Next Article in Journal
Visible-Light-Activated Carbon Dot Photocatalyst for ROS-Mediated Inhibition of Algae Growth
Next Article in Special Issue
Influence of Human Bone Marrow Mesenchymal Stem Cells Secretome from Acute Myeloid Leukemia Patients on the Proliferation and Death of K562 and K562-Lucena Leukemia Cell Lineages
Previous Article in Journal
New Insights into Photobiomodulation of the Vaginal Microbiome—A Critical Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Dual Role of Mesenchymal Stem Cells in Cancer Pathophysiology: Pro-Tumorigenic Effects versus Therapeutic Potential

1
Unité de Recherche Études Pharmaco-Immunologiques (EPI), Université de La Réunion, CHU de La Réunion, Allée des Topazes, 97400 Saint-Denis, La Réunion, France
2
Service de Radiothérapie, Clinique Sainte-Clotilde, Groupe Clinifutur, 127 Route de Bois de Nèfles, 97400 Saint-Denis, La Réunion, France
3
Laboratoire Interdisciplinaire de Recherche en Santé (LIRS), RunResearch, Clinique Sainte-Clotilde, 127 Route de Bois de Nèfles, 97400 Saint-Denis, La Réunion, France
4
Service d’Anatomie et Cytologie Pathologiques, CHU de La Réunion sites SUD—Saint-Pierre, Avenue François Mitterrand, 97448 Saint-Pierre Cedex, La Réunion, France
5
Service d’Oncologie Médicale, CHU de La Réunion sites SUD—Saint-Pierre, Avenue François Mitterrand, 97448 Saint-Pierre Cedex, La Réunion, France
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(17), 13511; https://doi.org/10.3390/ijms241713511
Submission received: 2 August 2023 / Revised: 29 August 2023 / Accepted: 30 August 2023 / Published: 31 August 2023
(This article belongs to the Special Issue Mesenchymal Stem Cells and Cancer)

Abstract

:
Mesenchymal stem/stromal cells (MSCs) are multipotent cells involved in numerous physiological events, including organogenesis, the maintenance of tissue homeostasis, regeneration, or tissue repair. MSCs are increasingly recognized as playing a major, dual, and complex role in cancer pathophysiology through their ability to limit or promote tumor progression. Indeed, these cells are known to interact with the tumor microenvironment, modulate the behavior of tumor cells, influence their functions, and promote distant metastasis formation through the secretion of mediators, the regulation of cell–cell interactions, and the modulation of the immune response. This dynamic network can lead to the establishment of immunoprivileged tissue niches or the formation of new tumors through the proliferation/differentiation of MSCs into cancer-associated fibroblasts as well as cancer stem cells. However, MSCs exhibit also therapeutic effects including anti-tumor, anti-proliferative, anti-inflammatory, or anti-oxidative effects. The therapeutic interest in MSCs is currently growing, mainly due to their ability to selectively migrate and penetrate tumor sites, which would make them relevant as vectors for advanced therapies. Therefore, this review aims to provide an overview of the double-edged sword implications of MSCs in tumor processes. The therapeutic potential of MSCs will be reviewed in melanoma and lung cancers.

1. Introduction

Mesenchymal stem/stromal cells (MSCs) have aroused great interest over the past 20 years. This interest is partly due to their implication in various physiological and pathological processes, comprising development, tissue repair, organ transplantation, auto-immunity, and cancer [1,2,3,4,5,6,7,8,9,10]. Initially, MSCs were isolated from adult’s bone marrow (BM) in the late 1960s and were defined as fibroblast-like cells [11,12]. Caplan and co-authors qualified them as “mesenchymal stem cells” in 1991 for their capacities of self-renewal and multilineage differentiation [13,14]. Indeed, MSCs can give rise to several mesenchymal tissues by differentiation into several cell types [15,16,17]. It is based on a publication of the International Society for Cellular Therapy (ISCT) from 2006 that MSCs were redefined.
The ISCT issued recommendations, in particular related to the designation of these cells as “multipotent mesenchymal stromal cells” [18,19]. However, the most commonly used terms remain “mesenchymal stromal cells” and “mesenchymal stem cells”. Thus, MSCs are named differently depending on their tissue location. Despite them having been initially described in BM, they were also isolated from many other tissues in different proportions [20,21,22,23,24]. Indeed, MSCs were detected in adipose [25,26], pulmonary [27], synovial [28], renal [22], hepatic [29], bone [30], and muscle [31,32] tissues as well as in the umbilical cord [3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35], placenta [36,37], or peripheral blood [33,38,39]. Most studies defined MSCs according to the minimum criteria proposed by the ISCT obtained on BM-derived MSCs, including cell properties to adhere to plastic under standard culture conditions, the expression of surface markers such as CD105 (Endoglin, TGFβR), CD73, and CD90 (Thy-1), the absence of surface markers such as CD45, CD34, CD14, CD11b, CD19, CD79a, and HLA-DR, as well as their plasticity as presented in Table 1 [40,41]. Plasticity is measured through the ability of MSCs to differentiate into osteocytes, adipocytes, and chondrocytes in response to appropriate growth factors under in vitro conditions [19].
Of note, there are substantial functional and phenotypic differences between MSCs depending on their tissue sources [42]. However, even today, a complete and precise definition for MSCs’ function and role is still missing. Unlike hematopoietic cell subpopulations, whose identity, developmental stage, and plasticity can be predicted based on the expression of transcription factors and a combination of cell surface markers, MSCs lack specific markers [43,44,45,46]. Consequently, the currently used definitions and classifications for MSCs are based on phenotypic and functional criteria common to other cell types, which are not specific enough. Of note is that MSCs may display several differentiation properties. Indeed, growth factors could promote the differentiation of a fraction of MSCs into osteocytes, adipocytes, or chondrocytes, leading to functional heterogeneity despite the expression of identical surface markers [47,48].
The literature data largely demonstrated that MSCs can both promote and limit tumor growth, with a pro-tumor predominance as recently described by Galland et al. [49]. In this present review, we aim to first review the function and role of MSCs in cancer progression and, secondly, the potential of MSCs as therapeutic tools.

2. Origin of Mesenchymal Stem Cells and Other Stromal Cells in the Tumor Microenvironment

The heterogeneity of the tumor microenvironment (TME) characterizes the tumor properties [50]. The TME that directly influences cancer progression is represented by a heterogeneous group of resident and infiltrating cells, which interact together through the secretion of mediators and the production of extracellular matrix (ECM) in order to support the dynamic ecosystem of the tumor [51]. Tumor parenchyma cells grow symbiotically by interacting with the surrounding stroma through paracrine and juxtacrine signaling events.
Cells and molecules composing the TME constitute a complex network involving cellular and physical changes within host tissue to support tumor growth and cancer progression [52,53]. The TME represents the non-neoplastic compartment surrounding tumor tissues. The exact composition of the TME varies depending on the type of tumors considered, but it notably comprises immune cells such as tumor-associated macrophages (TAMs), pericytes (PCs)/vascular smooth muscle cells (vSMCs), endothelial cells, myofibroblasts, sometimes adipocytes, MSCs, and fibroblasts, which are the most abundant cell type in the tumor microenvironment, also called tumor-associated fibroblasts (TAFs) or cancer-/carcinoma-associated fibroblasts (CAFs) [10,54].
MSCs in adult tissues have two main embryonic origins, deriving either from the mesoderm or the ectoderm neural crest cells, for example, in the central nervous system (CNS) [20,21,55,56,57]. During embryogenesis, MSCs can migrate along vessels and then reside in perivascular (p) niches to give rise to pMSCs in adult tissues, adopting almost similar features and cell surface markers to those of PCs (Figure 1) [21,58,59]. MSCs can also migrate along axons and express peripheral myelin markers, such as myelin P zero or myelin P zero-like 2 (MPZL2), generally found on Schwann cells of the peripheral nerves.
Commonly applied markers or genes (Table 1 and Table 2), such as NG2/Cspg4, CD13/Anpep, and desmin, are not specific to MSCs and their expression is not stable, particularly in disease conditions. Other markers, like CD248 (endosialin, also known as tumor endothelium marker 1, TEM1) and CD90/THY-1, are highly expressed by PCs but also by pMSCs, especially in the context of cancer [60,61]. It is established that PCs expressing high levels of CD248 maintain the interaction with CD93 on endothelial cells via the ECM protein multimerin-2 (Figure 1) [62].
Recent studies utilizing cell lineage tracing and single-cell RNA sequencing experiments have provided insights into the different cell types present in the stroma of healthy tissues (lung, heart, brain, skin, and colon) and the TME [60,63,64,65]. These studies have recently been reviewed comprehensively by Betsholtz et al. in 2022 [66]. All these studies support the emerging notion that periendothelial cells such as PCs (in close association with the basement membrane of endothelial cells) can give rise to pMSCs detaching from the vessels and infiltrating the tissue parenchyma [63,67]. Inversely, in zebrafish, pMSCs have been shown to give rise to PCs to support vessel survival/growth [68]. Interestingly, Kalluri and colleagues have long argued that the distinction between fibroblasts and MSCs is extremely difficult and the immunological and molecular collective evidence suggest that quiescent or resting fibroblasts are precursors of activated fibroblast called MSCs [69]. In the brain, this has been further evidenced by Garcia and colleagues, observing that PCs and MSCs (called vascular fibroblasts by the authors) have similar transcriptomic signatures [63].
Resting MSCs have the potential to proliferate when stimulated by growth factors, for example, transforming growth factor-β (TGF-β), platelet-derived growth factor (PDGF), and interleukin-6 (IL-6). Properties assigned to stromal fibroblasts are in fact properties associated with ‘activated’ fibroblasts, myofibroblasts, and MSCs. Once activated, their functions include the production of ECM cytokines and chemokines, recruiting immune cells, and exerting physical forces to modify tissue architecture. PCs maintain vascularization by providing structural support to blood vessels [59]. Their abundance in tumor stroma is directly related to their physiological dysregulation. Indeed, they constitute the pre-metastatic niche and participate in immunosuppression through M2 macrophage polarization and the growth of cancer cells [59,60,61,62,63,64,65,66,67,68,69,70]. Of note is that the PCs’ detach from endothelial cells to give rise to proliferating and infiltrating MSCs, which may help to promote tumor cell invasion (Figure 1) [71]. Angiogenesis is regulated by PCs/MSCs that mediate pro-angiogenic mediator secretions (e.g., VEGF).
Hence, during cancer development, cells composing the TME interact directly or indirectly with cancer cells and ECM through the secretion of pro-inflammatory cytokines, chemokines, and growth factors, in order to support tumor growth, proliferation, and metastatic tumor formation [72]. These interactions associated with pro-inflammatory secretions promote the alteration of the original properties and functions of TME partners. Then, new partners such as immune and stromal cells are recruited and come to enrich this complex network.
Figure 1. The proliferation and differentiation of perivascular (p) MSCs in cancer can give rise either to pericytes, perivascular fibroblasts, cancer-associated fibroblasts, as well as cancer stem cells. The vascular unit is formed by endothelial cells (CD93+), peridentothelial pericytes (PCs, CD248high) embedded into the extracellular matrix (ECM, e.g., multimerin-2) of the basement membrane, and perivascular mesenchymal stem cells (pMSCs CD248low/CD90+). Single-cell RNA profiling experiments have indicated that PCs (e.g., in the CNS) may give rise to subsets of pMSCs, which can proliferate, particularly in response to PDGF, and differentiate into either fibroblasts subsets (ECMhigh or VEGFhigh), cancer-associated fibroblasts (CAFs to support tumor growth), or cancer stem cells (CSCs), as recently evidenced in glioblastoma [63,73]. Fibroblasts stimulated by TGF-β1 may give rise to myofibroblasts (Smooth Muscle cell Actin/SMAhigh/ECMhigh). TAMs: tumor-associated macrophages. (Figure designed by BioRender).
Figure 1. The proliferation and differentiation of perivascular (p) MSCs in cancer can give rise either to pericytes, perivascular fibroblasts, cancer-associated fibroblasts, as well as cancer stem cells. The vascular unit is formed by endothelial cells (CD93+), peridentothelial pericytes (PCs, CD248high) embedded into the extracellular matrix (ECM, e.g., multimerin-2) of the basement membrane, and perivascular mesenchymal stem cells (pMSCs CD248low/CD90+). Single-cell RNA profiling experiments have indicated that PCs (e.g., in the CNS) may give rise to subsets of pMSCs, which can proliferate, particularly in response to PDGF, and differentiate into either fibroblasts subsets (ECMhigh or VEGFhigh), cancer-associated fibroblasts (CAFs to support tumor growth), or cancer stem cells (CSCs), as recently evidenced in glioblastoma [63,73]. Fibroblasts stimulated by TGF-β1 may give rise to myofibroblasts (Smooth Muscle cell Actin/SMAhigh/ECMhigh). TAMs: tumor-associated macrophages. (Figure designed by BioRender).
Ijms 24 13511 g001
Table 2. Canonical cell markers in healthy cells and in the tumor microenvironment.
Table 2. Canonical cell markers in healthy cells and in the tumor microenvironment.
Cell TypesMarkersCommentsReferences
Endothelial cellsCD31PECAM-1, homotypic cell adhesion[60]
CD93CD248 family member[60,74]
CLDN5Claudin 5[60]
CDH5Cadherin 5[60]
Pericytes (PCs)PDGFRβReceptor for platelet-derived growth factor[59]
NG2Encoded by Chondroitin sulfate proteoglycan 4 CSPG4 gene.[59,60]
Desmin [59]
CD13Alanyl Aminopeptidase N[59]
CD248Endosialin (TEM-1), highly expressed in tumor tissues (PCs > pMSCs)[59,61,75]
KCNJ8Kir6.1, Potassium inwardly rectifying channel, associates with Abcc9[59,60,76]
Vascular smooth muscle cells (vSMCs)PDGFRβLevels in PCs > vSMCs[77]
NG2 [77]
DesminMuscle class III intermediate filament[77]
CD13
RGS5Regulator of G protein signaling 5 GTPase activating protein[59,64,78]
CD146Melanoma cell adhesion molecule (MCAM)[59,78]
α-SMAAlpha smooth muscle cell actin encoded by ACTA2. Level of expression in vSMCs >> PCs[59]
TAGLNTransgelin, smooth muscle protein 22 alpha (SM22).[59]
Perivascular MSCs (also known as vascular fibroblasts)VIM
FSP
FAP
PDGFRβ
Vimentin, intermediate-filament ass. protein
intermediate-filament ass. protein
Fib activation protein alpha, serine protease
[79]
PDGFRα [60]
CD13 [79]
COL1A1pMSCs also express high levels of COL1A2, COL3A1, and COL4A1. Much less expressed by PCs’
LAMA-1laminin subunit alpha 1, also expressed by epithelial cells[60]
LUMLumican[60]
DCNDecorin[60]
MPZL2Myelin protein zero-like 2, adhesion[60]
SRPX2Sushi repeat-containing protein X-linked 2[60]
FNFibronectin, also expressed by PCs/scar tissue[80]
Tumor-associated Macrophages (TAMs)CD11b/CD18Complement receptor type 3 involved in phagocytosis of host cell debris[81,82]
CD163Scavenger receptor, M2 anti-inflammatory[83]
C1q/C3
CD68
Complement factors[84]
Adapted from Ah-Pine et al., 2023 [67]. Several markers are shared between the different subsets of periendothelial/perivascular cells, while others (in bold) are restricted (but not exclusive) to specific subsets. CD248, also known as TEM1: tumor endothelial marker 1, but subsequently found on pericytes around vessels but not expressed by endothelial cells.

2.1. Recruitment of Mesenchymal Stem Cells

As previously described, pMSCs exhibit multilineage differentiation and mainly reside in the perivascular niches of all organs [22,85]. Current evidence suggests that the MSCs present in tumor stroma come from both local and distant tissues [86,87,88]. Considering that MSCs are present in most body tissues, the population of tissue-resident MSCs found in the TME likely represents the first population to occupy the TME and support tumor development [89,90,91]. Subsequent recruitment of the distant MSC population is accompanied by a pro-inflammatory state at the tumor site [92]. This phenomenon, called “homing”, represents the migration of cells to the tissue, to graft themselves there and to exert local functional effects. Furthermore, it is well established that MSCs are recruited to tumor sites to support regenerative activity [93,94,95].
Dvorak defined cancer as “a wound that does not heal”, and MSCs are also recruited at tumor sites for tissue repair [93]. Then, MSCs are regulated by the pro-inflammatory environment of the TME, by mediators secreted by tumor cells, and may adopt a pro-tumorigenic phenotype [91,96]. Of note is that the migration of MSCs to tumor sites was observed in particular in Kaposi’s sarcoma [97]; colorectal tumors [98]; pancreatic tumors [99]; glioma [100]; breast, lung, and melanoma metastatic tumors [101,102]; gastric tumors [103]; melanoma [104]; or even ovarian tumors [105]. Many studies described progressive changes in the tumor stroma with an initial apparent expansion of MSCs which precedes the conversion to malignancy. MSCs are often observed circumscribing early or premalignant lesions.
Importantly, during cancer development, MSCs have the ability to migrate and be activated at the tumor sites [86,87,88]. Numerous factors are involved in these migration/activation processes, including growth factors (GF) (Epidermal GF (EGF), stem cell factor (SCF), PDGF, hepatocyte GF (HGF), and insulin GF (IGF-1)) [106,107], angiogenic factors (VEGF, hypoxia-inducible factor (HIF-1α), and Fibroblast GF (FGF-β)) [108,109], inflammatory factors such as chemokines (CXCL8, CCL2, CCL5, CCL22, and SDF-1/CXCL12) [107,108,109,110], and cytokines (TNF-α, TGF-β, IL-6, and IL-1β) [111,112,113,114]. Other factors such as reactive oxygen species (ROS), adhesion molecules (VCAM), matrix metalloproteases (MMPs) of the ECM, alarmins (e.g., HMGB1, ATP, extracellular DNA) produced by necrotic cancer cells, and the level of hypoxia (HIF-1) can also control this migration process (Figure 2) [115,116,117,118,119,120]. Indeed, it is now well established that MSCs express an armory of innate immune receptors (also called pattern recognition receptors, PRRs) for alarmins, which are also able to recognize microbial ligands [24].

2.2. Mesenchymal Stem Cell Activation and Multi-Lineage Differentiation

During cancerous pathophysiological process, MSCs are recruited into the TME where they are exposed to multiple pro-inflammatory mediators and other stimuli, which may change their behavior to promote tumor growth through immunomodulation [8].
The MSCs differentiate depending on the tumor stage of development, their tissue origin, the heterogeneity of TME, and the interactions with other cells composing the TME [121].
The MSCs’ lineage differentiation in the TME is presented in Figure 1, Figure 2 and Figure 3. The MSC population can give rise to several cell types, such as fibroblasts. Fibroblasts can differentiate or transdifferentiate into CAFs and play a critical role in cancer progression. The CAFs represent a heterogeneous population of activated fibroblasts, which can also potentially transdifferentiate from epithelial or endothelial cells [122,123].
MSCs are also able to differentiate into adipocytes that may give rise to cancer-associated adipocytes (CAAs) known to play an active role in the TME formation and function [124,125,126]. The CAAs modulate, for instance, the behavior of breast cancer cells via the release of pro-inflammatory mediators such as IL-6, TNF-α, CCL2, or CCL5, promoting angiogenesis, proliferation, and metastasis [126,127,128].
Importantly, MSCs could also give rise to CSCs characterized by their ability to renew, differentiate, mimic the behavior of tumor cells [7,129], or fuse with tumor cells, giving rise to a highly pro-tumor hybrid complex [15,130,131,132]. In the CNS, it has recently been established that PCs/MSCs originating from the neural crest can give rise to CSCs (also known as glioma-stem cells (GSCs)) in high-grade glioma (Glioblastoma multiforme) [73].

2.3. Interplays between Mesenchymal Stem Cells and Components of the Tumor Microenvironment

To establish the TME and build the primary and secondary sites of the tumor, cells of the TME communicate directly by cell–cell contact or indirectly. The indirect communication between cellular components of the TME contributes to the malignant properties of tumor cells as well as the ECM composition.
The pro- or anti-tumorigenic behavior of MSCs is regulated by this environment. As shown in Figure 4, MSCs communicate and regulate the TME components, resident and infiltrating cells promoting cancer progression.

2.3.1. Innate and Adaptative Immune Cells

Among the immune cells present in the TME, selective populations of T and B lymphocytes promote and/or suppress tumor growth [133,134]. A dual role in tumor progression has been attributed to tumor-infiltrating B cells. In the pathophysiology of breast cancer, the literature data reported different conclusions depending on cancer type and stage. In triple-negative breast cancer, a high level of tumor-infiltrating B cells and plasma cells is correlated with a favorable prognosis [135]. The immunosuppressive role of MSCs has already been reported, as well as the effect of T and B cells, as shown in Figure 4. During inflammatory disease, MSCs are known to drive a profile change for B and T cells by inducing regulatory B and T cell production [136]. During prostate cancer, MSCs suppress T cell proliferation, exhibiting immunosuppressive properties, participating in blocking immunologic recognition and the elimination of malignant cells [137]. Interestingly, the B and T cell profiles are influenced by an over secretion of anti-inflammatory mediators, such as IL-4 and IL-10, and a decrease in the secretion of pro-inflammatory mediators (TNF-α, IFN-γ) promoting B and Treg phenotypes [138,139,140]. Moreover, the immunosuppressive role of MSCs on B cells occurs through the modulatory activity of indoleamine-2,3-dioxygenase (IDO), cyclooxygenase-2 (Cox-2), and programmed cell death ligand 1 (PD-L1) [141]. The regulation of T cell activity by MSCs is regulated by iNOS, NO secretion, and prostaglandin E2 (PGE2) [142,143].
TAMs and tumor-associated neutrophils (TANs) are known to secrete pro-inflammatory mediators. Importantly, considering the immune role of macrophages, their capacity to phagocyte, and their cytotoxic properties (through the production of reactive oxygen species (ROS) and complement proteins), they could be associated with anti-tumor immunity. Nevertheless, high levels of TAMs are mainly associated with a poor prognosis in numerous cancers [144]. The recruitment of TAMs into tumors is known to be mediated by the cytokines and chemokines released by tumor cells and CAFs. These factors have been extensively studied and include CCL2, CCL7, glial-cell-line-derived neurotrophic factor (GDNF), VEGF, alarmin ATP (released from necrotic cells), macrophage colony-stimulating factor 1 (CSF1), granulocyte–macrophage colony-stimulating factor (GM–CSF), periostin, and complement anaphylatoxins (e.g., C3a) [145,146,147]. The MSCs promote macrophage polarization in favor of an anti-inflammatory phenotype (TAMs 2). Jia and co-authors found that MSCs express iNOS, CCL2, IL-6, and Cox-2 and recruit macrophages at tumor sites [148]. The IL-6 secreted by MSCs promotes the polarization of infiltrated TAMs into M2-like macrophages that will express an anti-inflammatory profile and play an immunosuppressive role on T cells. Moreover, TAMs participate in the development of metastases [149], while TANs promote angiogenesis and immune suppression [150,151]. MSCs drive TANs to adopt an anti-inflammatory profile and secrete IL-10 and PGE2. MSCs are able to modulate neutrophil activity. In a breast tumor model, an MSC/neutrophil (CD11b+Ly6G+) co-culture model allowed neutrophils to acquire immunosuppressive activity, downregulating T cell proliferation in an in vitro model and amplifying progression tumor in an in vivo model [152]. Then, TANs mediated MSCs’ differentiation into CAFs via the activation of AKT and p38 MAPK signaling pathways [153].
Importantly, the immunosuppressive role of MSCs also affects NK cell activity by decreasing their cytotoxicity capacity through mTOR inhibition in gastric cancer [154]. Consequently, the secretion of granzymes and perforin by NK is downregulated, and that promotes tumor cell survival and cancer progression. Moreover, a study based on a co-culture model of MSCs and NK cells co-exposed to IL-15 demonstrated an alteration of IFN-γ secretion by NK cells, a decrease in their proliferation, as well as a loss of their cytotoxicity linked to the alteration of cytolytic degranulation and perforin release [155]. Interestingly, authors detected important levels of IDO secreted by the MSCs participating in immune response suppression. The decrease in the cytotoxic activity of T and NK cells is also due to an increase in IL-10 and iNOS releases, a decrease in inflammatory effector production (TNF-α, IL-17, FASL, TRAIL), and a lack of immune cell responses to pro-inflammatory cytokines (IL-12, IL-18) leading to a loss of IFN-γ release through NFκB and STAT4 inhibition [138,156,157].

2.3.2. Tumor Cells

Interestingly, on tumor sites, MSCs interact with tumor cells to support tumor development. MSCs communicate with tumor cells indirectly through the secretion of cytokines, chemokines, and growth factors [52,158,159,160]. Among the indirect interactions, there are also interactions via metabolites, such as PGE2 in lymphoblastic leukaemia cells [161]. Importantly, numerous studies are interested in interactions via microparticles, such as functional mRNAs, proteins, or micro-RNAs called “miRs” or “miRNAs”. The MSCs released, via exosomes, miRNAs, play a crucial role in the regulation of cell death. Indeed, it has been demonstrated that the immunoregulatory functions of MSCs could be attributed to miRs such as miR-21 [162,163,164]. Direct interactions occurring between MSCs and tumor cells comprise interactions via Notch signaling, the formation of nanotubes, the fusion of cells, trogocytosis, or Gap junctions permitting exchanges of second messengers such as c-AMP, miRs, or ions [130,165]. Rafii et al. reported trogocytosis, which is an exchange of plasma membrane fragments between ovarian cancer cells and stromal cells, for the first time and demonstrated that it promoted the acquisition of chemoresistance [166].
The crosstalk between MSCs and tumor cells also involves the activation of pro-inflammatory and survival-signaling pathways, such as the NFκB, PI3K/Akt, Wnt, MAPK, and JAK/STAT signaling pathways that mediate tumor progression, metastasis formation, and angiogenesis [120,167,168]. Accordingly, the activation of the NFκB and PI3K/AKT signaling pathways promote immune suppression, EMT, and metastasis in breast and gastric cancers [169,170,171]. In the TME, naive MSCs exhibiting anti-tumor properties are reprogrammed by cancer cells and the TME into CAFs, which participate in tumor progression [172]. Interestingly, CAFs promote tumor growth, angiogenesis, immunosuppression, and metastasis via the secretion of growth factors such as VEGF [173,174], TGF-β [175], cytokines and chemokines (IL-6, CXCL12) [123,176,177], PDGF, and MMPs [178,179,180]. The CSCs are also regulated by MSCs through the secretion of pro-inflammatory mediators as well as growth factors [138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178,179,180,181].

2.3.3. Extracellular Matrix

The TME also comprises non-cellular components, such as ECM as collagen produced abundantly by MSCs [60]. This ECM is crucial for interactions between cellular components of the TME and for cancer expansion [181,182]. Indeed, ECM provides a structural backbone for the tumor stroma as well as active soluble factors such as growth, angiogenic factors, cytokines, chemokines, and hormones secreted by stromal and tumor cells regulating tumor behavior. An in vitro study on a three-dimensional model reported that ECM heterogeneity highly influenced invasion and metastasis efficiency [183]. During the process of tumor development, the ECM is disorganized and collagen deposit as well as cross-linking with other matrix proteins such as elastin, laminin, or fibronectin are associated with metastatic tumor formation [184]. However, depending on the ECM organization and composition, tumor cells will adopt different behavior and cancer will develop in different ways. All the interactions between cells and TME will be directly affected by ECM organization and composition [185,186]. More importantly, recruited immune cells’ behavior will be directly modulated by the ECM composition. Gonzalez et al. demonstrated in human MSCs isolated from breast cancer metastasis that stromal cells expressing a fibroblastic phenotype were able to upregulate collagen receptor expression, namely discoidin domain receptor 2 (DDR2), in breast cancer cells and increase collagen deposits promoting breast cancer cell alignment, migration, and metastasis [187]. Moreover, MSCs also modulate collagen 1, 4, 5, tenascin, and periostin deposits and modulate the ECM composition involving the ERK, Src, and NFκB signaling pathways [138,143,188]. High levels of MMPs expressed by MSCs/fibroblasts in pathological conditions will contribute to ECM degradation and facilitate the invasive behavior of tumor cells [168]. Importantly, as described by Gattazzo and co-authors, the ECM modulates stem cell differentiation and proliferation [185]. The molecular composition of the ECM constitutes an appropriate niche for stem cells that is directly regulated by mediators composing the ECM. The ECM is also regulated by cancer cells and can in turn regulate the behavior of MSCs.

3. Role of MSCs-1 and MSCs-2 Subtypes in Cancer Progression

The literature is extensive on the ability of MSCs to develop pro- and anti-tumor phenotypes [189,190]. Their role in tumor progression depends on the balance between their pro- and anti-inflammatory phenotypes [191]. Waterman and co-authors suggested that MSCs could undergo a functional switch from an anti-tumorigenic state, called MSCs-1, to a pro-tumorigenic state, called MSCs-2 [192,193].
Importantly, the immunosuppression capacity of MSCs involves the MSCs-2 phenotype that inhibits immune cells and turns off their anti-tumorigenic properties. Of note is that MSCs-1 inhibits tumor growth by promoting the infiltration of monocytes, granulocytes, and T cells into tumors, contributing to anti-cancer immunity. These immune cells, coupled with the pro-inflammatory state generated in the tumor parenchyma, induce the secretion of chemokines attracting activated lymphocytes to promote further anti-cancer immunity [194]. It has been demonstrated in an in vivo study on Kaposi’s sarcoma that MSCs were able to suppress tumor growth by inhibiting the Akt signaling pathway of tumor cells [97]. The MSCs were also able to suppress tumor proliferation in breast cancer through the inhibition of the Wnt pathway involved in tumor development [195]. Accordingly, studies demonstrated that MSCs upregulated the mRNA expression of a negative regulator of the cell cycle, namely p21, and caspase-3, which is a key protease of apoptosis activation, in tumor cells. This regulation is implicated in the inhibition of tumor cell growth through apoptosis activation and cell cycle arrest in the G(0)/G(1) phase of tumor cells [196]. Additionally, MSCs caused the inhibition of tumor angiogenesis through endothelial cell apoptosis and capillary degeneration, altering tumor growth [77]. Thus, MSCs produced exosomes which blocked hepatocellular carcinoma stem cell malignancy via the upregulation of long non-coding RNAs (IncRNAs), namely C5orf66-AS. Authors demonstrated that C5orf66-AS1 upregulated DUSP1 expression through the inhibition of the microRNA-127-3p (miR-127-3p) production controlled by ERK, significantly reducing the proliferation, migration, invasion, angiogenesis-stimulating, and self-renewal abilities of CSCs [197].
The deleterious role of MSCs in the tumorigenesis process is often attributed to diverse mechanisms promoted by MSCs-2 [198]: differentiation into pro-tumorigenic mediators; the suppression of the immune response [199]; angiogenesis [200,201]; EMT; the formation of CSCs; the tumor cells’ survival; and the metastatic process [6,172,202]. All these mechanisms, summarized in Figure 5, are driven by direct and indirect interplays between MSCs, cancer cells, and the TME and promote the resistance of tumor cells to therapeutic treatment [203].

3.1. The Epithelial-Mesenchymal Transition

Among the growth factors secreted by TME components, we find TGF-β. This growth factor is a cytokine regulating the cell proliferation, migration, and differentiation of a large number of cell types. At the onset of tumorigenicity, this growth factor acts as a tumor suppressor by inducing cytostasis and apoptosis of normal and pre-malignant cells. However, at an advanced stage, cancer cells become resistant to its cytostatic effects. By directly interacting with the cancer cells, TGF-β can promote tumorigenicity by inducing EMT [204]. In physiological conditions, EMT plays a crucial role in embryogenesis and tissue damage repair [205]. Nevertheless, this process can stimulate tumor development [206]. The EMT is characterized by changes in the production of three prominent biomarkers (E-cadherin, vimentin, and N-cadherin), and these changes can lead to decreased cell adhesion, the loss of tight junctions, and polarity. At the same time, epithelial cells adopt characteristics of a mesenchymal phenotype, including motility and the potential for invasion and metastasis [207]. Although the mechanisms of MSC tropism to the tumor site are not yet fully understood, this process was exploited to develop more specific and effective anti-tumor therapies [208].

3.2. Angiogenesis

Numerous studies demonstrated that MSCs influenced various mechanisms of angiogenesis. Indeed, MSCs produced high levels of cytokines and growth factors, such as IL-6, IL-8, VEGF, FGF-β, PDGF, TGF-β, and angiopoietin, thus promoting tumor angiogenesis [209,210]. It was also shown that CAFs strongly mediated tumor angiogenesis by secreting immunomodulatory and pro-angiogenic cytokines and chemokines, such as IL-4, IL-6, IL-8, TNF-α, TGF-β, VEGF, and SDF-1 [69]. An in vivo study reported that the IL-6 released by MSCs led to an increase in the levels of endothelin-1 secreted in tumor cells, involving endothelial cell activation via the Akt/ERK signaling pathway promoting angiogenesis [211].
Coffman and his team also observed, in the context of an ovarian cancer model, that MSCs induced angiogenesis via a paracrine signaling loop mediated by BMP4/HH [212]. Jeon et al., in the context of ovarian cancer, reported that MSCs, in response to lysophosphatidic acid secreted by the tumor, stimulated angiogenesis via an upregulation of VEGF and SDF-1 expression [213]. In the case of “oral squamous cell carcinoma”, MSCs mediated angiogenesis by driving the excessive production of MMP-1 [214].

3.3. Cell Survival

Tumor sites are characterized by, among other things, hypoxia, malnutrition, and inflammation. MSCs play a crucial role in maintaining tumor cells through the release of numerous pro-survival and anti-apoptotic factors, such as VEGF, FGF-β, PDGF, TGF-β, and SDF-1α [215]. Interestingly, PDGF and TGF-β can stimulate the expression of VEGF and FGF-β, which in turn mediate the production of Bcl-2, an anti-apoptotic protein [216,217]. Also, in BM cancer cells, MSCs upregulated Bcl-2 and boosted survival rates in the context of acute myeloid leukemia [218]. In colorectal cancer, tumor cells exposed to a conditioned medium of MSCs involved a downregulation of the expression of apoptosis-related proteins Bax and p53 and an upregulation of Bcl-2. The authors also observed a significant percentage of cancer cells in the S phase [219]. Finally, Burger et al. demonstrated that SDF-1α protected leukemic cells against spontaneous apoptosis [220].

3.4. Metastasis

The metastasis process, which is found in advanced-stage cancer, is often associated with a poor prognosis. Schaffer et al. discriminated two distinct phases occurring during the metastasis process: The first phase involves the physical translocation of a tumor cell from the primary site through the vasculature to a distant site and the second phase concerns the ability of the tumor cell to form metastasis at this distant site [221]. Numerous studies reported the implication of MSCs in the metastatic process [218,222,223]. These effects are mediated in particular by the growth factors, cytokines, and chemokines secreted by MSCs (Figure 4). Studies evaluating the role of local resident MSCs in the tumor microenvironment reported a pro-tumorigenic role. Indeed, MSCs act like PCs and promote the extravasation of melanoma cancer cells, particularly in BM and liver [224]. Importantly, it is known that during melanoma metastasis, the skeleton and liver are part of the preferred organs for cancer dissemination with, respectively, up to 20% of patients having liver metastasis and 17% having bone metastasis. MSCs seem to regulate tumor formation and growth as well as the process of metastasizing [225,226].
As demonstrated in a mouse xenograft model of melanoma and breast cancer, MSCs secreted chemokines, namely CCL2, CCL5, CCL9, and CXCL10, which activated tumor cell migration and invasion, promoting the development of lung, bone, and lymph node metastases [158,227,228,229]. In addition, CCL5 and CCL9 chemokines led to the activation of MMPs, which induced remodeling of the ECM, promoting the migration of tumor cells from the primary site [158]. Moreover, Rodini et al. reported that paracrine signaling via TGF-β1 increased cancer cell migration and metastasis in glioblastoma [230]. Accordingly, it has also been shown in an in vitro 3D co-culture model that MSCs caused the elongation and directional migration of invasive breast cancer cells [231]. These effects were mediated through the secretion of TGF-β1 by MSCs. Additionally, Lin and co-authors demonstrated on breast cancer cells (MCF7) that MSC-derived exosomes promoted cancer cell migration through an upregulation of the Wnt and β-catenin signaling pathways [232]. Moreover, a study on hepatocellular carcinoma (HCC) models showed that HCC cells exposed to a conditioned medium of MSCs extracted from HCC were associated with an increase in cell proliferation, invasion, and metastasis through the overexpression of DNM3OS (IncRNAs) and the DNM3OS interaction with the KDM6B/TAM1 axis [233]. Importantly, MSCs also promote metastasis by activating EMT through the secretion of growth factors and cytokines, including TNF-α, HGF, EGF, PDGF, and TGF-β. These factors induced transcriptional regulators, such as Snail, Slug, Twist, and Zeb1 [234,235]. Indeed, the TNF-α and TGF-β1 secreted by MSCs increased tumor cell EMT in colon carcinoma cells via the activation of the p38 MAPK and ERK signaling pathways [236]. The effect of MSCs on the phenotype of EMT tumor cells was also observed via the differentiation of MSCs into CAFs [123,237]. The MSCs notably produced fibroblast-derived factors such as α-SMA, tenascin-C, fibroblast surface protein, IL-6, CCL5, SDF-1, MMP-2, and MMP-9 [122,159].
Finally, MSCs were able to enhance the metastasis process by physically binding to tumor cells, forming a hetero-cellular metastatic unit. Indeed, the fusion of MSCs with the cancer cells of mammary, ovarian, and lung cells were confirmed in in vitro and in vivo models [132,238]. Importantly, Fan et al. identified MSCs as being able to form a complex with tumor cells in patients with ovarian cancer [239]. Thus, MSCs, by interacting with cancer cells, give rise to cancer cells with mesenchymal phenotyping, such as CSCs and CAFs, which strongly contributes to the development of metastases through various interplays with MSCs and other components of the TME [122,132,222].

4. Mesenchymal Stem Cells as a Therapeutic Tool against Cancer Progression

Currently, the literature data are divided on MSCs’ therapeutic potential. As previously explained, MSCs play a dual role in cancer progression, with a pro-tumorigenic phenotype (MSCs-2) and an anti-tumorigenic phenotype (MSCs-1). In the pathophysiological context of cancer, a growing interest is occurring in the therapeutic potential of MSCs to get over the limitations of current therapy or improve their effects. MSCs have some interesting properties, such as their ability to selectively migrate to local and distant tumoral sites and penetrate the tumor stroma after intravenous injection, as demonstrated in pancreas, melanoma, lung, or breast cancer animal models [86,87,88]. As a vehicle, MSCs may deliver drugs, genes, or other therapeutic agents such as cytokines in various cancers, promoting therapeutic properties, comprising anti-tumor, -proliferative, -inflammatory, -oxidative, or -metastatic effects [240,241]. Numerous studies enumerate the potential benefic effects of MSCs transduced/transfected with viral/non-viral constructs involving the overexpression of therapeutic agents [241,242,243].

4.1. Resistance to Treatments

Tumor resistance to current therapies is a real problem in the fight against cancer. Indeed, the management of cancerous diseases involves several approaches, including chemotherapies, radiotherapy, targeting therapies, and, more recently, immunotherapies. These approaches can be used alone, sequentially, or concurrently. Therapies aim to promote the death of tumor cells by various mechanisms. However, tumors are able to develop ways to protect themselves from most treatments, making them resistant. This tumor resistance is due to gene alterations in tumor cells, but also the activity of the TME [244,245].

4.1.1. Role of Mesenchymal Stem Cells in Tumor Resistance and Therapeutic Interest

MSCs have been shown to be partly responsible for treatment resistance through their activity in the TME [246,247,248]. It has been shown by Lis et al. that tumor-residing MSCs in ovarian cancer induced thermo-tolerance in cancer cells via activation of the SDF-1-CXCR4 axis. They also demonstrated that targeting the interaction between MSCs and cancer cells by blocking SDF-1 could restore the hyperthermia sensitivity of ovarian cancer cells [249]. Moreover, Dreuw and co-authors demonstrated that fibroblasts, resulting from the differentiation of MSCs, mediated IL-6 secretion, leading to the overexpression of Bcl-2 and Bcl-xL in tumor cells and an inhibition of apoptosis following anti-cancer treatment [250,251]. Accordingly, Gilbert et al. demonstrated that IL-6 contributed to resistance to cisplatin treatment [251]. Indeed, circulating MSCs were activated by cisplatin-based chemotherapy in tumor-bearing mouse models. The MSCs secreted two poly-unsaturated fatty acids (12-oxo-5,8,10-heptadecatrienoic acid [KHT] and hexadeca-4,7,10,13-tetraenoic acid [16:4 (n-3)]) that protect tumor cells from the cytotoxic effects of the drug [252]. Finally, Marithasan et al. examined tumors from a cohort of patients with metastatic urothelial cancer treated with an anti-PD-L1 agent (atezolizumab) and identified the main determinants of clinical outcome. The absence of response to drugs was associated with the release of TGF-β by peri-tumoral fibroblasts and an absence of CD8+ T cells in the tumor parenchyma. T cells were found in the peri-tumoral stroma, rich in fibroblasts and collagen. They also demonstrated, using a mouse model, that therapeutic co-administration of anti-TGF-β and anti-PD-L1 blocking antibodies reduced TGF-β signaling in stromal cells, allowing the release of TGF-β in the tumor center that promoted an anti-tumor immune response and tumor regression [253].
As mentioned above, MSCs have the ability to differentiate into CSCs. It is also suggested that tumor cells that undergo TEM are considered as CSCs or behave like CSCs [254,255]. These CSCs are a key factor in the resistance to therapies. Indeed, they can induce resistance to cisplatin and promote tumor relapse after the discontinuation of treatment [129]. MSCs are known to support a niche of CSCs contributing to drug resistance [256,257]. Of note is that PGE2, known to induce the autocrine expression of IL-6, IL-8, and CXCL1, promotes CSC formation [258].
Additionally, Skolekova and co-authors demonstrated that the pre-exposure of MSCs to cisplatin altered the phosphorylation of several tyrosine kinases, including STAT3, c-Jun, WNK-1, and p53, promoting the production of IL-6 and IL-8 and MSC survival, leading to tumor cell chemoresistance [259]. On the other hand, Kabashima et al. reported that CAFs maintained CSCs by secreting the Notch ligand Jagged-1, thereby promoting drug resistance [260].

4.1.2. Radioresistance and Mesenchymal Stem Cells

Radiotherapy is a treatment modality that can be used to treat about 60% of patients with cancerous diseases. Radiation therapy, known as a great technological progress over the past 20 years, allows a more precise delivery of radiation while sparing healthy tissue surrounding the tumor. Nevertheless, the radioresistance of certain cancers remains a major problem with this therapeutic application. A main factor leading to radioresistance is linked to CSCs’ activities in the TME. CSCs are dramatically implicated in the metastasis process, relapses, and the failure of radiotherapy, and are associated with a poor prognosis for patients [261,262]. The intrinsic radioresistance of CSCs can be attributed to several factors [263]. In the setting of breast cancer, Philipps and his team associated the radioresistance of CSCs with a lack of oxidative stress generated due to the high ability of CSCs to scavenge free radicals and activate mechanisms of DNA repair [264]. Also, other studies demonstrated a repopulation of CSCs after radiotherapy through the activation of the Wnt/β-catenin signaling axis that promotes self-renewal [265]. Moreover, in the context of mucoepidermoid carcinoma, CSCs’ radioresistance was associated with the activation of the NFκB signaling pathway [266]. In addition to these mechanisms, several studies reported the activation of survival signaling pathways, such as the anti-apoptotic Bcl-2 protein and PI3K/Akt/mTOR signaling pathways, which also contributed to CSCs’ radioresistance [266,267]. Radiation therapy can also induce EMT in tumor cells, thereby increasing their motility and invasive abilities in several cancers, including breast, lung, liver, and glioma [268,269,270].

4.2. Mesenchymal Stem Cells as a Therapeutic Vector

Over the past two decades, interest in MSCs was focused on their potential to migrate to tumor sites. Nevertheless, research teams remain cautious about their use due to their predominantly pro-tumorigenic characteristics. However, the inherent tumor tropism of MSCs gives them an interesting selective power for depositing anti-tumor agents thereon. Also, the development of genetic engineering techniques has allowed the control of the release of anti-tumor agents in situ [271,272,273]. These agents are, in particular, proteins, suicide genes, nanoparticles, or even oncolytic viruses, as presented in Figure 6. Of note is that an increasing number of studies reported that MSC-derived exosomes can also be used as anti-tumor agents. The approach around exosomes will not be developed in this present review.

4.2.1. Types of Vectoring Making Use of Mesenchymal Stem Cells

  • Therapeutic macromolecules
The therapy with bioactive molecules was used for the first time in 2002, with a modification involving an overexpression of IFN-β by engineered MSCs [274]. Afterwards, MSCs were modified to express cytokines, chemokines, interferons, and pro-apoptotic or anti-angiogenic factors, as presented on Figure 6. MSCs are able to deliver a variety of cytokines and growth factors to the tumor site that were identified as anti-tumor agents, comprising IFN-α, IFN-β, IFN-γ, IL-2, IL-7, IL-12, IL-15, IL-18, IL-25, and NK4, an antagonist of HGF [208,275,276]. Mirlekar and co-authors, in an in vivo study on renal cell carcinoma and cervical tumors, showed that MSCs expressed IL-12 and inhibited tumor proliferation [277,278]. Moreover, IL-12 is able to activate T cells and cytotoxic NK cells [279]. A particular interest was also taken in the anti-tumorigenic effect of the factor-related apoptosis-induced ligand (TRAIL) [99,280,281].
2.
Gene-directed enzyme prodrug therapy: suicide genes
A therapeutic approach using MSCs is based on genetic modifications allowing them to express suicide genes [282]. Among the strategies used for an anti-tumor role based on suicide genes and delivered by MSCs, we can cite three majors factors used: TK (GCV), cDy::UPRT (5-FC), and CD (5-FC) [243,275,282,283].
An in vivo study on prostate cancer used MSCs designed to express cytosine deaminase::uracil phosphoribosyltransferase (CD::UPRT), which is a transgene produced from suicide gene expression. They showed that modified MSCs inhibited tumor growth. Indeed, CD::UPRT was able to convert the relatively non-toxic 5-fluorocytosine (5-FC) into toxic 5-fluorouracil (5-FU) [284]. Another in vivo study on colon cancer demonstrated the same effectiveness concerning genetically modified MSCs [285]. Moreover, in the treatment of bone metastases, Segaliny and co-authors demonstrated that MSCs designed to overexpress P-selectin glycoprotein ligand-1 (PSGL-1)/Sialyl-Lewis X (SLEX) and modified versions of cytosine deaminase (CD) and osteoprotegerin (OPG), had better homing toward metastatic bone tumor cells, possibly due to interactions between the PSGL-1/SLEX expressed by MSCs and the P-selectin expressed by endothelial cells at the vascular level of the tumor. Thus, the authors showed that modified MSCs could effectively eliminate tumor cells while preserving bone integrity [286].
3.
Oncolytic virus
Another strategy in the development of anti-tumor MSCs consists of using cells to deliver an oncolytic virus (OV) [287]. Indeed, over the past 20 years, after that oncorine was validated as the first virus drug [288], OVs were studied as therapy in the fight against cancer [289,290,291]. This new anti-tumor approach is based on the use of viruses able to replicate inside tumor cells to kill them. Nevertheless, not all viruses can be used as OV; they need to have particular characteristics, such as the ability to specifically target tumor cells, low toxicity, a significant capacity to destroy tumor cells by viral proliferation, an oncolytic and targeting potential enhanced by structural and genetic modification, as well as the ability to carry transgenes in order to elevate their therapeutic potential [292,293]. Thus, several viruses from various viral families were studied [289,294]. As reported by Darestani et al., numerous preclinical studies demonstrated efficient features for MSCs as OV carriers. Among the OVs that can be transported by MSCs, we could cite adenovirus (Ad), herpes simplex virus (HSV), measles virus (MV), myxoma virus (MYXY), Newcastle disease virus (NDV), reovirus [295,296], as well as alphaviruses [24,297].
This therapeutic approach mainly exerts its effect through the direct oncolysis caused by the virus cell proliferation and tissue spreading [298]. Subsequently, the OV will promote antiviral immune responses (e.g., Interferon type I), leading to the death of infected tumor cells. Nevertheless, OV treatment largely depends on their ability to reach the tumor site without being recognized by the immune system [299,300,301,302,303,304].
In a study on glioma, the authors reported that NDV-carrying MSCs enhanced apoptosis and inhibited tumor cell proliferation and that the TRAIL protein played an important role in the cytotoxicity of the MSCs carrying NDV [305]. Moreover, a study on ovarian cancer demonstrated that groups treated with MSCs carrying MV had better survival rates compared to groups treated with the virus alone [306]. Interestingly, Zhan et al. also showed in glioma that MSCs carrying oncolytic recombinant adenovirus enriched with IL-24 and endostatin, and regulated by doxycycline, increased the anti-tumor response by inhibiting tumor growth [307]. It has also been demonstrated in glioma that MSCs carrying CRAD (CRAD-S-pk7), which is an adenovirus, potentialized anti-tumor activities and induced a significant decrease in tumor volume compared to the group administered with the virus alone [308,309]. M1 is an oncolytic alphavirus which has been validated for possible treatments of breast and brain cancers [310].
4.
Metallic nanoparticles
Recently, radiosensitizing agents coupled to MSCs demonstrated interesting anti-tumor properties [311]. Indeed, in order to sensitize tumor cells to ionizing radiation and to counteract cell radioresistance, metallic nanoparticles (NPs) exhibiting elements with a high atomic number (Z), such as gold (Au), hafnium (Hf), or bismuth (Bi), were used as radiosensitizers [312,313,314,315]. Nanoparticles could be used to potentialize radiation toxic effects. Indeed, they can accumulate in tumor areas by passive targeting through the enhanced permeability and retention (EPR) effect, by active targeting mediated by ligand–receptor interaction, or by a combination of passive and active targeting [316,317]. However, the therapeutic results of these new nano-radiosensitizers remain problematic in the case of inadequate targeting [318,319]. The non-viral strategy using nanoparticle-based gene delivery via bioengineer stem cells could offer a new theragnostic perspective. Interestingly, Huang and team used a magnetic ternary nanohybrid (MTN) system carried by superparamagnetic iron oxide nanoparticles to construct MSCs expressing TRAIL using magnetic force and receptor dual targeting. They imaged the MTN-transfected hMSCs using MRI and observed an apoptosis activation as well as a decrease in the cancer progression and an increase in the survival rate on the mice model of glioma [320].

4.2.2. Therapeutic Potential of Mesenchymal Stem Cells in the Case of Melanoma

Melanoma is a malignant tumor of melanocytes (derived from the neural crest) that represents the third most common cutaneous malignancy. Globocan estimated 325,000 new cases in 2020, 1.7% of global cancer diagnoses [321]. Melanoma management at an early stage leads to 93.5% 5-year relative survival, with the most common management coming via surgical resection. However, depending on the cancer stage at diagnosis, the prognosis can become poorer due to a high risk of melanoma metastasis beyond its primary site. Indeed, metastatic melanoma is the most mortal type of cancer with a 5-year relative survival reaching 4.7% in the worst cases, when the diagnosis is completed on older age patients for a distant melanoma [322]. This low survival rate is mainly due to anticancer agent resistance [323]. Indeed, it is important to consider the immunologic and genetic profile of patients to choose the most appropriate therapy or combination of therapies in order to keep the clinical response as long as possible [324,325].
Numerous studies reported the promising effects of genetically modified MSCs in order to overexpress therapeutic agents. A study reported that MSCs designed to express TNF-α could trigger the caspase-3/-7-dependent apoptosis of human A375 melanoma cells. In this study, the authors showed that a subcutaneous injection of A375 melanoma cells and genetically modified MSCs, in a ratio of 1:4, could induce 97.5% tumor regression in mice [326]. Interestingly, Petrov et al. used adenovirus comprising six vectors encoding IFN-α in a lung metastasis model of melanoma and demonstrated that modified MSCs significantly reduced the growth of melanoma cells and significantly prolonged the survival of mice through an increase in apoptosis and a decrease in proliferation and angiogenesis [327]. It has also been demonstrated that MSCs expressing thrombospondin-1 (TSP-1) in a metastasis melanoma model decrease nodule apparition and exerted an anti-tumor effect via an activation of T cells, an upregulation of Bax pro-apoptotic protein, and a downregulation of Bcl-2 anti-apoptotic protein [88]. Importantly, the interest in using MSCs as a vehicle is their capacities to target tumor stroma combined with their longevity in the tumor microenvironment as well as their lineage differentiation [52].
Another study reported the interesting role of MSCs derived from adipose tissue as a tool for delivering therapeutic genes to tumor cells. MSCs were used to deliver an anti-cancer drug, namely cisplatin, coupled with a lentiviral vector expressing IFN-β cytokine for its anti-proliferative, immunomodulatory, and anti-angiogenic activities [328]. MSCs genetically designed to produce IFN-β have been used against other cancers and demonstrated interesting effects. Authors reported anti-proliferative and pro-apoptotic effects on tumor cells [329]. It has been shown that IFN-β was able to inhibit the proliferation of melanoma tumor cells and to modulate the immune response [330]. In addition, a tumor-specific antibody coupling with IFN-β has been shown to be more effective than antibody treatment alone [331]. Finally, Yang et al. demonstrated that sequential IFN-β and cisplatin treatment improved the adaptive immune response [332].
Authors demonstrated in a mouse melanoma model that the association of therapeutic agents led to an increase in survival rate and inhibited the growth of melanoma compared to the control group and the group treated with cisplatin alone. Accordingly, other articles report the synergic effect between therapeutic agents and MSCs’ protective properties [283]. However, studies using MSCs alone exposed their beneficial effects, such as low immunogenicity and immunosuppressive, anti-tumor, anti-proliferative, anti-inflammatory, or antioxidant properties, depending on the model studied [7,333]. MSCs can induce a phenotype reprogramming of microenvironment resident cells, such as macrophages, in a mouse model of melanoma injected with xenogeneic bone marrow MSCs [334].
Çakır and co-authors reported similarities in the phenotype and function of MSCs and melanoma-associated fibroblasts (MAFs) [335]. Indeed, they observed an increase in macrophage-secreted IL-10 under the control of MAFs which were in intimate contact with macrophages in excised melanoma from two different patients. The MAFs caused immunosuppression by promoting the phenotype reprogramming of macrophages. The authors also showed that COX and IDO were implicated in this immunomodulation and demonstrated their critical role in cancer-progression-promoting immunosuppression. The inhibition of COX-1, 2, or IDO led to limited MAF–macrophage interactions and the modulation of intra-tumoral immune responses. The use of MSCs as vectors for COX and/or IDO inhibitors could be strategic for treating melanoma patients.
Importantly, MSCs as a deliverer of gene-directed enzyme prodrug molecular chemotherapy could also open a new era in therapeutic approaches using suicide genes, by avoiding the gastrointestinal toxicities related to chemotherapeutic drug absorption, such as the 5-Fluorouracil. Kucerova and co-authors explored the effect of a combination of cytosine deaminase with prodrug 5-fluorocytosine, delivered by MSCs derived from adipose tissue [336]. In their study, they selectively produced a cytotoxic agent in situ that caused a complete tumor regression in 89% of tumors and a tumor growth inhibition on a melanoma mice model co-injected with modified MSCs and tumor cells. In an in vivo model of lung metastases of melanoma, the MSCs carrying a recombinant oncolytic MYXY virus expressing IL-5 improved animal survival rate. Interestingly, the authors reported, concerning the anti-cancer mechanism of the proposed therapy, an increase in the level of natural killer (NK) cells recruited during the activation of CD8+ T cells. Moreover, mice who received MSCs coupled to the OV exhibited an increase in the release of pro-inflammatory cytokines, namely IL-15, IL-1β, IFN-γ, TNF-α, PD-1, and PD-L1, compared to mice who received virus alone [285].
Moreover, the therapeutic effects of BM-derived MSCs expressing CMV-thymidine kinase (TK) on pulmonary melanoma metastasis combined with the prodrug ganciclovir were investigated [243]. The MSCs were engineered through a non-viral gene vector. In vitro results displayed that the engineered MSCs caused a significant anti-tumor effect in the presence of ganciclovir in a dose-dependent manner. In vivo studies confirmed these beneficial effects on a metastasis tumor model.
However, the literature data report contradictory results about role of MSCs in tumor progression. The ambivalent role of MSCs in tumor suppression and progression needs to be understood before envisaging their use as cancer therapy. Indeed, depending on the cancer types or research models used, MSCs exhibit pro- and anti-tumorigenic potential depending on their interplays with resident cancer cells in the microenvironment. Importantly, MSCs injected as a vehicle for therapeutic agents or injected alone exhibit a protective role and anti-tumor effect in most in vivo and in vitro studies about melanoma cancer. These therapeutic strategies could be really promising for targeting tumor cells and help in the case of radioresistance for all types of resident cells in tumor parenchyma, particularly concerning CSCs.

4.2.3. Therapeutic Potential of Mesenchymal Stem Cells in the Case of Lung Cancer

The Globocan epidemiologic study reported 2,206,771 new cases of lung cancer in 2020, i.e., 11.4% of all identified cancer cases. Lung cancer is the most common cancer and the deadliest worldwide, with the number of deaths from lung cancer reaching 18% in 2020 [321]. There are different types of lung cancers, namely non-small-cell lung cancer (about 80% to 85% of all lung cancers), small-cell lung cancer (about 10% to 15% of cases), and other tumors occurring in the lungs (5% of cases). Currently, lung cancer management includes surgery, radiotherapy, chemotherapy, targeted therapy, or immunotherapy. Despite the combination of major therapeutic approaches, the 5-year survival of non-small-cell lung cancer remains at 8–15% [337]. The dramatic incidence of lung cancer, as well as the poor prognosis, pushes health professionals and researchers to imagine new therapeutic approaches to improve lung cancer management.
Various studies demonstrated that MSCs promoted tumor growth, metastasis, and neovascularization during lung cancer [156,338,339]. Thus, MSCs used as vehicles for therapeutic agent delivery represent a seductive strategy and a tool to target tumor sites, considering their homing capacity [340]. Importantly, engineered MSCs labeled with superparamagnetic iron oxide nanoparticles were used to track MSC localization via MRI [341]. It has been reported that after the intravenous injection of superparamagnetic-iron oxide-loaded MSCs in a multiple lung metastasis mouse model, modified MSCs were localized in tumor sites. The homing ability of MSCs makes them good candidates as vectors for therapeutic agents. It has been demonstrated in numerous studies that the expression of TRAIL by MSCs induced apoptosis of tumor cells in the context of colorectal, lung, breast, brain, and cervix cancers [342,343,344]. Interestingly, the most studies on the anti-tumor role of engineered MSCs in lung cancer relate to TRAIL. Loebinger et al. hypothesized that MSCs modified to produce and deliver TRAIL at tumor sites, a transmembrane protein that mediates apoptosis, could cause the death of tumor cells [343]. The authors engineered human MSCs transduced with TRAIL and the IRES-eGFP reporter gene, under the control of a tetracycline promoter using a lentiviral vector, and tested their anti-tumor effect on a lung cancer model. The MSC model homed to tumor sites, significantly reduced tumor growth, and completely cleared the metastatic disease in 38% of mice with tumors compared to 0% of controls.
Moreover, another study targeted MSCs to abrogate their deleterious initial role in cancer progression by silencing their TGF-β1 expression, which reverted their exosome-mediated EMT and enhanced the anti-proliferative and pro-apoptotic effect of MSCs in lung cancer [345]. By blocking TGF- β1 expression, they deactivated the Smad2/3, Akt/GSK-3β/β-catenin, NF-κB, ERK, JNK, and p38 MAPK signaling pathways, which are known to promote cancer progression. Moreover, mouse MSCs transduced with adenoviral vector expressing CX3CL1, injected into mice bearing lung metastases, mediated a decrease in the development of lung metastases and an improved survival rate [346]. The anti-tumor effects reported were related to both innate and adaptive immune modulation through an increase in infiltration of CD8+ lymphocytes, CD4 lymphocytes, and NK cells.
It has also been demonstrated that MSCs loaded with a doxorubicin (DOX)–polymer conjugate were able to penetrate tumor sites, inhibit tumor growth, and exhibit better efficacy than the drug alone [347]. Major studies on the anti-tumor role of engineered MSCs are related to pro-apoptotic protein delivery. Finally, Hakkarainen et al. explained that systemic adenoviral delivery lacked efficacy due to liver sequestration of the intravenously administered virus [348]. They used MSCs transduced by the OV to avoid liver sequestration. They noted in an in vivo model of lung cancer that intravenously injected-MSCs homed primarily to the lungs, that the virus delivered by the MSCs was released into advanced lung tumors enhancing mice survival rate, and that the same dose of the virus injected alone was found in the liver. MSCs are of high interest for their homing capacity as well as for improving the bioavailability of systemically administered oncolytic adenoviruses. Importantly, Ando and co-authors designed a suicide system via MSCs based on an inducible caspase-9, activated using a specific chemical inducer of dimerization (CID), for adenoviral-based delivery to lung tumors to sensitize non-small-cell lung cancer to apoptosis [349]. Their engineered MSCs coupled with CID-activated iC9 were combined or not with bortezomib, an inhibitor of the proteasome. The combination of the MSC-based delivery of the iC9 suicide gene to human non-small-cell lung cancer and bortezomib enhances the anti-tumor activity and the addition of bortezomib offered a greater reduction of tumor size.

5. Conclusions

MSCs play a major role in cancerous diseases. Understanding their behavior in this setting may be the key to the fight against cancer. Nevertheless, their use has long been debated due to all the diverging results about MSCs and cancer. The literature data describe both anti- and pro-tumorigenic effects. Moreover, the tumor microenvironment is a heterogeneous complex, and a dynamic system comprising stromal cells with different morphological and functional activities. Despite this reality, the latest therapies based on MSCs bring new hope to cancer patients by offering more effective anti-tumor treatments. Indeed, the use of MSCs as vectors for therapeutics represents an important advance for more specific and personalized treatments. Currently, it seems crucial to better understand the interactions between the different subsets of MSCs, cancer cells, as well as tumor microenvironment components to propose highly specific therapeutic approaches. From a pharmacological point of view, it may be important to address the drug treatment that may favor the phenotype switch between the different subsets of MSCs (e.g., MSCs-1 versus MSCs-2) and/or dedifferentiation into physiological PC-like cells.

Author Contributions

Conceptualization, Y.S., A.A. and P.G.; methodology, Y.S., F.A.-P., M.K., A.A., M.B., F.D. and P.G.; validation, Y.S., F.A.-P., M.K., A.A., M.B., F.D. and P.G.; writing—original draft preparation, Y.S., F.A.-P., M.K., A.A., M.B., F.D. and P.G.; writing—review and editing, Y.S., F.A.-P., M.K., A.A., M.B., F.D. and P.G.; visualization, Y.S., A.A. and P.G.; supervision, P.G.; funding acquisition, P.G. All authors have read and agreed to the published version of the manuscript.

Funding

PG and EPI team members received financial support from the EPIGEN project research (grant number CPER FEDER 2019221-0022768).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

Ad: adenovirus; AKT: protein kinase B; ANPEP: alanyl aminopeptidase; α-SMA: α-smooth muscle actin; Bcl: B-cell lymphoma/leukemia; BM: bone marrow; BMP4/HH: bone morphogenetic protein 4/hedgehog; C3a: cleaved complement component 3 anaphylatoxin; CAAs: cancer-associated adipocytes; CAFs: cancer-associated fibroblasts; c-AMP: cyclic adenosine monophosphate; CCL: CC chemokine ligand; CD: cluster of differentiation; CNS: central nervous system; COL: collagen; Cox-2: cyclooxygenase-2; CSCs: cancer stem cells; CSF1: macrophage colony-stimulating factor 1; Cspg4: chondroitin sulfate proteoglycan 4; CXCL: C-X-C motif ligand; CXCR: C-X-C motif chemokine receptor; DDR2: discoidin domain receptor tyrosine kinase 2; DUSP1: dual-specificity phosphatase 1; ECM: extracellular matrix; EGF: epidermal growth factor; EMT: epithelial–mesenchymal transition; EndEMT: endothelial–mesenchymal transition; EPR: enhanced permeability and retention; ERK; extracellular signal-regulated kinase; FASL: fas ligand; FGF-β: basic fibroblast growth factor; FSP1: fibroblast-specific protein 1; GDNF: glial-cell-line-derived neurotrophic factor; GF: growth factor; GM–CSF: granulocyte–macrophage colony-stimulating factor; HCC: hepatocellular carcinoma; HGF: hepatocyte growth factor; HIF-1: hypoxia-inducible factor-1; HLA-DR: human leukocyte antigen-DR isotype; HSV: herpes simplex virus; IDO: indoleamine-2,3-dioxygenase; IFN: interferon; IGF: insulin growth factor; IL: interleukin; iNOS: inducible nitric oxide synthase; ISCT: international society for cellular therapy; JAK: janus kinase; lncRNAs: long non-coding RNAs; MAFs: melanoma-associated fibroblasts; mIR/miRNA: microRNA; MMPs: matrix metalloproteinases; MSC(s): mesenchymal stem cell(s); MTN: magnetic ternary nanohybrid; mTOR: mammalian target of rapamycin; MV: measles virus; MYXY: myxoma virus; Mφ: macrophage; NDV: Newcastle disease virus; NFkB: nuclear factor-κB; NG2: neural/glial antigen 2; NK: natural killer; NO: nitric oxide; NPs: metallic nanoparticles; OV(s): oncolytic virus; p38 MAPK: p38 mitogen-activated protein kinase; PCs: pericytes; PD-1: programmed cell death-1; PDGF: platelet-derived growth factor; PDGFRβ: platelet-derived growth factor receptor β; PD-L1: programmed cell death ligand 1; PGE2: prostaglandin E2; PI3K: phosphoinositide 3-kinase; pMSCs: perivascular mesenchymal stem cells; PSGL-1: p-selectin glycoprotein ligand-1; RANTES: regulated on activation, normal T cell expressed and secreted; SCF: stem cell factor; SDF-1: stromal cell-derived factor 1; Smad: small mothers against decapentaplegic; Src: proto-oncogene tyrosine-protein kinase; STAT: signal transducers and activators of transcription; STRO-1: stromal cell surface marker-1; TAFs: tumor-associated fibroblasts; TAM1/2: tumor-associated macrophage 1/2; TAN1/2: tumor-associated neutrophil 1/2; TEM1: tumor endothelium marker 1; TGF-β: transforming growth factor-β; TGFBR: transforming growth factor b receptor; TME: tumor microenvironment; TNF-α: tumor necrosis factor-α; TNFR: tumor necrosis factor receptor; TRAIL: tumor-necrosis-factor-related apoptosis-inducing ligand; VCAM: vascular cell adhesion molecule; VEGF: vascular endothelial growth factor; vFB: vascular fibroblasts; vSMCs: vascular smooth muscle cells.

References

  1. Uccelli, A.; Moretta, L.; Pistoia, V. Mesenchymal Stem Cells in Health and Disease. Nat. Rev. Immunol. 2008, 8, 726–736. [Google Scholar] [CrossRef]
  2. Bianco, P.; Robey, P.G.; Simmons, P.J. Mesenchymal Stem Cells: Revisiting History, Concepts, and Assays. Cell Stem Cell 2008, 2, 313–319. [Google Scholar] [CrossRef] [PubMed]
  3. Tolar, J.; Le Blanc, K.; Keating, A.; Blazar, B.R. Concise Review: Hitting the Right Spot with Mesenchymal Stromal Cells. Stem Cells 2010, 28, 1446–1455. [Google Scholar] [CrossRef] [PubMed]
  4. Keating, A. Mesenchymal Stromal Cells: New Directions. Cell Stem Cell 2012, 10, 709–716. [Google Scholar] [CrossRef] [PubMed]
  5. Lazennec, G.; Lam, P.Y. Recent Discoveries Concerning the Tumor—Mesenchymal Stem Cell Interactions. Biochim. Biophys. Acta (BBA)—Rev. Cancer 2016, 1866, 290–299. [Google Scholar] [CrossRef]
  6. Ridge, S.M.; Sullivan, F.J.; Glynn, S.A. Mesenchymal Stem Cells: Key Players in Cancer Progression. Mol. Cancer 2017, 16, 31. [Google Scholar] [CrossRef]
  7. Papaccio, F.; Paino, F.; Regad, T.; Papaccio, G.; Desiderio, V.; Tirino, V. Concise Review: Cancer Cells, Cancer Stem Cells, and Mesenchymal Stem Cells: Influence in Cancer Development. Stem Cells Transl. Med. 2017, 6, 2115–2125. [Google Scholar] [CrossRef]
  8. Poggi, A.; Varesano, S.; Zocchi, M.R. How to Hit Mesenchymal Stromal Cells and Make the Tumor Microenvironment Immunostimulant Rather Than Immunosuppressive. Front. Immunol. 2018, 9, 262. [Google Scholar] [CrossRef]
  9. Leuning, D.G.; Beijer, N.R.M.; du Fossé, N.A.; Vermeulen, S.; Lievers, E.; van Kooten, C.; Rabelink, T.J.; de Boer, J. The Cytokine Secretion Profile of Mesenchymal Stromal Cells Is Determined by Surface Structure of the Microenvironment. Sci. Rep. 2018, 8, 7716. [Google Scholar] [CrossRef] [PubMed]
  10. Sahai, E.; Astsaturov, I.; Cukierman, E.; DeNardo, D.G.; Egeblad, M.; Evans, R.M.; Fearon, D.; Greten, F.R.; Hingorani, S.R.; Hunter, T.; et al. A Framework for Advancing Our Understanding of Cancer-Associated Fibroblasts. Nat. Rev. Cancer 2020, 20, 174–186. [Google Scholar] [CrossRef]
  11. Friedenstein, A.J.; Piatetzky-Shapiro, I.; Petrakova, K.V. Osteogenesis in Transplants of Bone Marrow Cells. J. Embryol. Exp. Morphol. 1966, 16, 381–390. [Google Scholar] [CrossRef] [PubMed]
  12. Friedenstein, A.J.; Chailakhjan, R.K.; Lalykina, K.S. The development of fibroblast colonies in monolayer cultures of guinea-pig bone marrow and spleen cells. Cell Prolif. 1970, 3, 393–403. [Google Scholar] [CrossRef] [PubMed]
  13. Caplan, A.I. Mesenchymal Stem Cells. J. Orthop. Res. 1991, 9, 641–650. [Google Scholar] [CrossRef] [PubMed]
  14. Goshima, J.; Goldberg, V.M.; Caplan, A.I. The Osteogenic Potential of Culture-Expanded Rat Marrow Mesenchymal Cells Assayed In Vivo in Calcium Phosphate Ceramic Blocks. Clin. Orthop. Relat. Res. 1991, 262, 298–311. [Google Scholar] [CrossRef]
  15. Spees, J.L.; Olson, S.D.; Ylostalo, J.; Lynch, P.J.; Smith, J.; Perry, A.; Peister, A.; Wang, M.Y.; Prockop, D.J. Differentiation, Cell Fusion, and Nuclear Fusion during Ex Vivo Repair of Epithelium by Human Adult Stem Cells from Bone Marrow Stroma. Proc. Natl. Acad. Sci. USA 2003, 100, 2397–2402. [Google Scholar] [CrossRef] [PubMed]
  16. Chen, Y.; Teng, F.Y.H.; Tang, B.L. Coaxing Bone Marrow Stromal Mesenchymal Stem Cells towards Neuronal Differentiation: Progress and Uncertainties. Cell. Mol. Life Sci. 2006, 63, 1649–1657. [Google Scholar] [CrossRef]
  17. Tohill, M.; Terenghi, G. Stem-Cell Plasticity and Therapy for Injuries of the Peripheral Nervous System. Biotechnol. Appl. Biochem. 2004, 40, 17–24. [Google Scholar] [CrossRef]
  18. Horwitz, E.M.; Le Blanc, K.; Dominici, M.; Mueller, I.; Slaper-Cortenbach, I.; Marini, F.C.; Deans, R.J.; Krause, D.S.; Keating, A. Clarification of the Nomenclature for MSC: The International Society for Cellular Therapy Position Statement. Cytotherapy 2005, 7, 393–395. [Google Scholar] [CrossRef]
  19. Dominici, M.; Le Blanc, K.; Mueller, I.; Slaper-Cortenbach, I.; Marini, F.C.; Krause, D.S.; Deans, R.J.; Keating, A.; Prockop, D.J.; Horwitz, E.M. Minimal Criteria for Defining Multipotent Mesenchymal Stromal Cells. The International Society for Cellular Therapy Position Statement. Cytotherapy 2006, 8, 315–317. [Google Scholar] [CrossRef]
  20. da Silva Meirelles, L.; Caplan, A.I.; Nardi, N.B. In Search of the In Vivo Identity of Mesenchymal Stem Cells. Stem Cells 2008, 26, 2287–2299. [Google Scholar] [CrossRef]
  21. Crisan, M.; Yap, S.; Casteilla, L.; Chen, C.-W.; Corselli, M.; Park, T.S.; Andriolo, G.; Sun, B.; Zheng, B.; Zhang, L.; et al. A Perivascular Origin for Mesenchymal Stem Cells in Multiple Human Organs. Cell Stem Cell 2008, 3, 301–313. [Google Scholar] [CrossRef]
  22. da Silva Meirelles, L.; Chagastelles, P.C.; Nardi, N.B. Mesenchymal Stem Cells Reside in Virtually All Post-Natal Organs and Tissues. J. Cell Sci. 2006, 119, 2204–2213. [Google Scholar] [CrossRef]
  23. Chamberlain, G.; Fox, J.; Ashton, B.; Middleton, J. Concise Review: Mesenchymal Stem Cells: Their Phenotype, Differentiation Capacity, Immunological Features, and Potential for Homing. Stem Cells 2007, 25, 2739–2749. [Google Scholar] [CrossRef]
  24. Lebeau, G.; Ah-Pine, F.; Daniel, M.; Bedoui, Y.; Vagner, D.; Frumence, E.; Gasque, P. Perivascular Mesenchymal Stem/Stromal Cells, an Immune Privileged Niche for Viruses? Int. J. Mol. Sci. 2022, 23, 8038. [Google Scholar] [CrossRef] [PubMed]
  25. Zuk, P.A.; Zhu, M.; Ashjian, P.; Ugarte, D.A.D.; Huang, J.I.; Mizuno, H.; Alfonso, Z.C.; Fraser, J.K.; Benhaim, P.; Hedrick, M.H. Human Adipose Tissue Is a Source of Multipotent Stem Cells. Mol. Biol. Cell 2002, 13, 4279–4295. [Google Scholar] [CrossRef] [PubMed]
  26. Zuk, P.A.; Zhu, M.; Mizuno, H.; Huang, J.; Futrell, J.W.; Katz, A.J.; Benhaim, P.; Lorenz, H.P.; Hedrick, M.H. Multilineage Cells from Human Adipose Tissue: Implications for Cell-Based Therapies. Tissue Eng. 2001, 7, 211–228. [Google Scholar] [CrossRef]
  27. Sabatini, F.; Petecchia, L.; Tavian, M.; de Villeroché, V.J.; Rossi, G.A.; Brouty-Boyé, D. Human Bronchial Fibroblasts Exhibit a Mesenchymal Stem Cell Phenotype and Multilineage Differentiating Potentialities. Lab. Investig. 2005, 85, 962–971. [Google Scholar] [CrossRef] [PubMed]
  28. De Bari, C.; Dell’Accio, F.; Tylzanowski, P.; Luyten, F.P. Multipotent Mesenchymal Stem Cells from Adult Human Synovial Membrane. Arthritis Rheum. 2001, 44, 1928–1942. [Google Scholar] [CrossRef]
  29. Wang, Y.; Yu, X.; Chen, E.; Li, L. Liver-Derived Human Mesenchymal Stem Cells: A Novel Therapeutic Source for Liver Diseases. Stem Cell Res. Ther. 2016, 7, 71. [Google Scholar] [CrossRef]
  30. Short, B.; Wagey, R. Isolation and Culture of Mesenchymal Stem Cells from Mouse Compact Bone. In Basic Cell Culture Protocols; Helgason, C.D., Miller, C.L., Eds.; Methods in Molecular Biology; Humana Press: Totowa, NJ, USA, 2013; Volume 946, pp. 335–347. ISBN 978-1-62703-127-1. [Google Scholar]
  31. Lecourt, S.; Marolleau, J.-P.; Fromigué, O.; Vauchez, K.; Andriamanalijaona, R.; Ternaux, B.; Lacassagne, M.-N.; Robert, I.; Boumédiene, K.; Chéreau, F.; et al. Characterization of Distinct Mesenchymal-like Cell Populations from Human Skeletal Muscle In Situ and In Vitro. Exp. Cell Res. 2010, 316, 2513–2526. [Google Scholar] [CrossRef]
  32. Jackson, W.M.; Nesti, L.J.; Tuan, R.S. Potential Therapeutic Applications of Muscle-Derived Mesenchymal Stem and Progenitor Cells. Expert Opin. Biol. Ther. 2010, 10, 505–517. [Google Scholar] [CrossRef]
  33. Trivanovic, D.; Kocic, J.; Mojsilovic, S.; Krstic, A.; Ilic, V.; Okic-Djordjevic, I.; Santibanez, J.; Jovcic, G.; Terzic, M.; Bugarski, D. Mesenchymal Stem Cells Isolated from Peripheral Blood and Umbilical Cord Wharton’s Jelly. Srp. Arh. Celok. Lek. 2013, 141, 178–186. [Google Scholar] [CrossRef] [PubMed]
  34. Hassan, G.; Kasem, I.; Soukkarieh, C.; Aljamali, M. A Simple Method to Isolate and Expand Human Umbilical Cord Derived Mesenchymal Stem Cells: Using Explant Method and Umbilical Cord Blood Serum. Int. J. Stem Cells 2017, 10, 184–192. [Google Scholar] [CrossRef] [PubMed]
  35. Beeravolu, N.; McKee, C.; Alamri, A.; Mikhael, S.; Brown, C.; Perez-Cruet, M.; Chaudhry, G.R. Isolation and Characterization of Mesenchymal Stromal Cells from Human Umbilical Cord and Fetal Placenta. J. Vis. Exp. 2017, 55224. [Google Scholar] [CrossRef]
  36. Vellasamy, S. Isolation and Characterisation of Mesenchymal Stem Cells Derived from Human Placenta Tissue. World J. Stem Cells 2012, 4, 53–61. [Google Scholar] [CrossRef]
  37. Pelekanos, R.A.; Sardesai, V.S.; Futrega, K.; Lott, W.B.; Kuhn, M.; Doran, M.R. Isolation and Expansion of Mesenchymal Stem/Stromal Cells Derived from Human Placenta Tissue. J. Vis. Exp. 2016, 54204. [Google Scholar] [CrossRef]
  38. Li, S.; Huang, K.-J.; Wu, J.-C.; Hu, M.S.; Sanyal, M.; Hu, M.; Longaker, M.T.; Lorenz, H.P. Peripheral Blood-Derived Mesenchymal Stem Cells: Candidate Cells Responsible for Healing Critical-Sized Calvarial Bone Defects. Stem Cells Transl. Med. 2015, 4, 359–368. [Google Scholar] [CrossRef]
  39. Williams, A.R.; Hare, J.M. Mesenchymal Stem Cells: Biology, Pathophysiology, Translational Findings, and Therapeutic Implications for Cardiac Disease. Circ. Res. 2011, 109, 923–940. [Google Scholar] [CrossRef]
  40. Lv, F.-J.; Tuan, R.S.; Cheung, K.M.C.; Leung, V.Y.L. Concise Review: The Surface Markers and Identity of Human Mesenchymal Stem Cells. Stem Cells 2014, 32, 1408–1419. [Google Scholar] [CrossRef]
  41. Fonseca, L.N.; Bolívar-Moná, S.; Agudelo, T.; Beltrán, L.D.; Camargo, D.; Correa, N.; Del Castillo, M.A.; Fernández de Castro, S.; Fula, V.; García, G.; et al. Cell Surface Markers for Mesenchymal Stem Cells Related to the Skeletal System: A Scoping Review. Heliyon 2023, 9, e13464. [Google Scholar] [CrossRef]
  42. Phinney, D.G.; Sensebé, L. Mesenchymal Stromal Cells: Misconceptions and Evolving Concepts. Cytotherapy 2013, 15, 140–145. [Google Scholar] [CrossRef] [PubMed]
  43. Wood, B. Multicolor Immunophenotyping: Human Immune System Hematopoiesis. In Methods in Cell Biology; Elsevier: Amsterdam, The Netherlands, 2004; Volume 75, pp. 559–576. ISBN 978-0-12-564170-8. [Google Scholar]
  44. Watt, S.M.; Gilmore, D.J.; Davis, J.M.; Clark, M.R.; Waldmann, H. Cell-Surface Markers on Haemopoietic Precursors. Reagents for the Isolation and Analysis of Progenitor Cell Subpopulations. Mol. Cell. Probes 1987, 1, 297–326. [Google Scholar] [CrossRef] [PubMed]
  45. Ng, A.P.; Alexander, W.S. Haematopoietic Stem Cells: Past, Present and Future. Cell Death Discov. 2017, 3, 17002. [Google Scholar] [CrossRef]
  46. Denu, R.A.; Nemcek, S.; Bloom, D.D.; Goodrich, A.D.; Kim, J.; Mosher, D.F.; Hematti, P. Fibroblasts and Mesenchymal Stromal/Stem Cells Are Phenotypically Indistinguishable. Acta Haematol. 2016, 136, 85–97. [Google Scholar] [CrossRef]
  47. Kuznetsov, S.A.; Krebsbach, P.H.; Satomura, K.; Kerr, J.; Riminucci, M.; Benayahu, D.; Robey, P.G. Single-Colony Derived Strains of Human Marrow Stromal Fibroblasts Form Bone after Transplantation In Vivo. J. Bone Miner. Res. 1997, 12, 1335–1347. [Google Scholar] [CrossRef] [PubMed]
  48. Colter, D.C.; Sekiya, I.; Prockop, D.J. Identification of a Subpopulation of Rapidly Self-Renewing and Multipotential Adult Stem Cells in Colonies of Human Marrow Stromal Cells. Proc. Natl. Acad. Sci. USA 2001, 98, 7841–7845. [Google Scholar] [CrossRef]
  49. Galland, S.; Stamenkovic, I. Mesenchymal Stromal Cells in Cancer: A Review of Their Immunomodulatory Functions and Dual Effects on Tumor Progression. J. Pathol. 2020, 250, 555–572. [Google Scholar] [CrossRef]
  50. Salmon, H.; Remark, R.; Gnjatic, S.; Merad, M. Host Tissue Determinants of Tumour Immunity. Nat. Rev. Cancer 2019, 19, 215–227. [Google Scholar] [CrossRef]
  51. Petitprez, F.; Sun, C.-M.; Lacroix, L.; Sautès-Fridman, C.; de Reyniès, A.; Fridman, W.H. Quantitative Analyses of the Tumor Microenvironment Composition and Orientation in the Era of Precision Medicine. Front. Oncol. 2018, 8, 390. [Google Scholar] [CrossRef]
  52. Hass, R. Role of MSC in the Tumor Microenvironment. Cancers 2020, 12, 2107. [Google Scholar] [CrossRef]
  53. Xuan, X.; Tian, C.; Zhao, M.; Sun, Y.; Huang, C. Mesenchymal Stem Cells in Cancer Progression and Anticancer Therapeutic Resistance. Cancer Cell Int. 2021, 21, 595. [Google Scholar] [CrossRef] [PubMed]
  54. Balkwill, F.R.; Capasso, M.; Hagemann, T. The Tumor Microenvironment at a Glance. J. Cell Sci. 2012, 125, 5591–5596. [Google Scholar] [CrossRef] [PubMed]
  55. Takashima, Y.; Era, T.; Nakao, K.; Kondo, S.; Kasuga, M.; Smith, A.G.; Nishikawa, S.-I. Neuroepithelial Cells Supply an Initial Transient Wave of MSC Differentiation. Cell 2007, 129, 1377–1388. [Google Scholar] [CrossRef]
  56. Le Douarin, N.M.; Creuzet, S.; Couly, G.; Dupin, E. Neural Crest Cell Plasticity and Its Limits. Development 2004, 131, 4637–4650. [Google Scholar] [CrossRef] [PubMed]
  57. Majesky, M.W. Developmental Basis of Vascular Smooth Muscle Diversity. Arterioscler. Thromb. Vasc. Biol. 2007, 27, 1248–1258. [Google Scholar] [CrossRef]
  58. Hungerford, J.E.; Little, C.D. Developmental Biology of the Vascular Smooth Muscle Cell: Building a Multilayered Vessel Wall. J. Vasc. Res. 1999, 36, 2–27. [Google Scholar] [CrossRef]
  59. Armulik, A.; Genové, G.; Betsholtz, C. Pericytes: Developmental, Physiological, and Pathological Perspectives, Problems, and Promises. Dev. Cell 2011, 21, 193–215. [Google Scholar] [CrossRef]
  60. Vanlandewijck, M.; He, L.; Mae, M.A.A.; Andrae, J.; Ando, K.; Del Gaudio, F.; Nahar, K.; Lebouvier, T.; Lavina, B.; Gouveia, L.; et al. A Molecular Atlas of Cell Types and Zonation in the Brain Vasculature. Nature 2018, 554, 475–480. [Google Scholar] [CrossRef]
  61. Brady, J.; Neal, J.; Sadakar, N.; Gasque, P. Human Endosialin (Tumor Endothelial Marker 1) Is Abundantly Expressed in Highly Malignant and Invasive Brain Tumors. J. Neuropathol. Exp. Neurol. 2004, 63, 1274–1283. [Google Scholar] [CrossRef]
  62. Khan, K.A.; Naylor, A.J.; Khan, A.; Noy, P.J.; Mambretti, M.; Lodhia, P.; Athwal, J.; Korzystka, A.; Buckley, C.D.; Willcox, B.E.; et al. Multimerin-2 Is a Ligand for Group 14 Family C-Type Lectins CLEC14A, CD93 and CD248 Spanning the Endothelial Pericyte Interface. Oncogene 2017, 36, 6097–6108. [Google Scholar] [CrossRef]
  63. Garcia, F.J.; Sun, N.; Lee, H.; Godlewski, B.; Mathys, H.; Galani, K.; Zhou, B.; Jiang, X.; Ng, A.P.; Mantero, J.; et al. Single-Cell Dissection of the Human Brain Vasculature. Nature 2022, 603, 893–899. [Google Scholar] [CrossRef]
  64. Travaglini, K.J.; Nabhan, A.N.; Penland, L.; Sinha, R.; Gillich, A.; Sit, R.; Chang, S.; Conley, S.D.; Mori, Y.; Seita, J.; et al. A Molecular Cell Atlas of the Human Lung from Single-Cell RNA Sequencing. Nature 2020, 587, 619–625. [Google Scholar] [CrossRef] [PubMed]
  65. Litviňuková, M.; Talavera-López, C.; Maatz, H.; Reichart, D.; Worth, C.L.; Lindberg, E.L.; Kanda, M.; Polanski, K.; Heinig, M.; Lee, M.; et al. Cells of the Adult Human Heart. Nature 2020, 588, 466–472. [Google Scholar] [CrossRef] [PubMed]
  66. Lendahl, U.; Muhl, L.; Betsholtz, C. Identification, Discrimination and Heterogeneity of Fibroblasts. Nat. Commun. 2022, 13, 3409. [Google Scholar] [CrossRef] [PubMed]
  67. Ah-Pine, F.; Khettab, M.; Bedoui, Y.; Slama, Y.; Daniel, M.; Doray, B.; Gasque, P. On the Origin and Development of Glioblastoma: Multifaceted Role of Perivascular Mesenchymal Stromal Cells. Acta Neuropathol. Commun. 2023, 11, 104. [Google Scholar] [CrossRef]
  68. Rajan, A.M.; Ma, R.C.; Kocha, K.M.; Zhang, D.J.; Huang, P. Dual Function of Perivascular Fibroblasts in Vascular Stabilization in Zebrafish. PLoS Genet. 2020, 16, e1008800. [Google Scholar] [CrossRef]
  69. Kalluri, R. The Biology and Function of Fibroblasts in Cancer. Nat. Rev. Cancer 2016, 16, 582–598. [Google Scholar] [CrossRef]
  70. Sun, R.; Kong, X.; Qiu, X.; Huang, C.; Wong, P.-P. The Emerging Roles of Pericytes in Modulating Tumor Microenvironment. Front. Cell Dev. Biol. 2021, 9, 676342. [Google Scholar] [CrossRef]
  71. Jiang, Z.; Zhou, J.; Li, L.; Liao, S.; He, J.; Zhou, S.; Zhou, Y. Pericytes in the Tumor Microenvironment. Cancer Lett. 2023, 556, 216074. [Google Scholar] [CrossRef]
  72. Bule, P.; Aguiar, S.I.; Aires-Da-Silva, F.; Dias, J.N.R. Chemokine-Directed Tumor Microenvironment Modulation in Cancer Immunotherapy. Int. J. Mol. Sci. 2021, 22, 9804. [Google Scholar] [CrossRef]
  73. Hu, Y.; Jiang, Y.; Behnan, J.; Ribeiro, M.M.; Kalantzi, C.; Zhang, M.-D.; Lou, D.; Häring, M.; Sharma, N.; Okawa, S.; et al. Neural Network Learning Defines Glioblastoma Features to Be of Neural Crest Perivascular or Radial Glia Lineages. Sci. Adv. 2022, 8, eabm6340. [Google Scholar] [CrossRef]
  74. Griffiths, M.R.; Botto, M.; Morgan, B.P.; Neal, J.W.; Gasque, P. CD93 Regulates Central Nervous System Inflammation in Two Mouse Models of Autoimmune Encephalomyelitis. Immunology 2018, 155, 346–355. [Google Scholar] [CrossRef] [PubMed]
  75. Teicher, B.A. CD248: A Therapeutic Target in Cancer and Fibrotic Diseases. Oncotarget 2019, 10, 993–1009. [Google Scholar] [CrossRef] [PubMed]
  76. Saunders, A.; Macosko, E.Z.; Wysoker, A.; Goldman, M.; Krienen, F.M.; de Rivera, H.; Bien, E.; Baum, M.A.; Bortolin, L.; Wang, S.; et al. Molecular Diversity and Specializations among the Cells of the Adult Mouse Brain. Cell 2018, 174, 1015–1030. [Google Scholar] [CrossRef]
  77. Bergers, G.; Song, S. The Role of Pericytes in Blood-Vessel Formation and Maintenance. Neuro-Oncology 2005, 7, 452–464. [Google Scholar] [CrossRef] [PubMed]
  78. Berger, M.; Bergers, G.; Arnold, B.; Hämmerling, G.J.; Ganss, R. Regulator of G-Protein Signaling-5 Induction in Pericytes Coincides with Active Vessel Remodeling during Neovascularization. Blood 2005, 105, 1094–1101. [Google Scholar] [CrossRef] [PubMed]
  79. Kalluri, R.; Zeisberg, M. Fibroblasts in Cancer. Nat. Rev. Cancer 2006, 6, 392–401. [Google Scholar] [CrossRef]
  80. Goritz, C.; Dias, D.O.; Tomilin, N.; Barbacid, M.; Shupliakov, O.; Frisen, J. A Pericyte Origin of Spinal Cord Scar Tissue. Science 2011, 333, 238–242. [Google Scholar] [CrossRef]
  81. Dorrier, C.E.; Aran, D.; Haenelt, E.A.; Sheehy, R.N.; Hoi, K.K.; Pintaric, L.; Chen, Y.; Lizama, C.O.; Cautivo, K.M.; Weiner, G.A.; et al. CNS Fibroblasts Form a Fibrotic Scar in Response to Immune Cell Infiltration. Nat. Neurosci. 2021, 24, 234–244. [Google Scholar] [CrossRef]
  82. Lapenna, A.; De Palma, M.; Lewis, C.E. Perivascular Macrophages in Health and Disease. Nat. Rev. Immunol. 2018, 18, 689–702. [Google Scholar] [CrossRef]
  83. Van Hove, H.; Martens, L.; Scheyltjens, I.; De Vlaminck, K.; Pombo Antunes, A.R.; De Prijck, S.; Vandamme, N.; De Schepper, S.; Van Isterdael, G.; Scott, C.L.; et al. A Single-Cell Atlas of Mouse Brain Macrophages Reveals Unique Transcriptional Identities Shaped by Ontogeny and Tissue Environment. Nat. Neurosci. 2019, 22, 1021–1035. [Google Scholar] [CrossRef] [PubMed]
  84. Dorrier, C.E.; Jones, H.E.; Pintarić, L.; Siegenthaler, J.A.; Daneman, R. Emerging Roles for CNS Fibroblasts in Health, Injury and Disease. Nat. Rev. Neurosci. 2022, 23, 23–34. [Google Scholar] [CrossRef] [PubMed]
  85. Oh, M.; Nör, J. The Perivascular Niche and Self-Renewal of Stem Cells. Front. Physiol. 2015, 6, 367. [Google Scholar] [CrossRef] [PubMed]
  86. Mohr, A.; Albarenque, S.M.; Deedigan, L.; Yu, R.; Reidy, M.; Fulda, S.; Zwacka, R.M. Targeting of XIAP Combined with Systemic Mesenchymal Stem Cell-Mediated Delivery of STRAIL Ligand Inhibits Metastatic Growth of Pancreatic Carcinoma Cells. Stem Cells 2010, 28, 2109–2120. [Google Scholar] [CrossRef]
  87. Stoff-Khalili, M.A.; Rivera, A.A.; Mathis, J.M.; Banerjee, N.S.; Moon, A.S.; Hess, A.; Rocconi, R.P.; Numnum, T.M.; Everts, M.; Chow, L.T.; et al. Mesenchymal Stem Cells as a Vehicle for Targeted Delivery of CRAds to Lung Metastases of Breast Carcinoma. Breast Cancer Res. Treat. 2007, 105, 157–167. [Google Scholar] [CrossRef] [PubMed]
  88. Bagheri-Mohammadi, S.; Moradian-Tehrani, R.; Noureddini, M.; Alani, B. Novel Application of Adipose-Derived Mesenchymal Stem Cells via Producing Antiangiogenic Factor TSP-1 in Lung Metastatic Melanoma Animal Model. Biologicals 2020, 68, 9–18. [Google Scholar] [CrossRef]
  89. Birbrair, A. (Ed.) Tumor Microenvironment: Non-Hematopoietic Cells; Advances in Experimental Medicine and Biology; Springer International Publishing: Cham, Switzerland, 2020; Volume 1234, ISBN 978-3-030-37183-8. [Google Scholar]
  90. Ren, G.; Zhao, X.; Wang, Y.; Zhang, X.; Chen, X.; Xu, C.; Yuan, Z.; Roberts, A.I.; Zhang, L.; Zheng, B.; et al. CCR2-Dependent Recruitment of Macrophages by Tumor-Educated Mesenchymal Stromal Cells Promotes Tumor Development and Is Mimicked by TNFα. Cell Stem Cell 2012, 11, 812–824. [Google Scholar] [CrossRef]
  91. Coffman, L.G.; Pearson, A.T.; Frisbie, L.G.; Freeman, Z.; Christie, E.; Bowtell, D.D.; Buckanovich, R.J. Ovarian Carcinoma-Associated Mesenchymal Stem Cells Arise from Tissue-Specific Normal Stroma. Stem Cells 2019, 37, 257–269. [Google Scholar] [CrossRef]
  92. Yang, L.; Lin, P.C. Mechanisms That Drive Inflammatory Tumor Microenvironment, Tumor Heterogeneity, and Metastatic Progression. Semin. Cancer Biol. 2017, 47, 185–195. [Google Scholar] [CrossRef]
  93. Dvorak, H.F. Tumors: Wounds That Do Not Heal—Redux. Cancer Immunol. Res. 2015, 3, 1–11. [Google Scholar] [CrossRef]
  94. Caplan, A.I.; Dennis, J.E. Mesenchymal Stem Cells as Trophic Mediators. J. Cell. Biochem. 2006, 98, 1076–1084. [Google Scholar] [CrossRef]
  95. Rustad, K.C.; Gurtner, G.C. Mesenchymal Stem Cells Home to Sites of Injury and Inflammation. Adv. Wound Care 2012, 1, 147–152. [Google Scholar] [CrossRef] [PubMed]
  96. McLean, K.; Gong, Y.; Choi, Y.; Deng, N.; Yang, K.; Bai, S.; Cabrera, L.; Keller, E.; McCauley, L.; Cho, K.R.; et al. Human Ovarian Carcinoma–Associated Mesenchymal Stem Cells Regulate Cancer Stem Cells and Tumorigenesis via Altered BMP Production. J. Clin. Investig. 2011, 121, 3206–3219. [Google Scholar] [CrossRef] [PubMed]
  97. Khakoo, A.Y.; Pati, S.; Anderson, S.A.; Reid, W.; Elshal, M.F.; Rovira, I.I.; Nguyen, A.T.; Malide, D.; Combs, C.A.; Hall, G.; et al. Human Mesenchymal Stem Cells Exert Potent Antitumorigenic Effects in a Model of Kaposi’s Sarcoma. J. Exp. Med. 2006, 203, 1235–1247. [Google Scholar] [CrossRef]
  98. O’Malley, G.; Heijltjes, M.; Houston, A.M.; Rani, S.; Ritter, T.; Egan, L.J.; Ryan, A.E. Mesenchymal Stromal Cells (MSCs) and Colorectal Cancer: A Troublesome Twosome for the Anti-Tumour Immune Response? Oncotarget 2016, 7, 60752–60774. [Google Scholar] [CrossRef] [PubMed]
  99. Moniri, M.R.; Sun, X.-Y.; Rayat, J.; Dai, D.; Ao, Z.; He, Z.; Verchere, C.B.; Dai, L.-J.; Warnock, G.L. TRAIL-Engineered Pancreas-Derived Mesenchymal Stem Cells: Characterization and Cytotoxic Effects on Pancreatic Cancer Cells. Cancer Gene Ther. 2012, 19, 652–658. [Google Scholar] [CrossRef]
  100. Nakamizo, A.; Marini, F.; Amano, T.; Khan, A.; Studeny, M.; Gumin, J.; Chen, J.; Hentschel, S.; Vecil, G.; Dembinski, J.; et al. Human Bone Marrow–Derived Mesenchymal Stem Cells in the Treatment of Gliomas. Cancer Res. 2005, 65, 3307–3318. [Google Scholar] [CrossRef]
  101. Ren, C.; Kumar, S.; Chanda, D.; Kallman, L.; Chen, J.; Mountz, J.D.; Ponnazhagan, S. Cancer Gene Therapy Using Mesenchymal Stem Cells Expressing Interferon-β in a Mouse Prostate Cancer Lung Metastasis Model. Gene Ther. 2008, 15, 1446–1453. [Google Scholar] [CrossRef]
  102. Studeny, M.; Marini, F.C.; Dembinski, J.L.; Zompetta, C.; Cabreira-Hansen, M.; Bekele, B.N.; Champlin, R.E.; Andreeff, M. Mesenchymal Stem Cells: Potential Precursors for Tumor Stroma and Targeted-Delivery Vehicles for Anticancer Agents. JNCI J. Natl. Cancer Inst. 2004, 96, 1593–1603. [Google Scholar] [CrossRef]
  103. Quante, M.; Tu, S.P.; Tomita, H.; Gonda, T.; Wang, S.S.W.; Takashi, S.; Baik, G.H.; Shibata, W.; DiPrete, B.; Betz, K.S.; et al. Bone Marrow-Derived Myofibroblasts Contribute to the Mesenchymal Stem Cell Niche and Promote Tumor Growth. Cancer Cell 2011, 19, 257–272. [Google Scholar] [CrossRef]
  104. Ahn, J.o.; Lee, H.w.; Seo, K.w.; Kang, S.k.; Ra, J.c.; Youn, H.y. Anti-Tumor Effect of Adipose Tissue Derived-Mesenchymal Stem Cells Expressing Interferon-β and Treatment with Cisplatin in a Xenograft Mouse Model for Canine Melanoma. PLoS ONE 2013, 8, e74897. [Google Scholar] [CrossRef] [PubMed]
  105. Komarova, S.; Kawakami, Y.; Stoff-Khalili, M.A.; Curiel, D.T.; Pereboeva, L. Mesenchymal Progenitor Cells as Cellular Vehicles for Delivery of Oncolytic Adenoviruses. Mol. Cancer Ther. 2006, 5, 755–766. [Google Scholar] [CrossRef]
  106. Becker, A.D.; Riet, I.V. Homing and Migration of Mesenchymal Stromal Cells: How to Improve the Efficacy of Cell Therapy? World J. Stem Cells 2016, 8, 73–87. [Google Scholar] [CrossRef] [PubMed]
  107. Ponte, A.L.; Marais, E.; Gallay, N.; Langonné, A.; Delorme, B.; Hérault, O.; Charbord, P.; Domenech, J. The In Vitro Migration Capacity of Human Bone Marrow Mesenchymal Stem Cells: Comparison of Chemokine and Growth Factor Chemotactic Activities. Stem Cells 2007, 25, 1737–1745. [Google Scholar] [CrossRef] [PubMed]
  108. Li, L.; Madu, C.O.; Lu, A.; Lu, Y. HIF-1α Promotes A Hypoxia-Independent Cell Migration. Open Biol. J. 2010, 3, 8–14. [Google Scholar] [CrossRef]
  109. Schmidt, A.; Ladage, D.; Schinköthe, T.; Klausmann, U.; Ulrichs, C.; Klinz, F.; Brixius, K.; Arnhold, S.; Desai, B.; Mehlhorn, U.; et al. Basic Fibroblast Growth Factor Controls Migration in Human Mesenchymal Stem Cells. Stem Cells 2006, 24, 1750–1758. [Google Scholar] [CrossRef]
  110. Dwyer, R.M.; Potter-Beirne, S.M.; Harrington, K.A.; Lowery, A.J.; Hennessy, E.; Murphy, J.M.; Barry, F.P.; O’Brien, T.; Kerin, M.J. Monocyte Chemotactic Protein-1 Secreted by Primary Breast Tumors Stimulates Migration of Mesenchymal Stem Cells. Clin. Cancer Res. 2007, 13, 5020–5027. [Google Scholar] [CrossRef]
  111. Chen, M.-S.; Lin, C.-Y.; Chiu, Y.-H.; Chen, C.-P.; Tsai, P.-J.; Wang, H.-S. IL-1 β -Induced Matrix Metalloprotease-1 Promotes Mesenchymal Stem Cell Migration via PAR1 and G-Protein-Coupled Signaling Pathway. Stem Cells Int. 2018, 2018, 3524759. [Google Scholar] [CrossRef]
  112. Dubon, M.J.; Yu, J.; Choi, S.; Park, K. Transforming Growth Factor β Induces Bone Marrow Mesenchymal Stem Cell Migration via Noncanonical Signals and N-cadherin. J. Cell. Physiol. 2018, 233, 201–213. [Google Scholar] [CrossRef]
  113. Lourenco, S.; Teixeira, V.H.; Kalber, T.; Jose, R.J.; Floto, R.A.; Janes, S.M. Macrophage Migration Inhibitory Factor–CXCR4 Is the Dominant Chemotactic Axis in Human Mesenchymal Stem Cell Recruitment to Tumors. J. Immunol. 2015, 194, 3463–3474. [Google Scholar] [CrossRef]
  114. Abarbanell, A.M.; Coffey, A.C.; Fehrenbacher, J.W.; Beckman, D.J.; Herrmann, J.L.; Weil, B.; Meldrum, D.R. Proinflammatory Cytokine Effects on Mesenchymal Stem Cell Therapy for the Ischemic Heart. Ann. Thorac. Surg. 2009, 88, 1036–1043. [Google Scholar] [CrossRef] [PubMed]
  115. Sun, Z.; Wang, S.; Zhao, R.C. The Roles of Mesenchymal Stem Cells in Tumor Inflammatory Microenvironment. J. Hematol. Oncol. 2014, 7, 14. [Google Scholar] [CrossRef] [PubMed]
  116. Hill, B.S.; Pelagalli, A.; Passaro, N.; Zannetti, A. Tumor-Educated Mesenchymal Stem Cells Promote pro-Metastatic Phenotype. Oncotarget 2017, 8, 73296–73311. [Google Scholar] [CrossRef]
  117. Ho, I.A.W.; Chan, K.Y.W.; Ng, W.-H.; Guo, C.M.; Hui, K.M.; Cheang, P.; Lam, P.Y.P. Matrix Metalloproteinase 1 Is Necessary for the Migration of Human Bone Marrow-Derived Mesenchymal Stem Cells toward Human Glioma. Stem Cells 2009, 27, 1366–1375. [Google Scholar] [CrossRef]
  118. Anton, K.; Banerjee, D.; Glod, J. Macrophage-Associated Mesenchymal Stem Cells Assume an Activated, Migratory, pro-Inflammatory Phenotype with Increased IL-6 and CXCL10 Secretion. PLoS ONE 2012, 7, e35036. [Google Scholar] [CrossRef] [PubMed]
  119. Barcellos-de-Souza, P.; Comito, G.; Pons-Segura, C.; Taddei, M.L.; Gori, V.; Becherucci, V.; Bambi, F.; Margheri, F.; Laurenzana, A.; Del Rosso, M.; et al. Mesenchymal Stem Cells Are Recruited and Activated into Carcinoma-Associated Fibroblasts by Prostate Cancer Microenvironment-Derived TGF-Β1. Stem Cells 2016, 34, 2536–2547. [Google Scholar] [CrossRef]
  120. Uchibori, R.; Tsukahara, T.; Mizuguchi, H.; Saga, Y.; Urabe, M.; Mizukami, H.; Kume, A.; Ozawa, K. NF-ΚB Activity Regulates Mesenchymal Stem Cell Accumulation at Tumor Sites. Cancer Res. 2013, 73, 364–372. [Google Scholar] [CrossRef]
  121. Papait, A.; Stefani, F.R.; Cargnoni, A.; Magatti, M.; Parolini, O.; Silini, A.R. The Multifaceted Roles of MSCs in the Tumor Microenvironment: Interactions with Immune Cells and Exploitation for Therapy. Front. Cell Dev. Biol. 2020, 8, 447. [Google Scholar] [CrossRef]
  122. Spaeth, E.L.; Dembinski, J.L.; Sasser, A.K.; Watson, K.; Klopp, A.; Hall, B.; Andreeff, M.; Marini, F. Mesenchymal Stem Cell Transition to Tumor-Associated Fibroblasts Contributes to Fibrovascular Network Expansion and Tumor Progression. PLoS ONE 2009, 4, e4992. [Google Scholar] [CrossRef]
  123. Mishra, P.J.; Mishra, P.J.; Humeniuk, R.; Medina, D.J.; Alexe, G.; Mesirov, J.P.; Ganesan, S.; Glod, J.W.; Banerjee, D. Carcinoma-Associated Fibroblast–Like Differentiation of Human Mesenchymal Stem Cells. Cancer Res. 2008, 68, 4331–4339. [Google Scholar] [CrossRef]
  124. Lee, K. Exosomes from Breast Cancer Cells Can Convert Adipose Tissue-Derived Mesenchymal Stem Cells into Myofibroblast-like Cells. Int. J. Oncol. 2011, 40, 130–138. [Google Scholar] [CrossRef]
  125. Miyazaki, Y.; Oda, T.; Inagaki, Y.; Kushige, H.; Saito, Y.; Mori, N.; Takayama, Y.; Kumagai, Y.; Mitsuyama, T.; Kida, Y.S. Adipose-Derived Mesenchymal Stem Cells Differentiate into Heterogeneous Cancer-Associated Fibroblasts in a Stroma-Rich Xenograft Model. Sci. Rep. 2021, 11, 4690. [Google Scholar] [CrossRef] [PubMed]
  126. Wu, Q.; Li, B.; Li, Z.; Li, J.; Sun, S.; Sun, S. Cancer-Associated Adipocytes: Key Players in Breast Cancer Progression. J. Hematol. Oncol. 2019, 12, 95. [Google Scholar] [CrossRef]
  127. Fujisaki, K.; Fujimoto, H.; Sangai, T.; Nagashima, T.; Sakakibara, M.; Shiina, N.; Kuroda, M.; Aoyagi, Y.; Miyazaki, M. Cancer-Mediated Adipose Reversion Promotes Cancer Cell Migration via IL-6 and MCP-1. Breast Cancer Res. Treat. 2015, 150, 255–263. [Google Scholar] [CrossRef] [PubMed]
  128. De Palma, M.; Biziato, D.; Petrova, T.V. Microenvironmental Regulation of Tumour Angiogenesis. Nat. Rev. Cancer 2017, 17, 457–474. [Google Scholar] [CrossRef] [PubMed]
  129. Teng, I.-W.; Hou, P.-C.; Lee, K.-D.; Chu, P.-Y.; Yeh, K.-T.; Jin, V.X.; Tseng, M.-J.; Tsai, S.-J.; Chang, Y.-S.; Wu, C.-S.; et al. Targeted Methylation of Two Tumor Suppressor Genes Is Sufficient to Transform Mesenchymal Stem Cells into Cancer Stem/Initiating Cells. Cancer Res. 2011, 71, 4653–4663. [Google Scholar] [CrossRef]
  130. Melzer, C.; Yang, Y.; Hass, R. Interaction of MSC with Tumor Cells. Cell Commun. Signal. 2016, 14, 20. [Google Scholar] [CrossRef]
  131. Melzer, C.; von der Ohe, J.; Hass, R. Enhanced Metastatic Capacity of Breast Cancer Cells after Interaction and Hybrid Formation with Mesenchymal Stroma/Stem Cells (MSC). Cell Commun. Signal. 2018, 16, 2. [Google Scholar] [CrossRef]
  132. Yang, Y.; Otte, A.; Hass, R. Human Mesenchymal Stroma/Stem Cells Exchange Membrane Proteins and Alter Functionality during Interaction with Different Tumor Cell Lines. Stem Cells Dev. 2015, 24, 1205–1222. [Google Scholar] [CrossRef] [PubMed]
  133. Fridman, W.H.; Pagès, F.; Sautès-Fridman, C.; Galon, J. The Immune Contexture in Human Tumours: Impact on Clinical Outcome. Nat. Rev. Cancer 2012, 12, 298–306. [Google Scholar] [CrossRef]
  134. Qin, Z.; Richter, G.; Schüler, T.; Ibe, S.; Cao, X.; Blankenstein, T. B Cells Inhibit Induction of T Cell-Dependent Tumor Immunity. Nat. Med. 1998, 4, 627–630. [Google Scholar] [CrossRef]
  135. Kuroda, H.; Jamiyan, T.; Yamaguchi, R.; Kakumoto, A.; Abe, A.; Harada, O.; Enkhbat, B.; Masunaga, A. Prognostic Value of Tumor-Infiltrating B Lymphocytes and Plasma Cells in Triple-Negative Breast Cancer. Breast Cancer 2021, 28, 904–914. [Google Scholar] [CrossRef]
  136. Liu, J.; Liu, Q.; Chen, X. The Immunomodulatory Effects of Mesenchymal Stem Cells on Regulatory B Cells. Front. Immunol. 2020, 11, 1843. [Google Scholar] [CrossRef] [PubMed]
  137. Krueger, T.E.; Thorek, D.L.J.; Meeker, A.K.; Isaacs, J.T.; Brennen, W.N. Tumor-Infiltrating Mesenchymal Stem Cells: Drivers of the Immunosuppressive Tumor Microenvironment in Prostate Cancer? Prostate 2019, 79, 320–330. [Google Scholar] [CrossRef] [PubMed]
  138. Harrell, C.R.; Volarevic, A.; Djonov, V.G.; Jovicic, N.; Volarevic, V. Mesenchymal Stem Cell: A Friend or Foe in Anti-Tumor Immunity. Int. J. Mol. Sci. 2021, 22, 12429. [Google Scholar] [CrossRef] [PubMed]
  139. Han, Z.; Tian, Z.; Lv, G.; Zhang, L.; Jiang, G.; Sun, K.; Wang, C.; Bu, X.; Li, R.; Shi, Y.; et al. Immunosuppressive Effect of Bone Marrow-Derived Mesenchymal Stem Cells in Inflammatory Microenvironment Favours the Growth of B16 Melanoma Cells. J. Cell. Mol. Med. 2011, 15, 2343–2352. [Google Scholar] [CrossRef]
  140. Guo, Y.; Chan, K.-H.; Lai, W.-H.; Siu, C.-W.; Kwan, S.-C.; Tse, H.-F.; Wing-Lok Ho, P.; Wing-Man Ho, J. Human Mesenchymal Stem Cells Upregulate CD1dCD5(+) Regulatory B Cells in Experimental Autoimmune Encephalomyelitis. Neuroimmunomodulation 2013, 20, 294–303. [Google Scholar] [CrossRef]
  141. Hermankova, B.; Zajicova, A.; Javorkova, E.; Chudickova, M.; Trosan, P.; Hajkova, M.; Krulova, M.; Holan, V. Suppression of IL-10 Production by Activated B Cells via a Cell Contact-Dependent Cyclooxygenase-2 Pathway Upregulated in IFN-γ-Treated Mesenchymal Stem Cells. Immunobiology 2016, 221, 129–136. [Google Scholar] [CrossRef]
  142. Ren, G.; Zhang, L.; Zhao, X.; Xu, G.; Zhang, Y.; Roberts, A.I.; Zhao, R.C.; Shi, Y. Mesenchymal Stem Cell-Mediated Immunosuppression Occurs via Concerted Action of Chemokines and Nitric Oxide. Cell Stem Cell 2008, 2, 141–150. [Google Scholar] [CrossRef]
  143. Frisbie, L.; Buckanovich, R.J.; Coffman, L. Carcinoma-Associated Mesenchymal Stem/Stromal Cells: Architects of the Pro-Tumorigenic Tumor Microenvironment. Stem Cells 2022, 40, 705–715. [Google Scholar] [CrossRef]
  144. Zhang, Q.; Liu, L.; Gong, C.; Shi, H.; Zeng, Y.; Wang, X.; Zhao, Y.; Wei, Y. Prognostic Significance of Tumor-Associated Macrophages in Solid Tumor: A Meta-Analysis of the Literature. PLoS ONE 2012, 7, e50946. [Google Scholar] [CrossRef] [PubMed]
  145. Zhou, W.; Ke, S.Q.; Huang, Z.; Flavahan, W.; Fang, X.; Paul, J.; Wu, L.; Sloan, A.E.; McLendon, R.E.; Li, X.; et al. Periostin Secreted by Glioblastoma Stem Cells Recruits M2 Tumour-Associated Macrophages and Promotes Malignant Growth. Nat. Cell Biol. 2015, 17, 170–182. [Google Scholar] [CrossRef]
  146. Ah-Pine, F.; Malaterre-Septembre, A.; Bedoui, Y.; Khettab, M.; Neal, J.W.; Freppel, S.; Gasque, P. Complement Activation and Up-Regulated Expression of Anaphylatoxin C3a/C3aR in Glioblastoma: Deciphering the Links with TGF-β and VEGF. Cancers 2023, 15, 2647. [Google Scholar] [CrossRef] [PubMed]
  147. Davidson, S.; Efremova, M.; Riedel, A.; Mahata, B.; Pramanik, J.; Huuhtanen, J.; Kar, G.; Vento-Tormo, R.; Hagai, T.; Chen, X.; et al. Single-Cell RNA Sequencing Reveals a Dynamic Stromal Niche That Supports Tumor Growth. Cell Rep. 2020, 31, 107628. [Google Scholar] [CrossRef] [PubMed]
  148. Jia, X.-H.; Feng, G.-W.; Wang, Z.-L.; Du, Y.; Shen, C.; Hui, H.; Peng, D.; Li, Z.-J.; Kong, D.-L.; Tian, J. Activation of Mesenchymal Stem Cells by Macrophages Promotes Tumor Progression through Immune Suppressive Effects. Oncotarget 2016, 7, 20934–20944. [Google Scholar] [CrossRef]
  149. Condeelis, J.; Pollard, J.W. Macrophages: Obligate Partners for Tumor Cell Migration, Invasion, and Metastasis. Cell 2006, 124, 263–266. [Google Scholar] [CrossRef] [PubMed]
  150. Nozawa, H.; Chiu, C.; Hanahan, D. Infiltrating Neutrophils Mediate the Initial Angiogenic Switch in a Mouse Model of Multistage Carcinogenesis. Proc. Natl. Acad. Sci. USA 2006, 103, 12493–12498. [Google Scholar] [CrossRef]
  151. Youn, J.-I.; Gabrilovich, D.I. The Biology of Myeloid-Derived Suppressor Cells: The Blessing and the Curse of Morphological and Functional Heterogeneity. Eur. J. Immunol. 2010, 40, 2969–2975. [Google Scholar] [CrossRef]
  152. Hu, X.; Zhou, Y.; Dong, K.; Sun, Z.; Zhao, D.; Wang, W.; Yu, G.; Liu, W.; Xu, G.; Han, Z.; et al. Programming of the Development of Tumor-Promoting Neutrophils by Mesenchymal Stromal Cells. Cell. Physiol. Biochem. 2014, 33, 1802–1814. [Google Scholar] [CrossRef]
  153. Zhang, J.; Ji, C.; Li, W.; Mao, Z.; Shi, Y.; Shi, H.; Ji, R.; Qian, H.; Xu, W.; Zhang, X. Tumor-Educated Neutrophils Activate Mesenchymal Stem Cells to Promote Gastric Cancer Growth and Metastasis. Front. Cell Dev. Biol. 2020, 8, 788. [Google Scholar] [CrossRef]
  154. Guo, S.; Huang, C.; Han, F.; Chen, B.; Ding, Y.; Zhao, Y.; Chen, Z.; Wen, S.; Wang, M.; Shen, B.; et al. Gastric Cancer Mesenchymal Stem Cells Inhibit NK Cell Function through MTOR Signalling to Promote Tumour Growth. Stem Cells Int. 2021, 2021, 9989790. [Google Scholar] [CrossRef] [PubMed]
  155. Pradier, A.; Passweg, J.; Villard, J.; Kindler, V. Human Bone Marrow Stromal Cells and Skin Fibroblasts Inhibit Natural Killer Cell Proliferation and Cytotoxic Activity. Cell Transpl. 2011, 20, 681–691. [Google Scholar] [CrossRef]
  156. Gazdic, M.; Simovic Markovic, B.; Jovicic, N.; Misirkic-Marjanovic, M.; Djonov, V.; Jakovljevic, V.; Arsenijevic, N.; Lukic, M.L.; Volarevic, V. Mesenchymal Stem Cells Promote Metastasis of Lung Cancer Cells by Downregulating Systemic Antitumor Immune Response. Stem Cells Int. 2017, 2017, 6294717. [Google Scholar] [CrossRef] [PubMed]
  157. Abbasi, B.; Shamsasenjan, K.; Ahmadi, M.; Beheshti, S.A.; Saleh, M. Mesenchymal Stem Cells and Natural Killer Cells Interaction Mechanisms and Potential Clinical Applications. Stem Cell Res. Ther. 2022, 13, 97. [Google Scholar] [CrossRef]
  158. Swamydas, M.; Ricci, K.; Rego, S.L.; Dréau, D. Mesenchymal Stem Cell-Derived CCL-9 and CCL-5 Promote Mammary Tumor Cell Invasion and the Activation of Matrix Metalloproteinases. Cell Adhes. Migr. 2013, 7, 315–324. [Google Scholar] [CrossRef] [PubMed]
  159. Mi, Z.; Bhattacharya, S.D.; Kim, V.M.; Guo, H.; Talbot, L.J.; Kuo, P.C. Osteopontin Promotes CCL5-Mesenchymal Stromal Cell-Mediated Breast Cancer Metastasis. Carcinogenesis 2011, 32, 477–487. [Google Scholar] [CrossRef]
  160. Gallo, M.; De Luca, A.; Lamura, L.; Normanno, N. Zoledronic Acid Blocks the Interaction between Mesenchymal Stem Cells and Breast Cancer Cells: Implications for Adjuvant Therapy of Breast Cancer. Ann. Oncol. 2012, 23, 597–604. [Google Scholar] [CrossRef]
  161. Naderi, E.H.; Skah, S.; Ugland, H.; Myklebost, O.; Sandnes, D.L.; Torgersen, M.L.; Josefsen, D.; Ruud, E.; Naderi, S.; Blomhoff, H.K. Bone Marrow Stroma-Derived PGE2 Protects BCP-ALL Cells from DNA Damage-Induced P53 Accumulation and Cell Death. Mol. Cancer 2015, 14, 14. [Google Scholar] [CrossRef] [PubMed]
  162. Loussouarn, C.; Pers, Y.-M.; Bony, C.; Jorgensen, C.; Noël, D. Mesenchymal Stromal Cell-Derived Extracellular Vesicles Regulate the Mitochondrial Metabolism via Transfer of MiRNAs. Front. Immunol. 2021, 12, 623973. [Google Scholar] [CrossRef]
  163. Jothimani, G.; Pathak, S.; Dutta, S.; Duttaroy, A.K.; Banerjee, A. A Comprehensive Cancer-Associated MicroRNA Expression Profiling and Proteomic Analysis of Human Umbilical Cord Mesenchymal Stem Cell-Derived Exosomes. Tissue Eng. Regen. Med. 2022, 19, 1013–1031. [Google Scholar] [CrossRef]
  164. Wu, T.; Liu, Y.; Fan, Z.; Xu, J.; Jin, L.; Gao, Z.; Wu, Z.; Hu, L.; Wang, J.; Zhang, C.; et al. MiR-21 Modulates the Immunoregulatory Function of Bone Marrow Mesenchymal Stem Cells Through the PTEN/Akt/TGF-Β1 Pathway. Stem Cells 2015, 33, 3281–3290. [Google Scholar] [CrossRef]
  165. Kandouz, M.; Batist, G. Gap Junctions and Connexins as Therapeutic Targets in Cancer. Expert Opin. Ther. Targets 2010, 14, 681–692. [Google Scholar] [CrossRef] [PubMed]
  166. Rafii, A.; Mirshahi, P.; Poupot, M.; Faussat, A.-M.; Simon, A.; Ducros, E.; Mery, E.; Couderc, B.; Lis, R.; Capdet, J.; et al. Oncologic Trogocytosis of an Original Stromal Cells Induces Chemoresistance of Ovarian Tumours. PLoS ONE 2008, 3, e3894. [Google Scholar] [CrossRef] [PubMed]
  167. Lan, T.; Luo, M.; Wei, X. Mesenchymal Stem/Stromal Cells in Cancer Therapy. J. Hematol. Oncol. 2021, 14, 195. [Google Scholar] [CrossRef]
  168. Bartsch, J.E.; Staren, E.D.; Appert, H.E. Matrix Metalloproteinase Expression in Breast Cancer. J. Surg. Res. 2003, 110, 383–392. [Google Scholar] [CrossRef] [PubMed]
  169. Li, J.; Deng, Z.; Wang, Z.; Wang, D.; Zhang, L.; Su, Q.; Lai, Y.; Li, B.; Luo, Z.; Chen, X.; et al. Zipper-Interacting Protein Kinase Promotes Epithelial-Mesenchymal Transition, Invasion and Metastasis through AKT and NF-ΚB Signaling and Is Associated with Metastasis and Poor Prognosis in Gastric Cancer Patients. Oncotarget 2015, 6, 8323–8338. [Google Scholar] [CrossRef]
  170. Yang, H.-L.; Thiyagarajan, V.; Shen, P.-C.; Mathew, D.C.; Lin, K.-Y.; Liao, J.-W.; Hseu, Y.-C. Anti-EMT Properties of CoQ0 Attributed to PI3K/AKT/NFKB/MMP-9 Signaling Pathway through ROS-Mediated Apoptosis. J. Exp. Clin. Cancer Res. 2019, 38, 186. [Google Scholar] [CrossRef]
  171. Zhang, M.; Liu, Z.Z.; Aoshima, K.; Cai, W.L.; Sun, H.; Xu, T.; Zhang, Y.; An, Y.; Chen, J.F.; Chan, L.H.; et al. CECR2 Drives Breast Cancer Metastasis by Promoting NF-ΚB Signaling and Macrophage-Mediated Immune Suppression. Sci. Transl. Med. 2022, 14, eabf5473. [Google Scholar] [CrossRef]
  172. Atiya, H.; Frisbie, L.; Pressimone, C.; Coffman, L. Mesenchymal Stem Cells in the Tumor Microenvironment. In Tumor Microenvironment: Non-Hematopoietic Cells; Birbrair, A., Ed.; Advances in Experimental Medicine and Biology; Springer International Publishing: Cham, Switzerland, 2020; pp. 31–42. ISBN 978-3-030-37184-5. [Google Scholar]
  173. Hayashi, Y.; Tsujii, M.; Kodama, T.; Akasaka, T.; Kondo, J.; Hikita, H.; Inoue, T.; Tsujii, Y.; Maekawa, A.; Yoshii, S.; et al. P53 Functional Deficiency in Human Colon Cancer Cells Promotes Fibroblast-Mediated Angiogenesis and Tumor Growth. Carcinogenesis 2016, 37, 972–984. [Google Scholar] [CrossRef]
  174. Monteran, L.; Erez, N. The Dark Side of Fibroblasts: Cancer-Associated Fibroblasts as Mediators of Immunosuppression in the Tumor Microenvironment. Front. Immunol. 2019, 10, 1835. [Google Scholar] [CrossRef]
  175. Meulmeester, E.; Ten Dijke, P. The Dynamic Roles of TGF-β in Cancer. J. Pathol. 2011, 223, 205–218. [Google Scholar] [CrossRef]
  176. Yang, J.; Lu, Y.; Lin, Y.-Y.; Zheng, Z.-Y.; Fang, J.-H.; He, S.; Zhuang, S.-M. Vascular Mimicry Formation Is Promoted by Paracrine TGF-β and SDF1 of Cancer-Associated Fibroblasts and Inhibited by MiR-101 in Hepatocellular Carcinoma. Cancer Lett. 2016, 383, 18–27. [Google Scholar] [CrossRef]
  177. Zhou, B.; Zhuang, X.; Wang, Y.; Lin, Z.; Zhang, D.; Fan, S.; Li, J.; Chen, W. Tumor Necrosis Factor α Induces Myofibroblast Differentiation in Human Tongue Cancer and Promotes Invasiveness and Angiogenesis via Secretion of Stromal Cell-Derived Factor-1. Oral Oncol. 2015, 51, 1095–1102. [Google Scholar] [CrossRef] [PubMed]
  178. Zigrino, P.; Kuhn, I.; Bäuerle, T.; Zamek, J.; Fox, J.W.; Neumann, S.; Licht, A.; Schorpp-Kistner, M.; Angel, P.; Mauch, C. Stromal Expression of MMP-13 Is Required for Melanoma Invasion and Metastasis. J. Investig. Dermatol. 2009, 129, 2686–2693. [Google Scholar] [CrossRef] [PubMed]
  179. Crawford, Y.; Kasman, I.; Yu, L.; Zhong, C.; Wu, X.; Modrusan, Z.; Kaminker, J.; Ferrara, N. PDGF-C Mediates the Angiogenic and Tumorigenic Properties of Fibroblasts Associated with Tumors Refractory to Anti-VEGF Treatment. Cancer Cell 2009, 15, 21–34. [Google Scholar] [CrossRef] [PubMed]
  180. Santos, A.M.; Jung, J.; Aziz, N.; Kissil, J.L.; Puré, E. Targeting Fibroblast Activation Protein Inhibits Tumor Stromagenesis and Growth in Mice. J. Clin. Investig. 2009, 119, 3613–3625. [Google Scholar] [CrossRef]
  181. Wang, M.; Zhao, J.; Zhang, L.; Wei, F.; Lian, Y.; Wu, Y.; Gong, Z.; Zhang, S.; Zhou, J.; Cao, K.; et al. Role of Tumor Microenvironment in Tumorigenesis. J. Cancer 2017, 8, 761–773. [Google Scholar] [CrossRef]
  182. Ahn, S.Y. The Role of MSCs in the Tumor Microenvironment and Tumor Progression. Anticancer Res. 2020, 40, 3039–3047. [Google Scholar] [CrossRef]
  183. Zhu, J.; Liang, L.; Jiao, Y.; Liu, L.; on behalf of the U.S.-China Physical Sciences-Oncology Alliance. Enhanced Invasion of Metastatic Cancer Cells via Extracellular Matrix Interface. PLoS ONE 2015, 10, e0118058. [Google Scholar] [CrossRef]
  184. Provenzano, P.P.; Inman, D.R.; Eliceiri, K.W.; Knittel, J.G.; Yan, L.; Rueden, C.T.; White, J.G.; Keely, P.J. Collagen Density Promotes Mammary Tumor Initiation and Progression. BMC Med. 2008, 6, 11. [Google Scholar] [CrossRef]
  185. Gattazzo, F.; Urciuolo, A.; Bonaldo, P. Extracellular Matrix: A Dynamic Microenvironment for Stem Cell Niche. Biochim. Biophys. Acta (BBA)—Gen. Subj. 2014, 1840, 2506–2519. [Google Scholar] [CrossRef] [PubMed]
  186. Konstantinov, A.S.; Kovaleva, O.V.; Samoilova, D.V.; Shelekhova, K.V. Role of Macrophages in Progression of Colorectal Cancer: A Contrast with the Traditional Paradigm. Int. J. Clin. Exp. Pathol. 2022, 15, 403–411. [Google Scholar] [PubMed]
  187. Gonzalez, M.E.; Martin, E.; Anwar, T.; Arellano-Garcia, C.; Medhora, N.; Lama, A.; Chen, Y.-C.; Tanager, K.S.; Yoon, E.; Kidwell, K.; et al. Mesenchymal Stem Cell Induced DDR2 Mediates Stromal-Breast Cancer Interactions and Metastasis Growth. Cell Rep. 2017, 18, 1215–1228. [Google Scholar] [CrossRef] [PubMed]
  188. Zeltz, C.; Primac, I.; Erusappan, P.; Alam, J.; Noel, A.; Gullberg, D. Cancer-Associated Fibroblasts in Desmoplastic Tumors: Emerging Role of Integrins. Semin. Cancer Biol. 2020, 62, 166–181. [Google Scholar] [CrossRef]
  189. Eiro, N.; Fraile, M.; Fernández-Francos, S.; Sánchez, R.; Costa, L.A.; Vizoso, F.J. Importance of the Origin of Mesenchymal (Stem) Stromal Cells in Cancer Biology: “Alliance” or “War” in Intercellular Signals. Cell Biosci. 2021, 11, 109. [Google Scholar] [CrossRef]
  190. Liang, W.; Chen, X.; Zhang, S.; Fang, J.; Chen, M.; Xu, Y.; Chen, X. Mesenchymal Stem Cells as a Double-Edged Sword in Tumor Growth: Focusing on MSC-Derived Cytokines. Cell. Mol. Biol. Lett. 2021, 26, 3. [Google Scholar] [CrossRef]
  191. Rhee, K.-J.; Lee, J.; Eom, Y. Mesenchymal Stem Cell-Mediated Effects of Tumor Support or Suppression. Int. J. Mol. Sci. 2015, 16, 30015–30033. [Google Scholar] [CrossRef]
  192. Waterman, R.S.; Henkle, S.L.; Betancourt, A.M. Mesenchymal Stem Cell 1 (MSC1)-Based Therapy Attenuates Tumor Growth Whereas MSC2-Treatment Promotes Tumor Growth and Metastasis. PLoS ONE 2012, 7, e45590. [Google Scholar] [CrossRef]
  193. Waterman, R.S.; Tomchuck, S.L.; Henkle, S.L.; Betancourt, A.M. A New Mesenchymal Stem Cell (MSC) Paradigm: Polarization into a Pro-Inflammatory MSC1 or an Immunosuppressive MSC2 Phenotype. PLoS ONE 2010, 5, e10088. [Google Scholar] [CrossRef]
  194. Ohlsson, L.B.; Varas, L.; Kjellman, C.; Edvardsen, K.; Lindvall, M. Mesenchymal Progenitor Cell-Mediated Inhibition of Tumor Growth In Vivo and In Vitro in Gelatin Matrix. Exp. Mol. Pathol. 2003, 75, 248–255. [Google Scholar] [CrossRef]
  195. Qiao, L.; Xu, Z.; Zhao, T.; Ye, L.; Zhang, X. Dkk-1 Secreted by Mesenchymal Stem Cells Inhibits Growth of Breast Cancer Cells via Depression of Wnt Signalling. Cancer Lett. 2008, 269, 67–77. [Google Scholar] [CrossRef] [PubMed]
  196. Lu, Y.; Yuan, Y.; Wang, X.; Wei, L.; Chen, Y.; Cong, C.; Li, S.; Long, D.; Tan, W.; Mao, Y.; et al. The Growth Inhibitory Effect of Mesenchymal Stem Cells on Tumor Cells In Vitro and In Vivo. Cancer Biol. Ther. 2008, 7, 245–251. [Google Scholar] [CrossRef]
  197. Gu, H.; Yan, C.; Wan, H.; Wu, L.; Liu, J.; Zhu, Z.; Gao, D. Mesenchymal Stem Cell-Derived Exosomes Block Malignant Behaviors of Hepatocellular Carcinoma Stem Cells through a LncRNA C5orf66-AS1/MicroRNA-127-3p/DUSP1/ERK Axis. Hum. Cell 2021, 34, 1812–1829. [Google Scholar] [CrossRef] [PubMed]
  198. Hanahan, D.; Weinberg, R.A. Hallmarks of Cancer: The Next Generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [PubMed]
  199. Poggi, A.; Giuliani, M. Mesenchymal Stromal Cells Can Regulate the Immune Response in the Tumor Microenvironment. Vaccines 2016, 4, 41. [Google Scholar] [CrossRef]
  200. Li, G.-C.; Zhang, H.-W.; Zhao, Q.-C.; Sun, L.; Yang, J.-J.; Hong, L.; Feng, F.; Cai, L. Mesenchymal Stem Cells Promote Tumor Angiogenesis via the Action of Transforming Growth Factor Β1. Oncol. Lett. 2016, 11, 1089–1094. [Google Scholar] [CrossRef]
  201. Zhang, T.; Lee, Y.W.; Rui, Y.F.; Cheng, T.Y.; Jiang, X.H.; Li, G. Bone Marrow-Derived Mesenchymal Stem Cells Promote Growth and Angiogenesis of Breast and Prostate Tumors. Stem Cell Res. Ther. 2013, 4, 70. [Google Scholar] [CrossRef]
  202. Fregni, G.; Quinodoz, M.; Möller, E.; Vuille, J.; Galland, S.; Fusco, C.; Martin, P.; Letovanec, I.; Provero, P.; Rivolta, C.; et al. Reciprocal Modulation of Mesenchymal Stem Cells and Tumor Cells Promotes Lung Cancer Metastasis. eBioMedicine 2018, 29, 128–145. [Google Scholar] [CrossRef]
  203. Bergfeld, S.A.; Blavier, L.; DeClerck, Y.A. Bone Marrow–Derived Mesenchymal Stromal Cells Promote Survival and Drug Resistance in Tumor Cells. Mol. Cancer Ther. 2014, 13, 962–975. [Google Scholar] [CrossRef]
  204. Peinado, H.; Quintanilla, M.; Cano, A. Transforming Growth Factor β-1 Induces Snail Transcription Factor in Epithelial Cell Lines. J. Biol. Chem. 2003, 278, 21113–21123. [Google Scholar] [CrossRef]
  205. Lamouille, S.; Xu, J.; Derynck, R. Molecular Mechanisms of Epithelial–Mesenchymal Transition. Nat. Rev. Mol. Cell Biol. 2014, 15, 178–196. [Google Scholar] [CrossRef] [PubMed]
  206. Nieto, M.A.; Huang, R.Y.-J.; Jackson, R.A.; Thiery, J.P. EMT: 2016. Cell 2016, 166, 21–45. [Google Scholar] [CrossRef]
  207. Tsai, J.H.; Yang, J. Epithelial–Mesenchymal Plasticity in Carcinoma Metastasis. Genes Dev. 2013, 27, 2192–2206. [Google Scholar] [CrossRef] [PubMed]
  208. Pawitan, J.A.; Bui, T.A.; Mubarok, W.; Antarianto, R.D.; Nurhayati, R.W.; Dilogo, I.H.; Oceandy, D. Enhancement of the Therapeutic Capacity of Mesenchymal Stem Cells by Genetic Modification: A Systematic Review. Front. Cell Dev. Biol. 2020, 8, 587776. [Google Scholar] [CrossRef] [PubMed]
  209. Beckermann, B.M.; Kallifatidis, G.; Groth, A.; Frommhold, D.; Apel, A.; Mattern, J.; Salnikov, A.V.; Moldenhauer, G.; Wagner, W.; Diehlmann, A.; et al. VEGF Expression by Mesenchymal Stem Cells Contributes to Angiogenesis in Pancreatic Carcinoma. Br. J. Cancer 2008, 99, 622–631. [Google Scholar] [CrossRef] [PubMed]
  210. Du, W.J.; Chi, Y.; Yang, Z.X.; Li, Z.J.; Cui, J.J.; Song, B.Q.; Li, X.; Yang, S.G.; Han, Z.B.; Han, Z.C. Heterogeneity of Proangiogenic Features in Mesenchymal Stem Cells Derived from Bone Marrow, Adipose Tissue, Umbilical Cord, and Placenta. Stem Cell Res. Ther. 2016, 7, 163. [Google Scholar] [CrossRef] [PubMed]
  211. Huang, W.-H.; Chang, M.-C.; Tsai, K.-S.; Hung, M.-C.; Chen, H.-L.; Hung, S.-C. Mesenchymal Stem Cells Promote Growth and Angiogenesis of Tumors in Mice. Oncogene 2013, 32, 4343–4354. [Google Scholar] [CrossRef]
  212. Coffman, L.G.; Choi, Y.-J.; McLean, K.; Allen, B.L.; di Magliano, M.P.; Buckanovich, R.J. Human Carcinoma-Associated Mesenchymal Stem Cells Promote Ovarian Cancer Chemotherapy Resistance via a BMP4/HH Signaling Loop. Oncotarget 2016, 7, 6916–6932. [Google Scholar] [CrossRef]
  213. Jeon, E.S.; Heo, S.C.; Lee, I.H.; Choi, Y.J.; Park, J.H.; Choi, K.U.; Park, D.Y.; Suh, D.-S.; Yoon, M.-S.; Kim, J.H. Ovarian Cancer-Derived Lysophosphatidic Acid Stimulates Secretion of VEGF and Stromal Cell-Derived Factor-1α from Human Mesenchymal Stem Cells. Exp. Mol. Med. 2010, 42, 280–293. [Google Scholar] [CrossRef]
  214. Li, S.; Han, Y.; Lu, M.; Liu, Z.; Jin, J.; Guo, Q.; Wang, Y.; Liu, H. Mesenchymal Stem Cell-exosome-mediated Matrix Metalloproteinase 1 Participates in Oral Leukoplakia and Carcinogenesis by Inducing Angiogenesis. J. Oral Pathol. Med. 2022, 51, 638–648. [Google Scholar] [CrossRef]
  215. Hung, S.-C.; Pochampally, R.R.; Chen, S.-C.; Hsu, S.-C.; Prockop, D.J. Angiogenic Effects of Human Multipotent Stromal Cell Conditioned Medium Activate the PI3K-Akt Pathway in Hypoxic Endothelial Cells to Inhibit Apoptosis, Increase Survival, and Stimulate Angiogenesis. Stem Cells 2007, 25, 2363–2370. [Google Scholar] [CrossRef]
  216. Brogi, E.; Wu, T.; Namiki, A.; Isner, J.M. Indirect Angiogenic Cytokines Upregulate VEGF and BFGF Gene Expression in Vascular Smooth Muscle Cells, Whereas Hypoxia Upregulates VEGF Expression Only. Circulation 1994, 90, 649–652. [Google Scholar] [CrossRef] [PubMed]
  217. Dias, S.; Choy, M.; Alitalo, K.; Rafii, S. Vascular Endothelial Growth Factor (VEGF)–C Signaling through FLT-4 (VEGFR-3) Mediates Leukemic Cell Proliferation, Survival, and Resistance to Chemotherapy. Blood 2002, 99, 2179–2184. [Google Scholar] [CrossRef] [PubMed]
  218. Konopleva, M.; Konoplev, S.; Hu, W.; Zaritskey, A.; Afanasiev, B.; Andreeff, M. Stromal Cells Prevent Apoptosis of AML Cells by Up-Regulation of Anti-Apoptotic Proteins. Leukemia 2002, 16, 1713–1724. [Google Scholar] [CrossRef] [PubMed]
  219. Wu, X.-B.; Liu, Y.; Wang, G.-H.; Xu, X.; Cai, Y.; Wang, H.-Y.; Li, Y.-Q.; Meng, H.-F.; Dai, F.; Jin, J.-D. Mesenchymal Stem Cells Promote Colorectal Cancer Progression through AMPK/MTOR-Mediated NF-ΚB Activation. Sci. Rep. 2016, 6, 21420. [Google Scholar] [CrossRef]
  220. Burger, J.A.; Tsukada, N.; Burger, M.; Zvaifler, N.J.; Dell’Aquila, M.; Kipps, T.J. Blood-Derived Nurse-like Cells Protect Chronic Lymphocytic Leukemia B Cells from Spontaneous Apoptosis through Stromal Cell–Derived Factor-1. Blood 2000, 96, 2655–2663. [Google Scholar] [CrossRef] [PubMed]
  221. Chaffer, C.L.; Weinberg, R.A. A Perspective on Cancer Cell Metastasis. Science 2011, 331, 1559–1564. [Google Scholar] [CrossRef]
  222. Yang, Y.; Bucan, V.; Baehre, H.; Von Der Ohe, J.; Otte, A.; Hass, R. Acquisition of New Tumor Cell Properties by MSC-Derived Exosomes. Int. J. Oncol. 2015, 47, 244–252. [Google Scholar] [CrossRef]
  223. Tu, B.; Zhu, J.; Liu, S.; Wang, L.; Fan, Q.; Hao, Y.; Fan, C.; Tang, T.-T. Mesenchymal Stem Cells Promote Osteosarcoma Cell Survival and Drug Resistance through Activation of STAT3. Oncotarget 2016, 7, 48296–48308. [Google Scholar] [CrossRef]
  224. Correa, D.; Somoza, R.A.; Lin, P.; Schiemann, W.P.; Caplan, A.I. Mesenchymal Stem Cells Regulate Melanoma Cancer Cells Extravasation to Bone and Liver at Their Perivascular Niche. Int. J. Cancer 2016, 138, 417–427. [Google Scholar] [CrossRef]
  225. Balch, C.M.; Houghton, A.N.; Sober, A.J.; Soong, S. Cutaneous Melanoma, 4th Edition. Dermatol. Surg. 2005, 31, 1715. [Google Scholar] [CrossRef]
  226. Kodet, O.; Kučera, J.; Strnadová, K.; Dvořánková, B.; Štork, J.; Lacina, L.; Smetana, K. Cutaneous Melanoma Dissemination Is Dependent on the Malignant Cell Properties and Factors of Intercellular Crosstalk in the Cancer Microenvironment (Review). Int. J. Oncol. 2020, 57, 619–630. [Google Scholar] [CrossRef]
  227. Chaturvedi, P.; Gilkes, D.M.; Wong, C.C.L.; Kshitiz; Luo, W.; Zhang, H.; Wei, H.; Takano, N.; Schito, L.; Levchenko, A.; et al. Hypoxia-Inducible Factor–Dependent Breast Cancer–Mesenchymal Stem Cell Bidirectional Signaling Promotes Metastasis. J. Clin. Investig. 2013, 123, 189–205. [Google Scholar] [CrossRef]
  228. Kudo-Saito, C.; Fuwa, T.; Murakami, K.; Kawakami, Y. Targeting FSTL1 Prevents Tumor Bone Metastasis and Consequent Immune Dysfunction. Cancer Res. 2013, 73, 6185–6193. [Google Scholar] [CrossRef]
  229. McGuire, J.J.; Frieling, J.S.; Lo, C.H.; Li, T.; Muhammad, A.; Lawrence, H.R.; Lawrence, N.J.; Cook, L.M.; Lynch, C.C. Mesenchymal Stem Cell-Derived Interleukin-28 Drives the Selection of Apoptosis Resistant Bone Metastatic Prostate Cancer. Nat. Commun. 2021, 12, 723. [Google Scholar] [CrossRef]
  230. Rodini, C.O.; Gonçalves da Silva, P.B.; Assoni, A.F.; Carvalho, V.M.; Okamoto, O.K. Mesenchymal Stem Cells Enhance Tumorigenic Properties of Human Glioblastoma through Independent Cell-Cell Communication Mechanisms. Oncotarget 2018, 9, 24766–24777. [Google Scholar] [CrossRef]
  231. McAndrews, K.M.; McGrail, D.J.; Ravikumar, N.; Dawson, M.R. Mesenchymal Stem Cells Induce Directional Migration of Invasive Breast Cancer Cells through TGF-β. Sci. Rep. 2015, 5, 16941. [Google Scholar] [CrossRef]
  232. Lin, R.; Wang, S.; Zhao, R.C. Exosomes from Human Adipose-Derived Mesenchymal Stem Cells Promote Migration through Wnt Signaling Pathway in a Breast Cancer Cell Model. Mol. Cell. Biochem. 2013, 383, 13–20. [Google Scholar] [CrossRef]
  233. Wang, W.; Wang, Q.; Huang, D.-B.; Sun, Q.-K.; Wu, S.-S.; Zhao, Y.-J.; Jia, W.; Hu, D.-S.; He, Y.-F. Tumor-Associated Mesenchymal Stem Cells Promote Hepatocellular Carcinoma Metastasis via a DNM3OS/KDM6B/TIAM1 Axis. Cancer Lett. 2021, 503, 19–31. [Google Scholar] [CrossRef]
  234. Martin, F.T.; Dwyer, R.M.; Kelly, J.; Khan, S.; Murphy, J.M.; Curran, C.; Miller, N.; Hennessy, E.; Dockery, P.; Barry, F.P.; et al. Potential Role of Mesenchymal Stem Cells (MSCs) in the Breast Tumour Microenvironment: Stimulation of Epithelial to Mesenchymal Transition (EMT). Breast Cancer Res. Treat. 2010, 124, 317–326. [Google Scholar] [CrossRef]
  235. Xue, Z.; Wu, X.; Chen, X.; Liu, Y.; Wang, X.; Wu, K.; Nie, Y.; Fan, D. Mesenchymal Stem Cells Promote Epithelial to Mesenchymal Transition and Metastasis in Gastric Cancer Though Paracrine Cues and Close Physical Contact: MSCs Promote EMT. J. Cell. Biochem. 2015, 116, 618–627. [Google Scholar] [CrossRef]
  236. Bates, R.C.; Mercurio, A.M. Tumor Necrosis Factor-Alpha Stimulates the Epithelial-to-Mesenchymal Transition of Human Colonic Organoids. Mol. Biol. Cell 2003, 14, 1790–1800. [Google Scholar] [CrossRef]
  237. De Wever, O.; Demetter, P.; Mareel, M.; Bracke, M. Stromal Myofibroblasts Are Drivers of Invasive Cancer Growth. Int. J. Cancer 2008, 123, 2229–2238. [Google Scholar] [CrossRef]
  238. Xu, M.-H.; Gao, X.; Luo, D.; Zhou, X.-D.; Xiong, W.; Liu, G.-X. EMT and Acquisition of Stem Cell-like Properties Are Involved in Spontaneous Formation of Tumorigenic Hybrids between Lung Cancer and Bone Marrow-Derived Mesenchymal Stem Cells. PLoS ONE 2014, 9, e87893. [Google Scholar] [CrossRef]
  239. Fan, X.-L.; Zhang, Y.; Li, X.; Fu, Q.-L. Mechanisms Underlying the Protective Effects of Mesenchymal Stem Cell-Based Therapy. Cell. Mol. Life Sci. 2020, 77, 2771–2794. [Google Scholar] [CrossRef]
  240. Szewc, M.; Radzikowska-Bűchner, E.; Wdowiak, P.; Kozak, J.; Kuszta, P.; Niezabitowska, E.; Matysiak, J.; Kubiński, K.; Masłyk, M. MSCs as Tumor-Specific Vectors for the Delivery of Anticancer Agents—A Potential Therapeutic Strategy in Cancer Diseases: Perspectives for Quinazoline Derivatives. Int. J. Mol. Sci. 2022, 23, 2745. [Google Scholar] [CrossRef]
  241. Li, Z.; Fan, D.; Xiong, D. Mesenchymal Stem Cells as Delivery Vectors for Anti-Tumor Therapy. Stem Cell Investig. 2015, 2, 6. [Google Scholar] [CrossRef]
  242. Mirzaei, H.; Sahebkar, A.; Avan, A.; Jaafari, M.R.; Salehi, R.; Salehi, H.; Baharvand, H.; Rezaei, A.; Hadjati, J.; Pawelek, J.M.; et al. Application of Mesenchymal Stem Cells in Melanoma: A Potential Therapeutic Strategy for Delivery of Targeted Agents. Curr. Med. Chem. 2016, 23, 455–463. [Google Scholar] [CrossRef]
  243. Zhang, T.-Y.; Huang, B.; Yuan, Z.-Y.; Hu, Y.-L.; Tabata, Y.; Gao, J.-Q. Gene Recombinant Bone Marrow Mesenchymal Stem Cells as a Tumor-Targeted Suicide Gene Delivery Vehicle in Pulmonary Metastasis Therapy Using Non-Viral Transfection. Nanomed. Nanotechnol. Biol. Med. 2014, 10, 257–267. [Google Scholar] [CrossRef]
  244. Katz, O.B.; Shaked, Y. Host Effects Contributing to Cancer Therapy Resistance. Drug Resist. Updates 2015, 19, 33–42. [Google Scholar] [CrossRef]
  245. Meads, M.B.; Gatenby, R.A.; Dalton, W.S. Environment-Mediated Drug Resistance: A Major Contributor to Minimal Residual Disease. Nat. Rev. Cancer 2009, 9, 665–674. [Google Scholar] [CrossRef] [PubMed]
  246. Houthuijzen, J.M.; Daenen, L.G.M.; Roodhart, J.M.L.; Voest, E.E. The Role of Mesenchymal Stem Cells in Anti-Cancer Drug Resistance and Tumour Progression. Br. J. Cancer 2012, 106, 1901–1906. [Google Scholar] [CrossRef] [PubMed]
  247. Borst, P. Cancer Drug Pan-Resistance: Pumps, Cancer Stem Cells, Quiescence, Epithelial to Mesenchymal Transition, Blocked Cell Death Pathways, Persisters or What? Open Biol. 2012, 2, 120066. [Google Scholar] [CrossRef] [PubMed]
  248. Rizzolio, S.; Giordano, S.; Corso, S. The Importance of Being CAFs (in Cancer Resistance to Targeted Therapies). J. Exp. Clin. Cancer Res. 2022, 41, 319. [Google Scholar] [CrossRef]
  249. Lis, R.; Touboul, C.; Mirshahi, P.; Ali, F.; Mathew, S.; Nolan, D.J.; Maleki, M.; Abdalla, S.A.; Raynaud, C.M.; Querleu, D.; et al. Tumor Associated Mesenchymal Stem Cells Protects Ovarian Cancer Cells from Hyperthermia through CXCL12. Int. J. Cancer 2011, 128, 715–725. [Google Scholar] [CrossRef]
  250. Dreuw, A.; Hermanns, H.M.; Heise, R.; Joussen, S.; Rodríguez, F.; Marquardt, Y.; Jugert, F.; Merk, H.F.; Heinrich, P.C.; Baron, J.M. Interleukin-6-Type Cytokines Upregulate Expression of Multidrug Resistance-Associated Proteins in NHEK and Dermal Fibroblasts. J. Investig. Dermatol. 2005, 124, 28–37. [Google Scholar] [CrossRef]
  251. Gilbert, L.A.; Hemann, M.T. DNA Damage-Mediated Induction of a Chemoresistant Niche. Cell 2010, 143, 355–366. [Google Scholar] [CrossRef]
  252. Roodhart, J.M.L.; Daenen, L.G.M.; Stigter, E.C.A.; Prins, H.-J.; Gerrits, J.; Houthuijzen, J.M.; Gerritsen, M.G.; Schipper, H.S.; Backer, M.J.G.; van Amersfoort, M.; et al. Mesenchymal Stem Cells Induce Resistance to Chemotherapy through the Release of Platinum-Induced Fatty Acids. Cancer Cell 2011, 20, 370–383. [Google Scholar] [CrossRef]
  253. Mariathasan, S.; Turley, S.J.; Nickles, D.; Castiglioni, A.; Yuen, K.; Wang, Y.; Kadel, E.E., III; Koeppen, H.; Astarita, J.L.; Cubas, R.; et al. TGFβ Attenuates Tumour Response to PD-L1 Blockade by Contributing to Exclusion of T Cells. Nature 2018, 554, 544–548. [Google Scholar] [CrossRef]
  254. Shibue, T.; Weinberg, R.A. EMT, CSCs, and Drug Resistance: The Mechanistic Link and Clinical Implications. Nat. Rev. Clin. Oncol. 2017, 14, 611–629. [Google Scholar] [CrossRef]
  255. Mani, S.A.; Guo, W.; Liao, M.-J.; Eaton, E.N.; Ayyanan, A.; Zhou, A.Y.; Brooks, M.; Reinhard, F.; Zhang, C.C.; Shipitsin, M.; et al. The Epithelial-Mesenchymal Transition Generates Cells with Properties of Stem Cells. Cell 2008, 133, 704–715. [Google Scholar] [CrossRef] [PubMed]
  256. Ramasamy, R.; Lam, E.W.-F.; Soeiro, I.; Tisato, V.; Bonnet, D.; Dazzi, F. Mesenchymal Stem Cells Inhibit Proliferation and Apoptosis of Tumor Cells: Impact on In Vivo Tumor Growth. Leukemia 2007, 21, 304–310. [Google Scholar] [CrossRef] [PubMed]
  257. Melzer, C.; von der Ohe, J.; Lehnert, H.; Ungefroren, H.; Hass, R. Cancer Stem Cell Niche Models and Contribution by Mesenchymal Stroma/Stem Cells. Mol. Cancer 2017, 16, 28. [Google Scholar] [CrossRef] [PubMed]
  258. Li, H.-J.; Reinhardt, F.; Herschman, H.R.; Weinberg, R.A. Cancer-Stimulated Mesenchymal Stem Cells Create a Carcinoma Stem Cell Niche via Prostaglandin E2 Signaling. Cancer Discov. 2012, 2, 840–855. [Google Scholar] [CrossRef] [PubMed]
  259. Skolekova, S.; Matuskova, M.; Bohac, M.; Toro, L.; Durinikova, E.; Tyciakova, S.; Demkova, L.; Gursky, J.; Kucerova, L. Cisplatin-Induced Mesenchymal Stromal Cells-Mediated Mechanism Contributing to Decreased Antitumor Effect in Breast Cancer Cells. Cell Commun. Signal. 2016, 14, 4. [Google Scholar] [CrossRef] [PubMed]
  260. Kabashima-Niibe, A.; Higuchi, H.; Takaishi, H.; Masugi, Y.; Matsuzaki, Y.; Mabuchi, Y.; Funakoshi, S.; Adachi, M.; Hamamoto, Y.; Kawachi, S.; et al. Mesenchymal Stem Cells Regulate Epithelial-Mesenchymal Transition and Tumor Progression of Pancreatic Cancer Cells. Cancer Sci. 2013, 104, 157–164. [Google Scholar] [CrossRef] [PubMed]
  261. Peitzsch, C.; Kurth, I.; Kunz-Schughart, L.; Baumann, M.; Dubrovska, A. Discovery of the Cancer Stem Cell Related Determinants of Radioresistance. Radiother. Oncol. 2013, 108, 378–387. [Google Scholar] [CrossRef]
  262. Kreso, A.; Dick, J.E. Evolution of the Cancer Stem Cell Model. Cell Stem Cell 2014, 14, 275–291. [Google Scholar] [CrossRef]
  263. Meacham, C.E.; Morrison, S.J. Tumor Heterogeneity and Cancer Cell Plasticity. Nature 2013, 501, 328–337. [Google Scholar] [CrossRef]
  264. Phillips, T.M.; McBride, W.H.; Pajonk, F. The Response of CD24−/Low /CD44 + Breast Cancer–Initiating Cells to Radiation. J. Natl. Cancer Inst. 2006, 98, 1777–1785. [Google Scholar] [CrossRef]
  265. Woodward, W.A.; Chen, M.S.; Behbod, F.; Alfaro, M.P.; Buchholz, T.A.; Rosen, J.M. WNT/β-Catenin Mediates Radiation Resistance of Mouse Mammary Progenitor Cells. Proc. Natl. Acad. Sci. USA 2007, 104, 618–623. [Google Scholar] [CrossRef]
  266. Wagner, V.P.; Martins, M.A.T.; Martins, M.D.; Warner, K.A.; Webber, L.P.; Squarize, C.H.; Nör, J.E.; Castilho, R.M. Overcoming Adaptive Resistance in Mucoepidermoid Carcinoma through Inhibition of the IKK-β/IκBα/NFκB Axis. Oncotarget 2016, 7, 73032–73044. [Google Scholar] [CrossRef]
  267. Bighetti-Trevisan, R.L.; Sousa, L.O.; Castilho, R.M.; Almeida, L.O. Cancer Stem Cells: Powerful Targets to Improve Current Anticancer Therapeutics. Stem Cells Int. 2019, 2019, 9618065. [Google Scholar] [CrossRef] [PubMed]
  268. Lee, S.Y.; Jeong, E.K.; Ju, M.K.; Jeon, H.M.; Kim, M.Y.; Kim, C.H.; Park, H.G.; Han, S.I.; Kang, H.S. Induction of Metastasis, Cancer Stem Cell Phenotype, and Oncogenic Metabolism in Cancer Cells by Ionizing Radiation. Mol. Cancer 2017, 16, 10. [Google Scholar] [CrossRef]
  269. Moncharmont, C.; Levy, A.; Guy, J.-B.; Falk, A.T.; Guilbert, M.; Trone, J.-C.; Alphonse, G.; Gilormini, M.; Ardail, D.; Toillon, R.-A.; et al. Radiation-Enhanced Cell Migration/Invasion Process: A Review. Crit. Rev. Oncol./Hematol. 2014, 92, 133–142. [Google Scholar] [CrossRef]
  270. Pei, J.; Park, I.-H.; Ryu, H.-H.; Li, S.-Y.; Li, C.-H.; Lim, S.-H.; Wen, M.; Jang, W.-Y.; Jung, S. Sublethal Dose of Irradiation Enhances Invasion of Malignant Glioma Cells through P53-MMP 2 Pathway in U87MG Mouse Brain Tumor Model. Radiat. Oncol. 2015, 10, 164. [Google Scholar] [CrossRef]
  271. Nowakowski, A.; Drela, K.; Rozycka, J.; Janowski, M.; Lukomska, B. Engineered Mesenchymal Stem Cells as an Anti-Cancer Trojan Horse. Stem Cells Dev. 2016, 25, 1513–1531. [Google Scholar] [CrossRef]
  272. Hagenhoff, A.; Bruns, C.J.; Zhao, Y.; von Lüttichau, I.; Niess, H.; Spitzweg, C.; Nelson, P.J. Harnessing Mesenchymal Stem Cell Homing as an Anticancer Therapy. Expert Opin. Biol. Ther. 2016, 16, 1079–1092. [Google Scholar] [CrossRef] [PubMed]
  273. Wu, H.-H.; Zhou, Y.; Tabata, Y.; Gao, J.-Q. Mesenchymal Stem Cell-Based Drug Delivery Strategy: From Cells to Biomimetic. J. Control. Release 2019, 294, 102–113. [Google Scholar] [CrossRef]
  274. Studeny, M.; Marini, F.C.; Champlin, R.E.; Zompetta, C.; Fidler, I.J.; Andreeff, M. Bone Marrow-Derived Mesenchymal Stem Cells as Vehicles for Interferon-β Delivery into Tumors1. Cancer Res. 2002, 62, 3603–3608. [Google Scholar]
  275. Ding, Y.; Wang, C.; Sun, Z.; Wu, Y.; You, W.; Mao, Z.; Wang, W. Mesenchymal Stem Cells Engineered by Nonviral Vectors: A Powerful Tool in Cancer Gene Therapy. Pharmaceutics 2021, 13, 913. [Google Scholar] [CrossRef] [PubMed]
  276. Zhu, Y.; Cheng, M.; Yang, Z.; Zeng, C.-Y.; Chen, J.; Xie, Y.; Luo, S.-W.; Zhang, K.-H.; Zhou, S.-F.; Lu, N.-H. Mesenchymal Stem Cell-Based NK4 Gene Therapy in Nude Mice Bearing Gastric Cancer Xenografts. Drug Des. Dev. Ther. 2014, 8, 2449–2462. [Google Scholar] [CrossRef] [PubMed]
  277. Gao, P.; Ding, Q.; Wu, Z.; Jiang, H.; Fang, Z. Therapeutic Potential of Human Mesenchymal Stem Cells Producing IL-12 in a Mouse Xenograft Model of Renal Cell Carcinoma. Cancer Lett. 2010, 290, 157–166. [Google Scholar] [CrossRef]
  278. Seo, S.H.; Kim, K.S.; Park, S.H.; Suh, Y.S.; Kim, S.J.; Jeun, S.-S.; Sung, Y.C. The Effects of Mesenchymal Stem Cells Injected via Different Routes on Modified IL-12-Mediated Antitumor Activity. Gene Ther. 2011, 18, 488–495. [Google Scholar] [CrossRef] [PubMed]
  279. Mirlekar, B.; Pylayeva-Gupta, Y. IL-12 Family Cytokines in Cancer and Immunotherapy. Cancers 2021, 13, 167. [Google Scholar] [CrossRef]
  280. Guiho, R.; Biteau, K.; Grisendi, G.; Taurelle, J.; Chatelais, M.; Gantier, M.; Heymann, D.; Dominici, M.; Redini, F. TRAIL Delivered by Mesenchymal Stromal/Stem Cells Counteracts Tumor Development in Orthotopic Ewing Sarcoma Models. Int. J. Cancer 2016, 139, 2802–2811. [Google Scholar] [CrossRef] [PubMed]
  281. Rossignoli, F.; Spano, C.; Grisendi, G.; Foppiani, E.M.; Golinelli, G.; Mastrolia, I.; Bestagno, M.; Candini, O.; Petrachi, T.; Recchia, A.; et al. MSC-Delivered Soluble TRAIL and Paclitaxel as Novel Combinatory Treatment for Pancreatic Adenocarcinoma. Theranostics 2019, 9, 436–448. [Google Scholar] [CrossRef]
  282. Tamura, R.; Miyoshi, H.; Yoshida, K.; Okano, H.; Toda, M. Recent Progress in the Research of Suicide Gene Therapy for Malignant Glioma. Neurosurg. Rev. 2021, 44, 29–49. [Google Scholar] [CrossRef]
  283. Zhang, T.-Y.; Huang, B.; Wu, H.-B.; Wu, J.-H.; Li, L.-M.; Li, Y.-X.; Hu, Y.-L.; Han, M.; Shen, Y.-Q.; Tabata, Y.; et al. Synergistic Effects of Co-Administration of Suicide Gene Expressing Mesenchymal Stem Cells and Prodrug-Encapsulated Liposome on Aggressive Lung Melanoma Metastases in Mice. J. Control. Release 2015, 209, 260–271. [Google Scholar] [CrossRef]
  284. Cavarretta, I.T.; Altanerova, V.; Matuskova, M.; Kucerova, L.; Culig, Z.; Altaner, C. Adipose Tissue–Derived Mesenchymal Stem Cells Expressing Prodrug-Converting Enzyme Inhibit Human Prostate Tumor Growth. Mol. Ther. 2010, 18, 223–231. [Google Scholar] [CrossRef]
  285. Kucerova, L.; Altanerova, V.; Matuskova, M.; Tyciakova, S.; Altaner, C. Adipose Tissue–Derived Human Mesenchymal Stem Cells Mediated Prodrug Cancer Gene Therapy. Cancer Res. 2007, 67, 6304–6313. [Google Scholar] [CrossRef]
  286. Segaliny, A.I.; Cheng, J.L.; Farhoodi, H.P.; Toledano, M.; Yu, C.C.; Tierra, B.; Hildebrand, L.; Liu, L.; Liao, M.J.; Cho, J.; et al. Combinatorial Targeting of Cancer Bone Metastasis Using MRNA Engineered Stem Cells. EBioMedicine 2019, 45, 39–57. [Google Scholar] [CrossRef] [PubMed]
  287. Parato, K.A.; Senger, D.; Forsyth, P.A.J.; Bell, J.C. Recent Progress in the Battle between Oncolytic Viruses and Tumours. Nat. Rev. Cancer 2005, 5, 965–976. [Google Scholar] [CrossRef]
  288. Liang, M. Oncorine, the World First Oncolytic Virus Medicine and Its Update in China. Curr. Cancer Drug Targets 2018, 18, 171–176. [Google Scholar] [CrossRef]
  289. Mondal, M.; Guo, J.; He, P.; Zhou, D. Recent Advances of Oncolytic Virus in Cancer Therapy. Hum. Vaccines Immunother. 2020, 16, 2389–2402. [Google Scholar] [CrossRef] [PubMed]
  290. Cao, G.; He, X.; Sun, Q.; Chen, S.; Wan, K.; Xu, X.; Feng, X.; Li, P.; Chen, B.; Xiong, M. The Oncolytic Virus in Cancer Diagnosis and Treatment. Front. Oncol. 2020, 10, 1786. [Google Scholar] [CrossRef] [PubMed]
  291. Goradel, N.H.; Baker, A.T.; Arashkia, A.; Ebrahimi, N.; Ghorghanlu, S.; Negahdari, B. Oncolytic Virotherapy: Challenges and Solutions. Curr. Probl. Cancer 2021, 45, 100639. [Google Scholar] [CrossRef] [PubMed]
  292. Maroun, J.; Muñoz-Alía, M.; Ammayappan, A.; Schulze, A.; Peng, K.-W.; Russell, S. Designing and Building Oncolytic Viruses. Future Virol. 2017, 12, 193–213. [Google Scholar] [CrossRef]
  293. Russell, S.J.; Peng, K.-W.; Bell, J.C. Oncolytic Virotherapy. Nat. Biotechnol. 2012, 30, 658–670. [Google Scholar] [CrossRef]
  294. Keshavarz, M.; Mohammad Miri, S.; Behboudi, E.; Arjeini, Y.; Dianat-Moghadam, H.; Ghaemi, A. Oncolytic Virus Delivery Modulated Immune Responses toward Cancer Therapy: Challenges and Perspectives. Int. Immunopharmacol. 2022, 108, 108882. [Google Scholar] [CrossRef]
  295. Pidelaserra-Martí, G.; Engeland, C.E. Mechanisms of Measles Virus Oncolytic Immunotherapy. Cytokine Growth Factor Rev. 2020, 56, 28–38. [Google Scholar] [CrossRef] [PubMed]
  296. Ghasemi Darestani, N.; Gilmanova, A.I.; Al-Gazally, M.E.; Zekiy, A.O.; Ansari, M.J.; Zabibah, R.S.; Jawad, M.A.; Al-Shalah, S.A.J.; Rizaev, J.A.; Alnassar, Y.S.; et al. Mesenchymal Stem Cell-Released Oncolytic Virus: An Innovative Strategy for Cancer Treatment. Cell Commun. Signal. 2023, 21, 43. [Google Scholar] [CrossRef] [PubMed]
  297. Lin, Y.; Zhang, H.; Liang, J.; Li, K.; Zhu, W.; Fu, L.; Wang, F.; Zheng, X.; Shi, H.; Wu, S.; et al. Identification and Characterization of Alphavirus M1 as a Selective Oncolytic Virus Targeting ZAP-Defective Human Cancers. Proc. Natl. Acad. Sci. USA 2014, 111, E4504–E4512. [Google Scholar] [CrossRef] [PubMed]
  298. Kaufman, H.L.; Kohlhapp, F.J.; Zloza, A. Oncolytic Viruses: A New Class of Immunotherapy Drugs. Nat. Rev. Drug Discov. 2015, 14, 642–662. [Google Scholar] [CrossRef]
  299. Toth, K.; Dhar, D.; Wold, W.S. Oncolytic (Replication-Competent) Adenoviruses as Anticancer Agents. Expert Opin. Biol. Ther. 2010, 10, 353–368. [Google Scholar] [CrossRef]
  300. Kwon, O.-J.; Kang, E.; Choi, J.-W.; Kim, S.W.; Yun, C.-O. Therapeutic Targeting of Chitosan–PEG–Folate-Complexed Oncolytic Adenovirus for Active and Systemic Cancer Gene Therapy. J. Control. Release 2013, 169, 257–265. [Google Scholar] [CrossRef]
  301. Alonso-Padilla, J.; Papp, T.; Kaján, G.L.; Benkő, M.; Havenga, M.; Lemckert, A.; Harrach, B.; Baker, A.H. Development of Novel Adenoviral Vectors to Overcome Challenges Observed with HAdV-5–Based Constructs. Mol. Ther. 2016, 24, 6–16. [Google Scholar] [CrossRef]
  302. Hendrickx, R.; Stichling, N.; Koelen, J.; Kuryk, L.; Lipiec, A.; Greber, U.F. Innate Immunity to Adenovirus. Hum. Gene Ther. 2014, 25, 265–284. [Google Scholar] [CrossRef]
  303. Taipale, K.; Liikanen, I.; Juhila, J.; Turkki, R.; Tähtinen, S.; Kankainen, M.; Vassilev, L.; Ristimäki, A.; Koski, A.; Kanerva, A.; et al. Chronic Activation of Innate Immunity Correlates with Poor Prognosis in Cancer Patients Treated with Oncolytic Adenovirus. Mol. Ther. 2016, 24, 175–183. [Google Scholar] [CrossRef]
  304. Msaouel, P.; Opyrchal, M.; Dispenzieri, A.; Peng, K.W.; Federspiel, M.J.; Russell, S.J.; Galanis, E. Clinical Trials with Oncolytic Measles Virus: Current Status and Future Prospects. Curr. Cancer Drug Targets 2018, 18, 177–187. [Google Scholar] [CrossRef]
  305. Kazimirsky, G.; Jiang, W.; Slavin, S.; Ziv-Av, A.; Brodie, C. Mesenchymal Stem Cells Enhance the Oncolytic Effect of Newcastle Disease Virus in Glioma Cells and Glioma Stem Cells via the Secretion of TRAIL. Stem Cell Res. Ther. 2016, 7, 149. [Google Scholar] [CrossRef] [PubMed]
  306. Mader, E.K.; Maeyama, Y.; Lin, Y.; Butler, G.W.; Russell, H.M.; Galanis, E.; Russell, S.J.; Dietz, A.B.; Peng, K.-W. Mesenchymal Stem Cell Carriers Protect Oncolytic Measles Viruses from Antibody Neutralization in an Orthotopic Ovarian Cancer Therapy Model. Clin. Cancer Res. 2009, 15, 7246–7255. [Google Scholar] [CrossRef] [PubMed]
  307. Zhang, J.; Chen, H.; Chen, C.; Liu, H.; He, Y.; Zhao, J.; Yang, P.; Mao, Q.; Xia, H. Systemic Administration of Mesenchymal Stem Cells Loaded with a Novel Oncolytic Adenovirus Carrying IL-24/Endostatin Enhances Glioma Therapy. Cancer Lett. 2021, 509, 26–38. [Google Scholar] [CrossRef]
  308. Ahmed, A.U.; Ulasov, I.V.; Mercer, R.W.; Lesniak, M.S. Maintaining and Loading Neural Stem Cells for Delivery of Oncolytic Adenovirus to Brain Tumors. In Oncolytic Viruses; Kirn, D.H., Liu, T.-C., Thorne, S.H., Eds.; Methods in Molecular Biology; Humana Press: Totowa, NJ, USA, 2012; Volume 797, pp. 97–109. ISBN 978-1-61779-339-4. [Google Scholar]
  309. Tyler, M.A.; Ulasov, I.V.; Sonabend, A.M.; Nandi, S.; Han, Y.; Marler, S.; Roth, J.; Lesniak, M.S. Neural Stem Cells Target Intracranial Glioma to Deliver an Oncolytic Adenovirus In Vivo. Gene Ther. 2009, 16, 262–278. [Google Scholar] [CrossRef] [PubMed]
  310. Song, D.; Jia, X.; Liu, X.; Hu, L.; Lin, K.; Xiao, T.; Qiao, Y.; Zhang, J.; Dan, J.; Wong, C.; et al. Identification of the Receptor of Oncolytic Virus M1 as a Therapeutic Predictor for Multiple Solid Tumors. Signal Transduct. Target. Ther. 2022, 7, 100. [Google Scholar] [CrossRef]
  311. Xiao, J.; Zeng, L.; Ding, S.; Chen, Y.; Zhang, X.; Bian, X.; Tian, G. Tumor-Tropic Adipose-Derived Mesenchymal Stromal Cell Mediated Bi2Se3 Nano-Radiosensitizers Delivery for Targeted Radiotherapy of Non-Small Cell Lung Cancer. Adv. Healthc. Mater. 2022, 11, 2200143. [Google Scholar] [CrossRef] [PubMed]
  312. Cihan, Y.B. Nanoparticle-Based Radiosensitizers in Radiotherapy Applications. Cancer Biother. Radiopharm. 2021, 36, 305–306. [Google Scholar] [CrossRef]
  313. Choi, J.; Kim, G.; Cho, S.B.; Im, H.-J. Radiosensitizing High-Z Metal Nanoparticles for Enhanced Radiotherapy of Glioblastoma multiforme. J. Nanobiotechnol. 2020, 18, 122. [Google Scholar] [CrossRef]
  314. Chen, Y.; Yang, J.; Fu, S.; Wu, J. Gold Nanoparticles as Radiosensitizers in Cancer Radiotherapy. Int. J. Nanomed. 2020, 15, 9407–9430. [Google Scholar] [CrossRef]
  315. Zhao, J.; Liu, P.; Ma, J.; Li, D.; Yang, H.; Chen, W.; Jiang, Y. Enhancement of Radiosensitization by Silver Nanoparticles Functionalized with Polyethylene Glycol and Aptamer As1411 for Glioma Irradiation Therapy. Int. J. Nanomed. 2019, 14, 9483–9496. [Google Scholar] [CrossRef]
  316. Yao, Y.; Li, P.; He, J.; Wang, D.; Hu, J.; Yang, X. Albumin-Templated Bi2Se3-MnO2 Nanocomposites with Promoted Catalase-like Activity for Enhanced Radiotherapy of Cancer. ACS Appl. Mater. Interfaces 2021, 13, 28650–28661. [Google Scholar] [CrossRef] [PubMed]
  317. Du, J.; Gu, Z.; Yan, L.; Yong, Y.; Yi, X.; Zhang, X.; Liu, J.; Wu, R.; Ge, C.; Chen, C.; et al. Poly(Vinylpyrollidone)- and Selenocysteine-Modified Bi 2 Se 3 Nanoparticles Enhance Radiotherapy Efficacy in Tumors and Promote Radioprotection in Normal Tissues. Adv. Mater. 2017, 29, 1701268. [Google Scholar] [CrossRef] [PubMed]
  318. Lammers, T.; Kiessling, F.; Ashford, M.; Hennink, W.; Crommelin, D.; Storm, G. Cancer Nanomedicine: Is Targeting Our Target? Nat. Rev. Mater. 2016, 1, 16069. [Google Scholar] [CrossRef]
  319. Wolfram, J.; Ferrari, M. Clinical Cancer Nanomedicine. Nano Today 2019, 25, 85–98. [Google Scholar] [CrossRef]
  320. Huang, R.-Y.; Lin, Y.-H.; Lin, S.-Y.; Li, Y.-N.; Chiang, C.-S.; Chang, C.-W. Magnetic Ternary Nanohybrids for Nonviral Gene Delivery of Stem Cells and Applications on Cancer Therapy. Theranostics 2019, 9, 2411–2423. [Google Scholar] [CrossRef]
  321. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  322. Song, X.; Zhao, Z.; Barber, B.; Farr, A.M.; Ivanov, B.; Novich, M. Overall Survival in Patients with Metastatic Melanoma. Curr. Med. Res. Opin. 2015, 31, 987–991. [Google Scholar] [CrossRef]
  323. Pollack, L.A.; Li, J.; Berkowitz, Z.; Weir, H.K.; Wu, X.-C.; Ajani, U.A.; Ekwueme, D.U.; Li, C.; Pollack, B.P. Melanoma Survival in the United States, 1992 to 2005. J. Am. Acad. Dermatol. 2011, 65, S78.e1–S78.e10. [Google Scholar] [CrossRef] [PubMed]
  324. Zaretsky, J.M.; Garcia-Diaz, A.; Shin, D.S.; Escuin-Ordinas, H.; Hugo, W.; Hu-Lieskovan, S.; Torrejon, D.Y.; Abril-Rodriguez, G.; Sandoval, S.; Barthly, L.; et al. Mutations Associated with Acquired Resistance to PD-1 Blockade in Melanoma. N. Engl. J. Med. 2016, 375, 819–829. [Google Scholar] [CrossRef]
  325. Zhang, G.; Frederick, D.T.; Wu, L.; Wei, Z.; Krepler, C.; Srinivasan, S.; Chae, Y.C.; Xu, X.; Choi, H.; Dimwamwa, E.; et al. Targeting Mitochondrial Biogenesis to Overcome Drug Resistance to MAPK Inhibitors. J. Clin. Investig. 2016, 126, 1834–1856. [Google Scholar] [CrossRef]
  326. Tyciakova, S.; Matuskova, M.; Bohovic, R.; Polakova, K.; Toro, L.; Skolekova, S.; Kucerova, L. Genetically Engineered Mesenchymal Stromal Cells Producing TNFα Have Tumour Suppressing Effect on Human Melanoma Xenograft: AT-MSC/HTNFα in Treatment of A375 Xenografts. J. Gene Med. 2015, 17, 54–67. [Google Scholar] [CrossRef] [PubMed]
  327. Ren, C.; Kumar, S.; Chanda, D.; Chen, J.; Mountz, J.D.; Ponnazhagan, S. Therapeutic Potential of Mesenchymal Stem Cells Producing Interferon-α in a Mouse Melanoma Lung Metastasis Model. Stem Cells 2008, 26, 2332–2338. [Google Scholar] [CrossRef] [PubMed]
  328. Seo, K.-W.; Lee, H.-W.; Oh, Y.-I.; Ahn, J.-O.; Koh, Y.-R.; Oh, S.-H.; Kang, S.-K.; Youn, H.-Y. Anti-Tumor Effects of Canine Adipose Tissue-Derived Mesenchymal Stromal Cell-Based Interferon-β Gene Therapy and Cisplatin in a Mouse Melanoma Model. Cytotherapy 2011, 13, 944–955. [Google Scholar] [CrossRef]
  329. Wong, V.L.Y.; Rieman, D.J.; Aronson, L.; Dalton, B.J.; Greig, R.; Anzano, M.A. Growth-Inhibitory Activity of Interferon-Beta against Human Colorectal Carcinoma Cell Lines. Int. J. Cancer 1989, 43, 526–530. [Google Scholar] [CrossRef] [PubMed]
  330. Audsley, K.M.; Wagner, T.; Ta, C.; Newnes, H.V.; Buzzai, A.C.; Barnes, S.A.; Wylie, B.; Armitage, J.; Kaisho, T.; Bosco, A.; et al. IFNβ Is a Potent Adjuvant for Cancer Vaccination Strategies. Front. Immunol. 2021, 12, 735133. [Google Scholar] [CrossRef]
  331. Yang, X.; Zhang, X.; Fu, M.L.; Weichselbaum, R.R.; Gajewski, T.F.; Guo, Y.; Fu, Y.-X. Targeting the Tumor Microenvironment with Interferon-β Bridges Innate and Adaptive Immune Responses. Cancer Cell 2014, 25, 37–48. [Google Scholar] [CrossRef] [PubMed]
  332. Yang, P.-M.; Hsieh, Y.-Y.; Du, J.-L.; Yen, S.-C.; Hung, C.-F. Sequential Interferon β-Cisplatin Treatment Enhances the Surface Exposure of Calreticulin in Cancer Cells via an Interferon Regulatory Factor 1-Dependent Manner. Biomolecules 2020, 10, 643. [Google Scholar] [CrossRef]
  333. Ahn, J.-O.; Coh, Y.-R.; Lee, H.-W.; Shin, I.-S.; Kang, S.-K.; Youn, H.-Y. Human Adipose Tissue-Derived Mesenchymal Stem Cells Inhibit Melanoma Growth In Vitro and In Vivo. Anticancer Res. 2015, 35, 159–168. [Google Scholar]
  334. Petrov, V.N.; Isaeva, E.V.; Ulyanenko, S.E.; Beketov, E.E.; Yatsenko, E.M.; Sayapina, E.V.; Lepekhina, L.A.; Nasedkina, N.V.; Grivtsova, L.Y.; Kaprin, A.D. In Vivo Effects of Human Bone Marrow Mesenchymal Stromal Cells on the Development of Experimental B16 Melanoma in Mice. Bull. Exp. Biol. Med. 2020, 168, 561–565. [Google Scholar] [CrossRef]
  335. Çakır, U.; Hajdara, A.; Széky, B.; Mayer, B.; Kárpáti, S.; Mezey, É.; Silló, P.; Szakács, G.; Füredi, A.; Pós, Z.; et al. Mesenchymal-Stromal Cell-like Melanoma-Associated Fibroblasts Increase IL-10 Production by Macrophages in a Cyclooxygenase/Indoleamine 2,3-Dioxygenase-Dependent Manner. Cancers 2021, 13, 6173. [Google Scholar] [CrossRef]
  336. Kucerova, L.; Matuskova, M.; Pastorakova, A.; Tyciakova, S.; Jakubikova, J.; Bohovic, R.; Altanerova, V.; Altaner, C. Cytosine Deaminase Expressing Human Mesenchymal Stem Cells Mediated Tumour Regression in Melanoma Bearing Mice. J. Gene Med. 2008, 10, 1071–1082. [Google Scholar] [CrossRef]
  337. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer Statistics, 2022. CA Cancer J. Clin. 2022, 72, 7–33. [Google Scholar] [CrossRef] [PubMed]
  338. Suzuki, K.; Sun, R.; Origuchi, M.; Kanehira, M.; Takahata, T.; Itoh, J.; Umezawa, A.; Kijima, H.; Fukuda, S.; Saijo, Y. Mesenchymal Stromal Cells Promote Tumor Growth through the Enhancement of Neovascularization. Mol. Med. 2011, 17, 579–587. [Google Scholar] [CrossRef]
  339. Gu, J.J.; Hoj, J.; Rouse, C.; Pendergast, A.M. Mesenchymal Stem Cells Promote Metastasis through Activation of an ABL-MMP9 Signaling Axis in Lung Cancer Cells. PLoS ONE 2020, 15, e0241423. [Google Scholar] [CrossRef]
  340. Kolluri, K.K.; Laurent, G.J.; Janes, S.M. Mesenchymal Stem Cells as Vectors for Lung Cancer Therapy. Respiration 2013, 85, 443–451. [Google Scholar] [CrossRef]
  341. Loebinger, M.R.; Kyrtatos, P.G.; Turmaine, M.; Price, A.N.; Pankhurst, Q.; Lythgoe, M.F.; Janes, S.M. Magnetic Resonance Imaging of Mesenchymal Stem Cells Homing to Pulmonary Metastases Using Biocompatible Magnetic Nanoparticles. Cancer Res. 2009, 69, 8862–8867. [Google Scholar] [CrossRef] [PubMed]
  342. Kim, S.-K.; Cargioli, T.G.; Machluf, M.; Yang, W.; Sun, Y.; Al-Hashem, R.; Kim, S.U.; Black, P.M.; Carroll, R.S. PEX-Producing Human Neural Stem Cells Inhibit Tumor Growth in a Mouse Glioma Model. Clin. Cancer Res. 2005, 11, 5965–5970. [Google Scholar] [CrossRef] [PubMed]
  343. Loebinger, M.R.; Eddaoudi, A.; Davies, D.; Janes, S.M. Mesenchymal Stem Cell Delivery of TRAIL Can Eliminate Metastatic Cancer. Cancer Res. 2009, 69, 4134–4142. [Google Scholar] [CrossRef]
  344. Mueller, L.P.; Luetzkendorf, J.; Widder, M.; Nerger, K.; Caysa, H.; Mueller, T. TRAIL-Transduced Multipotent Mesenchymal Stromal Cells (TRAIL-MSC) Overcome TRAIL Resistance in Selected CRC Cell Lines In Vitro and In Vivo. Cancer Gene Ther. 2011, 18, 229–239. [Google Scholar] [CrossRef]
  345. Zhao, X.; Wu, X.; Qian, M.; Song, Y.; Wu, D.; Zhang, W. Knockdown of TGF-Β1 Expression in Human Umbilical Cord Mesenchymal Stem Cells Reverts Their Exosome-Mediated EMT Promoting Effect on Lung Cancer Cells. Cancer Lett. 2018, 428, 34–44. [Google Scholar] [CrossRef]
  346. Bechet, D.; Auger, F.; Couleaud, P.; Marty, E.; Ravasi, L.; Durieux, N.; Bonnet, C.; Plénat, F.; Frochot, C.; Mordon, S.; et al. Multifunctional Ultrasmall Nanoplatforms for Vascular-Targeted Interstitial Photodynamic Therapy of Brain Tumors Guided by Real-Time MRI. Nanomed.: Nanotechnol. Biol. Med. 2015, 11, 657–670. [Google Scholar] [CrossRef] [PubMed]
  347. Yao, S.; Li, X.; Liu, J.; Sun, Y.; Wang, Z.; Jiang, Y. Maximized Nanodrug-Loaded Mesenchymal Stem Cells by a Dual Drug-Loaded Mode for the Systemic Treatment of Metastatic Lung Cancer. Drug Deliv. 2017, 24, 1372–1383. [Google Scholar] [CrossRef] [PubMed]
  348. Hakkarainen, T.; Särkioja, M.; Lehenkari, P.; Miettinen, S.; Ylikomi, T.; Suuronen, R.; Desmond, R.A.; Kanerva, A.; Hemminki, A. Human Mesenchymal Stem Cells Lack Tumor Tropism but Enhance the Antitumor Activity of Oncolytic Adenoviruses in Orthotopic Lung and Breast Tumors. Hum. Gene Ther. 2007, 18, 627–641. [Google Scholar] [CrossRef] [PubMed]
  349. Ando, M.; Hoyos, V.; Yagyu, S.; Tao, W.; Ramos, C.A.; Dotti, G.; Brenner, M.K.; Bouchier-Hayes, L. Bortezomib Sensitizes Non-Small Cell Lung Cancer to Mesenchymal Stromal Cell-Delivered Inducible Caspase-9-Mediated Cytotoxicity. Cancer Gene Ther. 2014, 21, 472–482. [Google Scholar] [CrossRef] [PubMed]
Figure 2. Recruitment of distant (BM)/resident MSCs, activation via diverse mechanisms, and differentiation to CAFs of the TME. Perivascular cells (MSCs/fibroblasts) as well as pericytes (blue) are resident cells, while MSCs from the bone marrow can be recruited in tumor sites. Factors derived mainly from necrotic tissues (notably in response to therapies) and infiltrating immune cells such as macrophages may contribute to the recruitment, proliferation, and activation of MSCs/Fib. MSCs will differentiate into CAFs with either restraining or promoting activities towards cancer cells. (Figure designed by BioRender).
Figure 2. Recruitment of distant (BM)/resident MSCs, activation via diverse mechanisms, and differentiation to CAFs of the TME. Perivascular cells (MSCs/fibroblasts) as well as pericytes (blue) are resident cells, while MSCs from the bone marrow can be recruited in tumor sites. Factors derived mainly from necrotic tissues (notably in response to therapies) and infiltrating immune cells such as macrophages may contribute to the recruitment, proliferation, and activation of MSCs/Fib. MSCs will differentiate into CAFs with either restraining or promoting activities towards cancer cells. (Figure designed by BioRender).
Ijms 24 13511 g002
Figure 3. Mesenchymal stem cells are highly plastic and exhibit multipotency (transdifferentiation) in the tumor microenvironment. CAFs: cancer-associated fibroblasts; EMT: epithelial–mesenchymal transition; EndMT: endothelial–mesenchymal transition; MSCs: mesenchymal stem cells. (Figure designed by BioRender).
Figure 3. Mesenchymal stem cells are highly plastic and exhibit multipotency (transdifferentiation) in the tumor microenvironment. CAFs: cancer-associated fibroblasts; EMT: epithelial–mesenchymal transition; EndMT: endothelial–mesenchymal transition; MSCs: mesenchymal stem cells. (Figure designed by BioRender).
Ijms 24 13511 g003
Figure 4. Interactions and regulations involving mesenchymal stem cells in the tumor microenvironment. AKT: protein kinase B; α-SMA: α-smooth muscle actin; C3a: complement component 3 anaphylatoxins; CAF: cancer-associated fibroblast; CCL: CC chemokine ligand; COL: collagen; Cox-2: cyclooxygenase-2; CSC: cancer stem cell; CXCL: C-X-C motif ligand; CXCR: C-X-C motif chemokine receptor; DDR2: discoidin domain receptor tyrosine kinase 2; ERK: extracellular signal-regulated kinase; FASL: fas ligand; FSP1: fibroblast-specific protein 1; HGF: hepatocyte growth factor; HIF-1: hypoxia-inducible factor-1; IDO: indoleamine-2,3-dioxygenase; IFN-γ: interferon-γ; IGF: insulin growth factor; IL: interleukin; iNOS: inducible nitric oxide synthase; JAK/STAT: janus kinase-signal transducer and activator of transcription; MMPs: matrix metalloproteinases; MSCs: mesenchymal stem cells; mTOR: mammalian target of rapamycin; NFκB: nuclear factor-κB; NK: natural killer; NO: nitric oxide; PD-1: programmed cell death protein 1; PDGF: platelet-derived growth factor; PD-L1: programmed cell death ligand 1; PGE2: prostaglandin E2; PI3K: phosphoinositide 3-kinase; RANTES: regulated on activation, normal T cell expressed and secreted; SDF-1: stromal cell-derived factor 1; Smad: small mothers against decapentaplegic; Src: proto-oncogene tyrosine-protein kinase; STAT: signal transducer and activator of transcription; TAM: tumor-associated macrophage; TAN: tumor-associated neutrophil; TGF-β: transforming growth factor-β; TNF-α: tumor necrosis factor-α; TNFR: tumor necrosis factor receptor; TRAIL: tumor-necrosis-factor-related apoptosis-inducing ligand; VEGF: vascular endothelial growth factor. (Figure designed by BioRender).
Figure 4. Interactions and regulations involving mesenchymal stem cells in the tumor microenvironment. AKT: protein kinase B; α-SMA: α-smooth muscle actin; C3a: complement component 3 anaphylatoxins; CAF: cancer-associated fibroblast; CCL: CC chemokine ligand; COL: collagen; Cox-2: cyclooxygenase-2; CSC: cancer stem cell; CXCL: C-X-C motif ligand; CXCR: C-X-C motif chemokine receptor; DDR2: discoidin domain receptor tyrosine kinase 2; ERK: extracellular signal-regulated kinase; FASL: fas ligand; FSP1: fibroblast-specific protein 1; HGF: hepatocyte growth factor; HIF-1: hypoxia-inducible factor-1; IDO: indoleamine-2,3-dioxygenase; IFN-γ: interferon-γ; IGF: insulin growth factor; IL: interleukin; iNOS: inducible nitric oxide synthase; JAK/STAT: janus kinase-signal transducer and activator of transcription; MMPs: matrix metalloproteinases; MSCs: mesenchymal stem cells; mTOR: mammalian target of rapamycin; NFκB: nuclear factor-κB; NK: natural killer; NO: nitric oxide; PD-1: programmed cell death protein 1; PDGF: platelet-derived growth factor; PD-L1: programmed cell death ligand 1; PGE2: prostaglandin E2; PI3K: phosphoinositide 3-kinase; RANTES: regulated on activation, normal T cell expressed and secreted; SDF-1: stromal cell-derived factor 1; Smad: small mothers against decapentaplegic; Src: proto-oncogene tyrosine-protein kinase; STAT: signal transducer and activator of transcription; TAM: tumor-associated macrophage; TAN: tumor-associated neutrophil; TGF-β: transforming growth factor-β; TNF-α: tumor necrosis factor-α; TNFR: tumor necrosis factor receptor; TRAIL: tumor-necrosis-factor-related apoptosis-inducing ligand; VEGF: vascular endothelial growth factor. (Figure designed by BioRender).
Ijms 24 13511 g004
Figure 5. Tumor support functions of mesenchymal stem cells. Adapted from Frisbie et al. [143]. CAAs: cancer-associated adipocytes; CAFs: cancer-associated fibroblasts; CAMs: cancer-associated myofibroblasts; CSCs: cancer stem cells; EMT: epithelial–mesenchymal transition; IL: interleukin; MSCs: mesenchymal stem cells; NK cells: natural killer cells; TME: tumor microenvironment; TNF-α: tumor necrosis factor-α. (Figure designed by BioRender).
Figure 5. Tumor support functions of mesenchymal stem cells. Adapted from Frisbie et al. [143]. CAAs: cancer-associated adipocytes; CAFs: cancer-associated fibroblasts; CAMs: cancer-associated myofibroblasts; CSCs: cancer stem cells; EMT: epithelial–mesenchymal transition; IL: interleukin; MSCs: mesenchymal stem cells; NK cells: natural killer cells; TME: tumor microenvironment; TNF-α: tumor necrosis factor-α. (Figure designed by BioRender).
Ijms 24 13511 g005
Figure 6. Mesenchymal stem cells as vectors for therapeutic agents. Ag: silver; Au: gold; CX3CL1: C-X3-C motif chemokine ligand 1; EMT: epithelial–mesenchymal transition; Fe: iron; Hf: hafnium; HSV-tk: herpes simplex virus thymidine kinase; IFN: interferon; IL: interleukin; MSC: mesenchymal stem cell; NK cell: natural killer cell; TNF-α: tumor necrosis factor-α; TRAIL: TNF-related apoptosis-inducing ligand; TSP-1: thrombospondin-1. (Figure designed by BioRender).
Figure 6. Mesenchymal stem cells as vectors for therapeutic agents. Ag: silver; Au: gold; CX3CL1: C-X3-C motif chemokine ligand 1; EMT: epithelial–mesenchymal transition; Fe: iron; Hf: hafnium; HSV-tk: herpes simplex virus thymidine kinase; IFN: interferon; IL: interleukin; MSC: mesenchymal stem cell; NK cell: natural killer cell; TNF-α: tumor necrosis factor-α; TRAIL: TNF-related apoptosis-inducing ligand; TSP-1: thrombospondin-1. (Figure designed by BioRender).
Ijms 24 13511 g006
Table 1. Canonical surface markers used to identify mesenchymal stem/stromal cells in vitro.
Table 1. Canonical surface markers used to identify mesenchymal stem/stromal cells in vitro.
Expressed
Surface Markers
Negative Selection
Surface Markers
Characteristics
CD105CD11bPlastic adherence
CD90/THY-1CD14Multi-lineage differentiation
CD73CD19Plasticity
CD44CD34
CD166CD45
CD29CD79a
STRO-1HLA-DR
CD146/MUCIN
CD248/ENDOSIALIN
CD140β/PDGFRβ
CD: cluster of differentiation; HLA-DR: human leukocyte antigen-DR isotype; PDGFRβ: platelet-derived growth factor receptor β; STRO-1: stromal cell surface marker-1.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Slama, Y.; Ah-Pine, F.; Khettab, M.; Arcambal, A.; Begue, M.; Dutheil, F.; Gasque, P. The Dual Role of Mesenchymal Stem Cells in Cancer Pathophysiology: Pro-Tumorigenic Effects versus Therapeutic Potential. Int. J. Mol. Sci. 2023, 24, 13511. https://doi.org/10.3390/ijms241713511

AMA Style

Slama Y, Ah-Pine F, Khettab M, Arcambal A, Begue M, Dutheil F, Gasque P. The Dual Role of Mesenchymal Stem Cells in Cancer Pathophysiology: Pro-Tumorigenic Effects versus Therapeutic Potential. International Journal of Molecular Sciences. 2023; 24(17):13511. https://doi.org/10.3390/ijms241713511

Chicago/Turabian Style

Slama, Youssef, Franck Ah-Pine, Mohamed Khettab, Angelique Arcambal, Mickael Begue, Fabien Dutheil, and Philippe Gasque. 2023. "The Dual Role of Mesenchymal Stem Cells in Cancer Pathophysiology: Pro-Tumorigenic Effects versus Therapeutic Potential" International Journal of Molecular Sciences 24, no. 17: 13511. https://doi.org/10.3390/ijms241713511

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