Next Article in Journal
Selective Degradation Permits a Feedback Loop Controlling Annexin A6 and Cholesterol Levels in Endolysosomes of NPC1 Mutant Cells
Previous Article in Journal
Managing the Oocyte Meiotic Arrest—Lessons from Frogs and Jellyfish
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Functional Roles of Matrix Metalloproteinases and Their Inhibitors in Melanoma

1
Department of Biomedical and Biotechnological Sciences, University of Catania, 95123 Catania, Italy
2
Research Center for Prevention, Diagnosis and Treatment of Cancer, University of Catania, 95123 Catania, Italy
3
Epidemiology Unit, IRCCS Istituto Nazionale Tumori “Fondazione G. Pascale”, 80131 Naples, Italy
*
Authors to whom correspondence should be addressed.
These authors are equally contributed to this work.
Cells 2020, 9(5), 1151; https://doi.org/10.3390/cells9051151
Submission received: 23 March 2020 / Revised: 1 May 2020 / Accepted: 6 May 2020 / Published: 7 May 2020
(This article belongs to the Section Intracellular and Plasma Membranes)

Abstract

:
The extracellular matrix (ECM) plays an important role in the regulation of the tissue microenvironment and in the maintenance of cellular homeostasis. Several proteins with a proteolytic activity toward several ECM components are involved in the regulation and remodeling of the ECM. Among these, Matrix Metalloproteinases (MMPs) are a class of peptidase able to remodel the ECM by favoring the tumor invasive processes. Of these peptidases, MMP-9 is the most involved in the development of cancer, including that of melanoma. Dysregulations of the MAPKs and PI3K/Akt signaling pathways can lead to an aberrant overexpression of MMP-9. Even ncRNAs are implicated in the aberrant production of MMP-9 protein, as well as other proteins responsible for the activation or inhibition of MMP-9, such as Osteopontin and Tissue Inhibitors of Metalloproteinases. Currently, there are different therapeutic approaches for melanoma, including targeted therapies and immunotherapies. However, no biomarkers are available for the prediction of the therapeutic response. In this context, several studies have tried to understand the diagnostic, prognostic and therapeutic potential of MMP-9 in melanoma patients by performing clinical trials with synthetic MMPs inhibitors. Therefore, MMP-9 may be considered a promising molecule for the management of melanoma patients due to its role as a biomarker and therapeutic target.

1. Introduction

The extracellular matrix (ECM) plays several roles in the regulation of cellular homeostasis and in the regulation of cell-cell interactions. ECM alteration leads to morphological and structural changes which are responsible for several pathological conditions [1]. Dysregulations in ECM structure and components are key events in tumor growth and cell migration favoring several processes, such as vascularization and loss of cell adhesion [2]. Indeed, the interaction between cells and ECM is mediated by several receptors and secretory molecules, capable not only of regulating connection and cell migration, but also in modulating the expression of the genes involved in cell growth and differentiation [3]. In this context, several studies have demonstrated that cell-ECM interactions regulate the molecular processes, which underlie cell differentiation [3], cellular homeostasis [4], wound healing [5], and cancer invasion [6,7].
ECM remodeling is finely regulated by several proteins involved in the cleavage and hydrolysis of the ECM component, favoring the pathological and physiological modification of ECM that have impacts on many cellular processes, as previously described [1,8].
Matrix Metalloproteinase (MMP) family is the protein family mainly associated with the degradation of the ECM. In humans, this family is composed of 24 proteins encoded by different genes and named MMP 1–28, although some MMPs are not human MMPs (MMP-4, MMP-5 and MMP-6) or are aliases of other MMPs (MMP-22). MMP family has numerous roles in physiology and fetal development, but it has been demonstrated that some of these proteinases, when dysregulated, can be precursor of pathological conditions [9].
MMP activities are strictly regulated by other proteins that ensure steady-state conditions between the degradative and reconstructive processes of the ECM [10]. In particular, MMPs are regulated by a class of endogenous inhibitors, named tissue inhibitors of metalloproteinase (TIMP), that are involved in both MMPs activation and inactivation [11]. The different structures characterizing the various MMPs allow the accomplishment of multiple functions influencing several processes like cell behavior, apoptosis and cell proliferation. Oncological studies have shown that MMPs play several functions favoring tumor progression through the degradation of surrounding tissues, the modulation of growth factors and membrane receptors, as well as inflammatory proteins, adhesion molecules and chemo-attractive proteins [12,13,14].
MMPs carry out an important role also in the modification of skin ECM. Indeed, MMPs are involved in skin matrix remodeling through the degradation and reconstruction of matrix components. Furthermore, several studies have demonstrated that MMPs play fundamental roles in melanoma, where tumor cells and tumor microenvironment alterations were associated with MMPs and TIMPs de-regulation [15].
Melanoma represents the most lethal tumor among all skin cancers because of its aggressiveness and its high metastatic potential. It represents about 3–5% of all skin cancers and it is often diagnosed in patients aged between 20–35 years or in patients with more than 65 years as a consequence of several risk factors, including chronic sunlight exposure, a clear phototype and specific somatic mutations [16,17,18]. Despite the evolution of cancer pharmacological treatments [19], several epidemiological studies have demonstrated that melanoma incidence is about 5%, while the mortality rate among all skin cancers is about 75%. In this regard, the last updated USA cancer statistics data shows that 20% of the population develops this tumor during their life [20,21].
Among the recognized risk factors for melanoma there are genetic susceptibility and gene polymorphisms [22]. Indeed, about 10–15% of melanomas are diagnosed in patients with a family history of melanoma. Another significant fraction of melanoma is associated with two different familial syndromes called familial atypical multiple mole melanoma syndrome (FAMMM syndrome) and the melanoma-astrocytoma syndrome (MAS) [23]. In addition, some germline mutations have been associated with a higher risk of melanoma. In particular, mutation occurring in the cyclin-dependent kinase inhibitor 2A (CDKN2A) and cyclin-dependent kinase 4 (CDK4) genes are the most frequent genetic abnormalities associated with melanoma risk [22,24].
As mentioned, besides these hereditary and familial risk factors, solar ultraviolet radiation (UV-A and UV-B rays) represents the most important trigger of melanocyte neoplastic transformation due to the induction of several genetic mutations [25,26,27].
According to Catalogue Of Somatic Mutations In Cancer (COSMIC), among these mutations the most frequently observed in melanoma are mutations affecting the BRAF (44%), NRAS (17%), TERT (25%), CDKN2A (18%), KIT (8%), GRIN2A (20%), TP53 (15%), PTPRT (20%), LRP1B (26%), NF1 (17%), and PTEN (10%) genes [28]. In particular, about 50% of melanoma is characterized by the presence of BRAF activating mutations. In 90% of cases, this mutation is represented by the BRAFV600E mutation where the valine 600 is substituted with the glutamic acid (p.V600E) that leads to the over-expression and hyperactivation of BRAF [29].
All these mutations are involved in the alteration of key molecular and signaling transduction pathways responsible for the abnormal proliferation of cancer cells and loss of apoptosis [30,31]. In addition, these somatic mutations, especially that of BRAF, are involved in the alteration of the tumor microenvironment through the regulation of the Mitogen-Activated Protein Kinase (MAPK) pathway. In particular, the hyperactivation of MAPKs pathways is associated with the over-expression of the transcription factor ERK that in turn leads to the over-expression of several genes involved in tumor development, including MMPs, TGFβ, and Osteopontin (OPN), whose role in melanoma is fundamental for the tumor invasion and metastatic processes [32,33,34].
Starting from these preliminary data on the physiological and pathological role of ECM and MMPs, and taking into account the epidemiological, molecular and mutational data of melanoma, this article aims to review the current knowledge on the involvement of MMPs. It will particularly focus on Matrix Metalloproteinase 9 (MMP-9), in the degradation of ECM and the consequent progression of melanoma, as well as the potential therapeutic implication of both endogenous and exogenous MMP inhibitors for the design of new therapeutic protocols for melanoma patients.

2. Matrix Metalloproteinases

The MMPs are a subgroup of metalloproteinases (including disintegrin, ADAM metalloproteinases and other proteolytic enzymes) containing a catalytic zinc ion site and an evolutionarily conserved methionine residue [35].
At present, 24 different MMPs were identified in humans and 23 in mice. All the MMPs are grouped in seven major categories according to their substrates or proteolytic functions: Collagenases, gelatinases, stromelysins, matrilysins, metalloelastase, membrane-type MMPs and other MMPs [36,37,38] (Table 1).

2.1. Collagenases

The group of MMPs with a collagenase activity contains three different MMPs: MMP-1, MMP-8, and MMP-13. MMP-1 is a collagenase found in the interstitial space able to cleave types I, II, and III collagens. Several studies showed that MMP-1 protein is over-expressed in several tumors and leads consequently to a more aggressive phenotype [39]. Although MMPs are generally considered as pro-oncogenic proteins, some MMPs, such as MMP-8, may have anti-tumor properties. Indeed, in different experimental models the neutrophil collagenase, MMP-8 showed anti-tumoral properties suppressing cell proliferation and the formation of metastases [12]. As described for MMP-1, also MMP-13 protein expression increased during the invasive processes of some tumors, including melanoma [41].

2.2. Gelatinases

This group contains MMP-2 and MMP-9, called also gelatinases A and B, respectively, with a gelatinolytic activity. In the following paragraphs, the function and regulation of MMP-9 will be widely described. Regarding MMP-2, it was demonstrated that MMP-2 protein is expressed in the stroma and cell cytoplasm. Furthermore, MMP-2 acts in the ECM by degrading the type I collagen, type IV collagen and other components, such as fibronectin and gelatin [42].

2.3. Stromelysins and Matrilysins

The group of stromelysin and matrilysin contains respectively MMP-3, MMP-10 and MMP-11 (Stromelysin) and MMP-7 and MMP-26 (Matrilysin). All these MMPs are involved in the degradation of different components of the ECM. In particular, stromelysins plays a major role in the degradation of proteoglycans, gelatin, and other constituents of the extracellular matrix [38], while the MMP-7 and MMP-26 matrilysins are associated with the cell membrane by binding to cholesterol sulphate and degrade substrates like fibronectin, laminin, collagen type IV, and gelatin. These MMPs are expressed in primary and metastatic melanoma [46].

2.4. Metalloelastase MMPs

The group of metalloelastase is only represented by MMP-12. This matrix metalloproteinase is able to hydrolyze soluble and insoluble elastin. It also degrades type IV collagen and fibronectin. It is involved in tissue remodeling after injuries [48]

2.5. Membrane-Type MMPs

Membrane-type matrix metalloproteinases (MT-MMPs) represents a subgroup of the MMP family composed of six MT-MMPs in humans [50,51]. There are two main groups of MT-MMPs: type I transmembrane-type (MT1-, MT2-, MT3- and MT5-MMPs) and glycosylphosphatidylinositol (GPI)-anchored type (MT4- and MT6-MMPs) [58]. All these MMPs are linked to the cell membrane, regulating key processes in the cell-cell and cell-ECM interactions [58]. The most investigated MT-MMPs is MT1-MMP (MMP-14). MMP-14 was the first discovered membrane-type MMP. This MMP is also involved in the invasive phenotype of some tumors, including melanoma [58], by interacting with MMP-2 in the stroma and promoting cancer progression [49]. In vivo, the cleaved form of laminin 332 was found in tumors and in tissues undergoing remodeling but not in quiescent tissues [38]. Cleavage fragments of the two chain of laminin 332 generated by MMP14 were detected in high invasive melanoma cells and played a crucial role in cell adhesion, migration and vasculogenic mimicry (VM) [59]. It was demonstrated also that BRAF and NRAS mutations positively regulate MMP-14 gene expression enhancing tumor growth and melanoma invasiveness in vivo [60].

2.6. Other MMPs

Other MMPs, like MMP-19, MMP-21, MMP-23A and B, etc., are not clustered in a specific group, therefore are named “other MMPs”. Of these, the most studied are the MMP-19 and MMP-21 whose expression in melanoma cell lines was correlated, respectively, to a higher invasive power (MMP-19) and to the malignant transformation of melanocytes (MMP-21), suggesting a possible use of these MMPs as predictive biomarkers of cancer progression [52,54].
As shown, several studies have described the involvement of almost all of these MMPs in the development of cutaneous melanoma. It is clear that the strong involvement of MMPs in melanoma development and other skin cancers depends mainly on the specific features of melanoma that is a tumor characterized by a high invasive power toward the surrounding tissues and high rates of metastases and recurrence [61].

3. MMP-9 Functions and Regulation in Melanoma

MMP-9 is a protease involved in extracellular matrix degradation. In humans, the MMP-9 gene is mapped in the genomic region 20q13.12 and codifies for a protein of 707 aa (92 kDa) secreted in the extracellular space as inactive pro-enzyme named pro-MMP-9. The pro-MMP-9 is inactive because of 80 aa residues at the N-terminal site where a cysteine switch motif coordinates the zinc ion forming the catalytic domain of the protein, thus maintaining it inactive [43,62].
In the extracellular space other proteinases, like MMP-3 or MMP-2, cleave the inactive form of pro-MMP-9 in the active form of 84 kDa [63,64] (Figure 1).
MMP-9 is expressed in neutrophils, macrophages, and fibroblasts. Proangiogenic factors, including the fibroblast growth factor (FGF) and vascular endothelial growth factor (VEGF), are activated when MMP-9 degrades the ECM [65].
It was demonstrated that melanoma MMP-9 and MMP-2 play a fundamental role in the degradation of the ECM, thus, favoring melanoma spreading towards the surrounding tissues until the formation of distant metastases [16,34,66]. Several studies have demonstrated that the up-regulation of MMP-9 is strictly influenced by several genetic alterations or modifications of the tumor microenvironments [65,67,68]. In particular, it was proved that the high levels of MMP-9 in melanoma patients might be due to the dysregulation of the TGFβ pathways where the alterations in the levels of NF-κB is able to induce the overexpression of MMP-9 via OPN activation [34]. Moreover, it has also been demonstrated that epigenetic modifications may lead to the up-regulation of MMP-9 in melanoma and other cancer types, as explained in the following paragraphs.
It was demonstrated that many molecular pathways, including Ras-Raf-MEK-ERK (MAPKs) and PI3K/PTEN/AKT/mTOR (PI3K/AKT), are associated with the regulation of MMP-9. Dysregulations in these pathways can lead to an aberrant over-expression of several inflammatory proteins and consequently to the increased expression of several ECM proteases, like MMP-9 [33].
The MAPKs pathway has a relevant role in the physiology of cells, initiating the phosphorylation cascade of different kinases proteins, yielding to cell growth and differentiation [69]. Activating mutations within this pathway can lead to the loss of negative feedback control, and in turn, to the over-activation of the pathway itself leading to uncontrolled cell proliferation [70]. For instance, BRAF mutations induce the constitutive activation of MAPKs pathway able to induce the hyperphosphorylation of the transcription factor ERK that in turn induce the up-regulation of genes involved in survival and proliferation of cancer cells [71]. Among these genes, there is also MMP-9 whose degradative action towards ECM was already widely discussed [72,73].
Other mechanisms of MMP-9 over-expression in melanoma are mediated by neural crest associated genes, i.e., FOXD1, via the RAC1B pathway. Wu et al. (2018) have demonstrated that the up-regulation of FOXD1 leads to the over-expression of MMP-9 mRNA and protein levels. In particular, the authors observed a decrease of MMP-9 expression by using a siRNA against FOXD1. In addition, the loss/down-regulation of RAC1B decreases the spreading of tumor cells and delays the Epithelial–Mesenchymal Transition, underlying the importance of the FOXD1-RAC1B axis in the regulation of MMP-9-mediated melanoma cells invasion [74].
Other studies have demonstrated that MMP-9 not only mediates the degradation of the ECM, but is also involved in neo-angiogenesis, cell migration and formation of metastases.
In this context, the most aggressive forms of melanoma are able to undergo VM, in which tumor cells behave like an endothelial cell forming microvascular channels, small interconnection in the ECM connecting different tissues, due to the action of proteinases and angiogenic factors [75]. Among these proteinases, MMP-2/9 showed both, in vitro and in vivo their implication in VM. In particular, the complex Rictor-mTORC2 is able to phosphorylate AKT, leading to the overexpression of MMP-2/9 and the formation of microvascular channels [76,77].
Finally, other regulatory mechanisms of MMP-9 expression are mediated by MMP-9 allelic variation (rs3918251GG (A>G) and by the over-expression of SOX4 responsible for the concomitant up-regulation of MMP-2/9 and NF-κB/p65 and, consequently, for a more aggressive tumor phenotype [78,79].

4. Epigenetic Modulation of MMP-9 Expression

Besides the well-recognized genetic alterations associated with the up-regulation of MMP-9, recent evidence suggests that epigenetic events may alter the expression levels of this gene.
The development of new high-throughput technologies in the field of molecular biology allowed the collection of a huge amount of bioinformatics data regarding the expression profiles of both mRNA and non-coding RNA (ncRNA) [80,81,82].
Through computational approaches, different studies have identified specific ncRNA associated with the development of cancer or with gene and protein alterations [66,83,84,85,86,87,88,89]. Specifically, through the analysis of bioinformatics data, it was possible to identify specific microRNAs (miRNAs) associated with melanoma or miRNAs able to selectively target and modulate MMP-9 [66,84]. Some studies have tried to determine the role of ncRNAs in the regulation of MMP-9 (Figure 2).
Zhou et al. (2018) have shown high levels of miR-155-5p in exosomes derived from a co-culture of melanoma cells and fibroblasts. The authors have demonstrated also that miR-155-5p is able to inhibit SOCS1 in fibroblast and in turn enhancing the expression levels of p-STAT3 responsible for VEGFa and MMP-9 transcriptional activation. The epigenetic alterations of both VEGFa and MMP-9 mediated by hsa-miR-155-5p were finally responsible for the phenotypical switch of fibroblasts towards a cancer-associated fibroblast phenotype known to be associated with angiogenesis [90].
In a study performed by Wang et al. (2018), two ncRNAs were computationally predicted to bind MMP-9 and MMP-2. These ncRNAs are the FOXC promoter upstream transcript (FOXCUT) and the miRNA hsa-miR-296-3p. In particular, the authors have shown in C918 cells, a model for choroidal malignant melanoma (CMM, an interocular type of melanoma), that both MMP-2/9 transcripts and proteins were downregulated when FOXCUT and/or miR-296 were transfected in C918 cells [91].

5. Approved Therapeutic Approaches for Melanoma

The therapeutic approaches for the treatment of melanoma vary significantly depending on the pathological features of tumor. Most of the dysplastic nevi and non-invasive superficial melanoma are treated by surgical resection with wide skin margins free from disease. This type of surgical resection is conclusive in most cases. However, in cases of advanced-stage melanoma, surgical resection is often not applicable or must be accompanied by drug therapy [92].
The therapeutic efficacy of pharmacological anti-cancer treatment in melanoma depends on different factors, including the genetic landscape of cancer cells, tumor microenvironment alterations and epigenetic alterations. At present, the best therapeutic approach for melanoma patients consists in a combination of drugs that inhibits the kinase activity of BRAF (BRAFi) and MEK (MEKi), downstream effectors of the MAPK pathway; in addition, new treatments with immune checkpoint inhibitors are now being used for advanced or metastatic melanomas [19,93].
On the basis of the latest evidence, the combination of BRAFi and MEKi results in a better therapeutic response of melanoma patients positive to BRAF mutations [94]. Currently, the combination therapies available for melanoma are three: Vemurafenib and Cobimetinib, Dabrafenib and Trametinib, Encorafenib and Binimetinib. In particular, Dabrafenib is a RAF inhibitor used for patients with advanced or metastatic melanoma positive to the BRAFV600E mutation [95]. On the other hand, Trametinib is a MEK-1/2 inhibitor and it is also used for the treatment of advanced or metastatic melanoma with BRAFV600E mutation [96].
The identification of drugs capable of targeting key mutation involved in melanoma raised the attention on the identification of potential biomarkers of therapeutic response. In this context, several studies have proposed MMP-9 as a candidate marker used for the assessment of Dabrafenib therapeutic response. In particular, Salemi R et al. (2018) have analyzed the presence of circulating-free (cfDNA) BRAFV600E mutation in liquid biopsy samples obtained from patients treated with BRAFi and/or MEKi by using the droplet digital PCR system. The authors showed that the baseline presence of cfDNA BRAFV600E mutation was associated with a worse prognosis. Moreover, the authors have demonstrated that during the therapy there was a positive correlation between the number of cfDNA of BRAFV600E mutation and MMP-9 protein levels, thus proposing MMP-9 as a new biomarker of response or resistance to the treatment [32]. In this context, recent studies highlight how the analysis of circulating tumor DNA and liquid biopsy samples coupled with the use of high-sensitive genetic and proteomics techniques may represent innovative strategies to identify molecular biomarkers, including MMPs and related proteins, in order to early detect pre-cancerous lesion, tumor or tumor relapse [97,98].
As anticipated, in the last years, new drugs have been developed for the management of melanoma patients. These monoclonal antibodies are directed towards the PD-1, a surface protein expressed on cancer cells capable to inhibit the T-lymphocyte activation. These new FDA and EMA approved monoclonal antibodies are Pembrolizumab and Nivolumab [99,100].
Finally, in the past years, some studies have provided evidence that new combination therapies, with both selective inhibitors and immune checkpoint inhibitors, have a higher therapeutic efficacy compared to monotherapy [101]. Furthermore, several studies have proved that the concomitant administration of anticancer treatment and supportive care regimens with probiotics and antioxidant compounds is associated with a higher response rate compared to the treatments with only anticancer agents [102,103].
Overall, despite the different drugs available for the treatment of melanoma, a significant fraction of patients is refractory to therapies. Therefore, it is necessary to identify new effective treatments for the management of melanoma patients. In this context, the development of new selective inhibitors for MMPs could represent a promising strategy to improve the efficacy of current treatments and inhibits tumor aggressiveness.

6. Clinical Application of MMPs Inhibitors

Different molecules are able to interact with MMPs modulating their functions; these are called MMPs inhibitors (MMPi), which can be divided into synthetic inhibitors and endogenous inhibitors [104]. There are several inhibitory mechanisms of MMPs activity, but the most common involves the binding of the molecule with the zinc atom of the catalytic domain of the protein [105]. The compounds and functional groups used to inhibit the MMPs activity are summarized in Table 2. The first synthetic inhibitor was developed with a structure similar to that of collagen where the hydroxylated domain was recognized by the zinc atom of the catalytic site of MMPs, thus, acting as a competitive inhibitor of collagen [106,107]. Pharmaceutical companies began developing MMPi drugs since the 1990 for the treatment of pathologies, in which the proteinases are dysregulated, like in cancers. The first were Marimastat (BB-2516) and Cipemastat (Ro 32-3555), both containing the hydroxamate group [108].
Chirivi et al. (1994) have studied the efficacy of Batimastat, a first-generation MMPi, in C57BL/6N mice injected with B16-BL6 melanoma cells to evaluate the inhibition of tumor growth. The results showed a reduction of lung metastasis and solid tumor dimensions when the drug was administrated after the inoculation of cells [109]. Another in vivo study on a melanoma animal model was performed to assess the efficacy of Batimastat when it is administrated in combination with IL-12; also in this case, the results showed strong antitumoral and antiangiogenetic effects [110]. In a study conducted by Wylie et al. (1999), Batimastat induced anti-angiogenic effects after inoculation of B16F1 melanoma cells in mice, thus reducing sensibly the size of liver metastases without interfering with the extravasation of circulating melanoma cells [111].
Another MMPi with promising in vivo results was MMI270. This is a synthetic hydroxamic acid derivate, capable of inhibiting the degradative action of MMPs at nanomolar concentration. It did not show anti-proliferative activity in the primary tumor, but it was effective in reducing the angiogenic activity. Kasaoka et al. (2002) have studied how the administration of MMI270 can affect the extravasation of melanoma cells injected i.v. in a murine model. They have hypothesized the lymphatic system as the main route of metastasis formation, proposing the possibility of the administrating MMI270 after lymphadenectomy to reduce the metastatic niches number [112].
Conway et al., in 1996, have conducted a study on four synthetic MMPi in mice injected s.c. with melanoma cells. The drugs were pumped continuously through a s.c. implanted micropump to reach high plasmatic concentration. Despite the high concentration of drugs, the results demonstrated no significant difference in metastasis formation, thus excluding these four molecules from being used in human clinical trials [113].
Another study on B16-BL2 melanoma cells injected intramuscularly in C57B2/7 mice using a Recombinant Human Tissue Inhibitors of Metalloproteinases-2 (r-hTIMP-2) shown a significant decrease in metastasis formation, especially when the drug was administered right after the tumor injection [114].
MMI-166, a third-generation inhibitor of MMP-2 and MMP-9, has shown to be significantly effective in reducing tumor cell proliferation and the formation of metastasis. In 2002, Hojo and colleagues have studied this drug injecting B16-BL6 melanoma cells in mice. Their results have shown that MMI-116 is able to reduce tumor proliferation, and the combination with paclitaxel or carboplatin yield to an increase of antitumoral activity of the drug [115].
The promising results that came from preclinical studies using MMPis led to the development of clinical studies on humans. Batimastat (B394) had no success in human, because of its low bioavailability. Indeed, the results obtained from phase I clinical studies have revealed that prolonged treatment with MPIs can lead to toxic effects such as muscular-skeletal pain and inflammation [116].
Consequently, Agouron and Bayer have produced two new drugs Prinomastat (AG 3340) [117] and Tanomastat [118], a thiol-based inhibitor (BAY 12-9566), with the aim of minimizing, or completely eliminating, the adverse effects of the previous generation. The efficacy of these molecules is particularly direct toward MMP-9 and MMP-2. Unfortunately, Prinomastat (AG 3340) showed no efficacy during the administration of phase III trial on patients with Non-Small Cell Lung Cancer (NSCLC) and prostate cancer, and this caused the anticipation in the shutting down of the study [119]. For this reason, there are no data about progression-free survival (PFS) or overall survival (OS). Anyway, a good portion of patients was affected by muscular-skeletal toxicity. Moreover, the patients with NSCLC, who received chemotherapy in association with the MPI, had a greater chance of developing venous thromboembolism [120].
For Tanomastat, two clinical trials were done, without showing any reduction of tumor growth or reduction of PFS and OS. The first trial, a multicentric randomized double-blind compared to placebo study, was performed on advanced epithelial ovarian cancer [121]. The second trial was performed on pancreatic cancer patients to assess the efficacy of Tanomastat compared to that of gemcitabine. In this second study, the effect of gemcitabine was higher compared to those observed for Tanomastat [122].
Other clinical studies were performed to test the efficacy of Marimastat compared to gemcitabine. In particular, the trials performed in more than 400 patients with non-resectable pancreatic cancer showed that Marimastat did not show a higher therapeutic effect compared to Gemcitabine [123]. In another trial, Marimastat was compared to a placebo in 369 patients with unresectable gastric cancer, in which a moderate increase of the progression-free survival, but not in the overall survival, was observed [123].
Another study performed by administering BMS-275291 has shown exciting preliminary results during the first step of a clinical trial on NSCLS patients where lower dose-dependent toxic effects were observed. However, at the end of the trial it has been demonstrated that the concomitant administration of BMS-275291, Paclitaxel and Carboplatin did not confer any advantage for the improvement of OS, PFS and relative risk. Moreover, toxicity was too high when compared to the control group, showing a higher incidence of hypersensitivity reaction, febrile neutropenia and skin rash. Despite the promising preclinical results, the study was stopped earlier together with other two trials: One on patients with breast cancer, and another on prostate cancer [124].
Table 2. Structure of synthetic MMPi compounds.
Table 2. Structure of synthetic MMPi compounds.
CompoundFunctional GroupActivitySource
Marimastathydroxamate group (-CONHOH)Binds to zinc domain of several MMPs[108]
Prinomastathydroxamate group (-CONHOH) and Aryl backboneSelectively inhibits MMP-2, 9, 13, and 14[117]
RebimastatThiolSelectively inhibits MMP-1, 2, 8, 9, and 14[118]
TanomastatThiolBinds to zinc and selectively inhibits of MMP-2, 3, and 9[118]
Ro 28-2653PyrimidineSelectively inhibitsMMP-2, MMP-9 and membrane type 1-MMP[125]
3-hydroxypyran-4-one (868368-30-3)Hydroxypyrone-based and Aryl backboneInhibits several MMPs[126]
No. 582311-81-7Phosphorus-based inhibitors and/or carbamoyl phosphonate zinc binding groupsSelectively inhibits MMP-1, 2, 3 8, 9, 14 and 13[127,128,129]
DoxycyclineTetracycline-basedSelectively inhibits MMP-2 and 9[104,130]

7. Endogenous MMP Inhibitors

In addition to synthetic inhibitors, there is a class of endogenous inhibitors of MMPs of which the Tissue Inhibitors of Metalloproteinases (TIMPs) and α2-macroglobulin are the most represented [131]. α2-macroglobulin is a serum protein [132] capable of binding MMPs forming an inactive complex [133]. In a pioneering study, Kancha and co-workers have analyzed the protein levels of the α2-macroglobulin receptor (LRP/α2-MR). Its levels were found less expressed in the invasive sub-clones compared to the non-invasive ones derived from PC-3 and DU-145 human prostatic cells and the A2058 melanoma cell line supporting the hypothesis that the down-regulation of LRP/α2-MR complex can increase tumor cell invasiveness [134].
The TIMP inhibitors are grouped in a family composed of four different TIMPs, which are able to block the proteolytic action of the MMPs [135,136]. These inhibitors are capable of inhibiting MMPs binding their catalytic sites [131]. The TIMP molecules have 12 cysteine residues that form six loops through disulfide bonds, a conformation necessary to carry out the inhibitory activity towards MMPs. Indeed, the N-terminal site of TIMPs is able to bind the majority of MMPs, while the C-terminal site of TIMP-1 and TIMP-2 bind respectively to the hemopexin domain of pro-MMP-2 and pro-MMP-9 [137].
Normally, TIMPs can interact with different MMPs, however, MMP-1 is not targetable by TIMP-1 [133]. Nonetheless, inhibitory preferences exist; an example is represented by TIMP-1 that favors the inhibition of MMP-9 [138]. Besides its inhibitory action, TIMP-1 has also shown to have the ability to promote cell proliferation and the inhibition of apoptosis in a large number of cell types [139]. Normally, TIMPs levels are in excess compared to MMPs levels in extracellular fluids and tissues, thus limiting their activity to focal peri-cellular sites. As previously mentioned, TIMPs are known for their inhibitory activity towards MMPs, nonetheless, it has been shown that the activation of MMP-2 is associated with a low concentration of TIMP-2 coupled with MMP-14 in the cell membrane [12].
Changes in TIMPs levels are supposed to be involved in pathological conditions associated with unbalanced MMPs activities. [138]. For example, studies have suggested that an imbalance in the MMP-9/TIMP-1 ratio could be involved in the pathogenesis of COPD (chronic obstructive pulmonary disease) [140], but conflicting results were obtained from other studies [141,142]. Some evidence also suggests the role of TIMPs in cancer. In particular, alterations in TIMP-1 and MMP-9 levels have been observed in subjects with lung cancer compared to controls, associated with a poor prognosis, indicating a possible role of TIMP-1 and MMP-9 as a prognostic marker [143]. Other alterations in MMP-9 and TIMP-1 levels have been observed in subjects with breast cancer compared to controls. These dysregulations were associated with lymph node metastases and lower overall survival rates, suggesting that the levels of MMP-9 and TIMP-1 could be further evaluated to predict prognosis and progression of breast cancer [144].
TIMPs could be used as therapeutic indicators even in cardiovascular diseases, for instance for the treatment of atherosclerosis and aneurysm formation where elevated levels of MMP-9 were found [145]. As previously mentioned, each TIMP is able to inhibit several MMPs; therefore, it would be appropriate to develop engineered TIMPs able to interact with specific MMPs that play a role in that pathological process [133].
Hofmann and colleagues (2000) have made a review showing why the balance in the concentrations of MMPs and TIMPs is important; their dysregulation can determine the progression of melanoma. In particular, it has been demonstrated that the overexpression of TIMPs using recombinant TIMPs is able to reduce the invasiveness of melanoma cells [146].
Other studies have shown the importance of TIMP-1 in the control of melanoma progression. In particular, Khokha has showed that the B16-F10 cells previously transfected with the over-expressing construct for TIMP-1 and subsequently injected in mice, shown inhibition of the metastatic potential and growth compared to the non-transfected counterpart [147].

8. MMP-9 and Osteopontin

Osteopontin, indicated also as OPN or SPP1, is a sialylated protein of 300 aa belonging to the family of SIBLING (Small Integrin-Binding Ligand, N-linked Glycoprotein) proteins. It is codified by the SPP1 gene (Secreted Phosphoprotein 1) located on the long arm of chromosome 4 region 22 (4q22.1) [148].
OPN was first isolated from bone tissue; then other studies demonstrated the presence of this protein in several tissues, e.g., epithelia, kidney, internal mucosae, carotid tissue [149,150,151].
OPN is involved in several functions, however, its major role is in bone remodeling. In particular, OPN is very important for bone replacement and resorption [152], because it is involved in the maturation of osteoclast and in the recruitment of osteoblast [153].
In addition, OPN is involved in immune processes, acting as an immunomodulator in several ways. It has been proved its involvement in chemotaxis, encouraging the migration of neutrophils in vitro and recruiting them into alcoholic liver disease [154,155]. Other functions, such as cytokine production, cell activation, and cell survival are mediated by OPN [156].
OPN is further involved in the block of apoptotic processes. Several harmful stimuli can induce apoptosis in both epithelial (fibroblasts and endothelial cells) and immune cells (macrophages and T-cells), but OPN is able to block these processes [157,158]. Several studies have been carried out to support the hypothesis that OPN improves the T cells survival [159]. These anti-apoptotic properties can lead to the survival of tumor cells [160]. This notion is supported by the fact that OPN is over-expressed in numerous tumors [161].
With respect to the role of OPN in cancer, several studies have shown its ability to stimulate neo-angiogenesis [162] and to promote the development of metastases [163]. Its expression in tumors is a negative prognostic factor and it is generally associated with an increase in tumor proliferation and invasiveness supported also by the OPN-induced MMP-9 over-expression. Studies have shown that OPN and MMP-9 interact indirectly through the MAPK and the IKK/IkB/NF-κB pathways [34,164,165]. In particular, the link between OPN and αvβ3 integrin promotes the phosphorylated active form of NIK that in turn binds IKKα/β that activates NF-κB through the degradation of IκBα. NIK, when is coupled with OPN, induces also the activation of the MAPKs pathway responsible for the further activation of NF-κB and its accessory proteins p65 and p50. Consequently, OPN, stimulating the protein complex formed by NIK and IKK/MAPK, is able to up-regulate MMP-9 transcription leading to the production of Pro-MMP-9. Finally, activated NF-κB stimulates also the transcription of uPA that in turn binds Pro-MMP-9 thus producing MMP-9 active form (Figure 3) [166].
The activation of MMP-9 mediated by OPN induces ECM degradation, cell invasion, tumor growth and metastasis. For example, the up-regulation of OPN, induced by (HIF-1)α and the activation of MMP-9, promotes gastric tumor metastasis. Finally, OPN is able to induce the activation of MAP3K1 signaling pathways regulated by MMP-9 resulting in the proliferation of melanoma cells and pulmonary metastases [167].
Youwen Zhou and co-workers have identified higher levels of OPN in melanoma cells than in normal nevi. To highlight the stimulatory effect of OPN on tumor progression, the KZ-28 melanoma cells were transfected with various concentrations of synthetic OPN small interference RNA (siRNA). Subsequently, cell viability assays have shown a decrease in the number of cells in a dose-dependent manner according to the increase of siRNA concentration. This is consistent with the stimulatory function of OPN on tumor cells [168].
On these bases, OPN and MMP-9 could be considered biomarkers of tumor progression [34].

9. Conclusions

The role of the MMP family and OPN have been widely described in order to understand their involvement in tumorigenesis, but the actual use of these proteins as biomarkers or therapeutic targets is still far away from happening. The paragraphs discussed above indicate that there is the need for discovering non-invasive biomarkers to assess the progression of melanoma both, during and after the therapy with small molecules, like Dabrafenib or other MAPK inhibitors. It has been also described in the use of MMPi drugs, with exciting preliminary results in vitro and in animal models, however, the expectation has declined rapidly because of the severe side effects observed in many clinical trials. Generally, the incongruences between the preclinical models and the clinical studies are probably due to the fact that in the preclinical models the drugs are administrated during the early phases of the disease when the conditions for severe and rapid worsening of tumor do not yet exist; while in humans, studies have been performed on patients with advanced tumors. So, it can be hypothesized that inhibition of MMPs could be effective in limiting tumor progression during its initial phase, keeping in mind that the efficacy of the therapy with MMPi drastically decreases with the progression of the disease.
On the other hand, the use of TIMPs, the endogenous inhibitors of several MMPs, has been suggested. An engineered TIMPs could be able to interact with a specific MMPs, and this is the reason why it has been hypothesized the use of these molecules to stop the tumor growth and migration.
Finally, the use of OPN and MMPs as biomarkers of progression in melanoma seems promising. Indeed, our recent evidence [32,33,34] showed that the OPN and MMP-9 are strongly involved in several neoplastic processes in melanoma, including ECM degradation, invasion of the surrounding tissues, metastasis formation and loss of apoptosis. Therefore, the evaluation of MMPs circulating levels, and in particular the use of OPN and MMP-9 as circulating biomarkers, may improve the current diagnostic strategies for melanoma and predict the aggressiveness of tumors, in order to personalize the therapeutic approach for the patients.
In this context, ongoing studies are attempting to screen melanoma patients for these biomarkers and further studies are needed to evaluate the predictive values of MMPs, TIMPs, OPN.

Author Contributions

Conceptualization, S.N., C.S., G.G. and L.F.; investigation, S.N., C.S., G.G., V.D.B., M.S.B.; writing—original draft preparation S.N., C.S., G.G., V.D.B., M.S.B.; writing—review and editing, M.L., and L.F.; visualization, M.L.; supervision, S.C., and M.L.; project administration, M.L. and L.F. All authors have read and agree to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The authors would like to thank the “Lega Italiana per la Lotta contro i Tumouri (LILT)”, section of Catania and the Research Center for Prevention, Diagnosis and Treatment of Cancer, University of Catania, Catania.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bonnans, C.; Chou, J.; Werb, Z. Remodelling the extracellular matrix in development and disease. Nat. Rev. Mol. Cell Biol. 2014, 15, 786–801. [Google Scholar] [CrossRef]
  2. Walker, C.; Mojares, E.; Del Río Hernández, A. Role of extracellular matrix in development and cancer progression. Int. J. Mol. Sci. 2018, 19, 3028. [Google Scholar] [CrossRef] [Green Version]
  3. Poltavets, V.; Kochetkova, M.; Pitson, S.M.; Samuel, M.S. The role of the extracellular matrix and its molecular and cellular regulators in cancer cell plasticity. Front. Oncol. 2018, 8, 431. [Google Scholar] [CrossRef] [Green Version]
  4. Watt, F.M.; Fujiwara, H. Cell-extracellular matrix interactions in normal and diseased skin. Cold Spring Harb. Perspect. Biol. 2011, 3, a005124. [Google Scholar] [CrossRef] [Green Version]
  5. Xue, M.; Jackson, C.J. Extracellular matrix reorganization during wound healing and its impact on abnormal scarring. Adv. Wound Care 2015, 4, 119–136. [Google Scholar] [CrossRef] [Green Version]
  6. Gkretsi, V.; Stylianopoulos, T. Cell adhesion and matrix stiffness: Coordinating cancer cell invasion and metastasis. Front. Oncol. 2018, 8, 145. [Google Scholar] [CrossRef]
  7. Jinka, R.; Kapoor, R.; Sistla, P.G.; Raj, T.A.; Pande, G. Alterations in cell-extracellular matrix interactions during progression of cancers. Int. J. Cell Biol. 2012, 2012, 219196. [Google Scholar] [CrossRef] [Green Version]
  8. Nagase, H.; Woessner, J.F. Matrix metalloproteinases. J. Biol. Chem. 1999, 274, 21491–21494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Visse, R.; Nagase, H. Matrix metalloproteinases and tissue inhibitors of metalloproteinases: Structure, function, and biochemistry. Circ. Res. 2003, 92, 827–839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Lu, P.; Takai, K.; Weaver, V.M.; Werb, Z. Extracellular matrix degradation and remodelling in development and disease. Cold Spring Harb. Perspect. Biol. 2011, 3, a005058. [Google Scholar] [CrossRef] [PubMed]
  11. Rohani, M.G.; Parks, W.C. Matrix remodelling by MMPs during wound repair. Matrix Biol. J. Int. Soc. Matrix Biol. 2015, 44–46, 113–121. [Google Scholar] [CrossRef]
  12. Bashir, O.; Cao, J.; Zucker, S. Clinical relevance of MMP and TIMP measurements in cancer tissue. In Matrix Proteases in Health and Disease; Behrendt, N., Ed.; Wiley: Hoboken, NJ, USA, 2012; pp. 345–371. [Google Scholar]
  13. Cathcart, J.; Pulkoski-Gross, A.; Zucker, S.; Cao, J. Therapeutics targeting matrix metalloproteinases. In Matrix Metalloproteinase Biology; Sagi, I., Gaffney, J.P., Eds.; Wiley: Hoboken, NJ, USA, 2015; pp. 85–101. [Google Scholar]
  14. Zucker, S.; Vacirca, J. Role of matrix metalloproteinases (MMPs) in colorectal cancer. Cancer Metastasis Rev. 2004, 23, 101–117. [Google Scholar] [CrossRef] [PubMed]
  15. Villanueva, J.; Herlyn, M. Melanoma and the tumour microenvironment. Curr. Oncol. Rep. 2008, 10, 439–446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Leonardi, G.C.; Falzone, L.; Salemi, R.; Zanghì, A.; Spandidos, D.A.; Mccubrey, J.A.; Candido, S.; Libra, M. Cutaneous melanoma: From pathogenesis to therapy (Review). Int. J. Oncol. 2018, 52, 1071–1080. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Falzone, L.; Marconi, A.; Loreto, C.; Franco, S.; Spandidos, D.A.; Libra, M. Occupational exposure to carcinogens: Benzene, pesticides and fibers (Review). Mol. Med. Rep. 2016, 14, 4467–4474. [Google Scholar] [CrossRef] [Green Version]
  18. Azoury, S.C.; Lange, J.R. Epidemiology, risk factors, prevention, and early detection of melanoma. Surg. Clin. N. Am. 2014, 94, 945–962. [Google Scholar] [CrossRef]
  19. Falzone, L.; Salomone, S.; Libra, M. Evolution of cancer pharmacological treatments at the turn of the third millennium. Front. Pharmacol. 2018, 9, 1300. [Google Scholar] [CrossRef] [Green Version]
  20. CDC. CDC Skin Cancer Statistics. Available online: https://www.cdc.gov/cancer/skin/statistics/index.htm (accessed on 1 April 2019).
  21. WHO. Skin Cancers. Available online: http://www.who.int/uv/faq/skincancer/en/ (accessed on 1 April 2019).
  22. Potrony, M.; Badenas, C.; Aguilera, P.; Puig-Butille, J.A.; Carrera, C.; Malvehy, J.; Puig, S. Update in genetic susceptibility in melanoma. Ann. Transl. Med. 2015, 3, 210. [Google Scholar]
  23. Soura, E.; Eliades, P.J.; Shannon, K.; Stratigos, A.J.; Tsao, H. Hereditary melanoma: Update on syndromes and management: Genetics of familial atypical multiple mole melanoma syndrome. J. Am. Acad. Dermatol. 2016, 74, 395–407; quiz 408–410. [Google Scholar] [CrossRef] [Green Version]
  24. Harland, M.; Cust, A.E.; Badenas, C.; Chang, Y.-M.; Holland, E.A.; Aguilera, P.; Aitken, J.F.; Armstrong, B.K.; Barrett, J.H.; Carrera, C.; et al. Prevalence and predictors of germline CDKN2A mutations for melanoma cases from Australia, Spain and the United Kingdom. Hered. Cancer Clin. Pract. 2014, 12, 20. [Google Scholar] [CrossRef] [Green Version]
  25. Candido, S.; Rapisarda, V.; Marconi, A.; Malaponte, G.; Bevelacqua, V.; Gangemi, P.; Scalisi, A.; McCubrey, J.A.; Maestro, R.; Spandidos, D.A.; et al. Analysis of the B-RafV600E mutation in cutaneous melanoma patients with occupational sun exposure. Oncol. Rep. 2014, 31, 1079–1082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Wangari-Talbot, J.; Chen, S. Genetics of melanoma. Front. Genet. 2012, 3, 330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Anna, B.; Blazej, Z.; Jacqueline, G.; Andrew, C.J.; Jeffrey, R.; Andrzej, S. Mechanism of UV-related carcinogenesis and its contribution to nevi/melanoma. Expert Rev. Dermatol. 2007, 2, 451–469. [Google Scholar] [PubMed]
  28. Tate, J.G.; Bamford, S.; Jubb, H.C.; Sondka, Z.; Beare, D.M.; Bindal, N.; Boutselakis, H.; Cole, C.G.; Creatore, C.; Dawson, E.; et al. COSMIC: The Catalogue Of Somatic Mutations In Cancer. Nucleic Acids Res. 2019, 47, D941–D947. [Google Scholar] [CrossRef] [Green Version]
  29. Long, G.V.; Menzies, A.M.; Nagrial, A.M.; Haydu, L.E.; Hamilton, A.L.; Mann, G.J.; Hughes, T.M.; Thompson, J.F.; Scolyer, R.A.; Kefford, R.F. Prognostic and clinicopathologic associations of oncogenic BRAF in metastatic melanoma. J. Clin. Oncol. 2011, 29, 1239–1246. [Google Scholar] [CrossRef] [PubMed]
  30. Piérard, G.E. Cell proliferation in cutaneous malignant melanoma: Relationship with neoplastic progression. ISRN Dermatol. 2012, 2012, 828146. [Google Scholar] [CrossRef] [Green Version]
  31. Broussard, L.; Howland, A.; Ryu, S.; Song, K.; Norris, D.; Armstrong, C.A.; Song, P.I. Melanoma cell death mechanisms. Chonnam. Med. J. 2018, 54, 135–142. [Google Scholar] [CrossRef] [Green Version]
  32. Salemi, R.; Falzone, L.; Madonna, G.; Polesel, J.; Cinà, D.; Mallardo, D.; Ascierto, P.A.; Libra, M.; Candido, S. MMP-9 as a candidate marker of response to BRAF inhibitors in melanoma patients with BRAFV600E mutation detected in circulating-free DNA. Front. Pharmacol. 2018, 9, 856. [Google Scholar] [CrossRef]
  33. Falzone, L.; Salemi, R.; Travali, S.; Scalisi, A.; McCubrey, J.A.; Candido, S.; Libra, M. MMP-9 overexpression is associated with intragenic hypermethylation of MMP9 gene in melanoma. Aging 2016, 8, 933–944. [Google Scholar] [CrossRef] [Green Version]
  34. Guarneri, C.; Bevelacqua, V.; Polesel, J.; Falzone, L.; Cannavò, P.S.; Spandidos, D.A.; Malaponte, G.; Libra, M. NF-κB inhibition is associated with OPN/MMP-9 downregulation in cutaneous melanoma. Oncol. Rep. 2017, 37, 737–746. [Google Scholar] [CrossRef] [Green Version]
  35. Cerdà-Costa, N.; Gomis-Rüth, F.X. Architecture and function of metallopeptidase catalytic domains. Protein Sci. 2014, 23, 123–144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Cui, N.; Hu, M.; Khalil, R.A. Biochemical and biological attributes of matrix metalloproteinases. Prog. Mol. Biol. Transl. Sci. 2017, 147, 1–73. [Google Scholar] [PubMed] [Green Version]
  37. Sagi, I.; Gaffney, J. Matrix Metalloproteinase Biology; Wiley: Hoboken, NJ, USA, 2015; pp. 1–216. [Google Scholar]
  38. Moro, N.; Mauch, C.; Zigrino, P. Metalloproteinases in melanoma. Eur. J. Cell Biol. 2014, 93, 23–29. [Google Scholar] [CrossRef] [PubMed]
  39. Jackson, B.C.; Nebert, D.W.; Vasiliou, V. Update of human and mouse matrix metalloproteinase families. Hum. Genomics 2010, 4, 194–201. [Google Scholar] [CrossRef] [Green Version]
  40. Rumbaugh, J.; Turchan-Cholewo, J.; Galey, D.; St Hillaire, C.; Anderson, C.; Conant, K.; Nath, A. Interaction of HIV Tat and matrix metalloproteinase in HIV neuropathogenesis: A new host defense mechanism. FASEB J. 2006, 20, 1736–1738. [Google Scholar] [CrossRef] [PubMed]
  41. Giricz, O.; Lauer, J.L.; Fields, G.B. Variability in melanoma metalloproteinase expression profiling. J. Biomol. Tech. 2010, 21, 194–204. [Google Scholar]
  42. Redondo, P.; Lloret, P.; Idoate, M.; Inoges, S. Expression and serum levels of MMP-2 and MMP-9 during human melanoma progression. Clin. Exp. Dermatol. 2005, 30, 541–545. [Google Scholar] [CrossRef]
  43. Vandooren, J.; Van den Steen, P.E.; Opdenakker, G. Biochemistry and molecular biology of gelatinase B or matrix metalloproteinase-9 (MMP-9): The next decade. Crit. Rev. Biochem. Mol. Biol. 2013, 48, 222–272. [Google Scholar] [CrossRef]
  44. Klein, T.; Bischoff, R. Physiology and pathophysiology of matrix metalloproteases. Amino Acids 2011, 41, 271–290. [Google Scholar] [CrossRef] [Green Version]
  45. Motrescu, E.R.; Blaise, S.; Etique, N.; Messaddeq, N.; Chenard, M.P.; Stoll, I.; Tomasetto, C.; Rio, M.C. Matrix metalloproteinase-11/stromelysin-3 exhibits collagenolytic function against collagen VI under normal and malignant conditions. Oncogene 2008, 27, 6347–6355. [Google Scholar] [CrossRef] [Green Version]
  46. Kawasaki, K.; Kawakami, T.; Watabe, H.; Itoh, F.; Mizoguchi, M.; Soma, Y. Expression of matrilysin (matrix metalloproteinase-7) in primary cutaneous and metastatic melanoma. Br. J. Dermatol. 2007, 156, 613–619. [Google Scholar] [CrossRef] [PubMed]
  47. De Coignac, A.B.; Elson, G.; Delneste, Y.; Magistrelli, G.; Jeannin, P.; Aubry, J.P.; Berthier, O.; Schmitt, D.; Bonnefoy, J.Y.; Gauchat, J.F. Cloning of MMP-26. A novel matrilysin-like proteinase. Eur. J. Biochem. 2000, 267, 3323–3329. [Google Scholar] [CrossRef] [PubMed]
  48. Martinez-Santibanez, G.; Singer, K.; Cho, K.W.; DelProposto, J.L.; Mergian, T.; Lumeng, C.N. Obesity-induced remodeling of the adipose tissue elastin network is independent of the metalloelastase MMP-12. Adipocyte 2015, 4, 264–272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Sato, H.; Takino, T. Coordinate action of membrane-type matrix metalloproteinase-1 (MT1-MMP) and MMP-2 enhances pericellular proteolysis and invasion. Cancer Sci. 2010, 101, 843–847. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Turunen, S.P.; Tatti-Bugaeva, O.; Lehti, K. Membrane-type matrix metalloproteases as diverse effectors of cancer progression. Biochim. Biophys. Acta Mol. Cell. Res. 2017, 1864, 1974–1988. [Google Scholar] [CrossRef]
  51. Zucker, S.; Pei, D.; Cao, J.; Lopez-Otin, C. Membrane type-matrix metalloproteinases (MT-MMP). Curr. Top. Dev. Biol. 2003, 54, 1–74. [Google Scholar] [PubMed]
  52. Müller, M.; Beck, I.M.; Gadesmann, J.; Karschuk, N.; Paschen, A.; Proksch, E.; Djonov, V.; Reiss, K.; Sedlacek, R. MMP19 is upregulated during melanoma progression and increases invasion of melanoma cells. Mod. Pathol. 2010, 23, 511–521. [Google Scholar] [CrossRef] [Green Version]
  53. Llano, E.; Pendas, A.M.; Knauper, V.; Sorsa, T.; Salo, T.; Salido, E.; Murphy, G.; Simmer, J.P.; Bartlett, J.D.; Lopez-Otin, C. Identification and structural and functional characterization of human enamelysin (MMP-20). Biochemistry 1997, 36, 15101–15108. [Google Scholar] [CrossRef]
  54. Kuivanen, T.; Ahokas, K.; Virolainen, S.; Jahkola, T.; Hölttä, E.; Saksela, O.; Saarialho-Kere, U. MMP-21 is upregulated at early stages of melanoma progression but disappears with more aggressive phenotype. Virchows Arch. Int. J. Pathol. 2005, 447, 954–960. [Google Scholar] [CrossRef]
  55. Galea, C.A.; Nguyen, H.M.; George Chandy, K.; Smith, B.J.; Norton, R.S. Domain structure and function of matrix metalloprotease 23 (MMP23): Role in potassium channel trafficking. Cell. Mol. Life Sci. 2014, 71, 1191–1210. [Google Scholar] [CrossRef]
  56. Cominelli, A.; Halbout, M.; N’Kuli, F.; Lemoine, P.; Courtoy, P.J.; Marbaix, E.; Tyteca, D.; Henriet, P. A unique C-terminal domain allows retention of matrix metalloproteinase-27 in the endoplasmic reticulum. Traffic 2014, 15, 401–417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Marchenko, G.N.; Strongin, A.Y. MMP-28, a new human matrix metalloproteinase with an unusual cysteine-switch sequence is widely expressed in tumors. Gene 2001, 265, 87–93. [Google Scholar] [CrossRef]
  58. Fillmore, H.L.; VanMeter, T.E.; Broaddus, W.C. Membrane-type matrix metalloproteinases (MT-MMPs): Expression and function during glioma invasion. J. Neurooncol. 2001, 53, 187–202. [Google Scholar] [CrossRef] [PubMed]
  59. Chung, H.; Suh, E.-K.; Han, I.-O.; Oh, E.-S. Keratinocyte-derived laminin-332 promotes adhesion and migration in melanocytes and melanoma. J. Biol. Chem. 2011, 286, 13438–13447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Bloethner, S.; Chen, B.; Hemminki, K.; Müller-Berghaus, J.; Ugurel, S.; Schadendorf, D.; Kumar, R. Effect of common B-RAF and N-RAS mutations on global gene expression in melanoma cell lines. Carcinogenesis 2005, 26, 1224–1232. [Google Scholar] [CrossRef] [Green Version]
  61. Kerkelä, E.; Saarialho-Kere, U. Matrix metalloproteinases in tumour progression: Focus on basal and squamous cell skin cancer. Exp. Dermatol. 2003, 12, 109–125. [Google Scholar] [CrossRef]
  62. Rowsell, S.; Hawtin, P.; Minshull, C.A.; Jepson, H.; Brockbank, S.M.V.; Barratt, D.G.; Slater, A.M.; McPheat, W.L.; Waterson, D.; Henney, A.M.; et al. Crystal structure of human MMP9 in complex with a reverse hydroxamate inhibitor. J. Mol. Biol. 2002, 319, 173–181. [Google Scholar] [CrossRef]
  63. Fridman, R.; Toth, M.; Peña, D.; Mobashery, S. Activation of progelatinase B (MMP-9) by gelatinase A (MMP-2). Cancer Res. 1995, 55, 2548–2555. [Google Scholar]
  64. Ogata, Y.; Enghild, J.J.; Nagase, H. Matrix metalloproteinase 3 (stromelysin) activates the precursor for the human matrix metalloproteinase 9. J. Biol. Chem. 1992, 267, 3581–3584. [Google Scholar]
  65. Kessenbrock, K.; Plaks, V.; Werb, Z. Matrix metalloproteinases: Regulators of the tumour microenvironment. Cell 2010, 141, 52–67. [Google Scholar] [CrossRef] [Green Version]
  66. Falzone, L.; Candido, S.; Salemi, R.; Basile, M.S.; Scalisi, A.; McCubrey, J.A.; Torino, F.; Signorelli, S.S.; Montella, M.; Libra, M. Computational identification of microRNAs associated with both epithelial to mesenchymal transition and NGAL/MMP-9 pathways in bladder cancer. Oncotarget 2016, 7, 72758–72766. [Google Scholar] [PubMed] [Green Version]
  67. Bera, A.; Zhao, S.; Cao, L.; Chiao, P.J.; Freeman, J.W. Oncogenic K-Ras and loss of Smad4 mediate invasion by activating an EGFR/NF-κB Axis that induces expression of MMP9 and uPA in human pancreas progenitor cells. PLoS ONE 2013, 8, e82282. [Google Scholar] [CrossRef] [Green Version]
  68. Zhao, X.; Benveniste, E.N. Transcriptional activation of human matrix metalloproteinase-9 gene expression by multiple co-activators. J. Mol. Biol. 2008, 383, 945–956. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Avruch, J.; Khokhlatchev, A.; Kyriakis, J.M.; Luo, Z.; Tzivion, G.; Vavvas, D.; Zhang, X.F. Ras activation of the Raf kinase: Tyrosine kinase recruitment of the MAP kinase cascade. Recent Prog. Horm. Res. 2001, 56, 127–155. [Google Scholar] [CrossRef]
  70. Santarpia, L.; Lippman, S.M.; El-Naggar, A.K. Targeting the MAPK-RAS-RAF signaling pathway in cancer therapy. Expert Opin. Ther. Targets 2012, 16, 103–119. [Google Scholar] [CrossRef] [Green Version]
  71. McCubrey, J.A.; Steelman, L.S.; Chappell, W.H.; Abrams, S.L.; Wong, E.W.T.; Chang, F.; Lehmann, B.; Terrian, D.M.; Milella, M.; Tafuri, A.; et al. Roles of the RAF/MEK/ERK pathway in cell growth, malignant transformation and drug resistance. Biochim. Biophys. Acta 2007, 1773, 1263–1284. [Google Scholar] [CrossRef] [Green Version]
  72. Wei, Y.; Zhao, L.; He, W.; Yang, J.; Geng, C.; Chen, Y.; Liu, T.; Chen, H.; Li, Y. Benzo[a]pyrene promotes gastric cancer cell proliferation and metastasis likely through the Aryl hydrocarbon receptor and ERK-dependent induction of MMP9 and c-myc. Int. J. Oncol. 2016, 49, 2055–2063. [Google Scholar] [CrossRef] [Green Version]
  73. Wu, Y.-J.; Neoh, C.-A.; Tsao, C.-Y.; Su, J.-H.; Li, H.-H. Sinulariolide suppresses human hepatocellular carcinoma cell migration and invasion by inhibiting matrix metalloproteinase-2/-9 through MAPKs and PI3K/Akt signaling pathways. Int. J. Mol. Sci. 2015, 16, 16469–16482. [Google Scholar] [CrossRef] [Green Version]
  74. Wu, H.; Larribère, L.; Sun, Q.; Novak, D.; Sachindra, S.; Granados, K.; Umansky, V.; Utikal, J. Loss of neural crest-associated gene FOXD1 impairs melanoma invasion and migration via RAC1B downregulation. Int. J. Cancer 2018, 143, 2962–2972. [Google Scholar] [CrossRef] [Green Version]
  75. Maniotis, A.J.; Folberg, R.; Hess, A.; Seftor, E.A.; Gardner, L.M.G.; Pe’er, J.; Trent, J.M.; Meltzer, P.S.; Hendrix, M.J.C. Vascular channel formation by human melanoma cells in vivo and in vitro: Vasculogenic mimicry. Am. J. Pathol. 1999, 155, 739–752. [Google Scholar] [CrossRef] [Green Version]
  76. Ge, H.; Luo, H. Overview of advances in vasculogenic mimicry—A potential target for tumour therapy. Cancer Manag. Res. 2018, 10, 2429–2437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Liang, X.; Sun, R.; Zhao, X.; Zhang, Y.; Gu, Q.; Dong, X.; Zhang, D.; Sun, J.; Sun, B. Rictor regulates the vasculogenic mimicry of melanoma via the AKT-MMP-2/9 pathway. J. Cell. Mol. Med. 2017, 21, 3579–3591. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Cheng, Q.; Wu, J.; Zhang, Y.; Liu, X.; Xu, N.; Zuo, F.; Xu, J. SOX4 promotes melanoma cell migration and invasion though the activation of the NF-κB signaling pathway. Int. J. Mol. Med. 2017, 40, 447–453. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Xu, Y.; Wang, Y.; Liu, H.; Shi, Q.; Zhu, D.; Amos, C.I.; Fang, S.; Lee, J.E.; Hyslop, T.; Li, X.; et al. Genetic variants in the metzincin metallopeptidase family genes predict melanoma survival. Mol. Carcinog. 2018, 57, 22–31. [Google Scholar] [CrossRef] [Green Version]
  80. Battaglia, R.; Palini, S.; Vento, M.E.; Ferlita, A.L.; Faro, M.J.L.; Caroppo, E.; Borzì, P.; Falzone, L.; Barbagallo, D.; Ragusa, M.; et al. Identification of extracellular vesicles and characterization of miRNA expression profiles in human blastocoel fluid. Sci. Rep. 2019, 9. [Google Scholar] [CrossRef]
  81. Hafsi, S.; Candido, S.; Maestro, R.; Falzone, L.; Soua, Z.; Bonavida, B.; Spandidos, D.A.; Libra, M. Correlation between the overexpression of Yin Yang 1 and the expression levels of miRNAs in Burkitt’s lymphoma: A computational study. Oncol. Lett. 2016, 11, 1021–1025. [Google Scholar] [CrossRef] [Green Version]
  82. Cancer Genome Atlas Research Network; Weinstein, J.N.; Collisson, E.A.; Mills, G.B.; Shaw, K.R.M.; Ozenberger, B.A.; Ellrott, K.; Shmulevich, I.; Sander, C.; Stuart, J.M. The Cancer Genome Atlas Pan-Cancer analysis project. Nat. Genet. 2013, 45, 1113–1120. [Google Scholar]
  83. Filetti, V.; Falzone, L.; Rapisarda, V.; Caltabiano, R.; Graziano, E.A.C.; Ledda, C.; Loreto, C. Modulation of microRNA expression levels after naturally occurring asbestiform fibers exposure as a diagnostic biomarker of mesothelial neoplastic transformation. Ecotoxicol. Environ. Saf. 2020, 198, 110640. [Google Scholar] [CrossRef]
  84. Falzone, L.; Romano, G.L.; Salemi, R.; Bucolo, C.; Tomasello, B.; Lupo, G.; Anfuso, C.D.; Spandidos, D.A.; Libra, M.; Candido, S. Prognostic significance of deregulated microRNAs in uveal melanomas. Mol. Med. Rep. 2019, 19, 2599–2610. [Google Scholar] [CrossRef] [Green Version]
  85. Falzone, L.; Scola, L.; Zanghì, A.; Biondi, A.; Di Cataldo, A.; Libra, M.; Candido, S. Integrated analysis of colorectal cancer microRNA datasets: Identification of microRNAs associated with tumour development. Aging 2018, 10, 1000–1014. [Google Scholar] [CrossRef]
  86. Polo, A.; Crispo, A.; Cerino, P.; Falzone, L.; Candido, S.; Giudice, A.; De Petro, G.; Ciliberto, G.; Montella, M.; Budillon, A.; et al. Environment and bladder cancer: Molecular analysis by interaction networks. Oncotarget 2017, 8, 65240–65252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. McCubrey, J.A.; Fitzgerald, T.L.; Yang, L.V.; Lertpiriyapong, K.; Steelman, L.S.; Abrams, S.L.; Montalto, G.; Cervello, M.; Neri, L.M.; Cocco, L.; et al. Roles of GSK-3 and microRNAs on epithelial mesenchymal transition and cancer stem cells. Oncotarget 2017, 8, 14221–14250. [Google Scholar] [CrossRef] [PubMed]
  88. Candido, S.; Lupo, G.; Pennisi, M.; Basile, M.S.; Anfuso, C.D.; Petralia, M.C.; Gattuso, G.; Vivarelli, S.; Spandidos, D.A.; Libra, M.; et al. The analysis of miRNA expression profiling datasets reveals inverse microRNA patterns in glioblastoma and Alzheimer’s disease. Oncol. Rep. 2019, 42, 911–922. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Falzone, L.; Lupo, G.; La Rosa, G.R.M.; Crimi, S.; Anfuso, C.D.; Salemi, R.; Rapisarda, E.; Libra, M.; Candido, S. Identification of novel microRNAs and their diagnostic and prognostic significance in oral cancer. Cancers (Basel) 2019, 11, E610. [Google Scholar] [CrossRef] [Green Version]
  90. Zhou, X.; Yan, T.; Huang, C.; Xu, Z.; Wang, L.; Jiang, E.; Wang, H.; Chen, Y.; Liu, K.; Shao, Z.; et al. Melanoma cell-secreted exosomal miR-155-5p induce proangiogenic switch of cancer-associated fibroblasts via SOCS1/JAK2/STAT3 signaling pathway. J. Exp. Clin. Cancer Res. CR 2018, 37, 242. [Google Scholar] [CrossRef] [Green Version]
  91. Wang, X.; Hu, Y.; Cui, J.; Zhou, Y.; Chen, L. Coordinated targeting of MMP-2/MMP-9 by miR-296-3p/FOXCUT exerts tumour-suppressing effects in choroidal malignant melanoma. Mol. Cell. Biochem. 2018, 445, 25–33. [Google Scholar] [CrossRef]
  92. Saranga-Perry, V.; Ambe, C.; Zager, J.S.; Kudchadkar, R.R. Recent developments in the medical and surgical treatment of melanoma. CA Cancer J. Clin. 2014, 64, 171–185. [Google Scholar] [CrossRef]
  93. Tran, K.A.; Cheng, M.Y.; Mitra, A.; Ogawa, H.; Shi, V.Y.; Olney, L.P.; Kloxin, A.M.; Maverakis, E. MEK inhibitors and their potential in the treatment of advanced melanoma: The advantages of combination therapy. Drug Des. Dev. Ther. 2016, 10, 43–52. [Google Scholar]
  94. Long, G.V.; Flaherty, K.T.; Stroyakovskiy, D.; Gogas, H.; Levchenko, E.; de Braud, F.; Larkin, J.; Garbe, C.; Jouary, T.; Hauschild, A.; et al. Dabrafenib plus trametinib versus dabrafenib monotherapy in patients with metastatic BRAF V600E/K-mutant melanoma: Long-term survival and safety analysis of a phase 3 study. Ann. Oncol. 2017, 28, 1631–1639. [Google Scholar] [CrossRef]
  95. Menzies, A.M.; Long, G.V.; Murali, R. Dabrafenib and its potential for the treatment of metastatic melanoma. Drug Des. Devel. Ther. 2012, 6, 391–405. [Google Scholar]
  96. Chopra, N.; Nathan, P.D. Trametinib in metastatic melanoma. Expert Rev. Anticancer Ther. 2015, 15, 749–760. [Google Scholar] [CrossRef] [PubMed]
  97. Tuaeva, N.O.; Falzone, L.; Porozov, Y.B.; Nosyrev, A.E.; Trukhan, V.M.; Kovatsi, L.; Spandidos, D.A.; Drakoulis, N.; Kalogeraki, A.; Mamoulakis, C.; et al. Translational application of circulating DNA in oncology: Review of the last decades achievements. Cells 2019, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Silantyev, A.S.; Falzone, L.; Libra, M.; Gurina, O.I.; Kardashova, K.S.; Nikolouzakis, T.K.; Nosyrev, A.E.; Sutton, C.W.; Mitsias, P.D.; Tsatsakis, A. Current and future trends on diagnosis and prognosis of glioblastoma: From molecular biology to proteomics. Cells 2019, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Lugowska, I.; Teterycz, P.; Rutkowski, P. Immunotherapy of melanoma. Contemp. Oncol. 2018, 22, 61–67. [Google Scholar] [CrossRef]
  100. Christofi, T.; Baritaki, S.; Falzone, L.; Libra, M.; Zaravinos, A. Current perspectives in cancer immunotherapy. Cancers (Basel) 2019, 11, E1472. [Google Scholar] [CrossRef] [Green Version]
  101. Deken, M.A.; Gadiot, J.; Jordanova, E.S.; Lacroix, R.; van Gool, M.; Kroon, P.; Pineda, C.; Geukes Foppen, M.H.; Scolyer, R.; Song, J.-Y.; et al. Targeting the MAPK and PI3K pathways in combination with PD1 blockade in melanoma. Oncoimmunology 2016, 5, e1238557. [Google Scholar] [CrossRef] [Green Version]
  102. Vivarelli, S.; Salemi, R.; Candido, S.; Falzone, L.; Santagati, M.; Stefani, S.; Torino, F.; Banna, G.L.; Tonini, G.; Libra, M. Gut microbiota and cancer: From pathogenesis to therapy. Cancers 2019, 11, 38. [Google Scholar] [CrossRef] [Green Version]
  103. Banna, G.L.; Torino, F.; Marletta, F.; Santagati, M.; Salemi, R.; Cannarozzo, E.; Falzone, L.; Ferraù, F.; Libra, M. Lactobacillus rhamnosus GG: An overview to explore the rationale of its use in cancer. Front. Pharmacol. 2017, 8, 603. [Google Scholar] [CrossRef] [Green Version]
  104. Acharya, M.R.; Venitz, J.; Figg, W.D.; Sparreboom, A. Chemically modified tetracyclines as inhibitors of matrix metalloproteinases. Drug Resist. Updat. 2004, 7, 195–208. [Google Scholar] [CrossRef]
  105. Woessner, J.F. Matrix metalloproteinase inhibition. From the Jurassic to the third millennium. Ann. N. Y. Acad. Sci. 1999, 878, 388–403. [Google Scholar] [CrossRef]
  106. Brown, P.D. Matrix metalloproteinase inhibitors in the treatment of cancer. Med. Oncol. 1997, 14, 1–10. [Google Scholar] [CrossRef] [PubMed]
  107. Whittaker, M.; Floyd, C.D.; Brown, P.; Gearing, A.J. Design and therapeutic application of matrix metalloproteinase inhibitors. Chem. Rev. 1999, 99, 2735–2776. [Google Scholar] [CrossRef] [PubMed]
  108. Underwood, C.K.; Min, D.; Lyons, J.G.; Hambley, T.W. The interaction of metal ions and Marimastat with matrix metalloproteinase 9. J. Inorg. Biochem. 2003, 95, 165–170. [Google Scholar] [CrossRef]
  109. Chirivi, R.G.; Garofalo, A.; Crimmin, M.J.; Bawden, L.J.; Stoppacciaro, A.; Brown, P.D.; Giavazzi, R. Inhibition of the metastatic spread and growth of B16-BL6 murine melanoma by a synthetic matrix metalloproteinase inhibitor. Int. J. Cancer 1994, 58, 460–464. [Google Scholar] [CrossRef] [PubMed]
  110. Dabrowska, A.; Giermasz, A.; Marczak, M.; Gołab, J.; Jakóbisiak, M. Potentiated antitumour effects of interleukin 12 and matrix metalloproteinase inhibitor batimastat against B16F10 melanoma in mice. Anticancer Res. 2000, 20, 391–394. [Google Scholar] [PubMed]
  111. Wylie, S.; MacDonald, I.C.; Varghese, H.J.; Schmidt, E.E.; Morris, V.L.; Groom, A.C.; Chambers, A.F. The matrix metalloproteinase inhibitor batimastat inhibits angiogenesis in liver metastases of B16F1 melanoma cells. Clin. Exp. Metastasis 1999, 17, 111–117. [Google Scholar] [CrossRef] [PubMed]
  112. Kasaoka, T.; Nishiyama, H.; Okada, M.; Nakajima, M. Matrix metalloproteinase inhibitor, MMI270 (CGS27023A) inhibited hematogenic metastasis of B16 melanoma cells in both experimental and spontaneous metastasis models. Clin. Exp. Metastasis 2008, 25, 827–834. [Google Scholar] [CrossRef]
  113. Conway, J.G.; Trexler, S.J.; Wakefield, J.A.; Marron, B.E.; Emerson, D.L.; Bickettf, D.M.; Deaton, D.N.; Garrison, D.; Elder, M.; McElroy, A.; et al. Effect of matrix metalloproteinase inhibitors on tumour growth and spontaneous metastasis. Clin. Exp. Metastasis 1996, 14, 115–124. [Google Scholar] [CrossRef]
  114. Oku, T.; Ata, N.; Yonezawa, K.; Tokai, H.; Fujii, H.; Shinagawa, A.; Ohuchi, E.; Saiki, I. Antimetastatic and antitumour effect of a recombinant human tissue inhibitor of metalloproteinases-2 in murine melanoma models. Biol. Pharm. Bull. 1997, 20, 843–849. [Google Scholar] [CrossRef] [Green Version]
  115. Hojo, K.; Maki, H.; Sawada, T.Y.; Maekawa, R.; Yoshioka, T. Augmented growth inhibition of B16-BL6 melanoma by combined treatment with a selective matrix metalloproteinase inhibitor, MMI-166, and cytotoxic agents. Anticancer Res. 2002, 22, 3253–3259. [Google Scholar]
  116. Coussens, L.M.; Fingleton, B.; Matrisian, L.M. Matrix metalloproteinase inhibitors and cancer: Trials and tribulations. Science 2002, 295, 2387–2392. [Google Scholar] [CrossRef] [PubMed]
  117. Griffioen, A.W. AG-3340 (Agouron Pharmaceuticals Inc). IDrugs Investig. Drugs J. 2000, 3, 336–345. [Google Scholar]
  118. Fisher, J.F.; Mobashery, S. Recent advances in MMP inhibitor design. Cancer Metastasis Rev. 2006, 25, 115–136. [Google Scholar] [CrossRef] [PubMed]
  119. Bissett, D.; O’Byrne, K.J.; von Pawel, J.; Gatzemeier, U.; Price, A.; Nicolson, M.; Mercier, R.; Mazabel, E.; Penning, C.; Zhang, M.H.; et al. Phase III study of matrix metalloproteinase inhibitor prinomastat in non-small-cell lung cancer. J. Clin. Oncol. Off. J. Am. Soc. Clin. Oncol. 2005, 23, 842–849. [Google Scholar] [CrossRef]
  120. Vitale, C.; D’Amato, M.; Calabrò, P.; Stanziola, A.A.; Mormile, M.; Molino, A. Venous thromboembolism and lung cancer: A review. Multidiscip. Respir. Med. 2015, 10, 28. [Google Scholar] [CrossRef] [Green Version]
  121. Hirte, H.; Vergote, I.B.; Jeffrey, J.R.; Grimshaw, R.N.; Coppieters, S.; Schwartz, B.; Tu, D.; Sadura, A.; Brundage, M.; Seymour, L. A phase III randomized trial of BAY 12-9566 (tanomastat) as maintenance therapy in patients with advanced ovarian cancer responsive to primary surgery and paclitaxel/platinum containing chemotherapy: A National Cancer Institute of Canada Clinical Trials Group Study. Gynecol. Oncol. 2006, 102, 300–308. [Google Scholar]
  122. Moore, M.J.; Hamm, J.; Dancey, J.; Eisenberg, P.D.; Dagenais, M.; Fields, A.; Hagan, K.; Greenberg, B.; Colwell, B.; Zee, B.; et al. Comparison of gemcitabine versus the matrix metalloproteinase inhibitor BAY 12-9566 in patients with advanced or metastatic adenocarcinoma of the pancreas: A phase III trial of the National Cancer Institute of Canada Clinical Trials Group. J. Clin. Oncol. 2003, 21, 3296–3302. [Google Scholar] [CrossRef]
  123. Bramhall, S.R.; Schulz, J.; Nemunaitis, J.; Brown, P.D.; Baillet, M.; Buckels, J.A.C. A double-blind placebo-controlled, randomised study comparing gemcitabine and marimastat with gemcitabine and placebo as first line therapy in patients with advanced pancreatic cancer. Br. J. Cancer 2002, 87, 161–167. [Google Scholar] [CrossRef]
  124. Leighl, N.B.; Paz-Ares, L.; Douillard, J.-Y.; Peschel, C.; Arnold, A.; Depierre, A.; Santoro, A.; Betticher, D.C.; Gatzemeier, U.; Jassem, J.; et al. Randomized phase III study of matrix metalloproteinase inhibitor BMS-275291 in combination with paclitaxel and carboplatin in advanced non-small-cell lung cancer: National Cancer Institute of Canada-Clinical Trials Group Study BR.18. J. Clin. Oncol. 2005, 23, 2831–2839. [Google Scholar] [CrossRef]
  125. Kilian, M.; Gregor, J.I.; Heukamp, I.; Hanel, M.; Ahlgrimm, M.; Schimke, I.; Kristiansen, G.; Ommer, A.; Walz, M.K.; Jacobi, C.A.; et al. Matrix metalloproteinase inhibitor RO 28-2653 decreases liver metastasis by reduction of MMP-2 and MMP-9 concentration in BOP-induced ductal pancreatic cancer in Syrian Hamsters: Inhibition of matrix metalloproteinases in pancreatic cancer. Prostaglandins Leukot. Essent. Fatty Acids 2006, 75, 429–434. [Google Scholar] [CrossRef]
  126. Yan, Y.-L.; Miller, M.T.; Cao, Y.; Cohen, S.M. Synthesis of hydroxypyrone- and hydroxythiopyrone-based matrix metalloproteinase inhibitors: Developing a structure-activity relationship. Bioorg. Med. Chem. Lett. 2009, 19, 1970–1976. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Breuer, E.; Salomon, C.J.; Katz, Y.; Chen, W.; Lu, S.; Röschenthaler, G.-V.; Hadar, R.; Reich, R. Carbamoylphosphonates, a new class of in vivo active matrix metalloproteinase inhibitors. 1. Alkyl- and cycloalkylcarbamoylphosphonic acids. J. Med. Chem. 2004, 47, 2826–2832. [Google Scholar] [CrossRef] [PubMed]
  128. Farkas, E.; Katz, Y.; Bhusare, S.; Reich, R.; Röschenthaler, G.-V.; Königsmann, M.; Breuer, E. Carbamoylphosphonate-based matrix metalloproteinase inhibitor metal complexes: Solution studies and stability constants. Towards a zinc-selective binding group. J. Biol. Inorg. Chem. 2004, 9, 307–315. [Google Scholar] [CrossRef]
  129. Pochetti, G.; Gavuzzo, E.; Campestre, C.; Agamennone, M.; Tortorella, P.; Consalvi, V.; Gallina, C.; Hiller, O.; Tschesche, H.; Tucker, P.A.; et al. Structural insight into the stereoselective inhibition of MMP-8 by enantiomeric sulfonamide phosphonates. J. Med. Chem. 2006, 49, 923–931. [Google Scholar] [CrossRef] [PubMed]
  130. Zakeri, B.; Wright, G.D. Chemical biology of tetracycline antibiotics. Biochem. Cell Biol. 2008, 86, 124–136. [Google Scholar] [CrossRef] [PubMed]
  131. Rundhaug, J.E. Matrix metalloproteinases and angiogenesis. J. Cell. Mol. Med. 2005, 9, 267–285. [Google Scholar] [CrossRef]
  132. Hua, H.; Li, M.; Luo, T.; Yin, Y.; Jiang, Y. Matrix metalloproteinases in tumourigenesis: An evolving paradigm. Cell. Mol. Life Sci. CMLS 2011, 68, 3853–3868. [Google Scholar] [CrossRef]
  133. Nagase, H.; Visse, R.; Murphy, G. Structure and function of matrix metalloproteinases and TIMPs. Cardiovasc. Res. 2006, 69, 562–573. [Google Scholar] [CrossRef] [Green Version]
  134. Kancha, R.K.; Stearns, M.E.; Hussain, M.M. Decreased expression of the low density lipoprotein receptor-related protein/alpha 2-macroglobulin receptor in invasive cell clones derived from human prostate and breast tumour cells. Oncol. Res. 1994, 6, 365–372. [Google Scholar]
  135. Mroczko, B.; Groblewska, M.; Okulczyk, B.; Kedra, B.; Szmitkowski, M. The diagnostic value of matrix metalloproteinase 9 (MMP-9) and tissue inhibitor of matrix metalloproteinases 1 (TIMP-1) determination in the sera of colorectal adenoma and cancer patients. Int. J. Colorectal Dis. 2010, 25, 1177–1184. [Google Scholar] [CrossRef] [Green Version]
  136. Nieuwesteeg, M.A.; Willson, J.A.; Cepeda, M.; Fox, M.A.; Damjanovski, S. Functional characterization of tissue inhibitor of metalloproteinase-1 (TIMP-1) N- and C-terminal domains during Xenopus laevis development. Sci. World J. 2014, 2014, 467907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Brew, K.; Nagase, H. The tissue inhibitors of metalloproteinases (TIMPs): An ancient family with structural and functional diversity. Biochim. Biophys. Acta 2010, 1803, 55–71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Nagase, H.; Murphy, G. Tailoring TIMPs for selective metalloproteinase inhibition. In The Cancer Degradome; Springer: Berlin, Germany, 2008; Chapter 37. [Google Scholar]
  139. Müller, V.; Riethdorf, S.; Rack, B.; Janni, W.; Fasching, P.A.; Solomayer, E.; Aktas, B.; Kasimir-Bauer, S.; Zeitz, J.; Pantel, K.; et al. Prospective evaluation of serum tissue inhibitor of metalloproteinase 1 and carbonic anhydrase IX in correlation to circulating tumour cells in patients with metastatic breast cancer. Breast Cancer Res. 2011, 13, R71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Montaño, M.; Sansores, R.H.; Becerril, C.; Cisneros, J.; González-Avila, G.; Sommer, B.; Ochoa, L.; Herrera, I.; Ramírez-Venegas, A.; Ramos, C. FEV1 inversely correlates with metalloproteinases 1, 7, 9 and CRP in COPD by biomass smoke exposure. Respir. Res. 2014, 15, 74. [Google Scholar] [CrossRef]
  141. Pinto-Plata, V.; Casanova, C.; Müllerova, H.; de Torres, J.P.; Corado, H.; Varo, N.; Cordoba, E.; Zeineldine, S.; Paz, H.; Baz, R.; et al. Inflammatory and repair serum biomarker pattern: Association to clinical outcomes in COPD. Respir. Res. 2012, 13, 71. [Google Scholar] [CrossRef] [Green Version]
  142. Yao, H.; Hwang, J.; Sundar, I.K.; Friedman, A.E.; McBurney, M.W.; Guarente, L.; Gu, W.; Kinnula, V.L.; Rahman, I. SIRT1 redresses the imbalance of tissue inhibitor of matrix metalloproteinase-1 and matrix metalloproteinase-9 in the development of mouse emphysema and human COPD. Am. J. Physiol. Lung Cell. Mol. Physiol. 2013, 305, L615–L624. [Google Scholar] [CrossRef] [Green Version]
  143. Ylisirniö, S.; Höyhtyä, M.; Turpeenniemi-Hujanen, T. Serum matrix metalloproteinases -2, -9 and tissue inhibitors of metalloproteinases -1, -2 in lung cancer--TIMP-1 as a prognostic marker. Anticancer Res. 2000, 20, 1311–1316. [Google Scholar]
  144. Wu, Z.-S.; Wu, Q.; Yang, J.-H.; Wang, H.-Q.; Ding, X.-D.; Yang, F.; Xu, X.-C. Prognostic significance of MMP-9 and TIMP-1 serum and tissue expression in breast cancer. Int. J. Cancer 2008, 122, 2050–2056. [Google Scholar] [CrossRef]
  145. Van den Steen, P.E.; Dubois, B.; Nelissen, I.; Rudd, P.M.; Dwek, R.A.; Opdenakker, G. Biochemistry and molecular biology of gelatinase B or matrix metalloproteinase-9 (MMP-9). Crit. Rev. Biochem. Mol. Biol. 2002, 37, 375–536. [Google Scholar] [CrossRef]
  146. Hofmann, U.B.; Westphal, J.R.; Van Muijen, G.N.; Ruiter, D.J. Matrix metalloproteinases in human melanoma. J. Investig. Dermatol. 2000, 115, 337–344. [Google Scholar] [CrossRef] [Green Version]
  147. Khokha, R. Suppression of the tumourigenic and metastatic abilities of murine B16-F10 melanoma cells in vivo by the overexpression of the tissue inhibitor of the metalloproteinases-1. J. Natl. Cancer Inst. 1994, 86, 299–304. [Google Scholar] [CrossRef] [PubMed]
  148. Nagoshi, S. Osteopontin: Versatile modulator of liver diseases. Hepatol. Res. 2014, 44, 22–30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Brown, L.F.; Berse, B.; Van de Water, L.; Papadopoulos-Sergiou, A.; Perruzzi, C.A.; Manseau, E.J.; Dvorak, H.F.; Senger, D.R. Expression and distribution of osteopontin in human tissues: Widespread association with luminal epithelial surfaces. Mol. Biol. Cell 1992, 3, 1169–1180. [Google Scholar] [CrossRef] [PubMed]
  150. Higgins, C.L.; Isbilir, S.; Basto, P.; Chen, I.Y.; Vaduganathan, M.; Vaduganathan, P.; Reardon, M.J.; Lawrie, G.; Peterson, L.; Morrisett, J.D. Distribution of alkaline phosphatase, osteopontin, RANK ligand and osteoprotegerin in calcified human carotid atheroma. Protein J. 2015, 34, 315–328. [Google Scholar] [CrossRef]
  151. Misawa, H.; Hara, M.; Tanabe, S.; Niikura, M.; Moriwaki, Y.; Okuda, T. Osteopontin is an alpha motor neuron marker in the mouse spinal cord. J. Neurosci. Res. 2012, 90, 732–742. [Google Scholar] [CrossRef]
  152. Chellaiah, M.A.; Kizer, N.; Biswas, R.; Alvarez, U.; Strauss-Schoenberger, J.; Rifas, L.; Rittling, S.R.; Denhardt, D.T.; Hruska, K.A. Osteopontin deficiency produces osteoclast dysfunction due to reduced CD44 surface expression. Mol. Biol. Cell 2003, 14, 173–189. [Google Scholar] [CrossRef] [Green Version]
  153. Rittling, S.R.; Matsumoto, H.N.; McKee, M.D.; Nanci, A.; An, X.R.; Novick, K.E.; Kowalski, A.J.; Noda, M.; Denhardt, D.T. Mice lacking osteopontin show normal development and bone structure but display altered osteoclast formation in vitro. J. Bone Miner. Res. 1998, 13, 1101–1111. [Google Scholar] [CrossRef]
  154. Banerjee, A.; Apte, U.M.; Smith, R.; Ramaiah, S.K. Higher neutrophil infiltration mediated by osteopontin is a likely contributing factor to the increased susceptibility of females to alcoholic liver disease. J. Pathol. 2006, 208, 473–485. [Google Scholar] [CrossRef]
  155. Koh, A.; da Silva, A.P.B.; Bansal, A.K.; Bansal, M.; Sun, C.; Lee, H.; Glogauer, M.; Sodek, J.; Zohar, R. Role of osteopontin in neutrophil function. Immunology 2007, 122, 466–475. [Google Scholar] [CrossRef]
  156. Wang, K.X.; Denhardt, D.T. Osteopontin: Role in immune regulation and stress responses. Cytokine Growth Factor Rev. 2008, 19, 333–345. [Google Scholar] [CrossRef] [Green Version]
  157. Denhardt, D.T.; Noda, M.; O’Regan, A.W.; Pavlin, D.; Berman, J.S. Osteopontin as a means to cope with environmental insults: Regulation of inflammation, tissue remodelling, and cell survival. J. Clin. Investig. 2001, 107, 1055–1061. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Standal, T.; Borset, M.; Sundan, A. Role of osteopontin in adhesion, migration, cell survival and bone remodelling. Exp. Oncol. 2004, 26, 179–184. [Google Scholar] [PubMed]
  159. Fan, K.; Dai, J.; Wang, H.; Wei, H.; Cao, Z.; Hou, S.; Qian, W.; Wang, H.; Li, B.; Zhao, J.; et al. Treatment of collagen-induced arthritis with an anti-osteopontin monoclonal antibody through promotion of apoptosis of both murine and human activated T cells. Arthritis Rheum. 2008, 58, 2041–2052. [Google Scholar] [CrossRef]
  160. Thalmann, G.N.; Sikes, R.A.; Devoll, R.E.; Kiefer, J.A.; Markwalder, R.; Klima, I.; Farach-Carson, C.M.; Studer, U.E.; Chung, L.W. Osteopontin: Possible role in prostate cancer progression. Clin. Cancer Res. 1999, 5, 2271–2277. [Google Scholar]
  161. Wai, P.Y.; Kuo, P.C. Osteopontin: Regulation in tumour metastasis. Cancer Metastasis Rev. 2008, 27, 103–118. [Google Scholar] [CrossRef]
  162. Wai, P.Y.; Kuo, P.C. The role of Osteopontin in tumour metastasis. J. Surg. Res. 2004, 121, 228–241. [Google Scholar] [CrossRef]
  163. Zhao, J.; Dong, L.; Lu, B.; Wu, G.; Xu, D.; Chen, J.; Li, K.; Tong, X.; Dai, J.; Yao, S.; et al. Down-regulation of osteopontin suppresses growth and metastasis of hepatocellular carcinoma via induction of apoptosis. Gastroenterology 2008, 135, 956–968. [Google Scholar] [CrossRef] [Green Version]
  164. Castellano, G.; Malaponte, G.; Mazzarino, M.C.; Figini, M.; Marchese, F.; Gangemi, P.; Travali, S.; Stivala, F.; Canevari, S.; Libra, M. Activation of the osteopontin/matrix metalloproteinase-9 pathway correlates with prostate cancer progression. Clin. Cancer Res. 2008, 14, 7470–7480. [Google Scholar] [CrossRef] [Green Version]
  165. Takafuji, V.; Forgues, M.; Unsworth, E.; Goldsmith, P.; Wang, X.W. An osteopontin fragment is essential for tumour cell invasion in hepatocellular carcinoma. Oncogene 2007, 26, 6361–6371. [Google Scholar] [CrossRef] [Green Version]
  166. Rangaswami, H.; Bulbule, A.; Kundu, G.C. Nuclear factor-inducing kinase plays a crucial role in osteopontin-induced MAPK/IkappaBalpha kinase-dependent nuclear factor kappaB-mediated promatrix metalloproteinase-9 activation. J. Biol. Chem. 2004, 279, 38921–38935. [Google Scholar] [CrossRef] [Green Version]
  167. Rangaswami, H.; Kundu, G.C. Osteopontin stimulates melanoma growth and lung metastasis through NIK/MEKK1-dependent MMP-9 activation pathways. Oncol. Rep. 2007, 18, 909–915. [Google Scholar] [CrossRef] [PubMed]
  168. Zhou, Y.; Dai, D.L.; Martinka, M.; Su, M.; Zhang, Y.; Campos, E.I.; Dorocicz, I.; Tang, L.; Huntsman, D.; Nelson, C.; et al. Osteopontin expression correlates with melanoma invasion. J. Investig. Dermatol. 2005, 124, 1044–1052. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Protein structure of MMP-9. MMP-9 is first produced as a pro-enzyme of 92 kDa, called pro-MMP-9, containing a pro-domain of 73 aa responsible for MMP-9 catalytic latency. Subsequently, other proteases cleave pro-MMP-9 pro-domain generating the active catalytic form of MMP-9 of 84 kDa.
Figure 1. Protein structure of MMP-9. MMP-9 is first produced as a pro-enzyme of 92 kDa, called pro-MMP-9, containing a pro-domain of 73 aa responsible for MMP-9 catalytic latency. Subsequently, other proteases cleave pro-MMP-9 pro-domain generating the active catalytic form of MMP-9 of 84 kDa.
Cells 09 01151 g001
Figure 2. Epigenetic regulation of MMP-9 in melanoma. (A) Melanoma cells secrete exosomes containing different proteins and ncRNAs, including hsa-miR-155-5p. These exosomes are internalized by cells of the tumor microenvironment, such as fibroblasts. Inside the cells the exosomal hsa-miR-155-5p is able to inhibit SOCS1 mRNA thus activating the JAK/STAT molecular pathway. Phosphorylated STAT3 binds the DNA inducing the transcription of MMP-9, VEGFa and other pro-angiogenetic factors (e.g., FGF2). These proteins will be responsible for ECM degradation, fibroblasts switch to cancer-associated fibroblast phenotype and neoangiogenesis; (B) ncRNAs may act also as inhibitors of MMP-9 and MMP-2. In particular, hsa-miR-296-3p is able to inhibit both MMP-9 and MMP-2 mRNA thus reducing their secretion in the extracellular space. Similarly, FOXCUT lncRNA is able to inhibit MMP-2 and MMP-9 transcription and translation into proteins. Both ncRNAs are responsible of the inhibition of melanoma cell proliferation, migration and invasion.
Figure 2. Epigenetic regulation of MMP-9 in melanoma. (A) Melanoma cells secrete exosomes containing different proteins and ncRNAs, including hsa-miR-155-5p. These exosomes are internalized by cells of the tumor microenvironment, such as fibroblasts. Inside the cells the exosomal hsa-miR-155-5p is able to inhibit SOCS1 mRNA thus activating the JAK/STAT molecular pathway. Phosphorylated STAT3 binds the DNA inducing the transcription of MMP-9, VEGFa and other pro-angiogenetic factors (e.g., FGF2). These proteins will be responsible for ECM degradation, fibroblasts switch to cancer-associated fibroblast phenotype and neoangiogenesis; (B) ncRNAs may act also as inhibitors of MMP-9 and MMP-2. In particular, hsa-miR-296-3p is able to inhibit both MMP-9 and MMP-2 mRNA thus reducing their secretion in the extracellular space. Similarly, FOXCUT lncRNA is able to inhibit MMP-2 and MMP-9 transcription and translation into proteins. Both ncRNAs are responsible of the inhibition of melanoma cell proliferation, migration and invasion.
Cells 09 01151 g002
Figure 3. Schematic representation of MMP-9 activation mediated by OPN. The binding between OPN and αvβ3 integrin induce the phosphorylation and activation of NIK protein. NIK is able to induce the activation of the protein complex NF-κB/p65/p50 through two different mechanisms: (1) NIK activates IKKα/β then IKKα/β phosphorylated is able to disrupt the binding between NF-κB/p65/p50 complex and IkBα, thus, activating NF-κB and inducing IkBα ubiquitination and degradation; (2) NIK is able to activate MEK1/ERK1/2 cascade through phosphorylation, phosphorylated ERK1/2 thus activate NF-κB by removing IkBα. Finally, NF-κB/p65/p50 activated complex moves within the nucleus inducing the transcription and production of Pro-MMP-9 and uPA. This latter protein is able to bind Pro-MMP-9 leading to the production of MMP-9 active form.
Figure 3. Schematic representation of MMP-9 activation mediated by OPN. The binding between OPN and αvβ3 integrin induce the phosphorylation and activation of NIK protein. NIK is able to induce the activation of the protein complex NF-κB/p65/p50 through two different mechanisms: (1) NIK activates IKKα/β then IKKα/β phosphorylated is able to disrupt the binding between NF-κB/p65/p50 complex and IkBα, thus, activating NF-κB and inducing IkBα ubiquitination and degradation; (2) NIK is able to activate MEK1/ERK1/2 cascade through phosphorylation, phosphorylated ERK1/2 thus activate NF-κB by removing IkBα. Finally, NF-κB/p65/p50 activated complex moves within the nucleus inducing the transcription and production of Pro-MMP-9 and uPA. This latter protein is able to bind Pro-MMP-9 leading to the production of MMP-9 active form.
Cells 09 01151 g003
Table 1. Classification of the MMP family members.
Table 1. Classification of the MMP family members.
Sub-FamilyGeneNameChr PositionCellular PositionMain Features [36,37]Refs
CollagenasesMMP-1Matrix Metallopeptidase 1 or
Interstitial collagenase
Chr 11 q22.2Extracellular spaceECM Substrates: Types I, II, and III, VII and X collagens.
Function: Cleaves HIV Tat protein decreasing Tat’s mediated neurotoxicity in HIV infections.
[39,40]
MMP-8Matrix Metallopeptidase 8 or
Neutrophil collagenase
Chr 11 q22.2Extracellular spaceECM Substrates: Fibrillar type I, II, and III collagens.[12]
MMP-13Matrix Metallopeptidase 13 or
Collagenase 3
Chr 11 q22.2Extracellular spaceECM Substrates: Fibrillar collagen, fibronectin, Aggrecan, Tenascin C, type I, III, IV, X, XIV collagens and especially soluble type II collagen.
Functions: embryonic development (TGFB and CCN2 degradation), wound healing, bone development and mineralization, cartilage degradation, etc.
[41]
GelatinasesMMP-2Matrix Metallopeptidase 2,
Gelatinase-A or 72 KDa
Type IV Collagenase
Chr 16 q12.2Extracellular spaceECM Substrates: Type I, II, III, IV, VII, X collagens and gelatin.
Functions: ECM degradation, angiogenesis, tumor invasion, tissue repair, atherosclerotic plaque rupture, regulating of myocardial functions (via GSK3beta cleavage).
[42]
MMP-9Matrix Metallopeptidase 9,
Gelatinase-B or 92 KDa
Type IV Collagenase
Chr 20 q13.12Extracellular spaceECM Substrates: Gelatin, fibronectin and type IV and V collagens.
Functions: Angiogenesis, tumor invasion, tissue repair, tissue remodeling, etc.
[43]
StromelysinsMMP-3Matrix Metallopeptidase 3 or
Stromelysin 1
Chr 11 q22.2Extracellular spaceECM Substrates: Fibronectin, laminin, types I, III, IV, and V gelatins and type III, IV, X, and IX collagens.
Functions: Degradation of cartilage proteoglycans and activation of procollagenase.
[44]
MMP-10Matrix Metallopeptidase 10 or
Stromelysin 2
Chr 11 q22.2Extracellular spaceECM Substrates: Fibronectin, elastin, type I, III, IV, and V gelatins.
Functions: Activation of protocollagenasse and weak action toward type III, IV, and V collagens.
[44]
MMP-11Matrix Metallopeptidase 11 or
Stromelysin 3
Chr 22 q11.23Extracellular spaceECM Substrates: Type IV collagen, fibronectin, laminin and aggrecan.
Function: Epithelial tumor invasiveness.
[45]
MatrilysinsMMP-7Matrix Metallopeptidase 7 or
Matrilysin
Chr 11 q22.2Extracellular spaceECM Substrates: Casein, types I, III, IV, and V gelatins and fibronectin.
Function: Activation of procollagenase.
[46]
MMP-26Matrix Metallopeptidase 26 or
Matrilysin-2
Chr 11p15.4Extracellular spaceECM Substrates: Fibronectin, fibrinogen, beta-casein, type IV collagen, type I gelatin and alpha-1 proteinase inhibitor.
Function: Activation of progelatinase B.
[47]
MetalloelastaseMMP-12Matrix Metallopeptidase 12 or
Macrophage metalloelastase
Chr 11 q22.2Extracellular spaceECM Substrates: Soluble and unsoluble elastin, type IV collagen and fibronectin.
Function: Tissue remodeling after injuries.
[48]
Membrane-type MMPsMMP-14Matrix Metallopeptidase 14 or
Membrane-Type-1
Matrix Metalloproteinase
Chr 14 q11.2Plasma membraneECM Substrates: Gelatin, fibronectin, laminin and collagen.
Functions: Activation of progelatinase A, skeletal and connective tissue development and remodeling, cell migration via pro-MMP-2 and MMP-15 binding.
[49]
MMP-15Matrix Metallopeptidase 15 or
Membrane-Type-2
Matrix Metalloproteinase
Chr 16 q21Plasma membraneECM Substrates: Gelatin, fibronectin, laminin and collagen.
Function: Activation of progelatinase A.
[50]
MMP-16Matrix Metallopeptidase 16 or
Membrane-Type-3
Matrix Metalloproteinase
Chr 8 q21.3Plasma membraneECM Substrates: Several types of collagens and fibronectin.
Functions: Activation of progelatinase A, angiogenesis and degradation and invasion of type I collagen by melanoma cells.
[44]
MMP-17Matrix Metallopeptidase 17 or
Membrane-Type-4
Matrix Metalloproteinase
Chr 12 q24.33Plasma membraneECM Substrates: Several ECM components, fibrinogen and fibrin.
Functions: Binding of growth factors and cytokines during inflammation, tumor progression via pro-TNF-alpha cleavage.
[44]
MMP-24Matrix Metallopeptidase 24 or
Membrane-Type-5
Matrix Metalloproteinase
Chr 20 q11.22Plasma membraneECM Substrates: Proteoglycanase, fibronectin, N-cadherin.
Functions: Induce degradation of dermatan sulfate and chondroitin sulfate proteoglycans, neurodevelopment and activation of progelatinase A.
[50]
MMP-25Matrix Metallopeptidase 25 or
Membrane-Type-6
Matrix Metalloproteinase
Chr 10 q26.2Extracellular space/Plasma membraneECM Substrates: Unknown
Function: It may activate progelatinase.
[51]
Other MMPsMMP-19Matrix Metallopeptidase 19,
Matrix Metallopeptidase RASI-1 or
Matrix Metallopeptidase 18
Chr 12 q13.2Extracellular spaceECM Substrates: collagen type IV, laminin, nidogen, nascin-C isoform, fibronectin, type I gelatin and aggrecan.
Function: Cartilage matrix degradation in physio-pathological conditions.
[52]
MMP-20Matrix Metallopeptidase 20 or
Enamelysin
Chr 11 q22.2Extracellular spaceECM Substrates: Amelogenin, aggrecan and cartilage oligomeric matrix protein (COMP).
Functions: Unknown
[53]
MMP-21Matrix Metallopeptidase 21Chr 10 q26.2Extracellular spaceECM Substrates: Alpha-1-antitrypsin
Function: Embryogenesis vis NOTCH-signaling pathway inhibition.
[54]
MMP-23AMatrix Metallopeptidase 23A or
Femalysin
Chr 1 p36.33ECM Substrates: Unknown.
Functions: Unknown.
[55]
MMP-23BMatrix Metallopeptidase 23B or
Femalysin
Chr 1 p36.33Extracellular spaceECM Substrates: Unknown.
Functions: Unknown.
[55]
MMP-27Matrix Metallopeptidase 27Chr 11 q22.2Endoplasmic reticulumECM Substrates: Fibronectin, laminin, gelatins and/or collagens.
Functions: Unknown.
[56]
MMP-28Matrix Metallopeptidase 28 or
Epilysin
Chr 17 q12Extracellular spaceSubstrates: Casein.
Functions: Tissues homeostasis and repair.
[57]

Share and Cite

MDPI and ACS Style

Napoli, S.; Scuderi, C.; Gattuso, G.; Di Bella, V.; Candido, S.; Basile, M.S.; Libra, M.; Falzone, L. Functional Roles of Matrix Metalloproteinases and Their Inhibitors in Melanoma. Cells 2020, 9, 1151. https://doi.org/10.3390/cells9051151

AMA Style

Napoli S, Scuderi C, Gattuso G, Di Bella V, Candido S, Basile MS, Libra M, Falzone L. Functional Roles of Matrix Metalloproteinases and Their Inhibitors in Melanoma. Cells. 2020; 9(5):1151. https://doi.org/10.3390/cells9051151

Chicago/Turabian Style

Napoli, Salvatore, Chiara Scuderi, Giuseppe Gattuso, Virginia Di Bella, Saverio Candido, Maria Sofia Basile, Massimo Libra, and Luca Falzone. 2020. "Functional Roles of Matrix Metalloproteinases and Their Inhibitors in Melanoma" Cells 9, no. 5: 1151. https://doi.org/10.3390/cells9051151

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