Open access peer-reviewed chapter

Deciphering and Targeting Epigenetics in Cancer Metastasis

Written By

Jie Huang, Aiping Lu and Chao Liang

Submitted: 07 June 2022 Reviewed: 15 July 2022 Published: 21 August 2022

DOI: 10.5772/intechopen.106584

From the Edited Volume

Cancer Metastasis - Molecular Mechanism and Clinical Therapy

Edited by Yusuf Tutar and Lütfi Tutar

Chapter metrics overview

148 Chapter Downloads

View Full Metrics

Abstract

Once cancer metastasizes to distant organs like the bone, liver, lung, and brain, it is in an advanced stage. Metastasis is a major contributor to cancer-associated deaths. Countless molecules and complex pathways are involved in the dissemination and colonization of cancer cells from a primary tumor at metastatic sites. Establishing the biological mechanisms of the metastatic process is crucial in finding open therapeutic windows for successful interventions. Emerging evidence suggested a variety of epigenetic regulations were identified to regulate cancer metastasis. Here we summarize the procedures and routes of cancer metastasis as well as the roles of epigenetics including ncRNA, DNA methylation, and histone modifications in common metastases. Then we further discuss the potentials and limitations of epigenetics-related target molecules in diagnosis, therapy, and prognosis.

Keywords

  • non-coding DNA
  • epigenetics
  • cancer metastasis
  • DNA methylation
  • histone modifications

1. Introduction

Cancer is the leading cause of death all over the world, accounting for nearly 10 million deaths in 2020 according to the World Health Organization (WHO). Metastatic cancer, the main contributor to high mortality, results in more than 90% of cancer death. This is because when cancer metastasizes to distant organs, especially the bone, liver, lung, and brain, this secondary tumor is formed. And this kind of tumor is difficult to remove despite the various systemic treatments including chemotherapy, screening, and immunotherapy. Efforts from doctors, researchers, and other aspects to promote cancer killing over the past years have paid off in some countries and in some cancers. Since 1991, the cancer death rate has fallen continuously in the United States. Up to 2018, the total mortality rate fell by 31%. Yet the mortality rate has been increasing in other places, such as China [1]. In most cases, many patients with metastatic cancer will face death within 5 years after their diagnosis, which is a horrible thing. Therefore, knowing the mechanism of cancer metastasis to treat metastasis is meaningful for patients, and is a challenging project for oncologists and clinical investigators.

Exploring the physiological mechanisms of the metastatic process is the foundation to find successful interventions. In the beginning, body fluids were thought to be responsible for tumor metastasis. In 1929, James Ewing proposed a theory that believed the anatomical structure of the vascular system contributes to metastasis and dissemination of cancer cells [2]. This view prevailed for decades. Nonetheless, the most classic and now popular is the “seed and soil” hypothesis proposed by Stephen Paget in 1889 [3]. Over the next few decades, cancer scientists gradually enhanced our knowledge of this mechanism based on molecular and cellular aspects. During this process of metastasis, countless molecules, and complex pathways, including epigenetic regulations, are involved in the dissemination and colonization of cancer cells from a primary tumor at metastatic sites. Epigenetics, a reversible process, refers to the study of heritable changes in gene expression without DNA sequence changes. Increasing studies of epigenetic regulation suggest that such regulations without altering the DNA sequence are critical for the normal physiological activities and the maintenance and development of tissue-specific gene expression in mammals [4]. The location of modified residue and the degree of methylation determines whether the transcriptional activation or repression. For example, the trimethylation of lysine 4 on histone H3 (H3K4me3) can be observed at the promoters of activated genes transcriptionally, yet trimethylation of H3K9 (H3K9me3) and H3K27 (H3K27me3) is enriched at repressed gene promoters transcriptionally [5].

Moreover, the importance of epigenetic changes in early tumorigenesis and cancer metastasis also has been shown, including non-coding RNA (ncRNA), DNA methylation, and histone modifications. Some such examples are increased N6-methyladenosine (m6A) modification of c-Myc mRNA enhances tumor cell growth, invasion, and tumorigenesis in animal models [6]. Upregulated Lysine Demethylase 6B (KDM6B) facilitates lung metastasis in osteosarcoma by modulating the H3K27me3 demethylation level of lactate dehydrogenase (LDHA) [7]. In addition, the enhancer of zeste homolog 2 (EZH2), the histone methyltransferase (HMT) of H3K27, is increased in cancers and promotes tumor metastasis [8, 9]. overexpressed long non-coding RNA (lncRNA) H19 enhances the migration of malignant cells and promotes the occurrence of epithelial to mesenchymal transition (EMT) in endometrial cancer [10].

Given the distinguished functions of epigenetics in cancer progression, and numerous crucial pathways and key biomarkers discovered by researchers, various potent and specific inhibitors targeting biomarkers have been studied and applied in clinics for treating cancer, since azacytidine, the first epigenetic drug approved by Food and Drug Administration (FDA) in 2004. In addition, as inhibitors of DNA methyltransferase (DNMT) enzymes (also termed hypomethylating agents), decitabine (5-aza-2′-deoxycytidine) and guadecitabin are the most extensively applied epigenetic therapies to kill various cancer cells, such as mutated monocyte in acute myeloid leukemia (AML) [11]. Several histone deacetylase (HDAC) inhibitors also have been extensively applied to anticancer (i.e., vorinostat, romidepsin, panobinostat, and belinostat), gaining the approval of FAD for hematological malignancies based on the activity of the single drug [12]. What’s more, histone methyltransferase (HMTs), like EZH2, protein arginine methyltransferase 6 (PRMT6), SET domain bifurcated 1 (SETDB1), SUV39H1, and disruptor of telomeric silencing 1-like (DOT1L), also are the targets of cancer treatment. Note that, several small-molecule inhibitors of EZH2 (i.e.,tazemetostat, SHR2554, MAK683) and DOT1L (i.e., EPZ-5676) have entered into clinic phases [13].

Based on the increasing knowledge about the mechanism of metastasis and drug development, the prognosis and survival in patients with cancer will gain an effective improvement in clinical outcomes. That is because using the vulnerabilities of metastatic cancer cells and the properties of metastatic tumor microenvironments are a great entry point to prevent cancer metastasis. As a result, several related drugs involved in cancer metastasis to treat cancer came into being. For instance, gefitinib and erlotinib are tyrosine kinase inhibitors (TKIs) that target activating epidermal growth factor receptor (EGFR) mutations and can improve overall survival by inhibiting metastasis in non-small cell lung cancer (NSCLC) [14, 15]. In general, cancer metastasizes to bone, liver, lung, and brain at an advanced stage, which is difficult for clinicians to destroy the secondary tumor. This is an urgent task and of great significance to patients. Thus, cancer biologists are working to deepen their understanding of epigenetic mechanisms in cancer metastasis to develop better therapy.

Although these drugs mentioned above contribute to clinical improvements in cancer patients, there exist some challenges. The first obstacle is that tumor heterogeneity, one of the characteristics of malignant tumors, which leads to the differences in immune characteristics, growth rate, aggressive ability, sensitivity to drugs, prognosis, and other phenotypic aspects after taking the same drugs. That means precision medicine and personalized medicine are the points of future medical development. Besides, the vast majority of genetic changes of epigenetics are inactivating mutations that are inherently difficult to treat, even though cancer biologists are designing drugs to interfere with adaptive mechanisms. Epigenetics provides a novel insight for researchers to improve the prognosis and survival of patients.

In this review, we summarize the procedures and routes of cancer metastasis as well as the roles of epigenetics including lncRNA, DNA methylation, and histone modifications in common metastases including bone, liver, lung, and brain, followed by discussing about potentials and limitations of epigenetics-related molecules in diagnosis, therapy, and prognosis.

Advertisement

2. Cancer metastasis

At present, metastasis is known as the result of a complex multistep cell-biological process collectively known as the invasion-metastasis cascade. It involves the changes of cancer cells in the physical position from the primary tumor to distant or adjacent sites of dissemination, and the colonization of the “seed cells” by adapting to the alien tissue microenvironments. More specifically, during metastatic progression, the first change is the cellular adhesion and morphology of cancer cells are reduced by EMT. Then, cancer cells improved the capability of invading the normal tissue surrounding (local invasion). Next, cancer cells make a way into (intravasation) and out of (extravasation) systemic circulation such as the lymphatic or circulatory system to land at distant sites. In this step, surviving cancer cells are termed circulating tumor cells. Lastly, the cancer cells proliferate and colonize in an unknown tissue microenvironment of different distant organs (Figure 1).

Figure 1.

Overview of metastatic Cascade. During metastatic progression, the first change is that the cellular adhesion and morphology of cancer cells are reduced by epithelial-mesenchymal transition (EMT). Then, the ability of cancer cells to invade the surrounding normal tissue (local invasion) is increased. Next, cancer cells make a way into (intravasation) and out of (extravasation) systemic circulation such as the lymphatic or circulatory system to land at a distant site. In this step, surviving cancer cells are termed circulating tumor cells. Lastly, the cancer cells proliferate and colonize an unknown tissue microenvironment of different distant organs. This figure was created with BioRender.com.

2.1 The invasion-metastasis cascade

2.1.1 Local invasion

The local invasion of cancer cells is the foundation for metastatic cancer process. Local invasion refers to the entry of cancer cells into the surrounding tumor-associated stroma, subsequently entering the adjacent normal parenchymal tissue. To invade the stroma, cancer cells must first break the basement membrane (BM) located at the interface of the epithelial tissue and connective tissue. The specialized extracellular matrix (ECM) can exchange material and regulate tissue growth, differentiation, and regeneration [16]. When these cancer cells invade the stroma, they have to be confronted with diverse cancer-associated stromal cells, including myofibroblasts, fibroblasts, adipocytes, endothelial cells, and plenty of bone marrow-derived cells (BMDCs) (i.e., macrophages, mesenchymal stem cells) [17]. Then, these stromal cells are able to enhance the aggressiveness of cancer cells through various cytokines. Secretion of interleukin-6 (IL-6) by cancer-associated fibroblasts (CAFs), stimulates the migration and invasiveness of colorectal cancer cells by the STAT3-LRG1 axis [18]. Increased IL-4 in endothelial cells can lead to enhanced invasiveness of liver cancer cells via the ERK-AKT signaling axis [19]. Besides, colorectal cancer cells secret IL-4 to promote M2-like tumor-associated macrophage (TAM) polarization [20]. These findings suggest there exists a positive-feedback loop in the tumor microenvironment, which is that cancer cells maintain high inflamed surroundings, and these stromal cells further enhance the malignant characteristics of cancer cells.

Researchers have observed various patterns of invasion when cancer cells infiltrate the substrates of adjacent tissues. Due to the dissemination of cancer cells, as individuals and collectives, these researchers divide migrations into individual cell migration and collective cell migration. Both types of migration are simultaneously present in many cancers [21]. In cancer progression, the plastic changes of numerous cancer cells are shown by morphological and phenotypical conversions, such as EMT and its reverse process the mesenchymal-epithelial transition (MET) [22], the collective-amoeboid transition (CAT) [23], the mesenchymal-amoeboid transition (MAT) [24]. Among these conversions, EMT has been increasingly considered a crucial and indispensable stage in the cancer metastatic process over the last decade [22], despite the studies of Seyfried et al. in the VM mouse model of systemic metastasis suggesting that EMT is unnecessary for the initial cancer metastasis [25, 26].

EMT is a cellular process activated by master transcription regulators, includingZEB1, ZEB2, Twist, Slug, and Snail, which enhance cell motility and migration ability to invade stroma. Besides, transforming growth factor (TGF)-β has proved to be a strong inducer of EMT by collaborating with other signaling pathways, especially the RAS-MAPK cascade [27]. Moreover, increasing emerging evidence shows the potent roles in invasion and EMT of many long noncoding RNAs (lncRNAs), such as lncRNA MEG3, lncRNA PNUTS, and lncRNA MIR100HG [28, 29, 30]. During this process, cells lost epithelial characteristics and markers like E-cadherin and cytokeratin, instead of gaining mesenchymal characteristics and markers like vimentin, N-cadherin, and fibronectin [22]. In addition to cancer metastasis, EMT has been involved in different cancer stages, including cancer initiation, malignant progression, cancer stemness, and drug resistance [31].

2.1.2 Intravasation

It is a vital and indispensable step for cancer cells to disseminate to distant organs during which cancer cells infiltrate into the vascular or lymphatic wall and then enter circulation, becoming circulating tumor cells (CTCs) and potential metastatic seeds. The formation of new blood vessels around cancer cells has a great influence on cancer cells entering the circulatory system, thus understanding the various mechanisms of neoangiogenesis stimulated by cancer cells in local microenvironment will help us comprehend intravasation. Vascular endothelial growth factor (VEGF), a highly bioactive functional glycoprotein, promotes blood vessel growth and lymphatic vessels, which plays an irreplaceable role in angiogenesis. However, the neo-vasculatures generated by cancer cells increase capillary permeability compared with the blood vessels produced by normal cells and tissues [32]. During the lung metastasis of breast cancer, VEGF/VEGF receptor 2 (VEGFR2) and its target proteins such as ERK1/2, Src, and FAK regulate neo-angiogenesis and blood vessel permeability to enhance metastasis [33].

On the other hand, a bunch of studies reveal intravasation can be improved by boosting the penetrability of cancer cells to pass the barrier of endothelial cells. For example, secretion of epidermal growth factor (EGF) by TAMs enhances the intravasation of breast cancer cells [34]. Additionally, the TGF-β enhances mammary cancer intravasation by increasing carcinoma cell penetration of micro-vessel walls or more generally strengthening invasiveness [35]. What’s more, in melanoma, the migration of cancer cells to endothelial cells and intravasation are promoted via endothelial-derived SLIT2 protein and its receptor ROBO1 [36]; activated Notch1 receptors (N1ICD) can promote neutrophil infiltration into the tumor, the intravasation of cancer cells and postsurgical metastasis [37]. In the study of Wei et al., increased IL-6 from TAMs is observed and can promote the invasiveness of cancer cells through the STAT3/miR-506-3p/FoxQ1 axis, then increases CCL2 level to boost the recruitment of macrophages. Besides, the authors suggested that there exists a feedback loop between TAMs and cancer cells, which was essential for the EMT and intravasation into the blood vessels [38].

2.1.3 Circulation

Once cancer cells have successfully entered lymph and blood, these malignant cells have the chance to disseminate throughout the body. In blood and lymphatic vessels, these cancer cells must escape the killing of immune cells and physical damage from hemodynamic shear forces to survive. In general, CTCs are in a dormant state that can cause relapse and poor prognosis for patients. This is because conventional surgery, radiotherapy, and chemotherapy are powerless against these CTCs in the blood, lymph, and body fluids, as well as dormant cancer cells, further leading to a decrease in immunity and the rapid growth and metastasis of hidden CTCs.

Many studies have verified the prognostic role and value of CTCs in the early and metastatic stages of cancer by measuring biomarkers [39]. An informative meta-analysis including 1847 patients with colorectal cancer under chemotherapy studied by Huang et al., demonstrated the high expression of CTCs in the bloodstream has a positive correlation with decreased progression-free survival (PFS) (hazard ratios = 2.500, 95% CI [1.746–3.580], P < 0.001) [40]. Moreover, CTCs in blood samples of 100 patients with head and neck squamous cell carcinoma were enriched and isolated and the PFS and overall survival of these patients were observed and recorded. The result showed a worse prognosis like decreased PFS and overall survival in CTCs-high patients [41]. Rink et al. also observed patients with ≥1CTCs C per 7.5 ml of blood in distant metastatic bladder cancer shortened the time of disease recurrence and cancer-specific death, resulting in worse clinical outcomes [42]. With the improvement of technologies and the depth of research, plenty of CTCs-related biomarkers are uncovered. At present, a set of biomarkers has been applied to detect CTCs in various cancers. Lin D et al. summarized the CTC-related biomarkers in different cancers [43]. EpCAM as the most common marker can be found in most cancer (i.e., breast cancer, liver cancer, prostate cancer, kidney cancer, melanoma, bladder cancer), which is because most cancers originate from the epithelium [44]. Just like EpCAM, human epidermal growth factor receptor-2 (HER-2), estrogen receptor (ER), prostate-specific membrane antigen (PSMA), and folate receptor (FR) also have been applied to detect CTCs in some cancers, with outstanding clinical significance [45, 46, 47, 48, 49].

In addition to CTCs-related biomarkers, the mechanism by which CTCs escape the detrimental shear stress and anoikis in the circulatory system is becoming clearer. There is evidence that CTCs in the blood can stay away from immune cells’ killing to increase survivability by bounding tightly to blood constituents like neutrophils, myeloid-derived suppressor cells (MDSCs), CAFs, or platelet [50, 51]. A few years ago, Szczerba et al. found the concentration of CTCs and neutrophils have a significant correlation in animal models and patients with breast cancer, which displays greater metastatic potential and higher gene expression involving cell proliferation. They thought the binding of CTC and neutrophil is possibly mediated by vascular cell adhesion molecule [52]. Besides, Spicer et al. suggested neutrophils could directly adhere to CTCs by the neutrophil Mac-1/ICAM-1, which becomes a bridge between cancer cells and the liver to accelerate CTCs extravasation and colonization [53]. Neutrophils can also enhance metastasis in an indirect manner by trapping CTCs in the circulation through neutrophil extracellular traps (NETs) [54]. In several in vivo experiments, liver or lung NETs were found to collect cancer cells to promote distant metastases by a transmembrane protein named coiled-coil domain containing 25 (CCDC25) to activate the ILK-β-parvin pathway, leading to enhance cell motility [55].

TAMs play crucial roles in the mechanical adhesiveness and endurance of CTCs, which contribute to the formation of protective cell clusters and the resistance to shear stress [56]. Liu et al. proposed that CTCs interacting with adhesive immune cells like MDSCs could create a defensive shield to allow evasion of immune surveillance, facilitating distant metastatic lesions [57]. Sprouse et al. found reactive oxygen species (ROS) from MDSCs could activate the Notch pathway in CTCs, promoting CTCs proliferation [58]. In addition, CAFs could protect CTCs from the fluid shear forces in the peripheral blood via intercellular contact and soluble derived factors in prostate cancer [59]. As an important component in blood, platelet also supports the survival and metastasis of CTCs in a CTCs-platelet cluster manner. Platelets have been shown to help CTCs evade attack by NK cells by creating a surface shield and normal MHC-I [60], or by downregulating natural killer group 2 member D (NKG2D) and its ligands, further stimulating glucocorticoid-induced tumor necrosis factor receptor-related protein (GITR) to exert functions in NK cells [61, 62, 63]. Furthermore, platelets involve the adhesion process of endothelial cells. The attachment between platelets and CTCs is enhanced by integrin αIIbβ3 and P-selectin (platelet adhesion receptors), in which supports the strong adherence of CTCs to the endothelial wall [64, 65, 66].

2.1.4 Extravasation

Cancer cells extravasate from a vascular lumen into tissues such as the lung, liver, and brain by passing through the endothelial cell and pericyte layers. Extravasation is comparable to intravasation in that it is morphologically similar to invadopodia but mechanistically different. Certain cell types in the primary tumor microenvironment, such as TAMs [34], can initiate intravasation, but these same cells do not have the same promoting function in the extravasation process of disseminated CTCs. Indeed, macrophage phenotype and function differ between primary and metastatic tumor locations. For example, macrophage seeding at distant places is VEGFR+, CCR2+, CXCR4-, Tie2-, and the subpopulation of macrophage at perivascular macrophage is phagocytic [67].

CTCs must overcome the physical barriers of the microvascular wall to extravasate. According to some studies, primary tumors have been shown to release substances that interfere with these distant microenvironments and cause vascular hyperpermeability. Secreted protein angiopoietin-like-4 (Angptl4), as well as the pleiotropically active proteins like NOX4, MMP-1, and MMP-9, disrupt pulmonary vascular endothelial cell-cell junctions, allowing colorectal cancer cells to extravasate into the lungs [68]. Angiopoietin2 (Angpt2), MMP-3, MMP-10, placental growth factor, and VEGF, all of which are secreted by many types of primary tumors, can induce pulmonary hyperpermeability prior to the arrival of cancer cells in the lungs, allowing CTCs to extravasate more easily [69, 70]. Finally, by secreting VEGF, inflammatory monocytes recruited to pulmonary metastases via CCL2-dependent processes increase breast cancer cell extravasation in the lung [71].

Interestingly, whereas Anglptl4 improved the extravasation of breast carcinoma cells in the lung, it did not increase the extravasation or intravasation efficiency of these same breast cancer cells in the bone [72]. As a result, Anglptl4 selectively and only increases extravasation within the lung tissue environment.

2.1.5 Colonization and metastatic growth

Cell-nonautonomous mechanisms required to transform a foreign microenvironment into a more friendly niche may be required for disseminated tumor cells to emerge from hibernation and begin active proliferation. For example, the growth of other inactivated disseminated tumor cells might need to stimulate BMDCs to enter the circulation system, as well as the followed recruitment of these cells to the metastatic location; in some situations, the process may be activated through systemic signals such as osteopontin (OPN) or SDF-1 produced by cancer cells, [73, 74].

Alternatively, because the body is in a constant state of homeostasis, dormant cancer cells could continue to proliferate without a net increase. The reasons driving such high rates of attrition are unknown, however, a lack of disseminated tumor cells to initiate neoangiogenesis has been hypothesized as one possible explanation. Prostate tumor cell-secreted prosaposin (Psap) may limit metastatic colonization by increasing the expression of the anti-angiogenic factor thrombo-spondin-1 in stromal cells, which is consistent with this theory [75]. Angpt2, on the other hand, promotes the metastatic colonization of breast and pancreatic cancer by improving the infiltrating capability of myeloid cells to support the vascularization of metastatic nodules [76].

Numerous genes promote the metastatic colonization of cells in breast cancer to bone, lung, brain, or liver, which have recently been discovered. These genes are able to adapt and overcome incompatibilities between the special development procedure of disseminated cancer cells and the demands from foreign tissue milieu, in parallel, researchers come up with the idea that these genes could control organ-specific metastatic tropism. The osteoclastic cytokine IL-11 is an excellent example of this, IL-11 works through a receptor activator for nuclear factor kB (RANK), which disrupts the normal crosstalk between osteoclasts and osteoblasts [77]. Moreover, it strengthens metastatic tumor growth in breast cancer and osteolysis by JAK1/STAT3/c-myc signaling pathway rather than in a RANKL-dependent manner [78].

Similarly, the Notch ligand Jagged1 enhances the osteolytic bone metastases in breast cancer cells by boosting osteoclast activity through IL-6 released by osteoblasts [79]. By encouraging osteoclast action, IL-11 and Jagged1 are able to cause osteolysis and release the rich deposits of growth factors from the bone matrix. The fact that genes identified as candidate mediators of breast cancer cell metastatic colonization in bone, lung, brain, or liver show very little overlap, illustrates the idea that different tissue microenvironments are needed to be organ-specific for metastatic colonization.

2.2 Routes of cancer metastasis

2.2.1 The circulatory system

Despite the fact that lymphatic diffusion of cancer cells is a key prognostic marker for cancer progression, spreading through the blood circulation seems to be the main mechanism of dispersal of metastatic carcinoma cells. Based on intravital imaging studies, tumor cells can travel toward blood arteries. Li et al. injected metastasized breast cancer cells in mice and found that these cells move toward arteries, illustrating that metastatic cells have the ability of directional migration toward blood streams [80]. Morphologically, compared with non-metastatic cells, metastatic cells are more round, and this kind of morphology boost both their ability to spread, enter, and colonic tumor vasculature [81, 82]. These findings show that tumor cells with an elongated morphology may need to change their shape to become more rounded in order to successfully intravasate and endure shear forces within blood arteries. It’s tempting to think that higher cortical acto–myosin contraction, which promotes the rounded morphology, also allows the cortical cytoskeleton to withstand more mechanical stress. Tumor cells have been found to form part of blood vessel walls in other imaging studies. GFP-labeled tumor cells have been shown to constitute part of the lumen of blood arteries using in situ imaging [80]. This behavior is likely to be linked to tumor cells expressing genes that are normally restricted to endothelial cells. In many circumstances, entering the bloodstream could be as simple as separating tumor cells from the walls of blood vessels.

2.2.2 The lymphatic system

Tumor cell entry into the lymphatic system is another major mechanism for tumor cell dissemination. Even in the early phases of tumor formation, changes in lymphatic artery architecture have been seen, under the aid of VEGFC lymphangiogensis can be established quickly using these cells [83]. The lymphatic vasculature can be examined and the interaction of tumor cells with lymphatics can be seen in live tissue by injecting dyes or fluorescent tracers [84]. Cancer cells may extend protrusions via holes in lymph vessel walls before entering the vessels, according to electron microscope photos [85, 86]. To learn more about the cytoskeletal architecture of intravasation cells and their interactions with the lymphatic endothelium, time-lapse imaging should be possible. Interstitial pressure within the tumor affects lymphatic outflow [87]. Interstitial flow can produce autocrine gradients that signal through CC chemokine receptor 7 (CCR7) to induce cell migration in the same direction as the interstitial flow, according to in vitro research. In vivo, however, it is uncertain whether lymphatic channel movement is influenced by interstitial pressure. Trapped breast cancer cells can be discovered in the subcapsular region of lymph nodes after entering lymphatic channels [88]. However, clinically, lymph node metastases might stay in this place or spread to other parts of the node. The admission of tumor cells into the central sections of lymph nodes, as well as their interactions with immune cells, has yet to be captured in high resolution, but when it is, it will likely provide fascinating discoveries.

Advertisement

3. Epigenetics in cancer metastasis

Due to the devilishness of cancer metastasis, understanding how cancer cells acquire and maintain metastatic characteristics is critical. However, metastasis-specific genetic alterations cannot be discovered in most exome or genome sequencing investigations. Reversible epigenetic pathways control important phases in metastasis, which can be targeted to prevent and treat metastatic illness in increasingly emerging data. ncRNA, DNA methylation, and histone changes are only a few of the epigenetic processes that have been discovered to modulate the cancer metastasis process. Large-scale chromatin structural changes, such as enhancer reprogramming and chromatin accessibility to transcription factors, have been revealed to be a possible driving force of cancer metastasis in diverse malignancies in recent years. Given that numerous researchers have reviewed the function of epigenetic markers in different stages of metastasis [89, 90, 91], we will concentrate on well-defined particular metastatic locations such as bone, liver, lung, and brain in various malignancies over the last 5 years.

3.1 Epigenetics in bone metastasis

Bone is one of the most common sites of metastasis for a variety of solid tumors, including lung, liver, breast, and prostate, with bone metastases being seen in 70% of metastatic prostate and breast cancer patients [92]. Unfortunately, once cancer has progressed to the bone, it is seldom treated and is associated with countless complications such as discomfort, increased fracture risk, and hypercalcemia. This finding has prompted scientists interested in bone and cancer biology to investigate the bone, revealing a number of mechanisms, including epigenetically related elements that promote cancer spread to the bone (Table 1).

Cancer typesEpigenetic typesBiomarkersPathwaysReferences
Lung cancerncRNAlncRNA-SOX2OTmiRNA-194-5p/RAC1 signaling axis[93]
miRNA-660-5pmiR-660-5p/SMARCA5/RANKL axis[94]
miR-574-5p, miR-328-3p, and miR-423-3pWnt/β-catenin pathway[95]
miR-106amiR-106a/TP53INP1[96]
miR-139-5pmiR-139-5p/Notch 1[97]
miR-17-5pPTEN/PI3K/Akt[98]
miR-365NKX2-1/EGFR/PI3K[99]
miR-192-5pmiR-192-5p/ TRIM-44[100]
miR-886-3pmiR-886-3p-PLK1/TGF-β1 pathway[101]
DNA methylationDNMTsmiR-886-3p-PLK1/TGF-β1 pathway[101]
DNMTsWIF-1[102]
Prostate cancerncRNAlncRNA NEAT1CYCLINL1/CDK19/NEAT1-1[103]
miR-940miR-940/ARHGAP1 and FAM134A[104]
miR-181b-5pmiR-181b-5p/ Oncostatin M axis[105]
lncRNA HOXA11-ASHOXB13/lncRNA HOXA11-AS/ IBSP[106]
lncRNA PCAT6PCAT6/IGF2BP2/IGF1R axis[107]
lncRNA NEAT1miR-205-5p/RUNX2/ SFPQ/PTBP2 axis[108]
miR-378a-3pmiR-378a-3p/ Dyrk1a/Nfatc1/Angptl2[109]
lncRNA MAYAROR1-HER3-lnc RNA MAYA/Hippo-YAP pathway[110]
miR-214miR-214/TRAF3[111]
miR-124miR-124/IL-11 axis[112]
miR-218miR-218/COL1A1[113]
miR-125bmiR-125b/ HIF1A/PTGS2[114]
miR-429miR-429/CrkL/MMP-9[115]
miR-21miR-21/PDCD4[116]
miR-19amiR-19a/IBSP[117]
lncRNA SNHG3miR-1273 g-3p/BMP3 axis[118]
DNA methylationDNMTsHGF/Met receptor signaling; E-cadherin, Twist transactivation[119]
Liver cancerncRNAlncRNA 34aPHB2/DNMT3a/ miR-34a; TGF-β/Smad4 pathway[120]
lncRNA H19H19/p38MPAK/OPG; H19/miR200b-3p/ZEB1[121]

Table 1.

Epigenetic biomarkers in bone metastases of various cancers.

Notes: ncRNA: non-coding RNA; TP53INP1: tumor protein 53-induced nuclear protein 1; NKX2-1: NKX homeobox-1, EGFR: epidermal growth factor receptor, PI3K: phosphoinositide-3-kinase; TRIM-44: tripartite motif-44; PHB2: Prohibitin 2; TGF-β: transforming growth factor β; MPAK: mitogen-activated protein kinase; OPG: osteoprotegerin; ROR1: receptor tyrosine kinase (RTKs)-like orphan receptor-1; YAP: yes-associated protein; TRAF-3: TNF receptor-associated factor-3; COL1A1: collagen type I alpha 1 chain; PTGS2: prostaglandin-endoperoxide synthase 2; HIF1A: hypoxia-inducible factor 1 alpha subunit; CrkL: v-crk avian sarcoma virus CT10 oncogene homolog-like; PDCD4: programmed cell death 4; IBSP: integrin-binding sialoprotein; RUNX2: runt-related transcription factor 2; SFPQ: splicing factor proline- and glutamine-rich; and PTBP2: polypyrimidine tract-binding protein 2.

3.1.1 Lung cancer

In the last year, Ni et al. extracted exosomes from the plasma of non-small cell lung cancer (NSCLC) patients with or without bone metastasis. They found exosomal lncRNA-SOX2OT enhanced bone metastasis of NSCLC by targeting the miRNA-194-5p/RAC1 signaling axis in osteoclasts [93]. In organ-specific metastatic lung cancer cells, Ai et al. observed miR-660-5p involved tumor progression and bone-specific metastasis by nm23-H1/miR-660-5p/SMARCA5/RANKL axis [94]. Yang et al. identified an exosomal microRNA cluster that has an association with bone metastasis. Specifically, in this cluster miR-574-5p was down-regulated, miR-328-3p and miR-423-3p were up-regulated in patients with bone metastasis, which suppressed or activated the Wnt/β-catenin pathway [95]. Han et al. observed that upregulated miR-106a promoted bone metastasis by targeting tumor protein 53-induced nuclear protein 1 (TP53INP1), including cell migration, death, and EMT [96]. Xu et al. found miR-139-5p was downregulated in serum to facilitate lytic bone metastasis by targeting Notch1 [97]. In addition, miR-17-5p promotes osteoclastogenesis through the PI3K/Akt pathway via targeting PTEN [98]. Liu et al. revealed that miR-365 was reduced in patients with bone metastasis of NSCLC, and miR-365 could suppress lung metastasis via NKX2-1/EGFR/PI3K axis [99]. Zou et al. demonstrated that increased miR-192-5p in patient serum inhibited lung cancer metastasis, possibly by reducing TRIM44 [100]. Loss of miR-886-3p expression was mediated by DNA hypermethylation of its promoter in both cultured small cell lung cancer (SCLC) cells and tumor samples. What’s more, upregulated miR-886-3p greatly inhibited bone metastasis [101]. The downregulation of Wnt inhibitory factor 1 (WIF-1) expression was linked to hypermethylation of its promoter, which increased lung metastasis [102].

3.1.2 Liver cancer

In liver cancer bone metastasis, Zhang et al. revealed the molecular function of lncRNA 34a regulated bone metastasis. Mechanistically, lncRNA 34a epigenetically suppressed miR-34a level via the recruitment of DNMT3a by Prohibitin 2 (PHB2) to methylate miR-34a promoter and histone deacetylase (HDAC) 1 to promote histones deacetylation. On the other hand, miR-34a regulated Smad4 through the transforming growth factor-β (TGF-β) pathway, impacting the downstream genes (i.e., connective tissue growth factor (CTGF) and IL-11) associated with bone metastasis [120]. Huang et al. identified lnRNA H19/p38 mitogen-activated protein kinase (MPAK)/ osteoprotegerin (OPG) and lncRNA H19/miR200b-3p/ZEB1 axes contributed to hepatocellular carcinoma bone metastasis [121].

3.1.3 Breast cancer

In breast cancer, the Hippo-YAP pathway was controlled by a ROR1-HER3-LncRNA signaling axis to govern bone metastases [110]. Both osteoclastic miR-214/ TNF receptor-associated factor-3 (TRAF-3) pathway and dysregulated miR-124/IL-11 axis were devoted to the understanding of breast cancer metastases to the bone [111, 112]. A study pointed to a concept in which cancer-derived miR-218 impairs osteoblast function by directly targeting collagen type I alpha 1 chain (COL1A1) and regulating inhibin βA expression [113]. miR-125b may reduce the effect of hypoxia-inducible factor 1 alpha subunit (HIF1A), which is known to enhance metastatic spread by upregulating prostaglandin-endoperoxide synthase 2 (PTGS2) [114]. To investigate the influence of miR-429 on the metastatic bone environment in vivo, Zhang et al. created an orthotopic bone degradation model and a left ventricle implantation paradigm. The levels of V-crk sarcoma virus CT10 oncogene homolog-like (CrkL) and MMP-9 were negatively influenced by miR-429 [115]. Exosomal miR-21 generates from breast cancer cells promotes osteoclastogenesis by modulating the levels of the protein programmed cell death 4 (PDCD4). Furthermore, the amount of miR-21 in breast cancer patients with bone metastases is considerably greater in serum exosomes [116]. Exosomal miR-19a and integrin-binding sialoprotein (IBSP) are highly increased and secreted from bone-tropic estrogen receptor-positive (ER+) breast cancer cells, resulting in a milieu conducive to colonization in the bone [117]. Teng et al. identified many key lncRNAs such as lncRNA RP11-317-J19.1 related to bone metastasis in breast cancer [122]. By influencing the miR-1273 g-3p/BMP3 axis, LncRNA SNHG3 regulates BMSC osteogenic development in breast cancer bone metastases [118].

In breast cancer, the level of DNA methylation is increased to further regulate Wwox, following to stimulate HGF/Met receptor signaling and E-cadherin, downregulating Twist transactivation, leading to bone metastasis [119].

3.1.4 Prostate cancer

Wen et al. analyzed the m6A status using patient samples and bone metastatic patient-derived xenografts (PDXs) with prostate cancer through m6A high-throughput sequencing, and they found 4 credible m6A sites on lncRNA NEAT1-1. Besides, NEAT1-1 acted as a bridge to strengthen the combination between CYCLINL1 and CDK19 and promoted the Pol II ser2 phosphorylation in the promoter of RUNX2, leading to the development of bone metastatic prostate cancer [103]. By targeting ARHGAP1 and FAM134A, miR-940 boosted osteogenic differentiation of human mesenchymal stem cells [104]. miR-181b/Oncostatin m axis also contributes to prostate cancer bone metastasis by altering osteoclast differentiation [105]. To promote the bone-specific metastasis of prostate cancer, HOXA11-AS controlled the expression of chemokines, integrins, and associated genes like IBSP in collaboration with HOXB13 [106]. lncRNA PCAT6 enhances prostate cancer bone metastasis and tumor growth by upregulating IGF1R expression via increasing IGF1R mRNA stability through the PCAT6/IGF2BP2/IGF1R pathway [107]. lncRNA NEAT1/miR-205-5p/RUNX2/SFPQ/PTBP2 axis and miR-378a-3p/ Dyrk1a/Nfatc1/Angptl2 axis are also devoted to bone metastasis [108, 109].

3.2 Epigenetics in liver metastasis

3.2.1 Esophageal squamous cell carcinoma

Tang et al. explored the function of lncRNA LOC146880 in esophageal squamous cell carcinoma (ESCC) progression. The result of in vivo and in vitro experiments showed LOC146880 sponged miR-328-5p to regulate fascin actin-bundling protein 1 (FSCN1) activating MAPK signaling pathway, resulting in liver metastasis [123] (Table 2).

Cancer typesEpigenetic typesBiomarkersPathwaysReferences
Liver metastases
ESCCncRNAlncRNA LOC146880LOC146880/miR-328-5p/FSCN1/MAPK axis[123]
Breast cancerncRNAmiR-190ZEB1-miR-190-SMAD2 axis[124]
miR-1204miR-1204/VDR[125]
cirRNA ciRS-7ciRS-7/miR-1299[126]
circROBO1circROBO1/ KLF5/FUS[127]
Colorectal cancerncRNAmiR-221; miR-222[128]
Brain metastases
Breast cancerncRNAlncRNA XISTEMT and MSN/c-Met[129]
miR-10b[130]
miR-576-3p[131]
lncRNA BCBMlncRNA BCBM /JAK2/STAT3[132]
circBCBM1circBCBM1/miR-125a/BRD4 axis[133]
lncRNA-CCRRlncRNA-CCRR/connexin 43[134]
miRNA let-7dPDGF/PDGFR axis[135]
miR-802-5p; miR-194-5p[136]
miR-132-3p; miR-199a-5p; miR-150-5p; miR-155-5p[137]
miR-211SOX11/NGN2 axis[138]

Table 2.

Epigenetic biomarkers in liver and brain metastases of various cancers.

Notes: ESCC: esophageal squamous cell carcinoma; FSCN1: fascin actin-bundling protein 1; VDR: vitamin D receptor; KLF5: Kruppel like factor 5; bHLH transcription factor 2; CDK6: Cyclin Dependent Kinase-6; STAT3: signal transducer and activator of transcription-3; HDAC: histone deacetylase 1; ZEB1: Zinc Finger E-Box Binding Homeobox 1; ROCK1: Rho associated coiled-coil containing protein kinase 1; CRYAB: αB-crystallin; DNMTs: DNA methyltransferase; ERα: estrogen receptor alpha; TET2: ten-eleven translocation 2; IRX1: iroquois homeobox 1; PDK1: phosphoinositide-dependent kinase-1; ST7L: suppression of tumorigenicity 7 like; and BRD4: bromodomain containing 4.

3.2.2 Breast cancer

In breast cancer, the TGF network in liver metastasis can be explained by the ZEB1-miR-190-SMAD2 axis [124]. miR-1204 inhibits vitamin D receptors (VDR), which promotes epithelial-mesenchymal transition and metastasis [125]. As a ceRNA of miR-1299, circular RNA ciRS-7 promotes lung and liver metastases by targeting MMPs [126]. Wang et al. found circROBO1 was upregulated to boost tumor development and liver metastasis in vivo. Further research revealed the mechanism that circROBO1 upregulated KLF5 by sponging miR-217-5p, allowing KLF5 to activate FUS transcription, hence promoting circROBO1 back splicing [127].

3.3 Epigenetics in lung metastasis

3.3.1 Osteosarcoma

LncRNA-CASC15 promotes lung metastasis in osteosarcoma by regulating EMT via the Wnt/β-catenin signaling pathway [139] (Table 3). MIR205HG also can drive lung metastatic osteosarcoma via regulating the axis of miR-2114-3p/twist family bHLH transcription factor 2 (TWIST2) [140]. miR-485-3p regulated by lncRNA MALAT1 inhibites osteosarcoma glycolysis and lung metastasis by directly suppressing c-MET and AKT3/mTOR signaling, meanwhile, MALAT1 also facilitated lung metastasis of osteosarcomas through miR-202 sponging [141, 142]. Besides, Chen et al. also observed that the LOC100129620/miR-335-3p/CDK6 signaling promoted the lung metastasis of osteosarcoma by mediating the osteosarcoma cells proliferation, macrophage polarization, and angiogenesis [144]. The lncRNA NEAT1/miR-483/STAT3 axis also exerts a crucial role in regulating the lung metastasis process in osteosarcoma, especially in EMT [145]. miR-326 inhibited by SP1/HDAC1 has a great impact on proliferation and metastasis of osteosarcoma through stimulating SMO/Hedgehog pathway [146]. The miR-19a/RhoB/AKT1 network and miR-491/αB-crystallin (CRYAB) axis also may help us to better know the lung metastatic mechanism of osteosarcoma [147, 149]. In Ewing sarcoma, miR-130b directly targets ARHGAP1 to activate a lung metastatic CDC42-PAK1-AP1 positive feedback loop [157].

Cancer typesEpigenetic typesBiomarkersPathwaysReferences
OsteosarcomancRNAlncRNA-CASC15Wnt/β-catenin signaling[139]
MIR205HGMIR205HG/miR-2114-3p/TWIST2 axis[140]
lncRNA MALAT1miR-485-3p/c-MET; miR-485-3p /AKT3/mTOR signaling; lncRNA MALAT1/miR-202; miR-129-5p/RET/ PI3K-Akt axis[141, 142, 143]
lncRNA LOC100129620LOC100129620/miR-335-3p/CDK6 signaling[144]
lncRNA NEAT1lncRNA NEAT1/miR-483/STAT3 axis[145]
miR-326Sp1/HDAC1/miR-326/SMO/Hedgehog axis[146]
miR-19amiR-19a/RhoB/AKT1[147]
lncRNA DANCRmiR-335-5p and miR-1972/ROCK1[148]
miR-491miR-491/CRYAB[149]
DNA methylationDNMTsERα[150]
DNMTsSPARCL1/ WNT/β-catenin signaling[151]
DNA demethylationTET2/IL-6[152]
IRX1/ CXCL14/NF-κB signaling[153]
Breast cancerncRNAlinc-ZNF469-3miR-574-5p-ZEB1 axis[154]
cirRNA ciRS-7ciRS-7/ miR-1299[126]
lncRNA MIR31HGmiR-575/ ST7L[155]
Colorectal cancerHistone acetylationCBPCBP-DOT1L/ RNF8/H3K79[156]
Ewing SarcomancRNAmiR-130bmiR-130b-AP-1/ CDC42-PAK1-AP1 axis[157]
Gastric cancerncRNAlncRNA GMANephrin A1[158]
Lnc RNA MIR17HGWnt/β-catenin signaling[159]
lncRNAs AC093818.1PDK1[160]

Table 3.

Epigenetic biomarkers in lung metastases of various cancers.

Notes: TWIST2: twist family bHLH transcription factor 2; CDK6: Cyclin Dependent Kinase-6; STAT3: signal transducer and activator of transcription-3; HDAC: histone deacetylase 1; ZEB1: Zinc Finger E-Box Binding Homeobox 1; ROCK1: Rho associated coiled-coil containing protein kinase 1; CRYAB: αB-crystallin; DNMTs: DNA methyltransferase; ERα: estrogen receptor alpha; TET2: ten-eleven translocation 2; IRX1: iroquois homeobox 1; PDK1: phosphoinositide-dependent kinase-1; and ST7L: suppression of tumorigenicity 7 like.

Additionally, Lillo et al. found estrogen receptor alpha (ERα) was not expressed in osteosarcoma due to promoter DNA methylation. They took Decitabine, a DNA methyltransferase (DNMTs) inhibitor to activate ERα, further inhibiting osteosarcoma growth and lung metastasis [150]. In primary osteosarcoma cells, increased IL-6 expression regulated by DNA demethylation of the promoter of ten-eleven translocation 2 (TET2) promotes lung metastasis in osteosarcoma [152]. Secreted protein acidic and rich in cysteine (SPARCL1) downregulated by epigenetic promoter DNA methylation in osteosarcoma promotes lung metastasis via canonical WNT/β-catenin signaling activated through stabilization of the WNT–receptor complex [151]. Hypomethylation of iroquois homeobox 1 (IRX1) in osteosarcoma cell lines substantially affected metastatic behavior in vitro, including migration, invasion, and resistance to anoikis and influenced lung metastasis in animal models by upregulating CXCL14/NF-B signaling, according to another study [153].

3.3.2 Breast cancer

In the study by Wang et al. they showed that linc-ZNF469-3 accelerated lung metastasis of triple-negative breast cancer (TNBC) via miR-574-5p-ZEB1, which may be acted as a potential and promising prognostic marker for TNBC patients [154]. MIR31HG, a long noncoding RNA that sponges miRNA-575 to control ST7L expression, suppresses hepatocellular carcinoma proliferation and metastasis [155].

3.3.3 Colorectal cancer

Colostrum basic protein (CBP), a histone acetyltransferase (HAT), mediates DOT1L K358 acetylation and has a positive correlation with colorectal cancer stages. DOT1L acetylation confers DOT1L stability by blocking RNF8 binding to the protein and subsequent proteasomal degradation, but it has no effect on the enzyme’s activity. DOT1L can catalyze the H3K79 methylation of genes involved in epithelial-mesenchymal transition, such as SNAIL and ZEB1, once stabilized [156].

3.3.4 Gastric cancer

GMAN, a long non-coding RNA, is upregulated in stomach cancer patients and is linked to overall survival and metastasis process. It inhibits the translation of ephrin-A1 mRNA by binding to GMAN-AS in a competitive manner [158]. Interferon regulatory factor-1 (IRF-1) suppresses gastric cancer spread by suppressing Wnt/−catenin signaling and downregulating the MIR17HG-miR-18a/miR-19a axis to inhibit gastric cancer lung metastasis [159].

3.4 Epigenetics in brain metastasis

Through activation of EMT- and MSN-mediated up-regulation of c-Met, the loss of lncRNA XIST increases breast cancer brain metastasis by boosting both stemness and aggressiveness of tumor cells [129] (Table 2). Yoo et al. proved the therapeutics function of miRNA-10b by targeting the brain metastases process in breast cancer [130]. JAK2-binding lncRNA BCBM promotes breast cancer brain metastasis by regulating STAT3 [132]. circBCBM1 is involved in breast cancer brain metastasis via circBCBM1/miR-125a/BRD4 axis [133]. By modulating connexin 43 expression, dysregulation of lncRNA-CCRR contributes to breast cancer brain metastases through intercellular coupling [134]. Loss of miRNA let-7d and active hypoxia- inducible factor-1 (HIF1) signaling enhances breast cancer brain metastasis via platelet-derived growth factor (PDGF), while pharmacologic inhibition of PDGF receptor (PDGFR) inhibits brain metastasis, implying new therapeutic possibilities [135]. miR-132-3p, miR-199a-5p, miR-150-5p, miR-155-5p, miR-802-5p and miR-194-5p from breast cancer cells also were identified the important role in brain metastasis [136, 137]. In triple-negative breast cancer, miR-211 regulates brain metastatic selectivity via the SOX11/NGN2 axis [138].

Advertisement

4. Therapeutic potentials and limitations

Epigenetic drugs are chemicals that alter DNA and chromatin structure, promoting the disruption of transcriptional and post-transcriptional modifications, primarily by regulating the enzymes required for their establishment and maintenance, and reactivating epigenetically silenced tumor-suppressor and DNA repair genes [161]. The development of treatment techniques incorporating epigenetic medicines, which focus on the cancer epigenome to generate pharmacological molecules that could restore a “normal” epigenetic landscape, is a developing field of drug discovery [161]. Epigenetic medicines target the enzymes that are required for the maintenance and establishment of epigenetic alterations, with the inhibition of DNMTs and HDACs being the most common technique [161]. The epigenetic alterations caused by these medications can regulate the temporal and spatial expression of genes [162], and they have ramifications for the regulation and dysregulation of physiological and pathological processes. Because epigenetic markings are tightly linked to the type of tumor and stage of disease, as well as individual genetic variation, such as in personalized medicine [163, 164], they have a lot of promise to give molecular biomarkers for diagnosis and treatment alternatives for cancer therapy [165].

The FDA has approved six new epigenetic medicines and multi-drug regimens for use in clinical cancer treatment. Some side effects will happen, so novel epigenetic therapeutic compounds are continually being tested in preclinical research, as well as clinical trials for the development and release of new medicines, for cytotoxicity, and pharmacological characteristics, and to better understand their mechanism of action. The majority of epigenetic medication studies are focused on cancer detection, therapy, and prognosis.

NcRNAs have shown new promise and insight as therapeutic targets for cancer treatment and preventing cancer metastasis in vivo preclinical models of metastatic illness. Research has shown that lncMAYA, MALAT1, and lncARSR have all been targeted for in vivo suppression using ASOs in mice models to alleviate the burden of metastatic disease [110, 166, 167, 168]. When targeting lncRNAs with ASO therapies, however, it will be vital to ensure minimal off-target effects [169, 170], which could offer additional challenges given the decreased quantity of lncRNA transcripts in vivo. ASOs disrupt target RNAs by premature transcriptional termination [171, 172], in addition to RNase H-mediated destruction of mature RNA, according to new findings, which should be taken into account when estimating the efficacy of ASO therapies. The biggest problem is the species conservatism in ncRNA, especially lncRNAs. Animal testing is also required before conducting clinical trials.

Advertisement

5. Conclusions

Cancer metastasis is a common cause of death. The role of epigenetics in the etiology of metastases cannot be ignored. To overcome cancer and its metastasis, many methods and technologies are applied to developing new drugs. In parallel with the development of specific and potent small-molecule inhibitors, some novel and cutting-edge technologies like proteolysis-targeting chimeras (PROTACs), RNA interference (RNAi), clustered regularly interspaced short palindromic repeats (CRISPR)/Cas9-based genome editing, and artificial intelligence (AI)-based drug design, in chemical biology show huge potentials for cancer treatment, which allows to screening therapeutics targeting almost all kinds of molecules, like proteins and epigenetic regulators [173].

Advertisement

Acknowledgments

This section is supported by the Natural Science Foundation Council of China (81700780 and 81922081), the Department of Education of Guangdong Province (2021KTSCX104), and the Guangdong Basic and Applied Basic Research Foundation (2022A1515012164), and the Science, Technology, and Innovation Commission of Shenzhen (JCYJ20210324104201005).

We thank the administrative assistant (Ms. Yufang Zuo) for providing help and support.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Siegel RL, Miller KD, Fuchs HE, Jemal A. Cancer statistics, 2021. CA: A Cancer Journal for Clinicians. 2021;71:7-33
  2. 2. Ewing J. Neoplastic Disease. A Treatment on Tumors. Philadelphia: Saunders; 1928
  3. 3. Paget S. The distribution of secondary growths in cancer of the breast. Cancer Metastasis Reviews. 1989;8:98-101
  4. 4. Zhao S, Allis CD, Wang GG. The language of chromatin modification in human cancers. Nature Reviews. Cancer. 2021;21:413-430
  5. 5. Zhao Z, Shilatifard A. Epigenetic modifications of histones in cancer. Genome Biology. 2019;20:245
  6. 6. Yang X et al. WNT/β-catenin-suppressed FTO expression increases m(6)a of c-Myc mRNA to promote tumor cell glycolysis and tumorigenesis. Cell Death & Disease. 2021;12:462
  7. 7. Jiang Y et al. KDM6B-mediated histone demethylation of LDHA promotes lung metastasis of osteosarcoma. Theranostics. 2021;11:3868-3881
  8. 8. Huang C et al. EZH2-triggered methylation of SMAD3 promotes its activation and tumor metastasis. The Journal of Clinical Investigation. 2022;132(5):e152394
  9. 9. Wen Y et al. EZH2 activates CHK1 signaling to promote ovarian cancer chemoresistance by maintaining the properties of cancer stem cells. Theranostics. 2021;11:1795-1813
  10. 10. Dong P et al. Long non-coding RNA DLEU2 drives EMT and glycolysis in endometrial cancer through HK2 by competitively binding with miR-455 and by modulating the EZH2/miR-181a pathway. Journal of Experimental & Clinical Cancer Research. 2021;40:216
  11. 11. Bates SE. Epigenetic therapies for cancer. The New England Journal of Medicine. 2020;383:650-663
  12. 12. Cappellacci L, Perinelli DR, Maggi F, Grifantini M, Petrelli R. Recent progress in histone deacetylase inhibitors as anticancer agents. Current Medicinal Chemistry. 2020;27:2449-2493
  13. 13. Michalak EM, Burr ML, Bannister AJ, Dawson MA. The roles of DNA, RNA and histone methylation in ageing and cancer. Nature Reviews Molecular Cell Biology. 2019;20:573-589
  14. 14. Zhong WZ et al. Gefitinib versus Vinorelbine plus cisplatin as adjuvant treatment for stage II-IIIA (N1-N2) EGFR-mutant NSCLC: Final overall survival analysis of CTONG1104 phase III trial. Journal of Clinical Oncology. 2021;39:713-722
  15. 15. Jassem J. Adjuvant EGFR tyrosine kinase inhibitors in EGFR-mutant non-small cell lung cancer: Still an investigational approach. Translational Lung Cancer Research. 2019;8:S387-s390
  16. 16. Gaiko-Shcherbak A et al. Cell force-driven basement membrane disruption fuels EGF- and stiffness-induced invasive cell dissemination from benign breast gland acini. International Journal of Molecular Sciences. 2021;22(8):3962-3981
  17. 17. Vella V et al. Microenvironmental determinants of breast cancer metastasis: Focus on the crucial interplay between estrogen and insulin/insulin-like growth factor signaling. Frontiers in Cell and Development Biology. 2020;8:608412
  18. 18. Zhong B et al. Colorectal cancer-associated fibroblasts promote metastasis by up-regulating LRG1 through stromal IL-6/STAT3 signaling. Cell Death & Disease. 2021;13:16
  19. 19. Kim SD et al. The malignancy of liver cancer cells is increased by IL-4/ERK/AKT signaling axis activity triggered by irradiated endothelial cells. Journal of Radiation Research. 2020;61:376-387
  20. 20. Lin X et al. miR-195-5p/NOTCH2-mediated EMT modulates IL-4 secretion in colorectal cancer to affect M2-like TAM polarization. Journal of Hematology & Oncology. 2019;12:20
  21. 21. Sigismund S, Lanzetti L, Scita G, Di Fiore PP. Endocytosis in the context-dependent regulation of individual and collective cell properties. Nature Reviews. Molecular Cell Biology. 2021;22:625-643
  22. 22. Mullins RDZ, Pal A, Barrett TF, Heft Neal ME, Puram SV. Epithelial-mesenchymal plasticity in tumor immune evasion. Cancer Research. 2022;82:2329-2343
  23. 23. Gerashchenko TS et al. Markers of cancer cell invasion: Are they good enough? Journal of Clinical Medicine. 2019;8(8):1092-1110
  24. 24. Čermák V et al. High-throughput transcriptomic and proteomic profiling of mesenchymal-amoeboid transition in 3D collagen. Scientific Data. 2020;7:160
  25. 25. Huysentruyt LC, Mukherjee P, Banerjee D, Shelton LM, Seyfried TN. Metastatic cancer cells with macrophage properties: Evidence from a new murine tumor model. International Journal of Cancer. 2008;123:73-84
  26. 26. Huysentruyt LC, Shelton LM, Seyfried TN. Influence of methotrexate and cisplatin on tumor progression and survival in the VM mouse model of systemic metastatic cancer. International Journal of Cancer. 2010;126:65-72
  27. 27. Su J et al. TGF-β orchestrates fibrogenic and developmental EMTs via the RAS effector RREB1. Nature. 2020;577:566-571
  28. 28. Zhang J, Liu X, Gao Y. The long noncoding RNA MEG3 regulates Ras-MAPK pathway through RASA1 in trophoblast and is associated with unexplained recurrent spontaneous abortion. Molecular Medicine. 2021;27:70
  29. 29. Zhao H et al. Alternatively-spliced lncRNA-PNUTS promotes HCC cell EMT via regulating ZEB1 expression. Tumori. 2022;34:3008916211072585. DOI: 10.1177/03008916211072585
  30. 30. Liu H et al. Interaction of lncRNA MIR100HG with hnRNPA2B1 facilitates m(6)A-dependent stabilization of TCF7L2 mRNA and colorectal cancer progression. Molecular Cancer. 2022;21:74
  31. 31. Dongre A, Weinberg RA. New insights into the mechanisms of epithelial-mesenchymal transition and implications for cancer. Nature Reviews. Molecular Cell Biology. 2019;20:69-84
  32. 32. Wautier JL, Wautier MP. Vascular permeability in diseases. International Journal of Molecular Sciences. 2022;23(7):3645-3660
  33. 33. Tian M et al. Asiatic acid inhibits angiogenesis and vascular permeability through the VEGF/VEGFR2 signaling pathway to inhibit the growth and metastasis of breast cancer in mice. Phytotherapy Research. 2021;35:6389-6400
  34. 34. Wyckoff JB et al. Direct visualization of macrophage-assisted tumor cell intravasation in mammary tumors. Cancer Research. 2007;67:2649-2656
  35. 35. Giampieri S et al. Localized and reversible TGFbeta signalling switches breast cancer cells from cohesive to single cell motility. Nature Cell Biology. 2009;11:1287-1296
  36. 36. Tavora B et al. Tumoural activation of TLR3-SLIT2 axis in endothelium drives metastasis. Nature. 2020;586:299-304
  37. 37. Wieland E et al. Endothelial Notch1 activity facilitates metastasis. Cancer Cell. 2017;31:355-367
  38. 38. Wei C et al. Crosstalk between cancer cells and tumor associated macrophages is required for mesenchymal circulating tumor cell-mediated colorectal cancer metastasis. Molecular Cancer. 2019;18:64
  39. 39. Ahn JC et al. Detection of circulating tumor cells and their implications as a biomarker for diagnosis, prognostication, and therapeutic monitoring in hepatocellular carcinoma. Hepatology. 2021;73:422-436
  40. 40. Huang X et al. Relationship between circulating tumor cells and tumor response in colorectal cancer patients treated with chemotherapy: A meta-analysis. BMC Cancer. 2014;14:976
  41. 41. Zhou S et al. Circulating tumor cells correlate with prognosis in head and neck squamous cell carcinoma. Technology in Cancer Research & Treatment. 2021;20:1533033821990037-1533033821990037
  42. 42. Rink M et al. Detection of circulating tumour cells in peripheral blood of patients with advanced non-metastatic bladder cancer. BJU International. 2011;107:1668-1675
  43. 43. Lin D et al. Circulating tumor cells: Biology and clinical significance. Signal Transduction and Targeted Therapy. 2021;6:404
  44. 44. Gires O, Pan M, Schinke H, Canis M, Baeuerle PA. Expression and function of epithelial cell adhesion molecule EpCAM: Where are we after 40 years? Cancer Metastasis Reviews. 2020;39:969-987
  45. 45. Wang C et al. Prognostic value of HER2 status on circulating tumor cells in advanced-stage breast cancer patients with HER2-negative tumors. Breast Cancer Research and Treatment. 2020;181:679-689
  46. 46. Forsare C et al. Evolution of estrogen receptor status from primary tumors to metastasis and serially collected circulating tumor cells. International Journal of Molecular Sciences. 2020;21(8):2885-1898
  47. 47. Yin C et al. Molecular profiling of pooled circulating tumor cells from prostate cancer patients using a dual-antibody-functionalized microfluidic device. Analytical Chemistry. 2018;90:3744-3751
  48. 48. Chen X et al. Folate receptor-positive circulating tumor cells as a predictive biomarker for the efficacy of first-line pemetrexed-based chemotherapy in patients with non-squamous non-small cell lung cancer. Annals of Translational Medicine. 2020;8:631
  49. 49. Jiang XH et al. Clinical significance of plasma anti-TOPO48 autoantibody and blood survivin-expressing circulating cancer cells in patients with early stage endometrial carcinoma. Archives of Gynecology and Obstetrics. 2019;299:229-237
  50. 50. Rejniak KA. Circulating tumor cells: When a solid tumor meets a fluid microenvironment. Advances in Experimental Medicine and Biology. 2016;936:93-106
  51. 51. Garrido-Navas C et al. Cooperative and escaping mechanisms between circulating tumor cells and blood constituents. Cell. 2019;8(11):1382-1392
  52. 52. Szczerba BM et al. Neutrophils escort circulating tumour cells to enable cell cycle progression. Nature. 2019;566:553-557
  53. 53. Spicer JD et al. Neutrophils promote liver metastasis via mac-1-mediated interactions with circulating tumor cells. Cancer Research. 2012;72:3919-3927
  54. 54. Masucci MT, Minopoli M, Del Vecchio S, Carriero MV. The emerging role of neutrophil extracellular traps (NETs) in tumor progression and metastasis. Frontiers in Immunology. 2020;11:1749
  55. 55. Yang L et al. DNA of neutrophil extracellular traps promotes cancer metastasis via CCDC25. Nature. 2020;583:133-138
  56. 56. Osmulski PA et al. Contacts with macrophages promote an aggressive Nanomechanical phenotype of circulating tumor cells in prostate cancer. Cancer Research. 2021;81:4110-4123
  57. 57. Liu Q, Liao Q, Zhao Y. Myeloid-derived suppressor cells (MDSC) facilitate distant metastasis of malignancies by shielding circulating tumor cells (CTC) from immune surveillance. Medical Hypotheses. 2016;87:34-39
  58. 58. Sprouse ML et al. PMN-MDSCs enhance CTC metastatic properties through reciprocal interactions via ROS/notch/nodal signaling. International Journal of Molecular Sciences. 2019;20(8):1916-1936
  59. 59. Ortiz-Otero N et al. Cancer associated fibroblasts confer shear resistance to circulating tumor cells during prostate cancer metastatic progression. Oncotarget. 2020;11:1037-1050
  60. 60. Liu Y, Zhang Y, Ding Y, Zhuang R. Platelet-mediated tumor metastasis mechanism and the role of cell adhesion molecules. Critical Reviews in Oncology/Hematology. 2021;167:103502
  61. 61. Maurer S et al. Platelet-mediated shedding of NKG2D ligands impairs NK cell immune-surveillance of tumor cells. Oncoimmunology. 2018;7:e1364827
  62. 62. Placke T, Salih HR, Kopp HG. GITR ligand provided by thrombopoietic cells inhibits NK cell antitumor activity. Journal of Immunology. 2012;189:154-160
  63. 63. Maurer S, Ferrari de Andrade L. NK cell interaction with platelets and myeloid cells in the tumor milieu. Frontiers in Immunology. 2020;11:608849
  64. 64. Bendas G, Borsig L. Cancer cell adhesion and metastasis: Selectins, integrins, and the inhibitory potential of heparins. International Journal of Cell Biology. 2012;2012:676731
  65. 65. Lonsdorf AS et al. Engagement of αIIbβ3 (GPIIb/IIIa) with ανβ3 integrin mediates interaction of melanoma cells with platelets: A connection to hematogenous metastasis. The Journal of Biological Chemistry. 2012;287:2168-2178
  66. 66. Coupland LA, Chong BH, Parish CR. Platelets and P-selectin control tumor cell metastasis in an organ-specific manner and independently of NK cells. Cancer Research. 2012;72:4662-4671
  67. 67. Hourani T et al. Tumor associated macrophages: Origin, recruitment, phenotypic diversity, and targeting. Frontiers in Oncology. 2021;11:788365
  68. 68. Shen CJ et al. Oleic acid-induced NOX4 is dependent on ANGPTL4 expression to promote human colorectal cancer metastasis. Theranostics. 2020;10:7083-7099
  69. 69. Na TY, Schecterson L, Mendonsa AM, Gumbiner BM. The functional activity of E-cadherin controls tumor cell metastasis at multiple steps. Proceedings of the National Academy of Sciences of the United States of America. 2020;117:5931-5937
  70. 70. Green D et al. Targeting the MAPK7/MMP9 axis for metastasis in primary bone cancer. Oncogene. 2020;39:5553-5569
  71. 71. Li X et al. A S100A14-CCL2/CXCL5 signaling axis drives breast cancer metastasis. Theranostics. 2020;10:5687-5703
  72. 72. Padua D et al. TGFbeta primes breast tumors for lung metastasis seeding through angiopoietin-like 4. Cell. 2008;133:66-77
  73. 73. Łukowicz K et al. Chemical compounds released from specific Osteoinductive bioactive materials stimulate human bone marrow mesenchymal stem cell migration. International Journal of Molecular Sciences. 2022;23(5):2598-2609
  74. 74. Tang Y et al. Correction to: Pre-metastatic niche triggers SDF-1/CXCR4 axis and promotes organ colonisation by hepatocellular circulating tumour cells via downregulation of Prrx1. Journal of Experimental & Clinical Cancer Research. 2021;40:234
  75. 75. Kang SY et al. Prosaposin inhibits tumor metastasis via paracrine and endocrine stimulation of stromal p53 and Tsp-1. Proceedings of the National Academy of Sciences of the United States of America. 2009;106:12115-12120
  76. 76. Mazzieri R et al. Targeting the ANG2/TIE2 axis inhibits tumor growth and metastasis by impairing angiogenesis and disabling rebounds of proangiogenic myeloid cells. Cancer Cell. 2011;19:512-526
  77. 77. Ren L, Yu Y, Wang X, Ge J, Wen L. Levels and clinical significances of interleukin 11 in breast tissue and serum of bone metastasis of breast cancer. Zhonghua Yi Xue Za Zhi. 2014;94:2656-2660
  78. 78. Liang M et al. IL-11 is essential in promoting osteolysis in breast cancer bone metastasis via RANKL-independent activation of osteoclastogenesis. Cell Death & Disease. 2019;10:353
  79. 79. Sethi N, Dai X, Winter CG, Kang Y. Tumor-derived JAGGED1 promotes osteolytic bone metastasis of breast cancer by engaging notch signaling in bone cells. Cancer Cell. 2011;19:192-205
  80. 80. Chang YS et al. Mosaic blood vessels in tumors: Frequency of cancer cells in contact with flowing blood. Proceedings of the National Academy of Sciences of the United States of America. 2000;97:14608-14613
  81. 81. Wyckoff JB, Pinner SE, Gschmeissner S, Condeelis JS, Sahai E. ROCK- and myosin-dependent matrix deformation enables protease-independent tumor-cell invasion in vivo. Current Biology. 2006;16:1515-1523
  82. 82. Sahai E, Garcia-Medina R, Pouysségur J, Vial E. Smurf1 regulates tumor cell plasticity and motility through degradation of RhoA leading to localized inhibition of contractility. The Journal of Cell Biology. 2007;176:35-42
  83. 83. Li J et al. ZKSCAN5 activates VEGFC expression by recruiting SETD7 to promote the Lymphangiogenesis, tumour growth, and metastasis of breast cancer. Frontiers in Oncology. 2022;12:875033
  84. 84. Hoshida T et al. Imaging steps of lymphatic metastasis reveals that vascular endothelial growth factor-C increases metastasis by increasing delivery of cancer cells to lymph nodes: Therapeutic implications. Cancer Research. 2006;66:8065-8075
  85. 85. Karaman S, Detmar M. Mechanisms of lymphatic metastasis. The Journal of Clinical Investigation. 2014;124:922-928
  86. 86. Xiao Z et al. Molecular mechanism underlying lymphatic metastasis in pancreatic cancer. BioMed Research International. 2014;2014:925845
  87. 87. He Y et al. Pyroelectric catalysis-based “Nano-lymphatic” reduces tumor interstitial pressure for enhanced penetration and hydrodynamic therapy. ACS Nano. 2021;15:10488-10501
  88. 88. Sereda EE et al. Relationship estimation of cell mobility proteins level with processes of proteolysis and Lymphogenic metastasis in breast cancer. Doklady. Biochemistry and Biophysics. 2021;499:211-214
  89. 89. Fares J, Fares MY, Khachfe HH, Salhab HA, Fares Y. Molecular principles of metastasis: A hallmark of cancer revisited. Signal Transduction and Targeted Therapy. 2020;5:28
  90. 90. Liu SJ, Dang HX, Lim DA, Feng FY, Maher CA. Long noncoding RNAs in cancer metastasis. Nature Reviews. Cancer. 2021;21:446-460
  91. 91. Li J, Meng H, Bai Y, Wang K. Regulation of lncRNA and its role in cancer metastasis. Oncology Research. 2016;23:205-217
  92. 92. Hernandez RK et al. Patients with bone metastases from solid tumors initiating treatment with a bone-targeted agent in 2011: A descriptive analysis using oncology clinic data in the US. Support Care Cancer. 2014;22:2697-2705
  93. 93. Ni J et al. Tumour-derived exosomal lncRNA-SOX2OT promotes bone metastasis of non-small cell lung cancer by targeting the miRNA-194-5p/RAC1 signalling axis in osteoclasts. Cell Death & Disease. 2021;12:662
  94. 94. Ai C et al. Nm23-H1 inhibits lung cancer bone-specific metastasis by upregulating miR-660-5p targeted SMARCA5. Thoracic Cancer. 2020;11:640-650
  95. 95. Yang X-R et al. Correlation of exosomal microRNA clusters with bone metastasis in non-small cell lung cancer. Clinical & Experimental Metastasis. 2021;38:109-117
  96. 96. Han L et al. MicroRNA-106a regulates autophagy-related cell death and EMT by targeting TP53INP1 in lung cancer with bone metastasis. Cell Death & Disease. 2021;12:1037
  97. 97. Xu S et al. Serum microRNA-139-5p is downregulated in lung cancer patients with lytic bone metastasis. Oncology Reports. 2018;39:2376-2384
  98. 98. Wang M, Zhao M, Guo Q, Lou J, Wang L. Non-small cell lung cancer cell-derived exosomal miR-17-5p promotes osteoclast differentiation by targeting PTEN. Experimental Cell Research. 2021;408:112834
  99. 99. Liu Y et al. Serum micro RNA-365 suppresses non-small-cell lung cancer metastasis and invasion in patients with bone meta stasis of lung cancer. The Journal of International Medical Research. 2020;48:300060520939718
  100. 100. Zou P et al. miR-192-5p suppresses the progression of lung cancer bone metastasis by targeting TRIM44. Scientific Reports. 2019;9:19619-19619
  101. 101. Cao J et al. DNA methylation-mediated repression of miR-886-3p predicts poor outcome of human small cell lung cancer. Cancer Research. 2013;73:3326-3335
  102. 102. Rubin EM et al. Wnt inhibitory factor 1 decreases tumorigenesis and metastasis in osteosarcoma. Molecular Cancer Therapeutics. 2010;9:731-741
  103. 103. Wen S et al. Long non-coding RNA NEAT1 promotes bone metastasis of prostate cancer through N6-methyladenosine. Molecular Cancer. 2020;19:171
  104. 104. Hashimoto K et al. Cancer-secreted hsa-miR-940 induces an osteoblastic phenotype in the bone metastatic microenvironment via targeting ARHGAP1 and FAM134A. Proceedings of the National Academy of Sciences of the United States of America. 2018;115:2204-2209
  105. 105. Han Z et al. miR-181b/Oncostatin m axis inhibits prostate cancer bone metastasis via modulating osteoclast differentiation. Journal of Cellular Biochemistry. 2020;121:1664-1674
  106. 106. Misawa A, Kondo Y, Takei H, Takizawa T. Long noncoding RNA HOXA11-AS and transcription factor HOXB13 modulate the expression of bone metastasis-related genes in prostate cancer. Genes (Basel). 2021;12:182
  107. 107. Lang C et al. M(6) a modification of lncRNA PCAT6 promotes bone metastasis in prostate cancer through IGF2BP2-mediated IGF1R mRNA stabilization. Clinical and Translational Medicine. 2021;11:e426
  108. 108. Mo C et al. LncRNA nuclear-enriched abundant transcript 1 shuttled by prostate cancer cells-secreted exosomes initiates osteoblastic phenotypes in the bone metastatic microenvironment via miR-205-5p/runt-related transcription factor 2/splicing factor proline- and glutamine-rich/polypyrimidine tract-binding protein 2 axis. Clinical and Translational Medicine. 2021;11:e493
  109. 109. Wang J et al. Tumor-derived miR-378a-3p-containing extracellular vesicles promote osteolysis by activating the Dyrk1a/Nfatc1/Angptl2 axis for bone metastasis. Cancer Letters. 2022;526:76-90
  110. 110. Li C et al. A ROR1-HER3-lncRNA signalling axis modulates the hippo-YAP pathway to regulate bone metastasis. Nature Cell Biology. 2017;19:106-119
  111. 111. Liu J et al. Osteoclastic miR-214 targets TRAF3 to contribute to osteolytic bone metastasis of breast cancer. Scientific Reports. 2017;7:40487
  112. 112. Cai WL et al. microRNA-124 inhibits bone metastasis of breast cancer by repressing Interleukin-11. Molecular Cancer. 2018;17:9
  113. 113. Liu X et al. Metastatic breast cancer cells overexpress and secrete miR-218 to regulate type I collagen deposition by osteoblasts. Breast Cancer Research: BCR. 2018;20:127-127
  114. 114. Maroni P, Bendinelli P, Matteucci E, Desiderio MA. The therapeutic effect of miR-125b is enhanced by the prostaglandin endoperoxide synthase 2/cyclooxygenase 2 blockade and hampers ETS1 in the context of the microenvironment of bone metastasis. Cell Death & Disease. 2018;9:472
  115. 115. Zhang X et al. MicroRNA-429 inhibits bone metastasis in breast cancer by regulating CrkL and MMP-9. Bone. 2020;130:115139
  116. 116. Yuan X et al. Breast cancer exosomes contribute to pre-metastatic niche formation and promote bone metastasis of tumor cells. Theranostics. 2021;11:1429-1445
  117. 117. Wu K et al. Exosomal miR-19a and IBSP cooperate to induce osteolytic bone metastasis of estrogen receptor-positive breast cancer. Nature Communications. 2021;12:5196
  118. 118. Sun Z et al. LncRNA SNHG3 regulates the BMSC osteogenic differentiation in bone metastasis of breast cancer by modulating the miR-1273g-3p/BMP3 axis. Biochemical and Biophysical Research Communications. 2022;594:117-123
  119. 119. Maroni P, Matteucci E, Bendinelli P, Desiderio MA. Functions and epigenetic regulation of Wwox in bone metastasis from breast carcinoma: Comparison with primary tumors. International Journal of Molecular Sciences. 2017;18:75
  120. 120. Zhang L et al. The molecular mechanism of LncRNA34a-mediated regulation of bone metastasis in hepatocellular carcinoma. Molecular Cancer. 2019;18:120
  121. 121. Huang Z et al. H19 promotes HCC bone metastasis through reducing Osteoprotegerin expression in a protein phosphatase 1 catalytic subunit alpha/p38 mitogen-activated protein kinase-dependent manner and sponging microRNA 200b-3p. Hepatology. 2021;74:214-232
  122. 122. Teng X et al. Bioinformatics analysis for the identification of key genes and long non-coding RNAs related to bone metastasis in breast cancer. Aging. 2021;13:17302-17315
  123. 123. Tang J et al. LncRNA LOC146880 promotes esophageal squamous cell carcinoma progression via miR-328-5p/FSCN1/MAPK axis. Aging (Albany NY). 2021;13:14198-14218
  124. 124. Yu Y et al. miR-190 suppresses breast cancer metastasis by regulation of TGF-β-induced epithelial-mesenchymal transition. Molecular Cancer. 2018;17:70
  125. 125. Liu X et al. miR-1204 targets VDR to promotes epithelial-mesenchymal transition and metastasis in breast cancer. Oncogene. 2018;37:3426-3439
  126. 126. Sang M et al. Circular RNA ciRS-7 maintains metastatic phenotypes as a ceRNA of miR-1299 to target MMPs. Molecular Cancer Research. 2018;16:1665-1675
  127. 127. Wang Z et al. The circROBO1/KLF5/FUS feedback loop regulates the liver metastasis of breast cancer by inhibiting the selective autophagy of afadin. Molecular Cancer. 2022;21:29
  128. 128. Iida M et al. Overexpression of miR-221 and miR-222 in the cancer stroma is associated with malignant potential in colorectal cancer. Oncology Reports. 2018;40:1621-1631
  129. 129. Xing F et al. Loss of XIST in breast cancer activates MSN-c-met and reprograms microglia via Exosomal miRNA to promote brain metastasis. Cancer Research. 2018;78:4316-4330
  130. 130. Yoo B, Ross A, Pantazopoulos P, Medarova Z. MiRNA10b-directed nanotherapy effectively targets brain metastases from breast cancer. Scientific Reports. 2021;11:2844
  131. 131. Curtaz CJ et al. Analysis of microRNAs in exosomes of breast cancer patients in search of molecular prognostic factors in brain metastases. International Journal of Molecular Sciences. 2022;23(7):3683
  132. 132. Wang S et al. JAK2-binding long noncoding RNA promotes breast cancer brain metastasis. The Journal of Clinical Investigation. 2017;127:4498-4515
  133. 133. Fu B et al. Circular RNA circBCBM1 promotes breast cancer brain metastasis by modulating miR-125a/BRD4 axis. International Journal of Biological Sciences. 2021;17:3104-3117
  134. 134. Li D et al. Dysregulation of lncRNA-CCRR contributes to brain metastasis of breast cancer by intercellular coupling via regulating connexin 43 expression. Journal of Cellular and Molecular Medicine. 2021;25:4826-4834
  135. 135. Wyss CB et al. Gain of HIF1 activity and loss of miRNA let-7d promote breast cancer metastasis to the brain via the PDGF/PDGFR Axis. Cancer Research. 2021;81:594-605
  136. 136. Sereno M et al. Downregulation of circulating miR 802-5p and miR 194-5p and upregulation of brain MEF2C along breast cancer brain metastasization. Molecular Oncology. 2020;14:520-538
  137. 137. Giannoudis A et al. A novel panel of differentially-expressed microRNAs in breast cancer brain metastasis may predict patient survival. Scientific Reports. 2019;9:18518
  138. 138. Pan JK et al. MiR-211 determines brain metastasis specificity through SOX11/NGN2 axis in triple-negative breast cancer. Oncogene. 2021;40:1737-1751
  139. 139. Wang H, Zhang P. lncRNA-CASC15 promotes osteosarcoma proliferation and metastasis by regulating epithelial-mesenchymal transition via the Wnt/β-catenin signaling pathway. Oncology Reports. 2021;45(5):76-87
  140. 140. Wang X, Yu X, Long X, Pu Q. MIR205 host gene (MIR205HG) drives osteosarcoma metastasis via regulating the microRNA 2114-3p (miR-2114-3p)/TWIST family bHLH transcription factor 2 (TWIST2) axis. Bioengineered. 2021;12:1576-1586
  141. 141. Wang Q et al. miR-485-3p regulated by MALAT1 inhibits osteosarcoma glycolysis and metastasis by directly suppressing c-MET and AKT3/mTOR signalling. Life Sciences. 2021;268:118925
  142. 142. Zhang J, Piao CD, Ding J, Li ZW. LncRNA MALAT1 facilitates lung metastasis of osteosarcomas through miR-202 sponging. Scientific Reports. 2020;10:12757
  143. 143. Chen Y et al. LncRNA MALAT1 promotes cancer metastasis in osteosarcoma via activation of the PI3K-Akt signaling pathway. Cellular Physiology and Biochemistry. 2018;51:1313-1326
  144. 144. Chen Y et al. LncRNA LOC100129620 promotes osteosarcoma progression through regulating CDK6 expression, tumor angiogenesis, and macrophage polarization. Aging (Albany NY). 2021;13:14258-14276
  145. 145. Chen Y et al. The lncRNA NEAT1 promotes the epithelial-mesenchymal transition and metastasis of osteosarcoma cells by sponging miR-483 to upregulate STAT3 expression. Cancer Cell International. 2021;21:90
  146. 146. Huang JH, Xu Y, Lin FY. The inhibition of microRNA-326 by SP1/HDAC1 contributes to proliferation and metastasis of osteosarcoma through promoting SMO expression. Journal of Cellular and Molecular Medicine. 2020;24:10876-10888
  147. 147. Zou Q et al. miR-19a-mediated downregulation of RhoB inhibits the dephosphorylation of AKT1 and induces osteosarcoma cell metastasis. Cancer Letters. 2018;428:147-159
  148. 148. Wang Y et al. Long noncoding RNA DANCR, working as a competitive endogenous RNA, promotes ROCK1-mediated proliferation and metastasis via decoying of miR-335-5p and miR-1972 in osteosarcoma. Molecular Cancer. 2018;17:89
  149. 149. Wang S-N et al. miR-491 inhibits osteosarcoma lung metastasis and Chemoresistance by targeting αB-crystallin. Molecular Therapy. 2017;25:2140-2149
  150. 150. Lillo Osuna MA et al. Activation of estrogen receptor alpha by Decitabine inhibits osteosarcoma growth and metastasis. Cancer Research. 2019;79:1054-1068
  151. 151. Zhao SJ et al. SPARCL1 suppresses osteosarcoma metastasis and recruits macrophages by activation of canonical WNT/β-catenin signaling through stabilization of the WNT-receptor complex. Oncogene. 2018;37:1049-1061
  152. 152. Itoh H et al. TET2-dependent IL-6 induction mediated by the tumor microenvironment promotes tumor metastasis in osteosarcoma. Oncogene. 2018;37:2903-2920
  153. 153. Lu J et al. IRX1 hypomethylation promotes osteosarcoma metastasis via induction of CXCL14/NF-κB signaling. The Journal of Clinical Investigation. 2015;125:1839-1856
  154. 154. Wang PS et al. A novel long non-coding RNA linc-ZNF469-3 promotes lung metastasis through miR-574-5p-ZEB1 axis in triple negative breast cancer. Oncogene. 2018;37:4662-4678
  155. 155. Yan S et al. Long noncoding RNA MIR31HG inhibits hepatocellular carcinoma proliferation and metastasis by sponging microRNA-575 to modulate ST7L expression. Journal of Experimental & Clinical Cancer Research. 2018;37:214
  156. 156. Liu C et al. CBP mediated DOT1L acetylation confers DOT1L stability and promotes cancer metastasis. Theranostics. 2020;10:1758-1776
  157. 157. Satterfield L et al. miR-130b directly targets ARHGAP1 to drive activation of a metastatic CDC42-PAK1-AP1 positive feedback loop in Ewing sarcoma. International Journal of Cancer. 2017;141:2062-2075
  158. 158. Zhuo W et al. Long noncoding RNA GMAN, up-regulated in gastric cancer tissues, is associated with metastasis in patients and promotes translation of Ephrin A1 by competitively binding GMAN-AS. Gastroenterology. 2019;156:676-691.e611
  159. 159. Yuan J et al. MIR17HG-miR-18a/19a axis, regulated by interferon regulatory factor-1, promotes gastric cancer metastasis via Wnt/β-catenin signalling. Cell Death & Disease. 2019;10:454-454
  160. 160. Ba M-C et al. LncRNA AC093818.1 accelerates gastric cancer metastasis by epigenetically promoting PDK1 expression. Cell Death & Disease. 2020;11:64-64
  161. 161. Rodríguez-Paredes M, Esteller M. Cancer epigenetics reaches mainstream oncology. Nature Medicine. 2011;17:330-339
  162. 162. Jones PA, Baylin SB. The fundamental role of epigenetic events in cancer. Nature Reviews. Genetics. 2002;3:415-428
  163. 163. Mund C, Lyko F. Epigenetic cancer therapy: Proof of concept and remaining challenges. BioEssays. 2010;32:949-957
  164. 164. Jones PA, Issa JP, Baylin S. Targeting the cancer epigenome for therapy. Nature Reviews. Genetics. 2016;17:630-641
  165. 165. Egger G, Liang G, Aparicio A, Jones PA. Epigenetics in human disease and prospects for epigenetic therapy. Nature. 2004;429:457-463
  166. 166. Arun G et al. Differentiation of mammary tumors and reduction in metastasis upon Malat1 lncRNA loss. Genes & Development. 2016;30:34-51
  167. 167. Gutschner T et al. The noncoding RNA MALAT1 is a critical regulator of the metastasis phenotype of lung cancer cells. Cancer Research. 2013;73:1180-1189
  168. 168. Qu L et al. Exosome-transmitted lncARSR promotes Sunitinib resistance in renal cancer by acting as a competing endogenous RNA. Cancer Cell. 2016;29:653-668
  169. 169. Yoshida T et al. Estimated number of off-target candidate sites for antisense oligonucleotides in human mRNA sequences. Genes to Cells. 2018;23:448-455
  170. 170. Yoshida T et al. Evaluation of off-target effects of gapmer antisense oligonucleotides using human cells. Genes to Cells. 2019;24:827-835
  171. 171. Lai F, Damle SS, Ling KK, Rigo F. Directed RNase H cleavage of nascent transcripts causes transcription termination. Molecular Cell. 2020;77:1032-1043.e1034
  172. 172. Lee JS, Mendell JT. Antisense-mediated transcript knockdown triggers premature transcription termination. Molecular Cell. 2020;77:1044-1054.e1043
  173. 173. Huang J et al. Promising therapeutic targets for treatment of rheumatoid arthritis. Frontiers in Immunology. 2021;12:686155

Written By

Jie Huang, Aiping Lu and Chao Liang

Submitted: 07 June 2022 Reviewed: 15 July 2022 Published: 21 August 2022