated with high serum AFP (alpha-fetoprotein) level and poor overall survival. Conversely, increased Hsp60 inhibited invasion and migration of HCC cells both in vitro and in vivo, correlating with a better prognosis [118].
All the above observations (Table 2) make clear that Hsp60 has different roles in carcinogenesis that deserve investigation to elucidate the molecular mechanisms involved, particularly those determining whether the chaperonin acts against or for the cancer cell.
Some data in the literature support the assumption that Hsp60 acts as a pro-tumor protein, considering its role in the modulation of anti-apoptotic factors. It is known that during tumorigenesis, cells undergo complex transcriptional events that lead to a dysregulation of numerous factors. In this scenario, transformed cells show the typical phenotype of the proteotoxic stress, in which the consequent Hsp60 overexpression is a key event because of its central role in the regulation of protein homeostasis. Conversely, the observation that, in several types of tumors, such as lung cancer, Hsp60 levels are reduced, further demonstrate that its functioning is not only related to protein folding and the maintenance of protein homeostasis, but is complex and dependent on the cell and tissue type, its molecular interactors and its localization inside or outside the mitochondria and the cell. Thus, the multifaceted and at times contradictory functions of Hsp60 in cancer are still poorly understood and deserve active investigation, considering the importance of the chaperonin for survival of cells, normal or malignant.
Table 2. Hsp60 in human cancers.
Table 2. Hsp60 in human cancers.
CancerHsp60 Level, Location and/or StatusEffectReference
Lung carcinomaPresence of Hsp60 on cancer-cell plasma-cell membrane and on membrane of cancer cell-derived exosomes Possible involvement in cell-to-cell communication and anti-tumor immune response stimulation[47,48]
Decrease of intracellular Hsp60 level and increase of Hsp60 acetylation level after doxorubicin treatmentHsp60/p53 complex dissociation and restoration of cancer-cells replicative senescence[69]
Decrease of intracellular Hsp60 level after CubipyOXA treatmentDissociation of the Hsp60/pro-Caspase-3 complex and cancer-cell apoptosis[116]
Increased Hsp60 levelPositive correlation with cancer progression and poor prognosis[94,96]
Oral cancerPresence of Hsp60 on cell surfaceInteraction with gamma-delta T cells and transduction of anti-cancer immune response[52]
OsteosarcomaHyperacetylation and loss of mitochondrial Hsp60 after Geldanamycin treatmentDecreased viability and augmented cancer-cell death[68]
Breast cancerIncrease of phosphorylated surface Hsp60 α3β1 integrin activation and enhancement of cancer cells motility and adhesion[70]
Increased cytosolic Hsp60Enhanced cancer-cell proliferation and reduced apoptosis; positive correlation with worse disease-free survival and poor prognosis[91,92]
Bronchial carcinomaDecreased Hsp60 levelPositive correlation with bronchial cancer development and progression[85,119]
Cervical cancerIncreased Hsp60 levelPositive correlation with cancer progression and malignancy[86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110]
Ovarian cancerIncreased Hsp60 levelPositive correlation with cancer progression and severity (poor prognosis and resistance to anti-cancer treatment)[89,90,105]
Colorectal cancerIncreased Hsp60 levelPositive correlation with cancer progression and malignancy[93,120]
IGFBP7-dependent down-regulation of intracellular and extracellular Hsp60 levelInvolvement in tumor suppressive activity of IGFBP7[108]
Prostate cancerIncreased Hsp60 levelPositive correlation with tumor progression and hormone resistance[96]
Gastric cancersIncreased Hsp60 level Positive correlation with cancer progression, invasiveness, and poor overall survival. [97,98]
LeukemiaPresence of Hsp60 on the cell surfaceActivation/maturation of dendritic cells and generation of potent anti-tumor T-cell response[53]
Increased Hsp60 levelPositive correlation with lower complete remission rate and shorter survival[101]
Glioblastoma multiformeIncreased Hsp60 levelCytoprotective and pro-survival role[102]
Decreased Hsp60 levelReduced cancer cell proliferation and tumor growth[109,112]
Hepatocellular carcinomaDecreased Hsp60 levelPositive correlation with cancer progression and poor prognosis[118]
Increased exosomal release of Hsp60 after anti-cancer treatmentActivation of anti-tumor immune response[121]
Pancreatic cancerIncreased Hsp60 levelPositive correlation with cancer-cell proliferation and tumor growth and progression[111]
Bladder carcinomaIncreased Hsp60 levelPositive correlation with resistance to anti-cancer treatment[105]
Decreased Hsp60 levelPositive correlation with higher tumor stage and cancer recurrence[122]
Renal cell carcinomaDecreased Hsp60Disruption of mitochondria homeostasis and positive correlation with cancer progression[117]
Large bowel cancerIncreased intracellular and exosomal Hsp60 levelPositive correlation with tumor development and progression[123]
Thyroid cancersIncreased intracellular and exosomal Hsp60 level Positive correlation with tumor progression[99,100]

3.1. Hsp60 in Extracellular Vesicles in Carcinogenesis

Extracellular vesicles (EVs) are membranous particles with a diameter of 30–150 nm, which are found in blood, urine, cerebrospinal fluid, breast milk, and saliva, and are released by normal and tumor cells. The International Society for Extracellular Vesicles (ISEV) classify EVs considering their biogenesis pathways and specific markers received from the cells in which they originate, into exosomes, microvesicles, oncosomes, and apoptotic bodies; and considering size into “small EVs” (sEVs) and “medium/large EVs” (m/lEVs) [124]. However, considerable morphological and biochemical heterogeneity exists among the EVs and, to complicate matters even more, many publications do not provide detailed descriptions and/or use non-standardized terminology. Because of this, we will use here EVs and exosomes as synonyms. EVs are involved in physiological and pathological processes as mediator of cell-to-cell communication by carrying proteins, lipids, and nucleic acids (DNA, mRNA, and miRNAs), that affect recipient cells and modify their functions [125,126,127,128,129,130,131]. Their content may vary depending on the cell type that produces them, but in general EVs carry a range of diverse proteins, such as tumor susceptibility gene 101 (TSG101); integrins; and tetraspanins such as CD9, CD53, CD63, CD81, and CD82 [132,133]. EVs are thought to play a role in the remodelling of the pericellular microenvironment that is crucial in maintaining tumor growth and recurrence [123,134]. MicroRNAs transported by EVs are being considered the main players in the modulation of target cell functions [135,136]. Therefore, tumor-derived EVs are attracting attention from scientists interested in molecules that might be used as biomarkers for diagnosis and patient follow up. In addition, EVs are candidates for delivering anti-cancer compounds to specific target tissues because their content can be modified; they show some tissue-specificity; and their immunogenicity is low when they are isolated from the same patient, thus presenting slow risk of generating anti-EV immunity [137,138,139].
Our research group provided the first evidence that tumor cells actively release Hsp60 via exosomes, through a secretion mechanism that requires the translocation of the chaperonin molecule to the plasma-cell membrane and its association with lipid rafts (Figure 2) [47,48]. The exosomal release of Hsp60 by human tumor cells can be enhanced by anti-cancer treatment. For instance, histone deacetylase inhibitors SAHA causes cell cycle arrest and death in a human lung-derived carcinoma cell line; this cytotoxic effect was associated with generation of oxidative stress, mitochondrial damage, and diminution of the intracellular level of Hsp60, which became nitrated and was released via exosomes [110]. It has been suggested that the released Hsp60 interacts with the immune system, generating an anti-tumor response that potentiates the effect of SAHA. In this manner, exosomal Hsp60 could modulate the tumor microenvironment and, from the practical standpoint it could be consider for use as a diagnostic and prognostic biomarker [123]. Indeed, it was found that the plasma levels of exosomes with Hsp60 diminished after anti-cancer treatment that caused tumor mass reduction in bowel cancer and thyroid papillary carcinoma [100,123]. Additionally, treating hepatocellular carcinoma with anti-cancer drugs caused the release of a larger amount of exosomes containing Hsps, including Hsp60, which was followed by an enhanced anti-tumor immune response mediated by natural killer cells [121].

3.2. Hsp60 and miRNAs Correlations and Implications for Carcinogenesis

MicroRNAs (miRNAs) are a class of small non-coding RNAs that regulate gene expression post-transcriptionally by translational inhibition and/or mRNAs destabilization [140]. MiRNAs regulate most protein coding genes, and thus they virtually control all biological processes [141,142]. MiRNAs dysregulation following amplification or deletion of miRNA genes, abnormal transcriptional control of miRNAs, dysregulated epigenetic changes, or defects in the miRNA biogenesis process, are associated with pathological conditions, including cancer [143,144]. Dysregulated, abnormal miRNAs can affect all hallmarks of cancer (e.g., continued proliferative signalling, evasion of growth suppressors, cell death resistance, invasiveness and metastasization, and angiogenesis) by acting either as onco-miRs or tumor suppressors, with inhibition of tumor suppressive mRNAs or oncogenic mRNAs, respectively [145,146]. For this reason, miRNAs are considered for use not only as diagnostic and prognostic biomarkers, but also as potential targets or agents in anti-cancer treatment [147].
MiR-9 and miR-221 have been classified as onco-miRs because an increase in their levels paralleled increased risk for tumorigenesis and resistance to chemotherapeutics in primary cancers [148,149,150,151]. In breast cancer, miR-9 and miR-221 increased levels correlated with poor outcome and promoted tumor progression and aggressiveness by favoring epithelial-mesenchymal transition (EMT) and breast cancer stem cell phenotypes [152]. Therefore, they have been suggested as potential biomarkers for breast cancer progression and targets for treatment.
miR-30 was found decreased in prostate cancer and acted as tumor suppressor by targeting the EMT-associated gene ERG (Ets-related gene) [153]. Conversely, when increased, miR-30 suppressed EMT and inhibited cell migration and invasion, suggesting it could be used as a therapeutic agent (Table 3) [153].
Hsps, including Hsp60, are also targets for regulation by miRNAs and this has been investigated in various cancers [154,155,156,157]. miR-1 and miR-206 increase in rat cardiomyocytes contributed to glucose-mediated apoptosis by diminishing Hsp60 and IGF-1 expression and inhibiting the IGF-1/IGF-1R/PI3 K/Akt pathway [158]. Conversely, treatment of cardiomyocytes with carvedilol, a non-selective β-adrenergic receptor antagonist, inhibited miR-1, which resulted in increased levels of Hsp60 and apoptosis prevention (Table 3) [159].
MiR-29a was found augmented in the serum of breast cancer patients [155]. Downregulation of miR-29a in a breast cancer cell line promoted apoptosis by causing an increase in the level of Hsp60 and a decrease of Hsp27, Hsp40, Hsp70, and Hsp90, suggesting that downregulation of this miRNA is a promising strategy to sensitize cancer cells to chemotherapy [155]. It is likely that the anti-cancer effects observed with the increase of Hsp60 were related to the known pro-apoptotic role of the chaperonin [56,57], or to its ability to stimulate an anti-tumor immune response when exposed on cell surface or released extracellularly [53,54,121]. The level of miR-644a l was found low in hepatocellular carcinoma tissues and cell lines, and negatively correlated with tumor diameter and TNM (Tumor-Node-Metastases) stages [156]. In vitro, an increase in miR-644a promoted cancer cell apoptosis by inhibiting HSF1, Hsp90, Hsp60, Bcl-2, and Bcl-xL proteins while increasing the levels of BID, BAD, BIM, SMAC, Apaf-1, and cleaved caspases-3 and -9, which are all mediators of cell apoptosis [156]. In view of all these observations, miR-644a was suggested as potential prognostic biomarker and therapeutic target in HCC (Table 3) [156]. It is likely that Hsp60 typically displays a pro-tumorigenic activity that can be suppressed by its miRNA-mediated inhibition [157,158,159]. A further confirmation of the role of miRNAs in the regulation of Hsp60 level in cancer emerged from a study of the pro-tumorigenic activity of miR-17a in gastric lymphoma [160]. The expression of miR-17 was found significantly higher in gastric lymphoma than in normal tissues, which promoted tumor development, progression, and metastasization by regulating the Hsp60/TNFR2 pathway [160]. Therefore, the latter pathway emerges as a potential target for the diagnosis and treatment of gastric lymphoma (Table 3).
Table 3. Examples of miRNAs involved in Hsp60 regulation.
Table 3. Examples of miRNAs involved in Hsp60 regulation.
TissuemiRNAs and miRNAs StatusEffectReference
Rat cardiomyocytesmiR-1 and miR-206 high-glucose-dependent up-regulationIncreased cell apoptosis induced by Hsp60 and IGF-1 down-regulation and IGF-1/IGF-1R/PI3 K/Akt pathway inhibition[158,159]
Breast cancermiR-29a in vitro down-regulationIncreased cancer cell apoptosis and sensitization to anti-cancer treatment induced by Hsp60 up-regulation[155]
Gastric LymphomamiR-17 higher levelIncreased malignancy via regulation of Hsp60/TNFR2 pathway[160]
Hepatocellular carcinomamiR-644a in vitro up-regulationIncreased cancer-cell apoptosis induced by Hsp60 inhibition[156]

4. Concluding Remarks

The involvement of molecular chaperones, including Hsp60, in carcinogenesis has been suggested by various findings [107,161,162]. Molecular chaperones have been found increased in tumor tissues and closely associated with tumor growth and aggressiveness. Along the same lines, decreased levels and expression of chaperones have been found associated with reduced cancer cell proliferation, motility, survival, and metastasization, and with decreased neoangiogenesis and resistance to anti-tumor immune response and treatment [163,164,165,166,167]. The quantitative variations of Hsp60 during carcinogenesis, especially its increase, may be simply the reflection of increased metabolic and proliferative activities of the cancer cell, which would require more chaperonin molecules than a normal cell at its physiological level of metabolism and proliferation. Nevertheless, the role of Hsp60 can still be considered pro-tumoral even if it is not a distinct carcinogenic factor. It would be an example of chaperonopathy by mistake or collaborationism, meaning that the human molecule helps the tumor to grow, proliferate, metastasize, and resist stressors such as anti-cancer drugs. It is crucial to visualize this situation because it then appears very clearly that therapeutic strategies aiming at inhibiting-blocking the chaperonin (negative chaperonotherapy) in tumor cells may offer efficacious ways to defeat cancer. Therefore, negative chaperonopathy, aiming to inhibit or block pro-tumorigenic molecular chaperones deserves investigation [63,168,169].
The role of Hsp60 in carcinogenesis is not yet fully understood, and its elucidation is complicated because it seems to vary depending on the tumor type, tissue affected, stage, and other unidentified factors. While some reports indicate that Hsp60 displays a protumor role others show the contrary [85,90,98,117,118,119,122]. While the exact role of Hsp60 in every tumor in which its levels and location changes is being studied, the quantitative variations and tissue distribution of the chaperonin may be used as biomarkers potentially useful for diagnosis, prognostication, and disease monitoring [120,123,170,171]. Additionally, investigations on Hsp60 chaperonotherapy in its various forms should be performed and aiming at standardizing the use the chaperonin as a therapeutic target or agent [63]. Some possibilities are to (i) identify compounds and/or methods that will induce a antitumor immune response by promoting the presence of Hsp60 on the surface of tumor-cells [172]; (ii) develop Hsp60 vaccines exploiting its potential for inducing the production of cytokines by interaction with monocytes and macrophages, and by manipulating its ability to bind and activate TLR-2 in CD4+CD25+ regulatory T cells and TLR-4 in B cells [173]; and (iii) identify compounds able to modulate Hsp60 expression, in order to reduce or enhance its activity in those tumors in which it displays a pro-or anti-tumorigenic role, respectively. Understanding the mode of action of Hsp60 and its regulation by miRNAs in different tissue and tumor types will help in the development of novel therapeutic strategies for cancer therapy [68,112,174]. Moreover, the utilization of advanced drug delivery systems, as exosome-like vesicles promise to be, to deliver molecules, such as miRNAs known to regulate Hsp60, and internalize them in the target cells, should be actively investigated to standardize the particles and their contents as well as their isolation from cancer patients.

Author Contributions

Conceptualization, S.D., A.M.V., A.F., E.C.d.M. and C.C.B.; original draft preparation, S.D., A.M.V., A.F., F.S., G.V., E.C.d.M., A.P. and C.C.B.; supervision and final writing, A.J.L.M. and C.C.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded in part by the Italian National Operational Programme (PON) «Imprese e Competitività» 2014–2020 FESR, grant awarded by the Italian Ministry of Economic Development to the project titled «Gestione di un servizio integrato multicentrico di diagnostica e terapia personalizzata in oncologia» (Project code: F/090012/01-02/X36). A.J.L.M. and E.C.d.M. were partially supported by IMET and IEMEST. This is IMET contribution number IMET 21-003.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Brock, T.D.; Freeze, H. Thermus aquaticus gen. n. and sp. n., a nonsporulating extreme thermophile. J. Bacteriol. 1969, 98, 289–297. [Google Scholar] [CrossRef] [Green Version]
  2. Moseley, B.E.B. Photobiology and Radiobiology of Micrococcus (Deinococcus) radiodurans. In Photochemical and Photobiological Reviews; Springer: Boston, MA, USA, 1983; pp. 223–274. [Google Scholar]
  3. Herbert, R.A. A perspective on the biotechnological potential of extremophiles. Trends Biotechnol. 1992, 10, 395–402. [Google Scholar] [CrossRef]
  4. Minton, K.W. DNA repair in the extremely radioresistant bacterium Deinococcus radiodurans. Mol. Microbiol. 1994, 13, 9–15. [Google Scholar] [CrossRef] [PubMed]
  5. Minton, K.W.; Daly, M.J. A model for repair of radiation-induced DNA double-strand breaks in the extreme radiophile Deinococcus radiodurans. BioEssays 1995, 17, 457–464. [Google Scholar] [CrossRef] [PubMed]
  6. Gilichinsky, D.; Vishnivetskaya, T.; Petrova, M.; Spirina, E.; Mamykin, V.; Rivkina, E. Bacteria in permafrost. In Psychrophiles: From Biodiversity to Biotechnology; Springer: Berlin/Heidelberg, Germany, 2008; pp. 83–102. ISBN 9783540743347. [Google Scholar]
  7. Sorokin, D.Y.; Kuenen, J.G.; Muyzer, G. The microbial sulfur cycle at extremely haloalkaline conditions of soda lakes. Front. Microbiol. 2011, 2, 44. [Google Scholar] [CrossRef] [Green Version]
  8. Dhakar, K.; Pandey, A. Wide pH range tolerance in extremophiles: Towards understanding an important phenomenon for future biotechnology. Appl. Microbiol. Biotechnol. 2016, 100, 2499–2510. [Google Scholar] [CrossRef] [PubMed]
  9. Macario, A.J.L.; Conway de Macario, E. Stress genes: An introductory overview. Stress 1997, 1, 123–134. [Google Scholar] [CrossRef]
  10. Hartl, F.U.; Bracher, A.; Hayer-Hartl, M. Molecular chaperones in protein folding and proteostasis. Nature 2011, 475, 324–332. [Google Scholar] [CrossRef]
  11. Macario, A.J.L.; Conway de Macario, E. Chapter 12—Chaperone proteins and chaperonopathies. In Handbook of Stress, Physiology, Biochemistry, and Pathology; Fink, G., Ed.; Elsevier/Academic Press: Cambridge, MS, USA, 2019; Volume 3, pp. 135–152. Available online: https://doi.org/10.1016/B978-0-12-813146-6.00012-6 (accessed on 18 December 2020).
  12. Lindquist, S.; Craig, E.A. The heat-shock proteins. Annu. Rev. Genet. 1988, 22, 631–677. [Google Scholar] [CrossRef]
  13. Becker, J.; Craig, E.A. Heat-shock proteins as molecular chaperones. Eur. J. Biochem. 1994, 219, 11–23. [Google Scholar] [CrossRef]
  14. Nollen, E.A.A.; Morimoto, R.I. Chaperoning signaling pathways: Molecular chaperones as stress-sensing “heat shock” proteins. J. Cell Sci. 2002, 115, 2809–2816. [Google Scholar]
  15. Jacob, P.; Hirt, H.; Bendahmane, A. The heat-shock protein/chaperone network and multiple stress resistance. Plant Biotechnol. J. 2017, 15, 405–414. [Google Scholar] [CrossRef] [PubMed]
  16. Hendrick, J.P.; Hartl, F.U. Molecular chaperone functions of heat-shock proteins. Annu. Rev. Biochem. 1993, 62, 349–384. [Google Scholar] [CrossRef] [PubMed]
  17. Welch, W.J. Heat shock proteins functioning as molecular chaperones: Their roles in normal and stressed cells. Philos. Trans. R. Soc. B Biol. Sci. 1993, 339, 327–333. [Google Scholar]
  18. Kampinga, H.H.; Hageman, J.; Vos, M.J.; Kubota, H.; Tanguay, R.M.; Bruford, E.A.; Cheetham, M.E.; Chen, B.; Hightower, L.E. Guidelines for the nomenclature of the human heat shock proteins. Cell Stress Chaperones 2009, 14, 105–111. [Google Scholar] [CrossRef] [Green Version]
  19. Gupta, R.S. Evolution of the chaperonin families (HSP60, HSP 10 and TCP-1) of proteins and the origin of eukaryotic cells. Mol. Microbiol. 1995, 15, 1–11. [Google Scholar] [CrossRef]
  20. Rowland, S.E.; Robb, F.T. Structure, Function and Evolution of the Hsp60 Chaperonins; Springer: Singapore, 2017; pp. 3–20. [Google Scholar]
  21. Vilasi, S.; Bulone, D.; Caruso Bavisotto, C.; Campanella, C.; Marino Gammazza, A.; San Biagio, P.L.; Cappello, F.; Conway de Macario, E.; Macario, A.J.L. Chaperonin of Group I: Oligomeric spectrum and biochemical and biological implications. Front. Mol. Biosci. 2018, 4, 99. [Google Scholar] [CrossRef]
  22. Georgopoulos, C.P.; Hohn, B. Identification of a host protein necessary for bacteriophage morphogenesis (the groE gene product). Proc. Natl. Acad. Sci. USA 1978, 75, 131–135. [Google Scholar] [CrossRef] [Green Version]
  23. Barraclough, R.; Ellis, R.J. Protein synthesis in chloroplasts IX. Assembly of newly-synthesized large subunits into ribulose bishopshate carboxylase in isolated intact pea chloroplasts. Biochim. Biophys. Acta 1980, 608, 19–31. [Google Scholar] [CrossRef]
  24. Hemmingsen, S.M.; Ellis, R.J. Purification and Properties of Ribulosebisphosphate Carboxylase Large Subunit Binding Protein. Plant Physiol. 1986, 80, 269–276. [Google Scholar] [CrossRef] [Green Version]
  25. Hemmingsen, S.M.; Woolford, C.; Van Der Vies, S.M.; Tilly, K.; Dennis, D.T.; Georgopoulos, C.P.; Hendrix, R.W.; Ellis, R.J. Homologous plant and bacterial proteins chaperone oligomeric protein assembly. Nature 1988, 333, 330–334. [Google Scholar] [CrossRef] [PubMed]
  26. Horwich, A.L.; Fenton, W.A.; Chapman, E.; Farr, G.W. Two families of chaperonin: Physiology and mechanism. Annu. Rev. Cell Dev. Biol. 2007, 23, 115–145. [Google Scholar] [CrossRef] [PubMed]
  27. Techtmann, S.M.; Robb, F.T. Archaeal-like chaperonins in bacteria. Proc. Natl. Acad. Sci. USA 2010, 107, 20269–20274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. An, Y.J.; Rowland, S.E.; Na, J.H.; Spigolon, D.; Hong, S.K.; Yoon, Y.J.; Lee, J.H.; Robb, F.T.; Cha, S.S. Structural and mechanistic characterization of an archaeal-like chaperonin from a thermophilic bacterium. Nat. Commun. 2017, 8, 827. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Spiess, C.; Meyer, A.S.; Reissmann, S.; Frydman, J. Mechanism of the eukaryotic chaperonin: Protein folding in the chamber of secrets. Trends Cell Biol. 2004, 14, 598–604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Leitner, A.; Joachimiak, L.A.; Bracher, A.; Mönkemeyer, L.; Walzthoeni, T.; Chen, B.; Pechmann, S.; Holmes, S.; Cong, Y.; Ma, B.; et al. The molecular architecture of the eukaryotic chaperonin TRiC/CCT. Structure 2012, 20, 814–825. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Willison, K.R. The structure and evolution of eukaryotic chaperonin-containing TCP-1 and its mechanism that folds actin into a protein spring. Biochem. J. 2018, 475, 3009–3034. [Google Scholar] [CrossRef] [PubMed]
  32. McMullin, T.W.; Hallberg, R.L. A highly evolutionarily conserved mitochondrial protein is structurally related to the protein encoded by the Escherichia coli groEL gene. Mol. Cell. Biol. 1988, 8, 371–380. [Google Scholar] [CrossRef] [Green Version]
  33. Sigler, P.B.; Xu, Z.; Rye, H.S.; Burston, S.G.; Fenton, W.A.; Horwich, A.L. Structure and function in GroEL-mediated protein folding. Annu. Rev. Biochem. 1998, 67, 581–608. [Google Scholar] [CrossRef] [Green Version]
  34. Xu, Z.; Sigler, P.B. GroEL/GroES: Structure and function of a two-stroke folding machine. J. Struct. Biol. 1998, 124, 129–141. [Google Scholar] [CrossRef]
  35. Levy-Rimler, G.; Viitanen, P.; Weiss, C.; Sharkia, R.; Greenberg, A.; Niv, A.; Lustig, A.; Delarea, Y.; Azem, A. The effect of nucleotides and mitochondrial chaperonin 10 on the structure and chaperone activity of mitochondrial chaperonin 60. Eur. J. Biochem. 2001, 268, 3465–3472. [Google Scholar] [CrossRef] [PubMed]
  36. Vilasi, S.; Carrotta, R.; Mangione, M.R.; Campanella, C.; Librizzi, F.; Randazzo, L.; Martorana, V.; Marino Gammazza, A.; Ortore, M.G.; Vilasi, A.; et al. Human Hsp60 with its mitochondrial import signal occurs in solution as heptamers and tetradecamers remarkably stable over a wide range of concentrations. PLoS ONE 2014, 9, e97657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Enriquez, A.S.; Rojo, H.M.; Bhatt, J.M.; Molugu, S.K.; Hildenbrand, Z.L.; Bernal, R.A. The human mitochondrial Hsp60 in the APO conformation forms a stable tetradecameric complex. Cell Cycle 2017, 16, 1309–1319. [Google Scholar] [CrossRef] [PubMed]
  38. Viitanen, P.V.; Lorimer, G.; Bergmeier, W.; Weiss, C.; Kessel, M.; Goloubinoff, P. Purification of mammalian mitochondrial chaperonin 60 through in vitro reconstitution of active oligomers. Methods Enzymol. 1998, 290, 203–217. [Google Scholar]
  39. Nielsen, K.L.; Cowan, N.J. A single ring is sufficient for productive chaperonin-mediated folding in vivo. Mol. Cell 1998, 2, 93–99. [Google Scholar] [CrossRef]
  40. Nisemblat, S.; Parnas, A.; Yaniv, O.; Azem, A.; Frolow, F. Crystallization and structure determination of a symmetrical “football” complex of the mammalian mitochondrial Hsp60-Hsp10 chaperonins. Acta Crystallogr. Sect. F Struct. Biol. Commun. 2014, 70, 116–119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Nisemblat, S.; Yaniv, O.; Parnas, A.; Frolow, F.; Azem, A. Crystal structure of the human mitochondrial chaperonin symmetrical football complex. Proc. Natl. Acad. Sci. USA 2015, 112, 6044–6049. [Google Scholar] [CrossRef] [Green Version]
  42. Gomez-Llorente, Y.; Jebara, F.; Patra, M.; Malik, R.; Nisemblat, S.; Chomsky-Hecht, O.; Parnas, A.; Azem, A.; Hirsch, J.A.; Ubarretxena-Belandia, I. Structural basis for active single and double ring complexes in human mitochondrial Hsp60-Hsp10 chaperonin. Nat. Commun. 2020, 11, 1916. [Google Scholar] [CrossRef] [Green Version]
  43. Soltys, B.J.; Gupta, R.S. Immunoelectron microscopic localization of the 60-kDa heat shock chaperonin protein (Hsp60) in mammalian cells. Exp. Cell Res. 1996, 222, 16–27. [Google Scholar] [CrossRef]
  44. Soltys, B.J.; Gupta, R.S. Cell surface localization of the 60 kDa heat shock chaperonin protein (hsp60) in mammalian cells. Cell Biol. Int. 1997, 21, 315–320. [Google Scholar] [CrossRef]
  45. Soltys, B.J.; Gupta, R.S. Mitochondrial-matrix proteins at unexpected locations: Are they exported? Trends Biochem. Sci. 1999, 24, 174–177. [Google Scholar] [CrossRef]
  46. Cechetto, J.D.; Soltys, B.J.; Gupta, R.S. Localization of mitochondrial 60-kD heat shock chaperonin protein (Hsp60) in pituitary growth hormone secretory granules and pancreatic zymogen granules. J. Histochem. Cytochem. 2000, 48, 45–56. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Campanella, C.; Bucchieri, F.; Merendino, A.M.; Fucarino, A.; Burgio, G.; Corona, D.F.V.; Barbieri, G.; David, S.; Farina, F.; Zummo, G.; et al. The odyssey of Hsp60 from tumor cells to other destinations includes plasma membrane-associated stages and Golgi and exosomal protein-trafficking modalities. PLoS ONE 2012, 7, e42008. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Merendino, A.M.; Bucchieri, F.; Campanella, C.; Marcianò, V.; Ribbene, A.; David, S.; Zummo, G.; Burgio, G.; Corona, D.F.V.; Conway de Macario, E.; et al. Hsp60 is actively secreted by human tumor cells. PLoS ONE 2010, 5, e9247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Cappello, F.; Conway de Macario, E.; Marasà, L.; Zummo, G.; Macario, A.J.L. Hsp60 expression, new locations, functions and perspectives for cancer diagnosis and therapy. Cancer Biol. Ther. 2008, 7, 801–809. [Google Scholar] [CrossRef] [PubMed]
  50. Henderson, B.; Fares, M.A.; Lund, P.A. Chaperonin 60: A paradoxical, evolutionarily conserved protein family with multiple moonlighting functions. Biol. Rev. 2013, 88, 955–987. [Google Scholar] [CrossRef]
  51. Caruso Bavisotto, C.; Alberti, G.; Vitale, A.M.; Paladino, L.; Campanella, C.; Rappa, F.; Gorska, M.; Conway de Macario, E.; Cappello, F.; Macario, A.J.L.; et al. Hsp60 Post-translational modifications: Functional and pathological consequences. Front. Mol. Biosci. 2020, 7, 95. [Google Scholar] [CrossRef]
  52. Laad, A.D.; Thomas, M.L.; Fakih, A.R.; Chiplunkar, S.V. Human gamma delta T cells recognize heat shock protein-60 on oral tumor cells. Int. J. Cancer 1999, 80, 709–714. [Google Scholar] [CrossRef]
  53. Feng, H.; Zeng, Y.; Graner, M.W.; Katsanis, E. Stressed apoptotic tumor cells stimulate dendritic cells and induce specific cytotoxic T cells. Blood 2002, 100, 4108–4115. [Google Scholar] [CrossRef] [Green Version]
  54. Osterloh, A.; Meier-Stiegen, F.; Veit, A.; Fleischer, B.; Von Bonin, A.; Breloer, M. Lipopolysaccharide-free heat shock protein 60 activates T cells. J. Biol. Chem. 2004, 279, 47906–47911. [Google Scholar] [CrossRef] [Green Version]
  55. Chandra, D.; Choy, G.; Tang, D.G. Cytosolic accumulation of HSP60 during apoptosis with or without apparent mitochondrial release: Evidence that its pro-apoptotic or pro-survival functions involve differential interactions with caspase-3. J. Biol. Chem. 2007, 282, 31289–31301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Samali, A.; Cai, J.; Zhivotovsky, B.; Jones, D.P.; Orrenius, S. Presence of a pre-apoptotic complex of pro-caspase-3, Hsp60 and Hsp10 in the mitochondrial fraction of Jurkat cells. EMBO J. 1999, 18, 2040–2048. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Xanthoudakis, S.; Sophie, R.; Rasper, D.; Hennessey, T.; Aubin, Y.; Cassady, R.; Tawa, P.; Ruel, R.; Rosen, A.; Nicholson, D.W. Hsp60 accelerates the maturation of pro-caspase-3 by upstream activator proteases during apoptosis. EMBO J. 1999, 18, 2049–2056. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Shan, Y.X.; Liu, T.J.; Su, H.F.; Samsamshariat, A.; Mestril, R.; Wang, P.H. Hsp10 and Hsp60 modulate Bcl-2 family and mitochondria apoptosis signaling induced by doxorubicin in cardiac muscle cells. J. Mol. Cell. Cardiol. 2003, 35, 1135–1143. [Google Scholar] [CrossRef]
  59. Gupta, S.; Knowlton, A.A. HSP60, Bax, apoptosis and the heart. J. Cell. Mol. Med. 2005, 9, 51–58. [Google Scholar] [CrossRef] [PubMed]
  60. Lin, K.M.; Lin, B.; Lian, I.Y.; Mestril, R.; Scheffler, I.E.; Dillmann, W.H. Combined and individual mitochondrial HSP60 and HSP10 expression in cardiac myocytes protects mitochondrial function and prevents apoptotic cell deaths induced by simulated ischemia-reoxygenation. Circulation 2001, 103, 1787–1792. [Google Scholar] [CrossRef] [Green Version]
  61. Chun, J.N.; Choi, B.; Lee, K.W.; Lee, D.J.; Kang, D.H.; Lee, J.Y.; Song, I.S.; Kim, H.I.; Lee, S.H.; Kim, H.S.; et al. Cytosolic Hsp60 Is Involved in the NF-kB-Dependent Survival of Cancer Cells via IKK Regulation. PLoS ONE 2010, 5, e9422. [Google Scholar] [CrossRef]
  62. Ghosh, J.C.; Dohi, T.; Byoung, H.K.; Altieri, D.C. Hsp60 regulation of tumor cell apoptosis. J. Biol. Chem. 2008, 283, 5188–5194. [Google Scholar] [CrossRef] [Green Version]
  63. Cappello, F.; Marino Gammazza, A.; Palumbo Piccionello, A.; Campanella, C.; Pace, A.; Conway de Macario, E.; Macario, A.J.L. Hsp60 chaperonopathies and chaperonotherapy: Targets and agents. Expert Opin. Ther. Targets 2014, 18, 185–208. [Google Scholar] [CrossRef]
  64. Macario, A.J.L.; Conway de Macario, E. Sick chaperones, cellular stress, and disease. N. Engl. J. Med. 2005, 353, 1489–1501. [Google Scholar] [CrossRef]
  65. Macario, A.J.L.; Conway de Macario, E. Chaperonopathies by defect, excess, or mistake. Ann. N. Y. Acad. Sci. 2007, 1113, 178–191. [Google Scholar] [CrossRef] [PubMed]
  66. Macario, A.J.L.; Conway de Macario, E.; Cappello, F. The Chaperonopathies. Diseases with Defective Molecular Chaperones; Springer: Dordrecht, The Netherlands; Heidelberg, Germany; New York, NY, USA; London, UK, 2013; Available online: https://www.springer.com/us/book/9789400746664 (accessed on 18 December 2020).
  67. Baron, B. Role of the Post-translational Modifications of HSP60 in Disease. In Heat Shock Protein 60 in Human Diseases and Disorder; Springer: Cham, Switzerland, 2019; Volume 18, pp. 69–94. [Google Scholar]
  68. Gorska, M.; Marino Gammazza, A.; Zmijewski, M.A.; Campanella, C.; Cappello, F.; Wasiewicz, T.; Kuban-Jankowska, A.; Daca, A.; Sielicka, A.; Popowska, U.; et al. Geldanamycin-Induced Osteosarcoma Cell Death Is Associated with Hyperacetylation and Loss of Mitochondrial Pool of Heat Shock Protein 60 (Hsp60). PLoS ONE 2013, 8, e71135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Marino Gammazza, A.; Campanella, C.; Barone, R.; Caruso Bavisotto, C.; Gorska, M.; Wozniak, M.; Carini, F.; Cappello, F.; D’Anneo, A.; Lauricella, M.; et al. Doxorubicin anti-tumor mechanisms include Hsp60 post-translational modifications leading to the Hsp60/p53 complex dissociation and instauration of replicative senescence. Cancer Lett. 2017, 385, 75–86. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Barazi, H.O.; Zhou, L.; Templeton, N.S.; Krutzsch, H.C.; Roberts, D.D. Identification of heat shock protein 60 as a molecular mediator of alpha 3 beta 1 integrin activation. Cancer Res. 2002, 62, 1541–1548. [Google Scholar]
  71. Leung, W.-H.; Vong, Q.P.; Lin, W.; Bouck, D.; Wendt, S.; Sullivan, E.; Li, Y.; Bari, R.; Chen, T.; Leung, W. PRL-3 Mediates the Protein Maturation of ULBP2 by Regulating the Tyrosine Phosphorylation of HSP60. J. Immunol. 2015, 194, 2930–2941. [Google Scholar] [CrossRef] [Green Version]
  72. Direskeneli, H.; Ekşioǧlu-Demiralp, E.; Yavuz, Ş.; Ergun, T.; Shinnick, T.; Lehner, T.; Akoǧlu, T. T cell responses to 60/65 kDa heat shock protein derived peptides in Turkish patients with Behcet’s disease. J. Rheumatol. 2000, 27, 708–713. [Google Scholar]
  73. Shimizu, J.; Izumi, T.; Suzuki, N. Aberrant activation of heat shock protein 60/65 reactive T cells in patients with Behcet’s disease. Autoimmune Dis. 2012, 1, 105205. [Google Scholar] [CrossRef] [Green Version]
  74. Horváth, L.; Cervenak, L.; Oroszlán, M.; Prohászka, Z.; Uray, K.; Hudecz, F.; Baranyi, É.; Madácsy, L.; Singh, M.; Romics, L.; et al. Antibodies against different epitopes of heat-shock protein 60 in children with type 1 diabetes mellitus. Immunol. Lett. 2002, 80, 155–162. [Google Scholar] [CrossRef]
  75. Dieudé, M.; Senécal, J.L.; Raymond, Y. Induction of endothelial cell apoptosis by heat-shock protein 60-reactive antibodies from anti-endothelial cell autoantibody-positive systemic lupus erythematosus patients. Arthritis Rheum. 2004, 50, 3221–3231. [Google Scholar] [CrossRef]
  76. Jamin, C.; Dugué, C.; Alard, J.É.; Jousse, S.; Saraux, A.; Guillevin, L.; Piette, J.C.; Youinou, P. Induction of endothelial cell apoptosis by the binding of anti-endothelial cell antibodies to Hsp60 in vasculitis-associated systemic autoimmune diseases. Arthritis Rheum. 2005, 52, 4028–4038. [Google Scholar] [CrossRef]
  77. Alard, J.E.; Dueymes, M.; Youinou, P.; Jamin, C. Modulation of endothelial cell damages by anti-Hsp60 autoantibodies in systemic autoimmune diseases. Autoimmun. Rev. 2007, 6, 438–443. [Google Scholar] [CrossRef] [PubMed]
  78. Alard, J.E.; Dueymes, M.; Youinou, P.; Jamin, C. HSP60 and anti-HSP60 antibodies in vasculitis: They are two of a kind. Clin. Rev. Allergy Immunol. 2008, 35, 66–71. [Google Scholar] [CrossRef] [PubMed]
  79. Zhu, J.; Quyyumi, A.A.; Rott, D.; Csako, G.; Wu, H.; Halcox, J.; Epstein, S.E. Antibodies to human heat-shock protein 60 are associated with the presence and severity of coronary artery disease: Evidence for an autoimmune component of atherogenesis. Circulation 2001, 103, 1071–1075. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Benagiano, M.; D’Elios, M.M.; Amedei, A.; Azzurri, A.; van der Zee, R.; Ciervo, A.; Rombolà, G.; Romagnani, S.; Cassone, A.; Del Prete, G. Human 60-kDa Heat Shock Protein Is a Target Autoantigen of T Cells Derived from Atherosclerotic Plaques. J. Immunol. 2005, 174, 6509–6517. [Google Scholar] [CrossRef] [Green Version]
  81. Mandal, K.; Foteinos, G.; Jahangiri, M.; Xu, Q. Role of antiheat shock protein 60 autoantibodies in atherosclerosis. Lupus 2005, 14, 742–746. [Google Scholar] [CrossRef]
  82. Bodolay, E.; Prohászka, Z.; Paragh, G.; Csipő, I.; Nagy, G.; Laczik, R.; Demeter, N.; Zöld, E.; Nakken, B.; Szegedi, G.; et al. Increased levels of anti-heat-shock protein 60 (anti-Hsp60) indicate endothelial dysfunction, atherosclerosis and cardiovascular diseases in patients with mixed connective tissue disease. Immunol. Res. 2014, 60, 50–59. [Google Scholar] [CrossRef]
  83. Mantej, J.; Polasik, K.; Piotrowska, E.; Tukaj, S. Autoantibodies to heat shock proteins 60, 70, and 90 in patients with rheumatoid arthritis. Cell Stress Chaperones 2019, 24, 283–287. [Google Scholar] [CrossRef]
  84. Meng, Q.; Li, B.X.; Xiao, X. Toward developing chemical modulators of Hsp60 as potential therapeutics. Front. Mol. Biosci. 2018, 5, 35. [Google Scholar] [CrossRef] [Green Version]
  85. Cappello, F.; Di Stefano, A.; D’Anna, S.E.; Donner, C.F.; Zummo, G. Immunopositivity of heat shock protein 60 as a biomarker of bronchial carcinogenesis. Lancet Oncol. 2005, 6, 816. [Google Scholar] [CrossRef] [Green Version]
  86. Cappello, F.; Bellafiore, M.; Palma, A.; Marciano, V.; Martorana, G.; Belfiore, P.; Martorana, A.; Farina, F.; Zummo, G.; Bucchieri, F. Expression of 60-kD Heat Shock Protein Increases during Carcinogenesis in the Uterine Exocervix. Pathobiology 2002, 70, 83–88. [Google Scholar] [CrossRef]
  87. Castle, P.E.; Ashfaq, R.; Ansari, F.; Muller, C.Y. Immunohistochemical evaluation of heat shock proteins in normal and preinvasive lesions of the cervix. Cancer Lett. 2005, 229, 245–252. [Google Scholar] [CrossRef] [PubMed]
  88. Hwang, Y.J.; Lee, S.P.; Kim, S.Y.; Choi, Y.H.; Kim, M.J.; Lee, C.H.; Lee, J.Y.; Kim, D.Y. Expression of heat shock protein 60 kDa is upregulated in cervical cancer. Yonsei Med. J. 2009, 50, 399–406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Hjerpe, E.; Egyhazi, S.; Carlson, J.; Stolt, M.F.; Schedvins, K.; Johansson, H.; Shoshan, M.; Åvall-Lundqvist, E. HSP60 predicts survival in advanced serous ovarian cancer. Int. J. Gynecol. Cancer 2013, 23, 448–455. [Google Scholar] [CrossRef] [PubMed]
  90. Guo, J.; Li, X.; Zhang, W.; Chen, Y.; Zhu, S.; Chen, L.; Xu, R.; Lv, Y.; Wu, D.; Guo, M.; et al. HSP60-regulated Mitochondrial Proteostasis and Protein Translation Promote Tumor Growth of Ovarian Cancer. Sci. Rep. 2019, 9, 12628. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Bodoor, K.; Abu-Sheikha, A.; Matalka, I.; Alzou’bi, H.; Batiha, O.; Abu-Awad, A.; Jalboush, S.A.; Fayyad, L.M.; Qadiri, E.; Jarun, Y.; et al. Immunohistochemical analysis of heat shock proteins in triple negative breast cancer: HSP60 expression is a marker of poor prognosis. Eur. J. Gynaecol. Oncol. 2018, 39, 926–934. [Google Scholar]
  92. Kim, S.K.; Kim, K.; Ryu, J.W.; Ryu, T.Y.; Lim, J.H.; Oh, J.H.; Min, J.K.; Jung, C.R.; Hamamoto, R.; Son, M.Y.; et al. The novel prognostic marker, EHMT2, is involved in cell proliferation via HSPD1 regulation in breast cancer. Int. J. Oncol. 2019, 54, 65–76. [Google Scholar] [CrossRef] [Green Version]
  93. Cappello, F.; Bellafiore, M.; Palma, A.; David, S.; Marcianò, V.; Bartolotta, T.; Sciumè, C.; Modica, G.; Farina, F.; Zummo, G.; et al. 60 KDa chaperonin (HSP60) is over-expressed during colorectal carcinogenesis. Eur. J. Histochem. 2003, 47, 105–109. [Google Scholar] [CrossRef]
  94. Xu, X.; Wang, W.; Shao, W.; Yin, W.; Chen, H.; Qiu, Y.; Mo, M.; Zhao, J.; Deng, Q.; He, J. Heat shock protein-60 expression was significantly correlated with the prognosis of lung adenocarcinoma. J. Surg. Oncol. 2011, 104, 598–603. [Google Scholar] [CrossRef]
  95. Aǧababaoǧlu, I.; Önen, A.; Demir, A.B.; Aktaş, S.; Altun, Z.; Ersöz, H.; Şanlii, A.; Özdemir, N.; Akkoçlu, A. Chaperonin (HSP60) and annexin-2 are candidate biomarkers for non-small cell lung carcinoma. Medicine 2017, 96, e5903. [Google Scholar] [CrossRef]
  96. Castilla, C.; Congregado, B.; Conde, J.M.; Medina, R.; Torrubia, F.J.; Japn, M.A.; Sáez, C. Immunohistochemical expression of Hsp60 correlates with tumor progression and hormone resistance in prostate cancer. Urology 2010, 76, 1017.e1–1017.e6. [Google Scholar] [CrossRef]
  97. Giaginis, C.; Daskalopoulou, S.S.; Vgenopoulou, S.; Sfiniadakis, I.; Kouraklis, G.; Theocharis, S.E. Heat Shock Protein-27, -60 and -90 expression in gastric cancer: Association with clinicopathological variables and patient survival. BMC Gastroenterol. 2009, 9, 14. [Google Scholar] [CrossRef] [Green Version]
  98. Li, X.S.; Xu, Q.; Fu, X.Y.; Luo, W.S. Heat shock protein 60 overexpression is associated with the progression and prognosis in gastric cancer. PLoS ONE 2014, 9, e107507. [Google Scholar] [CrossRef] [PubMed]
  99. Pitruzzella, A.; Paladino, L.; Vitale, A.M.; Martorana, S.; Cipolla, C.; Graceffa, G.; Cabibi, D.; David, S.; Fucarino, A.; Bucchieri, F.; et al. Quantitative immunomorphological analysis of heat shock proteins in thyroid follicular adenoma and carcinoma tissues reveals their potential for differential diagnosis and points to a role in carcinogenesis. Appl. Sci. 2019, 9, 4324. [Google Scholar] [CrossRef] [Green Version]
  100. Caruso Bavisotto, C.; Cipolla, C.; Graceffa, G.; Barone, R.; Bucchieri, F.; Bulone, D.; Cabibi, D.; Campanella, C.; Marino Gammazza, A.; Pitruzzella, A.; et al. Immunomorphological pattern of molecular chaperones in normal and pathological thyroid tissues and circulating exosomes: Potential use in clinics. Int. J. Mol. Sci. 2019, 20, 4496. [Google Scholar] [CrossRef] [Green Version]
  101. Thomas, X.; Campos, L.; Mounier, C.; Cornillon, J.; Flandrin, P.; Le, Q.H.; Piselli, S.; Guyotat, D. Expression of heat-shock proteins is associated with major adverse prognostic factors in acute myeloid leukemia. Leuk. Res. 2005, 29, 1049–1058. [Google Scholar] [CrossRef]
  102. Ghosh, J.C.; Siegelin, M.D.; Dohi, T.; Altieri, D.C. Heat shock protein 60 regulation of the mitochondrial permeability transition pore in tumor cells. Cancer Res. 2010, 70, 8988–8993. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Caruso Bavisotto, C.; Graziano, F.; Rappa, F.; Marino Gammazza, A.; Logozzi, M.; Fais, S.; Maugeri, R.; Bucchieri, F.; Conway de Macario, E.; Macario, A.J.L.; et al. Exosomal chaperones and miRNAs in gliomagenesis: State-of-art and theranostics perspectives. Int. J. Mol. Sci. 2018, 19, 2626. [Google Scholar] [CrossRef] [Green Version]
  104. Graziano, F.; Caruso Bavisotto, C.; Marino Gammazza, A.; Rappa, F.; Conway de Macario, E.; Macario, A.J.L.; Cappello, F.; Campanella, C.; Maugeri, R.; Iacopino, D.G. Chaperonology: The third eye on brain gliomas. Brain Sci. 2018, 8, 110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Abu-Hadid, M.; Wilkes, J.D.; Elakawi, Z.; Pendyala, L.; Perez, R.P. Relationship between heat shock protein 60 (HSP6O) mRNA expression and resistance to platinum analogues in human ovarian and bladder carcinoma cell lines. Cancer Lett. 1997, 119, 63–70. [Google Scholar] [CrossRef]
  106. Tsai, Y.P.; Yang, M.H.; Huang, C.H.; Chang, S.Y.; Chen, P.M.; Liu, C.J.; Teng, S.C.; Wu, K.J. Interaction between HSP60 and β-catenin promotes metastasis. Carcinogenesis 2009, 30, 1049–1057. [Google Scholar] [CrossRef]
  107. Wu, J.; Liu, T.; Rios, Z.; Mei, Q.; Lin, X.; Cao, S. Heat Shock Proteins and Cancer. Trends Pharmacol. Sci. 2017, 38, 226–256. [Google Scholar] [CrossRef]
  108. Ruan, W.; Wang, Y.; Ma, Y.; Xing, X.; Lin, J.; Cui, J.; Lai, M. HSP60, a protein downregulated by IGFBP7 in colorectal carcinoma. J. Exp. Clin. Cancer Res. 2010, 29, 41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Tang, H.; Li, J.; Liu, X.; Wang, G.; Luo, M.; Deng, H. Down-regulation of HSP60 Suppresses the Proliferation of Glioblastoma Cells via the ROS/AMPK/mTOR Pathway. Sci. Rep. 2016, 6, 28388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Campanella, C.; D’Anneo, A.; Marino Gammazza, A.; Caruso Bavisotto, C.; Barone, R.; Emanuele, S.; Lo Cascio, F.; Mocciaro, E.; Fais, S.; Conway de Macario, E.; et al. The histone deacetylase inhibitor SAHA induces HSP60 nitration and its extracellular release by exosomal vesicles in human lung-derived carcinoma cells. Oncotarget 2016, 7, 28849–28867. [Google Scholar] [CrossRef] [PubMed]
  111. Zhou, C.; Sun, H.; Zheng, C.; Gao, J.; Fu, Q.; Hu, N.; Shao, X.; Zhou, Y.; Xiong, J.; Nie, K.; et al. Oncogenic HSP60 regulates mitochondrial oxidative phosphorylation to support Erk1/2 activation during pancreatic cancer cell growth. Cell Death Dis. 2018, 9, 161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Caruso Bavisotto, C.; Marino Gammazza, A.; Lo Cascio, F.; Mocciaro, E.; Vitale, A.M.; Vergilio, G.; Pace, A.; Cappello, F.; Campanella, C.; Palumbo Piccionello, A. Curcumin affects Hsp60 folding activity and levels in neuroblastoma cells. Int. J. Mol. Sci. 2020, 21, 661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Kim, W.; Ryu, J.; Kim, J.E. CCAR2/DBC1 and Hsp60 positively regulate expression of survivin in neuroblastoma cells. Int. J. Mol. Sci. 2019, 20, 131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Chaiwatanasirikul, K.A.; Sala, A. The tumour-suppressive function of CLU is explained by its localisation and interaction with HSP60. Cell Death Dis. 2011, 2, e219. [Google Scholar] [CrossRef] [Green Version]
  115. Campanella, C.; Bucchieri, F.; Ardizzone, N.M.; Marino Gammazza, A.; Montalbano, A.; Ribbene, A.; Di Felice, V.; Bellafiore, M.; David, S.; Rappa, F.; et al. Upon oxidative stress, the antiapoptotic Hsp60/procaspase-3 complex persists in mucoepidermoid carcinoma cells. Eur. J. Histochem. 2008, 52, 221–228. [Google Scholar] [CrossRef] [Green Version]
  116. Caruso Bavisotto, C.; Nikolic, D.; Marino Gammazza, A.; Barone, R.; Lo Cascio, F.; Mocciaro, E.; Zummo, G.; Conway de Macario, E.; Macario, A.J.L.; Cappello, F.; et al. The dissociation of the Hsp60/pro-Caspase-3 complex by bis(pyridyl)oxadiazole copper complex (CubipyOXA) leads to cell death in NCI-H292 cancer cells. J. Inorg. Biochem. 2017, 170, 8–16. [Google Scholar] [CrossRef] [Green Version]
  117. Tang, H.; Chen, Y.; Liu, X.; Wang, S.; Lv, Y.; Wu, D.; Wang, Q.; Luo, M.; Deng, H. Downregulation of HSP60 disrupts mitochondrial proteostasis to promote tumorigenesis and progression in clear cell renal cell carcinoma. Oncotarget 2016, 7, 38822–38834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Zhang, J.; Zhou, X.; Chang, H.; Huang, X.; Guo, X.; Du, X.; Tian, S.; Wang, L.; Lyv, Y.; Yuan, P.; et al. Hsp60 exerts a tumor suppressor function by inducing cell differentiation and inhibiting invasion in hepatocellular carcinoma. Oncotarget 2016, 7, 68976–68989. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Cappello, F.; Di Stefano, A.; David, S.; Rappa, F.; Anzalone, R.; La Rocca, G.; D’Anna, S.E.; Magno, F.; Donner, C.F.; Balbi, B.; et al. Hsp60 and Hsp10 down-regulation predicts bronchial epithelial carcinogenesis in smokers with chronic obstructive pulmonary disease. Cancer 2006, 107, 2417–2424. [Google Scholar] [CrossRef]
  120. Hamelin, C.; Cornut, E.; Poirier, F.; Pons, S.; Beaulieu, C.; Charrier, J.P.; Haïdous, H.; Cotte, E.; Lambert, C.; Piard, F.; et al. Identification and verification of heat shock protein 60 as a potential serum marker for colorectal cancer. FEBS J. 2011, 278, 4845–4859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Lv, L.H.; Wan, Y.L.; Lin, Y.; Zhang, W.; Yang, M.; Li, G.N.; Lin, H.M.; Shang, C.Z.; Chen, Y.J.; Min, J. Anticancer drugs cause release of exosomes with heat shock proteins from human hepatocellular carcinoma cells that elicit effective natural killer cell antitumor responses in vitro. J. Biol. Chem. 2012, 287, 15874–15885. [Google Scholar] [CrossRef] [Green Version]
  122. Lebret, T.; Watson, R.W.G.; Molinié, V.; O’Neill, A.; Gabriel, C.; Fitzpatrick, J.M.; Botto, H. Heat shock proteins HSP27, HSP60, HSP70, and HSP90: Expression in bladder carcinoma. Cancer 2003, 98, 970–977. [Google Scholar] [CrossRef]
  123. Campanella, C.; Rappa, F.; Sciumè, C.; Marino Gammazza, A.; Barone, R.; Bucchieri, F.; David, S.; Curcurù, G.; Caruso Bavisotto, C.; Pitruzzella, A.; et al. Heat shock protein 60 levels in tissue and circulating exosomes in human large bowel cancer before and after ablative surgery. Cancer 2015, 121, 3230–3239. [Google Scholar] [CrossRef]
  124. Théry, C.; Witwer, K.W.; Aikawa, E.; Alcaraz, M.J.; Anderson, J.D.; Andriantsitohaina, R.; Antoniou, A.; Arab, T.; Archer, F.; Atkin-Smith, G.K.; et al. Minimal information for studies of extracellular vesicles 2018 (MISEV2018): A position statement of the International Society for Extracellular Vesicles and update of the MISEV2014 guidelines. J. Extracell. Vesicles 2018, 7, 1535750. [Google Scholar] [CrossRef] [Green Version]
  125. Stoorvogel, W.; Kleijmeer, M.J.; Geuze, H.J.; Raposo, G. The biogenesis and functions of exosomes. Traffic 2002, 3, 321–330. [Google Scholar] [CrossRef]
  126. Camussi, G.; Deregibus, M.C.; Cantaluppi, V. Role of stem-cell-derived microvesicles in the paracrine action of stem cells. Biochem. Soc. Trans. 2013, 41, 283–287. [Google Scholar] [CrossRef]
  127. Liu, Q.; Rojas-Canales, D.M.; Divito, S.J.; Shufesky, W.J.; Stolz, D.B.; Erdos, G.; Sullivan, M.L.G.; Gibson, G.A.; Watkins, S.C.; Larregina, A.T.; et al. Donor dendritic cell-derived exosomes promote allograft-targeting immune response. J. Clin. Investig. 2016, 126, 2805–2820. [Google Scholar] [CrossRef] [PubMed]
  128. Long, Q.; Upadhya, D.; Hattiangady, B.; Kim, D.K.; An, S.Y.; Shuai, B.; Prockop, D.J.; Shetty, A.K. Intranasal MSC-derived A1-exosomes ease inflammation, and prevent abnormal neurogenesis and memory dysfunction after status epilepticus. Proc. Natl. Acad. Sci. USA 2017, 114, E3536–E3545. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Whiteside, T.L. Exosome and mesenchymal stem cell cross-talk in the tumor microenvironment. Semin. Immunol. 2018, 35, 69–79. [Google Scholar] [CrossRef] [PubMed]
  130. Tai, Y.L.; Chen, K.C.; Hsieh, J.T.; Shen, T.L. Exosomes in cancer development and clinical applications. Cancer Sci. 2018, 109, 2364–2374. [Google Scholar] [CrossRef] [Green Version]
  131. Caruso Bavisotto, C.; Marino Gammazza, A.; Rappa, F.; Fucarino, A.; Pitruzzella, A.; David, S.; Campanella, C. Exosomes: Can doctors still ignore their existence? EuroMediterr. Biomed. J. 2013, 8, 136–139. [Google Scholar]
  132. Vesiclepedia: Browse Results. Available online: http://microvesicles.org/browse_results?org_name=Homosapiens&cont_type=&tissue=&gene_symbol=&ves_type= (accessed on 10 January 2021).
  133. ExoCarta: Browse Results. Available online: http://exocarta.org/browse_results?org_name=Homosapiens&cont_type=&tissue=&gene_symbol= (accessed on 10 January 2021).
  134. Maacha, S.; Bhat, A.A.; Jimenez, L.; Raza, A.; Haris, M.; Uddin, S.; Grivel, J.C. Extracellular vesicles-mediated intercellular communication: Roles in the tumor microenvironment and anti-cancer drug resistance. Mol. Cancer 2019, 18, 55. [Google Scholar] [CrossRef] [Green Version]
  135. Dickman, C.T.; Lawson, J.; Jabalee, J.; MacLellan, S.A.; LePard, N.E.; Bennewith, K.L.; Garnis, C. Selective extracellular vesicle exclusion of miR-142-3p by oral cancer cells promotes both internal and extracellular malignant phenotypes. Oncotarget 2017, 8, 15252–15266. [Google Scholar] [CrossRef] [Green Version]
  136. Lawson, J.; Dickman, C.; MacLellan, S.; Towle, R.; Jabalee, J.; Lam, S.; Garnis, C. Selective secretion of microRNAs from lung cancer cells via extracellular vesicles promotes CAMK1D-mediated tube formation in endothelial cells. Oncotarget 2017, 8, 83913–83924. [Google Scholar] [CrossRef] [Green Version]
  137. Fais, S.; O’Driscoll, L.; Borras, F.E.; Buzas, E.; Camussi, G.; Cappello, F.; Carvalho, J.; Da Silva, A.C.; Del Portillo, H.; El Andaloussi, S.; et al. Evidence-Based Clinical Use of Nanoscale Extracellular Vesicles in Nanomedicine. ACS Nano 2016, 10, 3886–3899. [Google Scholar] [CrossRef]
  138. Campanella, C.; Caruso Bavisotto, C.; Logozzi, M.; Marino Gammazza, A.; Mizzoni, D.; Cappello, F.; Fais, S. On the Choice of the Extracellular Vesicles for Therapeutic Purposes. Int. J. Mol. Sci. 2019, 20, 236. [Google Scholar] [CrossRef] [Green Version]
  139. Caruso Bavisotto, C.; Cappello, F.; Macario, A.J.L.; Conway de Macario, E.; Logozzi, M.; Fais, S.; Campanella, C. Exosomal HSP60: A potentially useful biomarker for diagnosis, assessing prognosis, and monitoring response to treatment. Expert Rev. Mol. Diagn. 2017, 17, 815–822. [Google Scholar] [CrossRef] [PubMed]
  140. Bartel, D.P. MicroRNAs: Genomics, Biogenesis, Mechanism, and Function. Cell 2004, 116, 281–297. [Google Scholar] [CrossRef] [Green Version]
  141. Lewis, B.P.; Burge, C.B.; Bartel, D.P. Conserved seed pairing, often flanked by adenosines, indicates that thousands of human genes are microRNA targets. Cell 2005, 120, 15–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Friedman, R.C.; Farh, K.K.H.; Burge, C.B.; Bartel, D.P. Most mammalian mRNAs are conserved targets of microRNAs. Genome Res. 2009, 19, 92–105. [Google Scholar] [CrossRef] [Green Version]
  143. Kumar, M.S.; Lu, J.; Mercer, K.L.; Golub, T.R.; Jacks, T. Impaired microRNA processing enhances cellular transformation and tumorigenesis. Nat. Genet. 2007, 39, 673–677. [Google Scholar] [CrossRef]
  144. Liu, M.; Zhu, K.; Qian, X.; Li, W. Identification of miRNA/mRNA-negative regulation pairs in nasopharyngeal carcinoma. Med. Sci. Monit. 2016, 22, 2215–2234. [Google Scholar] [CrossRef] [Green Version]
  145. Peng, Y.; Croce, C.M. The role of microRNAs in human cancer. Signal Transduct. Target. Ther. 2016, 1, 15004. [Google Scholar] [CrossRef] [Green Version]
  146. Svoronos, A.A.; Engelman, D.M.; Slack, F.J. OncomiR or tumor suppressor? The duplicity of MicroRNAs in cancer. Cancer Res. 2016, 76, 3666–3670. [Google Scholar] [CrossRef] [Green Version]
  147. Lan, H.; Lu, H.; Wang, X.; Jin, H. MicroRNAs as potential biomarkers in cancer: Opportunities and challenges. BioMed Res. Int. 2015, 2015, 125094. [Google Scholar] [CrossRef] [Green Version]
  148. Callegari, E.; Elamin, B.K.; Giannone, F.; Milazzo, M.; Altavilla, G.; Fornari, F.; Giacomelli, L.; D’Abundo, L.; Ferracin, M.; Bassi, C.; et al. Liver tumorigenicity promoted by microRNA-221 in a mouse transgenic model. Hepatology 2012, 56, 1025–1033. [Google Scholar] [CrossRef]
  149. Gwak, J.M.; Kim, H.J.; Kim, E.J.; Chung, Y.R.; Yun, S.; Seo, A.N.; Lee, H.J.; Park, S.Y. MicroRNA-9 is associated with epithelial-mesenchymal transition, breast cancer stem cell phenotype, and tumor progression in breast cancer. Breast Cancer Res. Treat. 2014, 147, 39–49. [Google Scholar] [CrossRef]
  150. Tanaka, R.; Tomosugi, M.; Horinaka, M.; Sowa, Y.; Sakai, T. Metformin causes G1-phase arrest via down-regulation of MIR-221 and enhances TRAIL sensitivity through DR5 up-regulation in pancreatic cancer cells. PLoS ONE 2015, 10, e0125779. [Google Scholar] [CrossRef] [Green Version]
  151. Munoz, J.L.; Rodriguez-Cruz, V.; Rameshwar, P. High expression of miR-9 in CD133 + glioblastoma cells in chemoresistance to temozolomide. J. Cancer Stem Cell Res. 2015, 3, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Cheng, C.-W.; Yu, J.-C.; Hsieh, Y.-H.; Liao, W.-L.; Shieh, J.-C.; Yao, C.-C.; Lee, H.-J.; Chen, P.-M.; Wu, P.-E.; Shen, C.-Y. Increased Cellular Levels of MicroRNA-9 and MicroRNA-221 Correlate with Cancer Stemness and Predict Poor Outcome in Human Breast Cancer. Cell. Physiol. Biochem. 2018, 48, 2205–2218. [Google Scholar] [CrossRef] [PubMed]
  153. Kao, C.J.; Martiniez, A.; Shi, X.B.; Yang, J.; Evans, C.P.; Dobi, A.; Devere White, R.W.; Kung, H.J. MiR-30 as a tumor suppressor connects EGF/Src signal to ERG and EMT. Oncogene 2014, 33, 2495–2503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Ozgur, A.; Tutar, L.; Tutar, Y. Regulation of Heat Shock Proteins by miRNAs in Human Breast Cancer. MicroRNA 2014, 3, 118–135. [Google Scholar] [CrossRef]
  155. Choghaei, E.; Khamisipour, G.; Falahati, M.; Naeimi, B.; Mossahebi-Mohammadi, M.; Tahmasebi, R.; Hasanpour, M.; Shamsian, S.; Hashemi, Z.S. Knockdown of microRNA-29a changes the expression of heat shock proteins in breast carcinoma MCF-7 cells. Oncol. Res. 2016, 23, 69–78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Liang, W.; Liao, Y.; Li, Z.; Wang, Y.; Zheng, S.; Xu, X.; Ran, F.; Tang, B.; Wang, Z. MicroRNA-644a promotes apoptosis of hepatocellular carcinoma cells by downregulating the expression of heat shock factor 1. Cell Commun. Signal. 2018, 16, 30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Ban, H.S.; Han, T.S.; Hur, K.; Cho, H.S. Epigenetic alterations of Heat Shock Proteins (HSPs) in cancer. Int. J. Mol. Sci. 2019, 20, 4758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Shan, Z.X.; Lin, Q.X.; Deng, C.Y.; Zhu, J.N.; Mai, L.P.; Liu, J.L.; Fu, Y.H.; Liu, X.Y.; Li, Y.X.; Zhang, Y.Y.; et al. MiR-1/miR-206 regulate Hsp60 expression contributing to glucose-mediated apoptosis in cardiomyocytes. FEBS Lett. 2010, 584, 3592–3600. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Hu, Y.; Chen, X.; Li, X.; Li, Z.; Diao, H.; Liu, L.; Zhang, J.; Ju, J.; Wen, L.; Liu, X.; et al. MicroRNA-1 downregulation induced by carvedilol protects cardiomyocytes against apoptosis by targeting heat shock protein 60. Mol. Med. Rep. 2019, 49, 3527–3536. [Google Scholar] [CrossRef]
  160. Wang, L.L.; Dong, J.J.; An, B.Z.; Liang, J.; Cai, K.R.; Jin, Z.S.; Jin, H.S.; Hu, J.P. Has-miR-17 increases the malignancy of gastric lymphoma by HSP60/TNFR2 pathway. J. Biol. Regul. Homeost. Agents 2020, 34, 1317–1324. [Google Scholar] [PubMed]
  161. Ciocca, D.R.; Calderwood, S.K. Heat shock proteins in cancer: Diagnostic, prognostic, predictive, and treatment implications. Cell Stress Chaperones 2005, 10, 86–103. [Google Scholar] [CrossRef]
  162. Lianos, G.D.; Alexiou, G.A.; Mangano, A.; Mangano, A.; Rausei, S.; Boni, L.; Dionigi, G.; Roukos, D.H. The role of heat shock proteins in cancer. Cancer Lett. 2015, 360, 114–118. [Google Scholar] [CrossRef]
  163. Qian-Lin, Z.; Ting-Feng, W.; Qi-Feng, C.; Min-Hua, Z.; Ai-Guo, L. Inhibition of cytosolic chaperonin CCTζ-1 expression depletes proliferation of colorectal carcinoma in vitro. J. Surg. Oncol. 2010, 102, 419–423. [Google Scholar] [CrossRef]
  164. Dong, D.; Stapleton, C.; Luo, B.; Xiong, S.; Ye, W.; Zhang, Y.; Jhaveri, N.; Zhu, G.; Ye, R.; Liu, Z.; et al. A critical role for GRP78/BiP in the tumor microenvironment for neovascularization during tumor growth and metastasis. Cancer Res. 2011, 71, 2848–2857. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Wang, J.; Cui, S.; Zhang, X.; Wu, Y.; Tang, H. High Expression of Heat Shock Protein 90 Is Associated with Tumor Aggressiveness and Poor Prognosis in Patients with Advanced Gastric Cancer. PLoS ONE 2013, 8, e62876. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Jagadish, N.; Agarwal, S.; Gupta, N.; Fatima, R.; Devi, S.; Kumar, V.; Suri, V.; Kumar, R.; Suri, V.; Sadasukhi, T.C.; et al. Heat shock protein 70-2 (HSP70-2) overexpression in breast cancer. J. Exp. Clin. Cancer Res. 2016, 35, 1–14. [Google Scholar] [CrossRef] [Green Version]
  167. Huang, C.-Y.; Wei, P.-L.; Chen, W.-Y.; Chang, W.-C.; Chang, Y.-J. Silencing Heat Shock Protein 27 Inhibits the Progression and Metastasis of Colorectal Cancer (CRC) by Maintaining the Stability of Stromal Interaction Molecule 1 (STIM1) Proteins. Cells 2018, 7, 262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Rappa, F.; Farina, F.; Zummo, G.; David, S.; Campanella, C.; Carini, F.; Tomasello, G.; Damiani, P.; Cappello, F.; Conway de Macario, E.; et al. HSP-molecular chaperones in cancer biogenesis and tumor therapy: An overview. AntiCancer Res. 2012, 32, 5139–5150. [Google Scholar]
  169. Chatterjee, S.; Burns, T.F. Targeting heat shock proteins in cancer: A promising therapeutic approach. Int. J. Mol. Sci. 2017, 18, 1978. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Desmetz, C.; Bibeau, F.; Boissière, F.; Bellet, V.; Rouanet, P.; Maudelonde, T.; Mangé, A.; Solassol, J. Proteomics-based identification of HSP60 as a tumor-associated antigen in early-stage breast cancer and ductal carcinoma in situ. J. Proteome Res. 2008, 7, 3830–3837. [Google Scholar] [CrossRef] [PubMed]
  171. Abdalla, M.A.K.; Haj-Ahmad, Y. Promising urinary protein biomarkers for the early detection of hepato-cellular carcinoma among high-risk hepatitis C virus egyptian patients. J. Cancer 2012, 3, 390–403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Chang, C.-L.; Hsu, Y.-T.; Wu, C.-C.; Yang, Y.-C.; Wang, C.; Wu, T.-C.; Hung, C.-F. Immune Mechanism of the Antitumor Effects Generated by Bortezomib. J. Immunol. 2012, 189, 3209–3220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Murshid, A.; Gong, J.; Stevenson, M.A.; Calderwood, S.K. Heat shock proteins and cancer vaccines: Developments in the past decade and chaperoning in the decade to come. Expert Rev. Vaccines 2011, 10, 1553–1568. [Google Scholar] [CrossRef] [Green Version]
  174. Wiechmann, K.; Müller, H.; König, S.; Wielsch, N.; Svatoš, A.; Jauch, J.; Werz, O. Mitochondrial Chaperonin HSP60 Is the Apoptosis-Related Target for Myrtucommulone. Cell Chem. Biol. 2017, 24, 614–623.e6. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Hsp60 plays multiple roles intra- and extracellularly. Its canonical functions pertain to maintenance of protein homeostasis inside mitochondria, whereas its non-canonical roles are unrelated to protein quality control and are played in a variety of locations (shown with a yellow star). The chaperonin (small dark-pink cylinder) is typically in the mitochondrion matrix (1) in which it assists in the folding of intra-mitochondrial proteins, for instance those of the electron transport system, and therefore it is a vital molecule for maintaining cell viability and organismal physiology under normal conditions and in the face of stress. Consequently, Hsp60 chaperonopathies are usually serious conditions. The non-canonical functions of Hsp60 displayed in a variety of locations beyond the mitochondria, for example in the cytosol (2), are also vital. One example is regulation of apoptosis, which can be in either direction pro- or anti-apoptotic with implications for carcinogenesis. The anti-apoptotic effect of Hsp60 helps cancer cells to become immortal. In the plasma-cell membrane (3), Hsp60 can be recognized by immune cells and antibodies and generate immune reactions that damage the cell, which is a welcome event if the target cell is a cancerous one, but it is a pathogenic reaction when the cell is a normal one, a vascular epithelial cell, for instance, generating autoimmune conditions. Hsp60 exits cells (4) via different mechanisms and can, thus, reach molecules in the extracellular space and other cells nearby and far away, the latter via blood free or on particles, such as platelets, red cells, and microvesicles, like exosomes. In this manner, Hsp60 plays a role in intercellular communication, as illustrated by the microglia activation that occurs in some neurodegenerative diseases via the Hsp60-TLR-4-NF-κB signaling pathway.
Figure 1. Hsp60 plays multiple roles intra- and extracellularly. Its canonical functions pertain to maintenance of protein homeostasis inside mitochondria, whereas its non-canonical roles are unrelated to protein quality control and are played in a variety of locations (shown with a yellow star). The chaperonin (small dark-pink cylinder) is typically in the mitochondrion matrix (1) in which it assists in the folding of intra-mitochondrial proteins, for instance those of the electron transport system, and therefore it is a vital molecule for maintaining cell viability and organismal physiology under normal conditions and in the face of stress. Consequently, Hsp60 chaperonopathies are usually serious conditions. The non-canonical functions of Hsp60 displayed in a variety of locations beyond the mitochondria, for example in the cytosol (2), are also vital. One example is regulation of apoptosis, which can be in either direction pro- or anti-apoptotic with implications for carcinogenesis. The anti-apoptotic effect of Hsp60 helps cancer cells to become immortal. In the plasma-cell membrane (3), Hsp60 can be recognized by immune cells and antibodies and generate immune reactions that damage the cell, which is a welcome event if the target cell is a cancerous one, but it is a pathogenic reaction when the cell is a normal one, a vascular epithelial cell, for instance, generating autoimmune conditions. Hsp60 exits cells (4) via different mechanisms and can, thus, reach molecules in the extracellular space and other cells nearby and far away, the latter via blood free or on particles, such as platelets, red cells, and microvesicles, like exosomes. In this manner, Hsp60 plays a role in intercellular communication, as illustrated by the microglia activation that occurs in some neurodegenerative diseases via the Hsp60-TLR-4-NF-κB signaling pathway.
Applsci 11 01175 g001
Figure 2. Diagram representing the Hsp60 secreting mechanism. Hsp60 (pink cylinder) is encoded in a single nuclear gene (1). After translation (2), Hsp60 is translocated to the mitochondrion matrix, in which it performs its canonical chaperoning function (3). In tumor cells, Hsp60 accumulates in the cytosol (4), and/or reaches the plasma-cell membrane, near lipid rafts (yellow) (5). The membrane-associated Hsp60 is internalized through a mechanism of endocytosis into early endosomes (6), and then secreted via exosomes after the fusion of multivesicular bodies (MVBs) with the plasma-cell membrane (7). Hsp60 is also exported extracellularly via the classic Golgi-mediated secretion pathway (8). Exosomal Hsp60 (9) and free Hsp60 (10) can reach other cells, e.g., tumor cells and immune cells, nearby or in distant sites via circulation.
Figure 2. Diagram representing the Hsp60 secreting mechanism. Hsp60 (pink cylinder) is encoded in a single nuclear gene (1). After translation (2), Hsp60 is translocated to the mitochondrion matrix, in which it performs its canonical chaperoning function (3). In tumor cells, Hsp60 accumulates in the cytosol (4), and/or reaches the plasma-cell membrane, near lipid rafts (yellow) (5). The membrane-associated Hsp60 is internalized through a mechanism of endocytosis into early endosomes (6), and then secreted via exosomes after the fusion of multivesicular bodies (MVBs) with the plasma-cell membrane (7). Hsp60 is also exported extracellularly via the classic Golgi-mediated secretion pathway (8). Exosomal Hsp60 (9) and free Hsp60 (10) can reach other cells, e.g., tumor cells and immune cells, nearby or in distant sites via circulation.
Applsci 11 01175 g002
Table 1. Classification of chaperonopathies according to pathogenic mechanism.
Table 1. Classification of chaperonopathies according to pathogenic mechanism.
ChaperonopathiesMechanismExampleReferences
ExcessQuantitative variation, in which a gene is dysregulated or overexpressed.
Qualitative variation, with a gain of function.
e.g., Alzheimer’s disease, Parkinson’s disease, Huntington’s disease. [65]
DefectQuantitative variation with gene downregulation.
Qualitative variation, due to structural defect (genetic or acquired)
e.g., Charcot–Marie–Tooth disease, Spastic paraplegia, Hypo-myelinating leukodystrophy[65]
Mistake or collaborationismThe chaperone is normal but the pathway in which it is involved may promote cell pathology.e.g., certain tumor types, autoimmune conditions, prion disease[65]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

David, S.; Vitale, A.M.; Fucarino, A.; Scalia, F.; Vergilio, G.; Conway de Macario, E.; Macario, A.J.L.; Caruso Bavisotto, C.; Pitruzzella, A. The Challenging Riddle about the Janus-Type Role of Hsp60 and Related Extracellular Vesicles and miRNAs in Carcinogenesis and the Promises of Its Solution. Appl. Sci. 2021, 11, 1175. https://doi.org/10.3390/app11031175

AMA Style

David S, Vitale AM, Fucarino A, Scalia F, Vergilio G, Conway de Macario E, Macario AJL, Caruso Bavisotto C, Pitruzzella A. The Challenging Riddle about the Janus-Type Role of Hsp60 and Related Extracellular Vesicles and miRNAs in Carcinogenesis and the Promises of Its Solution. Applied Sciences. 2021; 11(3):1175. https://doi.org/10.3390/app11031175

Chicago/Turabian Style

David, Sabrina, Alessandra Maria Vitale, Alberto Fucarino, Federica Scalia, Giuseppe Vergilio, Everly Conway de Macario, Alberto J. L. Macario, Celeste Caruso Bavisotto, and Alessandro Pitruzzella. 2021. "The Challenging Riddle about the Janus-Type Role of Hsp60 and Related Extracellular Vesicles and miRNAs in Carcinogenesis and the Promises of Its Solution" Applied Sciences 11, no. 3: 1175. https://doi.org/10.3390/app11031175

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