ArticlePDF AvailableLiterature Review

Recent advances in regenerative medicine strategies for cancer treatment

Authors:

Abstract and Figures

Cancer stands as one of the most leading causes of death worldwide, while one of the most significant challenges in treating it is revealing novel alternatives to predict, diagnose, and eradicate tumor cell growth. Although various methods, such as surgery, chemotherapy, and radiation therapy, are used today to treat cancer, its mortality rate is still high due to the numerous shortcomings of each approach. Regenerative medicine field, including tissue engineering, cell therapy, gene therapy, participate in cancer treatment and development of cancer models to improve the understanding of cancer biology. The final intention is to convey fundamental and laboratory research to effective clinical treatments, from the bench to the bedside. Proper interpretation of research attempts helps to lessen the burden of treatment and illness for patients. The purpose of this review is to investigate the role of regenerative medicine in accelerating and improving cancer treatment. This study examines the capabilities of regenerative medicine in providing novel cancer treatments and the effectiveness of these treatments to clarify this path as much as possible and promote advanced future research in this field.
Content may be subject to copyright.
Biomedicine & Pharmacotherapy 141 (2021) 111875
0753-3322/© 2021 Published by Elsevier Masson SAS. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
Review
Recent advances in regenerative medicine strategies for cancer treatment
Vahid Mansouri
a
,
f
, Nima Beheshtizadeh
b
,
c
,
f
,
*
, Maliheh Gharibshahian
d
,
f
, Leila Sabouri
f
,
Mohammad Varzandeh
e
,
f
, Nima Rezaei
g
,
h
,
i
,
**
a
Gene Therapy Research Center, Digestive Diseases Research Institute, Shariati Hospital, Tehran University of Medical Sciences, Tehran, Iran
b
Department of Tissue Engineering and Applied Cell Sciences, School of Advanced Technologies in Medicine, Tehran University of Medical Sciences, Iran
c
School of Engineering, Faculty of Science and Engineering, Macquarie University, Sydney, NSW 2109, Australia
d
Department of Tissue Engineering, School of Medicine, Shahroud University of Medical Sciences, Shahroud, Iran
e
Department of Materials Engineering, Isfahan University of Technology, Isfahan 84156-83111, Iran
f
Regenerative Medicine group (REMED), Universal Scientic Education and Research Network (USERN), Tehran, Iran
g
Research Center for Immunodeciencies, Childrens Medical Center, Tehran University of Medical Sciences, Tehran, Iran
h
Department of Immunology, School of Medicine, Tehran University of Medical Sciences, Tehran, Iran
i
Network of Immunity in Infection, Malignancy and Autoimmunity (NIIMA), Universal Scientic Education and Research Network (USERN), Tehran, Iran
ARTICLE INFO
Keywords:
Cancer treatment
Regenerative medicine
Tissue engineering
Cell therapy
Gene therapy
ABSTRACT
Cancer stands as one of the most leading causes of death worldwide, while one of the most signicant challenges
in treating it is revealing novel alternatives to predict, diagnose, and eradicate tumor cell growth. Although
various methods, such as surgery, chemotherapy, and radiation therapy, are used today to treat cancer, its
mortality rate is still high due to the numerous shortcomings of each approach. Regenerative medicine eld,
including tissue engineering, cell therapy, gene therapy, participate in cancer treatment and development of
cancer models to improve the understanding of cancer biology. The nal intention is to convey fundamental and
laboratory research to effective clinical treatments, from the bench to the bedside. Proper interpretation of
research attempts helps to lessen the burden of treatment and illness for patients. The purpose of this review is to
investigate the role of regenerative medicine in accelerating and improving cancer treatment. This study ex-
amines the capabilities of regenerative medicine in providing novel cancer treatments and the effectiveness of
these treatments to clarify this path as much as possible and promote advanced future research in this eld.
1. Introduction
Regenerative medicine and its sub-disciplines, as emerging elds, try
to address various medical challenges. Regenerative medicine is an
approach to restore normal function and health to the organs and pa-
tients, commonly directed towards utilizing various stem cells. In recent
years, several regenerative approaches were used to ght against can-
cers. Although considering the broad range of other treatments that
could be counted as a subset of regenerative medicine, we encounter
many innovative, FDA-approved products for previously-known incur-
able disorders along with a mass of promising therapeutics in the eld of
cell and gene therapy [1].
Abbreviations: AAVs, Adeno-associated vectors; APC, Antigen-presenting cells; ADA, Adenosine deaminase; CAR T Cell, Chimeric antigen receptor T cells; CAFs,
Cancer-associated broblasts; CSCs, Cancer stem cells; CRISPR, Clustered regularly interspaced short palindromic repeats; CNT, Carbon nanotube; DMD, Deschene
muscular dystrophy; DCs, Dendritic cells; ESCs, Embryonic stem cells; ECM, Extracellular matrix; HA, Hyaluronic acid; HSCs, Hematopoietic stem cells; GVT, Graft-
versus-tumor; ICM, Inner cell mass; iPSCs, Induced pluripotent stem cells; LLC, Lewis lung carcinoma; MMP, Matrix-metalloproteinase; MSCs, Mesenchymal stem
cells; MPLA, Monophosphoryl lipid A; MSR, Mesoporous silica microrods; NK cells, Natural killer cells; NSCs, Neural stem cells; NPs, Nanoparticles; PAD, peripheral
arterial disease; PTT, Photothermal therapy; PDT, Photodynamic therapy; P I:C, Polyinosinic: polycytidylic acid; PAA, Polyacrylamide; PEG, Polyethylene glycol;
PLG, Poly(D,L-lactide-co-glycolide); PCL, Polycaprolactone; PLGA, Poly (lactic-co-glycolic acid); SCID, Severe combined immunodeciency; SMA, Spinal muscular
atrophy; SMN1, Survival motor neuron 1; scFv, Single-chain variable fragment; TME, Tumor microenvironment; TCRs, T-cell receptors; 3D, Three-dimensional; 2D,
Two-dimensional; VEGF, Vascular endothelial growth factor.
* Corresponding author at: Department of Tissue Engineering and Applied Cell Sciences, School of Advanced Technologies in Medicine, Tehran University of
Medical Sciences, Iran.
** Correspondence to: Childrens Medical Center Hospital, Dr. Qarib St, Keshavarz Blvd, Tehran 14194, Iran.
E-mail addresses: N-Beheshtizadeh@Razi.tums.ac.ir, nima.beheshtizadeh@hdr.mq.edu.au (N. Beheshtizadeh), Rezaei_Nima@tums.ac.ir (N. Rezaei).
Contents lists available at ScienceDirect
Biomedicine & Pharmacotherapy
journal homepage: www.elsevier.com/locate/biopha
https://doi.org/10.1016/j.biopha.2021.111875
Received 8 May 2021; Received in revised form 23 June 2021; Accepted 28 June 2021
Biomedicine & Pharmacotherapy 141 (2021) 111875
2
One of these challenges is to provide effective treatments for cancers
with fewer side effects [2]. Increasing our knowledge of the cancer tissue
structure and function makes regenerative medicine scientists more
effective and purposeful in providing treatment alternatives. Technol-
ogies related to cell engineering and cell therapy, as well as scaffolds and
biomaterials, are two major categories that underlie many approaches
[35]. Depending on the issue, they can be used separately or in com-
bination to solve the challenge [6,7].
Cancer is one of the most leading causes of death worldwide, and one
of the biggest challenges in treating it is developing better ways to
predict, diagnose, and eradicate tumor cell growth [810]. Although
various methods, such as surgery, chemotherapy, and radiation therapy,
are used today to treat cancer, this disease mortality rate is still high due
to the numerous shortcomings of each of these methods [8].
Surgical removal of the tumor is one of the most common methods
used to treat cancer for patients with early stages and small tumors. In
this method, due to vascular invasion, tumor spread and metastasis, and
new tumor formation after tumor resection, subsequent recurrences
occur, which has made this method efcient for only a small number of
patients [1114].
This method usually requires secondary radiation therapy, which
itself is associated with many problems. Since high doses of radiation
therapy are used to treat cancer, the surrounding healthy cells are also
damaged. Side effects of radiation therapy depend on radiation location
and sometimes continue for several weeks or months after treatment.
Common side effects include fatigue, hair loss, and the possibility of
secondary cancer, dryness, itching, blisters, and other skin problems
[1518].
Chemotherapy is useful for patients with advanced and metastatic
stages. Most chemotherapy drugs, including Sorafenib, Lenvatinib, and
Regorafenib, have anti-inammatory and anti-proliferative properties
[19]. They increase a patients life expectancy and delayed cancer
progression. Since cancer cells grow so fast, the action mechanism of
these drugs is to affect fast-growing cells. The problem with this function
is the effect of chemotherapy drugs on healthy fast-growing cells, such as
hair follicles, oral cells, and cells of the reproductive system. Side effects
of these drugs, such as diarrhea, anorexia, high blood pressure, weight
loss, skin reactions, and hair loss, limit their use [2023].
Accumulation of cells genetic mutations causes them to overgrow
and form tumors [24]. Tumor cells come into close contact with the
stroma, while genetic alteration in these cells causes the stroma to
change, leading to the development of an environment called TME
(tumor microenvironment), which helps cancer survive and progress
[25].
Interactions among tumor cells and stromal cells in various forms,
including cell-cell communication or paracrine and endocrine connec-
tions, cause structural and functional changes that lead to tumor sur-
vival and growth [25,26]. During their continuous expansion and
proliferation, tumor cells reach a stage where they lack oxygen and need
to form a vascular network. Studies show that hypoxia has no role,
although genetic changes caused by tumor cells cause overexpression of
angiogenesis. Nevertheless, angiogenesis occurs early in the develop-
ment of tumors [26].
Tumor cells secrete FGF, which calls broblasts to the site, and in
turn, metaplasia and become tumor and endothelial cells. Tumor-
associated broblasts modify tumor ECM by producing a specic
collagen content [26,27]. Also, tumor cells secrete chemokines, leading
to the accumulation of monocytes, neutrophils, and regulatory lym-
phocytes. This cell inltration in tumors occurs from the early stages of
tumor growth and is inuenced by tumor-secreting factors such as
TGF-B, IL 4, 10, other exosomes, and cytokines. Taken together, this cell
complex with the existing ECM, which is the result of their interactions,
creates a tumor prole, or TME, in-situ [26]. Studies showed that there
are at least six types of TMEs in tumors [24,28]. TME also varies at
various tumor sites, tumor growth stages, and even between different
patients [24].
Attempts to decipher the complexities of TME have led to the
development of new approaches to estimating the specic expression
proles of that type of tumor-associated cells, immune cells, and tumor
stroma. Targeting non-cancerous TME cells without considering the
cancer cells causes the TMEs to rell after treatment [24,29].
The only exception reported is immunotherapy, which includes
checkpoint therapies and has produced dramatic therapeutic effects.
There are several essential points in this treatment method. The rst is a
more understanding of the basis for activating and inhibiting immune
and stromal cells, although we have reached an acceptable under-
standing of T cells. The next item is to use biomarkers-based therapy [24,
30]. Given the high heterogeneity of TME, the development of reliable
biomarkers to guide targeted TME therapies will be essential to achieve
clinical efcacy. Finally, the use of hybrid approaches is recommended.
Despite the signicant and lasting effect of immunotherapy, most
cancer patients do not currently benet from it. It is recognized that
combination therapy is essential for improving the immune response
[24]. Therefore, it is essential to further identify TME as a tool for
developing new approaches to cancer treatment [26].
Among these, one of the most comprehensive approaches to target
therapy is the use of various regenerative medicine approaches,
including targeted drug delivery, anti-metastatic and restorative scaf-
folds, and preclinical modeling for the development of new drugs can be
considered with the concepts of TMEs.
Using different approaches, including gene therapy, cell therapy, and
tissue engineering, regenerative medicine can participate in the treat-
ment of cancer and understanding of cancer biology (Fig. 1). In addition
to the traditional application of tissue engineering scaffolds in encour-
aging the restoration of tissue structure and function of lost tissue due to
the architecture and materials used in them, they play a signicant role
in the development of cancer tissue in our body or as an immune system
adjuvant factor and be effective in modulating the immune system
[3134].
Generally, research in the regenerative medicine applications in
cancer treatment refers to the multiple major elds, including stem cell
transplantation to recover immune system, production of cancer-specic
T cell from induced pluripotent stem cells (iPSCs), stem cells as vector
carrying therapeutic to tumors, cancer stem cells and vaccines for can-
cers, and cancer microenvironment and tissue engineering [35].
The majority of attempts in regenerative therapy, such as gene
therapy and cell engineering, cell therapy, and tissue engineering scaf-
folds, have reached from the bench to the bedside. Although regenera-
tive medicine suffers from major challenges in cancer treatment, several
clinical trials resulted in the commercialization of approved products
available in the market. Despite all the successes and failures in utilizing
available products, future perspectives, including the vaccination and
Fig. 1. Various approaches in regenerative medicine participate in the treat-
ment of cancer.
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
3
research in microenvironment modication, were widened. Finally, the
goal is to bring fundamental and laboratory research to effective clinical
treatments. Proper interpretation of these researches will help to reduce
the burden of treatment and illness for patients. The purpose of this
review is to investigate the role of regenerative medicine in accelerating
and improving cancer treatment. This study examines the capabilities of
regenerative medicine in providing new cancer treatments and the
effectiveness of these treatments to clarify this path as much as possible
and help advance future research in this eld.
2. Novel advanced therapeutic strategies, based on regenerative
medicine approaches
2.1. Gene therapy and cell engineering
Gene therapy corresponds to transferring genetic materials into cells
to induce new features or manipulate the characteristics of the cells
[36]. Gene therapy is considered a novel therapeutic method compared
to other regenerative medicine facilities [37]. However, the number of
gene therapy applications is emerging since the rst gene therapy was
done in 1990 [38]. The primary elds that gene therapy has entered are
cancers and genetic disorders, contributing to more than 90% of the
clinical trials in this eld [39].
Generally, therapeutic gene therapies include transferring genetic
materials into cells to reverse an abnormal condition or induce a new
feature. This approach can use different strategies according to the ge-
netic problem, such as addition, edition (repair), and deletion/knockout
(inactivation) [40]. Sometimes gene therapy is used to add a normal and
functional copy of an allele to make the gene expression higher. For
example, adding a human clotting factor IX allele produces a higher
amount of blood coagulation factor XI in hemophilia type B [41].
Sometimes gene therapy is used to add a modied allele into the cells to
show new features, such as the widely used chimeric antigen receptor
(CAR) T cells [42].
Also, gene therapy is used to repair/edit a mutation or defective
allele, such as editing the survival motor neuron 1 (SMN1) gene in spinal
muscular atrophy [43]. Moreover, gene therapy is used for inactivating
abnormal or defective genes. For instance, using siRNA (anti-sense oli-
gonucleotides) to degrade TTR mRNA and decrease TTR protein pro-
duction in the treatment of hereditary transthyretin-mediated
amyloidosis [44].
Noteworthy the fact that incorporating and employing two actions of
the mentioned mechanisms is possible. The examples for these ap-
proaches include silencing the mutated allele and simultaneously
introduce a normal copy of the allele. This approach was used for dis-
eases in which an abnormal product interferes with the normal function
of the patient, such as thalassemia [45] and sickle cell disease [46]. Also,
clustered regularly interspaced short palindromic repeats (CRISPR)
could act as a new valuable tool regarding these mechanisms [47].
Vectors are one of the essential factors in transferring genetic ma-
terials. Commonly there are two types of vectors, including viral and no-
viral vectors [48]. Various vectors with their specic characteristics
have been using for various applications. The most common vectors are
viral vectors, which can effectively transfer their gene into their host
cells. Genes responsible for replicating, assembling, or infecting the
virus are substituted with therapeutic genes to make them
replication-decient, specic, and safe therapeutic viral vectors [49].
Retroviruses and Adeno-(associated)-viruses are the most common
viral vectors, allowing for stable viral genome integration [49].
Although retroviral vectors only infect dividing cells, lentiviral vectors,
as a complex type of retroviral vectors, can also infect non-proliferating
cells, result in their extensive usage [50]. Adenovirus vectors have a
high transgene expression and the ability to infect various quiescent to
proliferating cells. Therefore, they have a high reputation for use as
oncolytic viruses and viral vaccines.
Among viral vectors, Adeno-associated vectors (AAVs) are one of the
best candidates and widely-used in-vivo applications. AAVs can infect
both dividing and non-dividing cells and contribute to the life-long
expression of the transgene. Besides, it contains no virulence genes
preventing an undesirable immune reaction. Due to their low integra-
tion frequency, their risk of insertional mutagenesis is low [40,51].
Non-viral vectors are the most straightforward route of delivering
genetic material, are less popular, mainly due to their off-targets and risk
of degradation. The traditional form of non-viral vectors are Plasmids,
DNA, and RNAs (siRNA and miRNA) [52]. The rst two are used for the
gain of function to replace or to substitute a specic genetic function in a
target cell, and the last one is used for loss of function for reducing the
expression of endogenous genes in a sequence-specic manner purpose
[52].
Plasmids have currently entered the market as a vector in the form of
Neovasculgen, which consists of a plasmid DNA encoding Vascular
endothelial growth factor (VEGF) under the control of a CMV promoter
for treatment of atherosclerotic peripheral arterial disease (PAD) [53].
Two major methods for transferring genetic materials into the cells
include ex-vivo and in-vivo gene therapy. The rst one, ex-vivo gene
therapy, involves laboratory-based transduction of cells. After harvest-
ing specic cells either from the patient (autologous) or a donor (allo-
geneic), a vector is used to carry the therapeutic gene into the cells. The
transduced cells are then returned to the patient after possible activation
or expansion [54]. In-vivo gene therapy involves vector delivery into a
patient without any cellular vehicle so that the gene transfer process
takes place inside the body [55]. By ex-vivo gene transferring, we could
specically target the obtained appropriate cells; also, the ability to
select and rene the cells of interest before transplantation enables us to
incorporate weaker vectors and ensure the cells desired characteristics
[56].
However, ex-vivo genetic manipulation of cells requires a high-tech
laboratory, strong-background researches, and advanced equipment.
Unlike ex-vivo gene therapy, the in-vivo method needs much fewer
laboratory works since the vector is directly administered into the body,
and the rest is done inside [57]. After the entrance of the vector into the
body, based on the injection route and vector tropism and specicity (e.
g., specied tropism in some AAV vectors and usage of target
tissue-specic promoters), the vector would infect the desired cells and
transformed it using one of the aforementioned strategies. The direct
administration of the products cuts the necessary in-vitro process for
manufacturing the personalized products and enables the product to be
prepared off-the-shelf, therefore fasten up the treatment process.
However, this method is limited to accessible tissues. In-vivo gene
manipulation is the primary method of gene therapy, when cells cannot
be isolated from the body or are unable to live a long time in ex-vivo.
Also, in-vivo gene therapies often target long-lived non-mitotic cells
(e.g., retinal cells), in which non-integrative (episomal) gene transfer
could be safely expressed for years [58]. However, non-specic cell
targeting and a higher risk for immune reactions have limited the use of
this method [54]. The most common vector used for in-vivo gene ther-
apies is AAVs, which their minimal immunogenicity along with effective
gene transfer, especially in non-dividing cells, makes it the pioneer of
current gene therapy vectors. Several current gene therapy products
used the in-vivo method for gene delivery, including Spinraza, Zolge-
nesma (for SMA) and Luxturna (for retinitis pigmentosa) [58]. It should
be noted that clinical trials from 2010 to 2020 were rather equally used
the in-vivo and ex-vivo methods; however, it seems that in-vivo gene
therapies would excel in the future, even in gene therapy of cancers
[58].
To this date, There are several gene therapy products indicated for
cancer treatment, including Gendicine® (2003, indicated for head and
neck cancer), Oncorine® (2005, indicated for nasopharyngeal carci-
noma), Rexin-G® (2007, indicated for Soft tissue sarcoma and osteo-
sarcoma), Imlygic® (2015, indicated for melanoma), and ve CAR T
cells, which are discussed in the next sections, naming Kymriah® (2017,
indicated for relapsed B cell acute lymphoblastic leukemia), Yescarta®
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
4
(2017, indicated for relapsed or refractory large B cell lymphoma),
Tecartus® (2020, indicated for relapsed/refractory mantle cell lym-
phoma), ABECMA® (2021, indicated for multiple myeloma) and
BREYANZI® (2021, indicated for relapsed or refractory (R/R) large B-
cell lymphoma).
Gendicine®, the rst gene therapy product approved for clinical use
[59], was shown acceptable safety and efcacy in treating head and
neck cancer, primarily when used in combination treatments with
radiotherapy. Approximately 93% of patients with head and neck
squamous cell carcinoma had clinical response compared to 79% when
radiotherapy was used alone [60]. Oncorine®, the rst oncolytic virus
product approved for clinical use, was designed against solid cancers
and used safely for glioma, head and neck, pancreatic, and ovarian
cancers; however, weak efcacy causes its limited use nowadays [59].
Rexin-G®, a pathotropic (disease-seeking) tumor matrix-targeted
retrovector, was rst evaluated for pancreatic cancer [61]. However,
its safety and efcacy were approved in several studies for osteosar-
coma, soft tissue sarcoma, and pancreatic cancer [62,63]. Imlygic®, an
HSV-1 oncolytic virus armed with GM-CSF cassette, showed promising
therapeutic effects on melanoma and also other cancers, either alone or
in combination with pembrolizumab or ipilimumab [64]. About 16.3%
of patients with melanoma receiving Imlygic® got a durable response in
comparison with the control groups just receiving GM-CSF [65]. During
recent years, gene therapy of cancers continued in the form of a new
generation of genetically modied effector cells, called CAR-T cells,
which growing numbers of them were being approved or expanding
their previous indications.
2.2. Cell therapy
Cell therapy corresponds to various cellsusage to reverse or repair
the diseased state of tissues or organs. Cell therapy can be administered
either locally or systemic and is almost always minimally invasive [66].
Generally, cell therapy started with the rst blood infusion early in the
1940s [67]. Later in the 1950s, it followed with bone marrow trans-
plantation, enabling humankind to utilize different cellscharacteristics
and functions to identify and treat diseases [68].
Cell therapies include a wide range of therapeutics, from cell-based
vaccines to engineered cells, which in recent years showed promising
results, especially in treating hematological cancers [69,70]. Among
different types of cellular therapies, stem cells, chimeric antigen re-
ceptor (CAR)-T cells, and natural killer (NK) cells get the most attention
in cancer treatment (Fig. 2) [71,72].
2.2.1. Stem cells
Stem cells are non-specialized cells with the ability of self-renewal,
unlimited proliferation, and differentiation into various cell types.
They can migrate to the site of cancer cells and play a signicant role in
suppressing and modulating the immune system by secreting multiple
factors, such as growth factors, chemokines, and cytokines [73,74]. By
targeting and destroying primary and metastatic cancer cells, stem cells
reduce tumor volume, increase patient survival, and prevent cancer
recurrence.
The use of stem cells in cancer treatment is one of the newest and
most non-invasive methods for cancer therapy that uses different ap-
proaches for this purpose. Some of the main application of stem cells in
cancer therapy include regeneration of damaged tissues after traditional
treatments, immunotherapy, inducing immune cells, targeting and
destruction of cancer stem cells (CSCs) by stem cell-based-anti-cancer
vaccines, anti-cancer drug screening, therapeutic carriers, and targeted
delivery of oncolytic viruses and nanoparticles [7578]. Depending on
the stem cellstypes, transplantation method, transplantation time, and
the number of transplanted cells, the efciency of stem cell therapy and
its effect on tumor suppression or tumorigenesis is determined.
The type of stem cells determines the therapeutic potential and how
they function. In general, stem cells have two primary categories of
embryonic (ESCs) and adult stem cells [77]. Neural stem cells (NSCs),
mesenchymal stem cells (MSCs), and hematopoietic stem cells (HSCs)
are some of the adult stem cells that have been studied to treat cancer
[7981].
ESCs are pluripotent cells that are isolated from the inner cell mass
(ICM) of the embryo. These cells can differentiate into all cells except
placental cells. However, the use of ESCs has some ethical obstacles due
to their primary source of isolation and embryo destruction. Hence,
induced pluripotent stem cells (iPSCs) were introduced [82].
Somatic cells can be reprogrammed through the transferring of Oct
3/4, Sox2, Klf4, and c-Myc, and produce iPSCs [83]. Based on previous
studies, there is no ethical concern about iPSCs, and the lack of immu-
nogenicity in these cells has led to their higher efciency [84,85]. They
can induce a variety of immune cells and anti-tumor-vaccines to target
tumor cells. IPSCs can be used to repair and replace damaged cells with
chemotherapy, surgery, and radiation therapy. These cells are also
useful in evaluating anti-cancer drugs for end-stage patients. Simulta-
neously, iPSCs-based vaccines effectively prevent tumor relapse and
increase immune cellsproliferation and their secretions [86,87]. These
cells can be efcient in immunotherapy by inducing T and NK cells. The
possibility of teratoma formation and autoimmunity are some major
issues using them [8789].
NSCs are multipotent stem cells with the expression of classical
markers, such as Sox2 and Nestin. NSCs have the ability to self-renewal
and differentiate into astrocytes, oligodendrocytes, and neurons [90,
91]. They can migrate to tumor sites, and their homing could effectively
deliver anti-tumor therapeutic agents, proteins, enzymes, prodrugs, and
nanoparticles to tumor cells. The migration of these cells to the tumor
site is caused by hypoxia [81,9294]. These cells can cross the
blood-brain barrier and are particularly useful in the treatment of brain
cancer. Besides, studies show that NSCs are useful for the treatment of
breast [92], ovarian [95], prostate [93], pancreatic [74], and lung [94]
tumors. Isolation and purication of these cells are generally arduous.
Moreover, MSCs are multipotent cells that, like other stem cells,
possess self-renewal capacity and differentiate into different cell types.
They are derived from bone marrow, adipose tissue, placenta, umbilical
cord, peripheral blood, and muscle [96]. These cells have valuable
properties, such as tropism, anti-tumor effects, anti-angiogenic effects,
reduction of tumor cell proliferation, migration, and homing ability at
the tumor site. The CXCL12 / CXCR4 signaling pathway is one of the
signicant ways, which affects the movement of MSCs cells toward the
tumor.
These cellssecretions can activate the patients immune system and
regulate the tumor microenvironment (TME) [97]. MSCs can also
release drugs at the tumor site, deliver therapeutic agents, and regen-
erate cancerous tissues [98101]. Previous studies showed that MSCs
are effective in treating various tumors, such as the brain [99], mela-
noma [100], breast [102], and lung [101] tissues. However, many
Fig. 2. Various cells in cancer therapy, include stem cells, NK cells, and CAR-
T cells.
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
5
studies have suggested dual roles for MSCs in stimulating and sup-
pressing cancer [103,104]. Therefore, the function of MSCs depends on
their origin, the number of them, and the method of cell transplantation.
HSCs are multipotent cells that can form mature blood cells and are
mainly present in the bone marrow [105]. The anti-tumor effects of HSC
transplantation are very signicant in the treatment of hematologic
malignancies. HSCs are also valuable in immunotherapy by introducing
genes encoding CARs and T-cell receptors (TCRs) against the tumor
antigen. Transplantation of these cells in solid tumors can increase
survival by the graft-versus-tumor (GVT) since the new donor T lym-
phocytes react with tumor antigens and stimulate the immune response
against them. Transplantation of HSCs after chemotherapy and radiation
therapy is useful and helps facilitate hematological recovery. The
migration of these cells to the tumor depends on the interaction of
CXCL12 and its receptor CXCR4 [106108]. Properly adjusting the niche
factors and expressed factors on the cell surface (CXCR4-CXCL12)
binding helps to better homing and stabilizing these cells. CD26 express
in many solid tumors and cells, such as HSCs can cleave CXCL12 and
reduce its activity. By reducing the amount of CD26 via various inor-
ganic and biological inhibitors, CXCL12 binds to CXCR4 and improves
homing. To improve homing, these cells or their target tissue must be
manipulated to create higher levels of chemokines and their receptors
[109].
In addition to the above cells, CSCs are multipotent stem-like cells
possessing self-renewal, unlimited proliferation, differentiation into
other cancer cells, and resistance to chemotherapy drugscapabilities.
These cells are the foremost cause of metastasis, progression, and cancer
recurrence [110112]. There is always a small population of CSCs in the
heterogeneous environment of the tumor. Conventional cancer treat-
ment methods do not destroy these cells, and these cells can initiate
various tumors.
Targeting and killing CSCs is critical to successful cancer treatment.
Since stem cells are adsorbed to CSCs, stem cell therapy and cell-cell
interactions can be used to reduce tumor proliferation, metastasis, and
angiogenesis [111,113]. Besides, to remove these cells, involved signals
inhibition could be used in CSCs self-renewal, destruction of CSCs
microenvironment, and targeting CSCs surface markers [114,115]. CSCs
in different cancers have various surface markers such as CD44, IL-3R,
CD133, and CD13. For example, since CD13 is a marker of hepatic
CSCs, Haraguchi et al. [116] used anti-CD13 antibodies to suppress
dormant CSCs and prevent their tumorigenicity.
CSCs are located within specic microenvironments capable of
maintaining their phenotype, drug resistance, and facilitating their
metastasis [117]. Therefore, changes in this microenvironment, such as
reduced vascularization, changes in hypoxia, and inhibition of CSCs
growth in this microenvironment, can also prevent the growth of tumor
cells. Besides, various signaling pathways, such as Wnt and Notch1, are
involved in maintaining the basic properties of CSCs and their differ-
entiation and self-renewal, which can be controlled and targeted to help
eliminate CSCs and treat cancer [118].
Another parameter affecting the efciency of cell therapy is the
method used for cell transplantation. Stem cells are usually delivered to
the patient by controlled injection at the tumor site, intranasal, and
using semi-solid substrates and hydrogels. In choosing the trans-
plantation method, the degree of invasion, and the risk for patients, the
type of therapeutic agent, and the location and tumor type are vital
parameters. For instance, intracranial injection is an invasive and non-
repeatable procedure for brain tumors, while encapsulating cells in
different substrates helps to accurately deliver therapeutic agents to the
desired location and increase the success and durability in the treatment
of brain cancer [119,120]. Other factors affecting the anti-tumor or
tumorigenesis function of stem cells are the number of transplanted cells
and the time of transplantation. The insufcient and the excess number
of transplanted stem cells will result in cancer recurrence and teratoma
formation, respectively [121].
Stem cell therapy is usually used in combination with other methods.
According to the general protocol chosen for treatment, the time of cell
entry into the tumor must be carefully estimated. For instance, NSCs that
reach the brain tumor before radiation and Temozolomide (XRT-TMZ)
treatments can increase survival relative to the reverse [121,122].
Besides, various modications can be done on stem cells to increase
their effectiveness in treating cancer. Genetically modied stem cells
can convert non-toxic prodrugs into toxic products by expressing spe-
cic enzymes after transplantation and migration to the tumor site,
causing tumor cellsdeath. This method is called Enzyme/prodrug
therapy or suicide gene therapy, which can increase the accuracy of
control over time, place, and amount of drug. Up to now, suicide gene
therapy has entered clinical phases I and II [123].
2.2.2. CAR-T cells
CAR-T cells are one of the prominent successful applications of gene
therapy in immunotherapy. Genetically engineered CAR-T cells were
produced by reconstructing chimeric receptors on T cells using a gene
sequence of monoclonal antibodies that could identify specic mole-
cules, particularly tumor-associated antigens [124]. Moreover, along
with the promotion of CAR-T cells across its generations, they occupy
some intracellular co-stimulatory domain responsible for fully acti-
vating T cells after encountering the specic antigen. The focal gath-
ering of activation and expansion potential of several co-stimulatory
domains in cytotoxic T cells with the specicity of monoclonal anti-
bodies make the CAR-T cells one of the most promising tools to ght
cancers [124]. Fig. 3 shows the cancer therapy procedure utilizing
CAR-T cells.
Car-T cellspotential anti-cancer effects lie under the incorporation
of two main segments together with a hinge. These two segments
include an extracellular antigen-targeting part, such as a single-chain
variable fragment (scFv), and an intracellular domain featuring helper
co-stimulatory domains triggered by antigen detection. The scFv mod-
ule, with a structure similar to a small part of antibody-containing the
variable segments of heavy and light chain fused with a linker, could
easily be generated against every antigen.
This module is responsible for the specic reactivity of CAR-T cells,
which is inuenced by the antigens binding afnity, expression level,
antigen density on the target cell, and epitope proximity [125]. Intra-
cellular domains of CAR are responsible for its potential efcacy and
durability. Incorporating one or more co-stimulatory domains could
enhance CAR-T cells inuence while reducing their exhaustion and
improving their persistence. These cells have shown promising results in
treating liquid malignancies, especially CD19-positive cells, contrib-
uting to the approval of two of their leading product-Kymriah and
Yescarta [126]. Recent studies demonstrate that CAR-T cells could also
be impressive in treating solid tumors. It is under discussion whether
their homing in solid tumors is reliable or not [127129].
2.2.3. NK cells
NK cells are active cells in the innate immune system and are
considered as a part of the lymphocyte family. These cells can respond to
many pathological conditions and are commonly known as controllers of
cancerous microenvironments [130]. For this reason, they have many
functional interactions with T cells from adaptive immunity [131]. In
contrast, natural killer cells do not need antigen-presenting cells for
identication and directly identify tumor cells [132].
In addition to the direct function, NK cells indirectly affect other
immune cells, such as macrophages and dendritic cells, with their
cytokine secretions, such as INF Ɣ and TNF
α
, enhancing the immune
response and further the success of immunotherapy [133]. Therefore, by
increasing the activity of NK Cells, a more robust immune response can
be induced in pathological conditions such as cancer [133].
Researchers are working on NK cell engineering and indirect
methods to strengthen the immune system, including the use of small
molecules, cytokines, and even genetically engineered stem cells
secreting specic cases of cytokines [134]. Stem cells can also
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
6
differentiate into NK cells in-vitro conditions and provide ready-use
stocks of such immune cells for research and therapeutic purposes
(Fig. 4a) [135]. However, NK cell response rate depends on interactions
with other immune system cells, such as T cells, dendritic cells, and
macrophages. Type 1 IFNƔ, IL-12, IL-18, and IL-15, are potent activators
of NK cell function [132]. IL-2 has also been shown to stimulate NK cells
proliferation, toxicity, and cytokine secretion [133].
For achieving more effective immunotherapy, genetically engineer-
ing NK cells and modifying them to identify tumor targets could be
considered novel and accurately potent approaches [136]. NK cells, like
T cells, can be engineered using CARs. The preparation and production
of chimeric antigen receptor-natural killer (CAR-NK) cells as
off-the-shelf products in immunotherapies can increase the access of NK
cells to cancerous microenvironments (Fig. 4b) [133]. NK cell engi-
neering can increase the cytotoxic ability of NK cells and detect tumor
antigens faster and greater. Although studies in this area are still in the
preclinical stages, CAR-NK cells can be considered as a new immuno-
therapeutic approach [136]. Overall, it can be concluded from studies
that a combination of methods should be used in order for immuno-
therapy and its engineering to obtain more effective results in treatment
[137,138].
2.3. Tissue engineering scaffolds
What is known as scaffolds in tissue engineering are the ordered
structures of various materials adapted to the bodys physiological
environment that have been constructed and processed according to the
desired task and application [4]. Various biomaterials have been
approved as scaffolds to regenerate the structure and function of the
specic tissues. The particular physical and chemical properties of these
substances allow their application in the body.
They support cell attachment and migration, differentiation,
increased angiogenesis in the case of need, crosstalk with the sur-
rounding tissues, and integration with them [4,5]. So far, the concept of
tissue engineering scaffolds in cancer therapy has meant using them in
the repair of lesions leftover from surgical treatments and tissue tumor
resection site, adjuvants, targeted drug delivery, immunotherapy, and
in-vitro three-dimensional (3D) modeling.
The most common approach to using scaffolds in cancer therapy, as
in other chronic and degenerative diseases, is to reconstruct the lost
tissues and organsstructure and function during the disease period and
improve the patients quality of life. It is necessary to mention that the
use of scaffolds in cancer treatment is a supportive approach.
Manufacturing and designing various types of bone scaffolds to repair
Fig. 3. Cancer therapy procedure via employing CAR-T cells.
Fig. 4. (a) The potential of iPSC-NK cells as a template for engineering. (b) The preparation and production procedure of off-the-shelf NK cells.
(a) Reproduced with permission from [134].
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
7
critical defects are examples of this approach [5,139].
The next approach is to use adjuvants, which have nally evolved
into modern vaccinations [140]. Adjuvants are compounds that stimu-
late the immune system to ght cancer [141]. In other words, they in-
crease the presentation possibility of tumor cells to the
antigen-presenting cells and prevent hiding tumor cells from the im-
mune system [33,141]. Researchers have used a wide range of bio-
materials as adjuvants in cancer immunotherapy, including various
mineral and organic materials, having specic characteristics, such as
size, shape, surface, and molecular weight, affecting immune response
(Fig. 5a) [32,142,143]. These materials are used in combination or
separately; thus, scaffolds made of these materials can stimulate the
immune system against cancer. The activity of immune system cells is a
pre-inammatory spectrum, which can be regulated by the inherent and
designed properties of implanted materials in the body. One of these
characteristics is the materials size, affecting drug trafcking rate,
uptaking, and exposure to immune stations in the body. Thus, they can
play an essential role in further activating the bodys immune system or
keeping it off. Besides, the shape of the implanted biomaterials and their
surface characteristics also affect the type and manner of immune re-
sponses, and the immune system signaling cascade can be easily
manipulated by deforming the biomaterials. For instance, biomaterials
with rougher levels have been shown to stimulate pro-inammatory and
inammatory mechanisms in the body (Fig. 5b and c).
On the other hand, biomaterials with a higher surface-to-volume
ratio can interact with more immune cells and move structurally later
in the body, stimulating the immune system more than others. Even the
release of by-products from the degradation of implanted biomaterials
and their chemical and physical properties can affect the intensity of
immune responses and how they are performed.
Scaffolds have developed another approach to cancer immuno-
therapy. This treatments basis is the specic function of the cancer
microenvironment in regulating the immune system and suppressing it
[6]. Studies have shown that this particular microenvironment has
factors that stimulate myeloid precursor cells to travel to tissues, such as
the lungs and liver, where they differentiate into myeloid-derived sup-
pressive cells [6,146,147]. Therefore, they prepare the body for metas-
tasis and secondary establishment of cells. Such tissues that provide the
conditions for tumor cell metastasis are called pre-metastatic tissue
(Fig. 6) [148,149].
Scaffolds have now been developed that trap tumor cells by simu-
lating such pre-metastatic sites and preventing cell proliferation and
tumor recurrence. These scaffolds are implanted near the primary tumor
site and, if equipped with image-detecting agents, will track tumor
function and initiate tumor recurrence by periodic imaging. Among the
scaffolds used to play this role are Poly (lactic-co-glycolic acid) (PLGA)
and Polycaprolactone (PCL) porous scaffolds, and scaffolds containing
collagen, brin, hyaluronic acid, and alginate [31,150152].
Using each of these approaches or combining them can be helpful
during the cancer therapy stages, along with other traditional treat-
ments such as surgery, radiation therapy, and chemotherapy [153,154].
However, optimizing the application of these methods in different
conditions and tissues requires simulation in laboratory models.
Scaffolds also are used to upgrade two-dimensional (2D) cultures and
are a branch of 3D culture modeling [7,155,156]. Laboratory models
increase our knowledge of cancers evolutionary pathway and lead to
more effective treatments. Scaffolds are also efcient in these areas, with
the bodys 3D environment simulation to prevent information from
being lost and getting research closer to reality. Biomaterials can also
play the carriersrole for drug, gene sequencing, and stem cell delivery
to cancer tissue and boost direct targeted therapy [3,32,157].
Fig. 5. (a) Intrinsic properties of materials inuence immune
responses. Biomaterials commonly used in the vaccine,
immunotherapy, and tissue engineering approaches present
features including size, shape, surface charge, hydrophobicity,
and molecular weight that modify interactions with the im-
mune system. Confronting components of the innate and
adaptive immune system with scaffolds and biomaterials re-
sults in the formation of the brotic capsule to isolate the
material, differential activation of dendritic cells and macro-
phages, recognition and removal by antibody and complement
proteins, and even manipulating the adaptive immune
response. (b) Particle shape dictates immune cell uptake and
activation. Spherical polymeric particles fabricated from
polystyrene polyethylene oxide that exhibit rough surfaces
(left) were preferentially taken up by macrophages and
induced a pro-inammatory response compared to smooth
particles (right) (Scale bar, 10 mm; inset scale bar, 5 mm). (c)
Electron micrographs of gold nano constructs with spherical
(left), cube (center), or rod-like (right) shapes. Incubating with
dendritic cells, rod-like particles induced inammatory IL-1b
and activated the inammasome, while sphere and cubes
caused secretion of TNF-a (Scale bar, 40 nm).
(a) Reproduced with permission from [32]. (b) Reprinted with
permission from [144]. (c) Reprinted with permission from
[145].
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
8
3. Regenerative medicine in cancer; from bench to bedside
3.1. Clinical trials
Today, gene therapy is used as an emerging treatment for cancers
and genetic disorders. Although gene therapy is considered a way to
treat genetic disorders, cancer gets rst place among the diseases treated
with gene therapy [39]. Cancers constituted a signicant part of the
trials each year from 2016 to 2020 [39]. Among the various cancers,
hematological cancers with broad-expressed tumor-associated antigen
with excellent cell-cell interactions were the most investigated [158].
Moreover, looking at current trials addressing cancer diseases, the ma-
jority of them are in phase I [39]. Although the number of phase II
clinical trials is also high, it seems that few of them reach phase III and
IV. Notably, to this date, there are ve phase-IV trials are focusing on
leukemia and lymphoma, solid tumors, especially prostate cancer [39].
Most genetic disorders, unlike cancers, are congenital, such as
Thalassemia, Deschene muscular dystrophy (DMD), and cystic brosis
[159]. In principle, this group of diseases is based on the transferring
genetic materialsconcept to treat diseases. It rst started in 1985 for
Adenosine deaminase (ADA) deciency resulting in severe combined
immunodeciency (SCID) [160]. Metabolic disorders and blood coag-
ulation disorders accounted for about half of the trials in this group.
Looking in more detail, Hemophilia (either type A or B), DMD, and
retinitis pigmentosa were the most common genetic disorders with the
highest number of clinical trials [161].
Furthermore, cell therapy clinical trials are running in cancer treat-
ment. The main focus of more than 600 ongoing clinical trials is on some
types of CAR T cells. However, CAR T cells targeting hematologic cancer
accounts for more than half of all clinical trials. After then, gastroin-
testinal cancer, nervous system cancer, and breast cancer have the
highest counts, which collectively account for more than three-quarters
of all CAR T cell trials. Most of the CAR T cells used in the trials were
directed against CD19, CD 20, CD30, CD22, EGFR, Mesothelin, and
BMCA [162].
3.2. Approved products in the market
There are currently two FDA-approved CAR T cell-based products for
CD19+B cell acute lymphoblastic leukemia, named Kymriah®
(Novartis, Swiss) and Yescarta® (Kite Pharma, US), which revolution-
ized the use of gene therapy in the eld of cancers [126]. However, the
number of approved products is higher for genetic disorders (ten for
genetic disorders versus six for cancer). To date, approved products for
the various diseases have entered the market, including ADA deciency,
DMD, spinal muscular atrophy (SMA), retinitis pigmentosa, B-Thalas-
semia, and several metabolic disorders, such as Familial lipoprotein
lipase deciency, Homozygous familial hypercholesterolemia, Heredi-
tary Transthyretin Amyloidosis, and Adult Familial Chylomicronemia
syndrome. Table 1 showed the approved gene therapy products along
with their indication and approval year.
Moreover, Table 2 shows the approved cellular products. First
approved cell therapy products dated back to 1998, the year of approval
of the rst gene therapy product [191]. However, unlike gene therapy
products, cellular products are focused on donor hematopoietic stem/-
progenitor cells. Only one product targeted against cancers: Provenge®
(Dendreon, US), whereas others mainly focused on the healing process
in soft tissues such as skin, cartilage, and wounds.
3.3. Current challenges in cancer treatment via regenerative medicine
approaches
Regenerative medicine as a promising emerging interdisciplinary
eld has faced many ups and downs since it was established in the late
90s [204]. At rst, the simplistic nature of replacing tissues with cells
cultured in a matrix raises many hopes for the regeneration and sus-
tainability of tissue and organs. Although tissue engineering worked
well in avascular and tissues with low metabolism, critical issues hamper
its success when applied to more complex tissues. Complex and 3D tis-
sues, unlike their ancestors, need accurate control of cell-cell in-
teractions as a signicant factor for their proper function. Various tissue
components like blood vessels, parenchymal, and non-parenchymal cells
need to be adjusted spatially at microscales. However, tissue
Fig. 6. Pre-metastatic microenvironments formation process.
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
9
engineering provides spectacular abilities for basic tissue engineers, but
in the eye of medical practitioners, tissue-engineered products are still
far away from being a reliable choice in clinics. Therefore, regenerative
medicine needs to utilize appropriate strategies to overcome its current
challenges in cancer treatment via regenerative medicine approaches
[205]. Fig. 7 demonstrates the principle challenges of regenerative
medicine in cancer treatment.
Regulatory issues
The main challenge that regenerative medicine faced nowadays is
regulatory issues. Although regenerative medicine approaches involved
a simple rationale for repairing and restoring body function, it involves a
wide variety of formulations, bio-manufacturing, biological
interactions, unstable nature of products, storage difculty and
personalized mode of utilization and administration. From this point of
view, until the recent last year, categorizing these products in the prior
existing classication cause signicant problems during regulatory
consideration. Innovative regenerative approaches could hardly be put
into any medical drugs or devices categories, requiring a unique regu-
latory review process and authorization.
Applying the current regulatory approach used for other pharma-
ceutical drugs/devices is incompatible with the novel, cutting-edge
nature of regenerative medicine products and prevent considering
comprehensive aspects of these products [206]. Using specic review
guidelines and assessment themes for these drugs in special reviewing
board committees could be one promising way for easing the appro-
priate assessment of investigated products [207].
Table 1
Approved gene therapy products.
No. Trade name (General name) Year of
Approval
Indication Manufacturer Price Refs.
1 Vitravene (Fomivirsen) 1998 cytomegalovirus retinitis in
immunocompromised patients
Isis Pharmaceuticals with
Novartis Ophthalmics
withdrawn [163]
2 Gendicine (Recombinant
adenoviral vector expressing
p53)
2003 Head and neck cancer Shenzhen
SiBionoGeneTech
$387 per injection [164]
3 Macugen (Pegaptanib) 2004 Wet age-related macular degeneration Eyetech Pharmaceuticals
and Pzer. Inc
$765 per dose [165]
4 Oncorine (recombinant human
adenovirus type 5 injection)
2005 Nasopharyngeal carcinoma Shanghai Sunway Biotech [68]
5 Rexin-G (Mx-dnG1) 2007 Soft tissue sarcoma and osteosarcoma Epeius Biotechnologies $5000 per bag 500,000 $ per
treatment
[166]
6 Neovasculgen
(Cambiogenplasmid)
2011 Peripheral vascular disease and limb
ischemia
Human Stem Cells
Institute
$6600 for treatment course [167]
7 Glybera 2012 Familial lipoprotein lipase deciency Amsterdam Molecular
Therapeutics
over $1.2m per patient [168]
8 Kynamro (Mipomersen) 2013 Homozygous familial
hypercholesterolemia
ISIS Pharmaceuticals $6910 For 1-ml vial [163]
9 Imlygic (Talimogene
Laherparevec)
2015 Melanoma Amgen $65000 per treatment [169,170]
10 Zalmoxis 2016 Restoring the immune system of the
patient after hematopoietic stem cell
transplantation
MolMed SPAA
149000 Dec 2017
163900 EUR
in Germany: ex-factory price
[171]
11 Strimvelis (Modied
Autologous CD34 +cells)
2016 Severe combined immunodeciency
(SCID) due to ADA deciency
GlaxoSmithKline (GSK) 594,000 euros, or $648000 [172174]
12 Spinraza (Nusinersen) 2016 Spinal Muscular Atrophy (SMA) Biogen Ionis
Pharmaceuticals, Inc.
$125000 per injection [175,176]
13 Exondys 51 (Eteplirsen) 2016 Duchenne Muscular Dystrophy (DMD) Sarepta Therapeutics $1678 For 2 ml vial [176,177]
14 Kymriah (Tisagenlecleucel) 2017 Relapsed B cell acute lymphoblastic
leukemia
Novartis Pharmaceuticals
Corporation
$475000 Feb 2018 [178]
15 Yescarta (Axicabtagene
Ciloleucel)
2017 Relapsed or Refractory large B cell
lymphoma
Kite Pharma,
Incorporated
$373000 April 2018 [179]
16 Luxturna (Voretigene
Neparvovec-rzyl)
2017 RPE65 mutation-associated retinal
dystrophy
Novartis Inc. $850000 or $425000 per eye [180]
17 Invossa (chondrocytes
transduced with TGF-ß1)
2017 Moderate Knee Arthritis TissueGene (now called
KolonTissueGene)
[181,182]
18 Onpattro (Patisiran) 2018 Hereditary Transthyretin Amyloidosis Alnylam Pharmaceuticals
Inc.
$345000 per 2 mg/ml [44]
19 Collategene (Beperminogene
perplasmid)
2019 Critical Limb Ischemia AnGes Inc. $18,000 to $27,000 per patient [183]
20 Zolgensma (Onasemnogene
Abeparvovec-xioi)
2019 Pediatric SMA Novartis Inc. Range ($2.1255.0M/
treatment), the average lifetime
cost/patient is $4.26.6M
[184]
21 Waylivra (Volanesorsen) 2019 Adult Familial Chylomicronemia
syndrome
Ionis Pharmaceuticals,
Inc.
17381.15 for 1.5 ml vial [185]
22 Zynteglo 2019 Adult transfusion-dependent ß-
thalassemia
bluebird bio, Inc. 1.58 million euros ($1.78
million)
[186]
23 Vyondys 53 (Golodirsen) 2019 Duchenne Muscular Dystrophy Sarepta Therapeutics $1679.90 for 2 ml vial [187]
24 Tecartus (brexucabtagene
autoleucel)
2020 relapsed/refractory mantle cell
lymphoma
Kite, A Gilead Company $373,000 for a one-time infusion [188]
25 ABECMA (Idecabtagene
vicleucel)
2021 Adult patients with multiple myeloma
who have received at least three prior
therapies
bluebird bio, Inc. $419,500 [189]
26 BREYANZI (lisocabtagene
maraleucel)
2021 adult patients with relapsed or
refractory (R/R) large B-cell lymphoma
(LBCL)
Bristol Myers Squibb $410,300 for a one-time infusion [190]
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
10
Potential tumorigenesis concerns
This challenge is related to the stem cell therapy eld. Stem cells
used in regenerative medicine approaches share some common
characteristics with cancer stem cells, including self-renewal, dif-
ferentiation, and epithelial to mesenchymal transition capacities,
which raise concerns about their potential ability for tumorigenesis
[208]. There were several conicting results about the tumor growth
promotion or tumor formation of these cells, although the exact
tumorigenesis ability of these cells is still under doubt [209211].
Also, there is a difference between the rates of malignant transform
of mesenchymal, neural, and hematopoietic stem cells with ESCs and
iPSCs.
Although it has been shown that the long-term rate of malignant
transformation of bone-marrow-derived MSC was about 45% [210],
ESC and iPSCs are less safe for clinical use, and most studies
demonstrated the highly tumorigenic capability of these cells [212].
Therefore, several strategies have been proposed for their ex-vivo
elimination prior to administration [213]. These strategies include
excluding these cells using antibodies targeting their specic surface
biomarkers using cell sorters, forcing them to differentiate into
specic cell lineage expressing certain reporter protein used for their
identication, using toxic antibodies or antibody-guided toxins,
using cytotoxic agents against undifferentiated pluripotent cells,
using prodrugs after incorporating suicide genes, and using novel
DNase expression induction following continuous proliferation
instead of differentiation. Further studies still would be necessary to
comprehensively assess the tumorigenesis potential of these cells and
ways to mitigate it [212,213].
Public perceptions of gene therapy
Although gene therapy and other regenerative approaches raise
the hope for curing the disease previously known as incurable, there
are some public concerns towards genome editing, transgenic crea-
tures, and regenerative approaches [214]. In general, people support
the benets of gene therapy outweighing its risks, as shown in the
study by Robillard et al. [215], demonstrating a positive attitude
toward gene therapy in almost three-forth of the people in the United
States. However, following the adverse effects of gene therapy seen
in the early clinical trials of ornithine transcarbamylase, deciency
(OTCD) (rst death attributable to gene therapy because of massive
Table 2
Approved cell therapy products.
Category Trade name (Proper name) Year Details Indication Refs.
Adult cell
therapy
Azcel-T (Laviv) 2011 Autologous source of skin broblast cells, First
cellular therapy autologous drug
Aesthetic therapy [192]
Maci (Autologous Cultured Chondrocytes on a Porcine
Collagen Membrane)
2016 Full-thickness cartilage
defects of the knee under
55 years
[193]
APLIGRAF (tissue-engineered biological wound
dressing mat)
1998 The rst cell contained and matrix composite tissue
analog
wound healing process [191]
Gintuit (Allogeneic Cultured Keratinocytes and
Fibroblasts in Bovine Collagen)
2012 Fibroblast and keratinocyte-containing sheet, the rst
FDA approved cell therapy product for the treatment
of wounds in the oral soft tissue defects, especially
gum surgery
Wounds of the oral soft
tissue defects
[194]
Stem cell
therapy
Clevecord (HPC, Cord Blood) 2016 Unrelated donor hematopoietic stem/progenitor cell Disorders affecting the
hematopoietic system
[195]
Hemacord (HEMACORD (HPC, Cord Blood)) 2011 Unrelated donor hematopoietic stem/progenitor cell,
rst-ever cellular drug approved
Disorders affecting the
hematopoietic system
[196]
Ducord (HPC, Cord Blood) 2012 Unrelated donor hematopoietic stem/progenitor cell Disorders affecting the
hematopoietic system
[197]
HPC, Cord Blood (Clinimmune Labs; University of
Colorado Cord Blood Bank, LifeSouth Community
Blood Centers, Inc; Bloodworks; MD Anderson Cord
Blood Bank)
2012 Unrelated donor hematopoietic stem/progenitor cell Disorders affecting the
hematopoietic system
[198201]
Allocord (HPC, Cord Blood) 2011 Unrelated donor hematopoietic stem/progenitor cell Disorders affecting the
hematopoietic system
[202]
Immuno-
cell
therapy
Provenge (Autologous Cellular Immunotherapy) 2010 First autologous cell therapy cancer vaccine Advanced prostate cancer [203]
Fig. 7. Principle challenges of regenerative medicine in cancer treatment.
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
11
immunological response and multi-organ failure) [216], SCIDX1
(insertional mutagenesis following the insertion of vector upstream
of LMO protooncogene, resulting in the development of T cell acute
lymphoblastic leukemia in 25% of participants) [217,218], public
concerns about the risk and side effects of these approaches, chal-
lenged its sustainable progression. Although nowadays these risks
are better identied and understood and supposed very unlikely in
gene therapy, the situation is better for organ transplantation and
stem cell therapy has given a vast amount of information and
well-identied risks [219]. Finally, despite extensive progression in
the safety of gene therapy vectors and strict regulatory supervision,
strategies should be used to build up the necessary public trust in
these approaches [220].
Establishing a balance between the benets and risks would be one
of the hurdles of gene therapy in the future. Improving the standards
of informed consent, using minimally invasive methods, developing
greater public trust, acquiring appropriate regulatory frameworks,
providing better education of gene therapy to patients and stake-
holders, and personalized discussion about the individual risk/be-
nets of gene therapy, would be the future steps to make gene
therapy more appealing for people to accept.
Ethical and moral issues
Signicant ethical and moral considerations should be taken for
implementing regenerative medicine approaches. The sense of
interfering with natural or genetic manipulating the natural diversity
for longevity and enhancements faced this eld of treatment with
ethical issues [220]. However, less risky techniques have shown to be
more morally acceptable, demonstrating its close relationship with
treatment risks [221], some other serious issues including germ cell
(versus somatic cell) genetic modication, administration of gene
therapy to children and fetus and the conict of interests for selecting
participants who would be enrolled in high cost, high-risk clinical
trials of gene therapy still under debate [220].
Implementation
As these products are basically novel, clinicians and specialists
need to be trained to use them. Clinicians should be aware of the
specic methods for decision making, treatment implementing,
monitoring, and expected results of these products to manage the
patients according to a clear clinical workow [222]. Also, necessary
information about the long-term outcomes of these products, which
is essential for advocacy and getting the support of trusts/clinics, is
still relatively scarce. Coupling tissue engineers with clinicians in
centers of excellence could pave the way for this kind of approach
[223].
Treatment durability
Treatment durability is one of the most essential factors for clini-
cians, especially for cancer treatment in which relapses are common.
Cellular treatment showed promising results in cancer treatment;
however, treatment efciency and durability are still under concern.
Similar to other treatment modalities, like chemotherapy or radio-
therapy, single-agent treatments usually fail to achieve a complete
durable response, encouraging combination treatments [81]. In
combination with immunotherapies or chemo- or radio-therapies,
stem cell-based treatments showed more enhanced durable treat-
ment outcomes [224226]. Also, scaffolds could be used as cancer
treatment supportive platforms, providing necessary growth factors,
cytokines, and co-stimulatory molecules [212].
Scalability
Tissue engineering for clinical applications needs further beyond
laboratory-scaled engineered products. They need to have a suitable
size, appropriate transport (vascular) system, good physical speci-
cations according to the host environment, and considerable homo-
geneity across the engineered tissue [205]. In addition, translational
aspects of novel regenerative medicine approaches should be
considered in their development and large-scale manufacturing
[227,228].
Materials and manufacturing protocols
The interaction between the tissue matrix and the corresponding
cells stresses out the role of a reliable selection of tissue components
in its fully functional state. The tissue component sources like cells or
matrix used for culture or the scaffold material used plays a pivotal
role in the reliability of the regenerative medicine approach [205,
227229].
Host toleration
One of the key steps for these approaches is the host toleration
against the engineered products, in which without successful toler-
ation and engraftment, the host body will reject the corresponding
tissue, which results in more consequences. Several strategies for
overcoming this challenge were used, including modulation of the
local immune response, use of a preconditioning regimen that eases
the metabolic shock of going from culture medium to plasma, or
stimulation of angiogenesis [205].
Costs
Like any novel treatment facilities, such as transplants and ro-
botics, regenerative medicine approaches, including cell therapy,
gene therapies, or tissue engineering, have high up-front costs [230].
Moving towards large-scale manufacturing, off-the-shelf products,
health economic models, insurance coverage could help these
productsaccessibility and affordability [223,231].
4. Future perspectives
4.1. Vaccination
Biomaterial-based vaccines in cancer treatment
Vaccination is a promising way to treat cancer by recruiting the
patients immune system [232]. However, problems associated with
the specied T-cell response to cancer and the low inltration
capability of effector T-cells into tumors remain unsolved. Several
approaches using biomaterials and nanomedicine concepts are
recently pursued to achieve a better outcome in treating cancers
[233,234].
Regarding biomaterial-based vaccines, the inherent
antigen-presenting cells (APC) are recruited by releasing
chemo-attractants, such as GM-CSF, and fed with the specic tumor
antigens and adjuvants [235]. GM-CSF is a promising chemotactic
agent for recruiting dendritic cells (DCs) and inducing MHC-II and
CD86(+) expression [236]. In other words, the biomaterials in the
shape of the scaffold provide an appropriate setting to maturate the
inltrated immature DCs. The resulted DCs migrate to the lymph
node to deliver the antigens to the T-cells leading to the lymphocyte
activation and the specied antitumor reaction [237]. The porous
structure of the biomaterials alongside the sustained-release ability
of cytokines is necessary to design a biomaterial-based vaccine.
Bencherif et al. [238] created an injectable microporous alginate
cryogel and loaded it with CPG-ODN and GM-CSF as adjuvant and
chemoattractant, respectively. The system considerably enhanced
the fraction of CD86+MHCII+cells and the amount of IL-12
in-vitro. Besides, this system signicantly increased cytokines
levels, including the interleukin family, as proof of successful
vaccination. In the next research, they used tough alginate to
maintain the porous structure post-injection, which has enhanced
the fraction of CD86+and CD11+compared to chemically
cross-linked [239].
Ali et al. [240] assessed several Toll-like receptor agonists com-
bined with GM-CSF in the poly(D, L-lactide-co-glycolide) PLG scaf-
fold for eliminating melanoma cancer. The CpG-ODN,
monophosphoryl lipid A (MPLA), and polyinosinic: polycytidylic
acid (P (I: C)) showed 15, 20, and 23-fold DCs, respectively. More-
over, CPG-ODN and P(I: C) enhanced the migration of the DCs to the
lymph nodes by 10-fold, which is consistent with the anti-cancer
effect [240]. Combining PLG-based vaccines with CTLA-4 and
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
12
anti-PD-1 checkpoint antibodies increased their effect against
B16-F10 melanoma cancer [241].
Recently, ceramic-based biomaterials have been introduced in
vaccine cancer therapy [242]. In this way, mesoporous silica
microrods (MSR) mixed with PEI and loaded with CPG, Ovalbumin,
and GM-CSF to form a scaffold-shaped structure following the in-
jection for hosting the attracted DCs [242]. This system was suc-
cessful in DCs maturation and CD8 cytotoxic T-cells activation,
which exerts tumor regression. Additionally, the combinational
approach with anti-CTLA4 shows more tumor growth reduction
versus an alone method [242].
Stem cell-based vaccine therapy
Similarly, expressed antigens over ESCs and tumors bring the
chance of immunotherapy. Using a whole-cell as a vaccine agent
instead of specied or cell lysate brings the availability of a large
number of known and unknown antigens [243]. This property is
beyond the ability of monoclonal antibodies, which necessitates the
identication of dened antigens. The role of ESCs in inducing an
immune response against tumors was initially reported on the colon
[244] and lung cancers [245]. Zhang et al. [246] showed that the
human embryonic stem cells (H9) and murine mesenchymal stem
cells (IVP-ES1) effectively suppressed cancer in a mouse bearing ID8
ovarian cancer.
Yaddanapudi et al. [247] reported that the synergistic effect be-
tween ESCs and GM-CSF administration decelerated the rate of Lewis
lung carcinoma (LLC) tumor growth. Later, the same group used
GM-CSF bearing exosomes extracted from engineered ESCs to mini-
mize tumor outgrowth [248]. The results showed the activation of
effector CD8+T-cells and a suppressive effect on the migration of
immunosuppressive Treg. Li et al. [249] used a simulated micro-
gravity process on MSCs to increase the Treg cells ratio and apoptosis
mediated by CD8+cytotoxic T lymphocytes. Elevated expression of
MHC I and HSP proteins was consistent with the ultimate immune
response.
As mentioned before, iPSCs are an emerging technology in stem
cell therapy and regenerative medicine. These cells are a suitable
candidate for resolving the ethical problems associated with ESCs
utilization. Therefore, Kooreman et al. [88] injected CpG and iPSCs
in a breast cancer model via activating the cytotoxic and helper T
cells. The immunosuppressive microenvironment made this system
unable to elicit antitumor activity on the melanoma model.
However, tumor manipulation via resection led to B cells and
myeloid cell activation [88]. Combination methods using immune
checkpoint inhibitors (ICI) or radiotherapy with iPSC may syner-
gistically improve the treatment. In this way, Mashima et al. [250]
introduced GM-CSF secreting myeloid cells derived from induced
pluripotent stem cells (iPSC-pMCs) as antigen-presenting cells. The
combination of iPSC-pMC and ICI increased the inltrated cytotoxic
T-cells and NK cells. Most recently, Wang et al. [251] found a similar
gene expression among iPSC derived from human broblast and
cancer stem cells from lung adenocarcinoma. The iPSC and CpG
simultaneous administration signicantly increase effector/cyto-
toxic T-cells, resulting in reduced tumor size.
4.2. Microenvironment modication
The TME is a sophisticated medium in both chemical and mechanical
aspects [252]. The progressive TME stiffens over time due to the per-
manent extracellular matrix (ECM) production and rearrangement
[252]. Also, the tumors physical modication leads to malignancy and
progression due to the altered cell-matrix interaction and
epithelial-mesenchymal transition [253,254]. In brief, cell interaction
with matrix conventionally applies using cell surface-expressed pro-
teins, such as integrin, which coordinates with the intracellular cyto-
skeleton [255]. Therefore, the cytoskeleton directly modulates cell
migration, and integrin-dominated mechano-sensing affects cancer cell
invasion [256]. In a study by Peng et al. [257], the stiff polyacrylamide
(PAA) gel represents higher migration owing to the b1-FAK activation
mediated RhoA/ROCK1/p-MLC and RhoA/ROCK2/p-colin pathway in
the breast cancer cell. Epithelial-transition is a hallmark of metastasis
that identies via various mechanisms, such as cadherin-mediated
cell-cell adhesion loss [258]. Surprisingly a meaningful decrease in
E-cadherin content is proved in stiff materials that bring
epithelial-mesenchymal transition [259].
MSCs showed a wide range of results from tumoricidal to metastasis
induction by secreting cytokines in the TME. Moreover, the interplay
among MSCs and cancer cells in the TME produces carcinoma-associated
MSC and regulates their function [260,261]. MSCs protect cancer cells
via interaction with the immune system, modulating their antitumor
effects [262]. Namely, TGF-β1 secretion by MSCs increases the regula-
tory T cells (Tregs) with a diminished amount of antitumor cytokines
[263]. In other words, the MSCs provide the circumstance for the breast
cancer cellsdormancy.
Also, the dormancy effect of the MSCs on metastatic breast cancer
cells was dominated under exosomes secreted carrying microRNAs is
proved [264,265]. Noteworthy exosomes from MSCs arrange TME by
angiogenesis suppression [266,267] and increased migration [268]. In
contrast, several exosomes from cancer cells regulate the environment
by transforming MSCs into supportive stromal cells [269].
Cholangiocarcinoma-related research has evidenced increased met-
astatic behavior and proliferation in cancer cells mixed with human
umbilical cord-derived MSCs through the Wnt/ β-catenin pathway
[270]. Besides, MSCs secrete several chemokines, such as CCL5, which
activate the AKT/NF-κB pathway and enhance metastasis under
inammation [271]. Also, CCL5 secretion by MSCs was responsible for
enhancing metastasis of the low metastatic breast cancer cells [209].
Breast cancer cells have two main categories, hormone-dependent
and independent; the rst demands hormones in the medium to form
a tumor. However, MCF-7 cells expressing estrogen receptors could form
tumors in assistance with adult human MSCs [272]. This report inten-
sied the role of MSCs as tumoral stromal cells, such as
cancer-associated broblasts (CAFs) and the remained cells for cancer
progression [272].
Studies proved the migration and homing of MSCs into the TME and
used it for targeted delivery [273275]. Jung et al. [276] showed that
prostate tumors recruit MSCs by releasing CXCL16 chemokine and
inducing their transformation to CAFs. The transformed cells secrete
CXCL12 for increasing the metastatic of prostate cancer cells via
epithelial-to-mesenchymal transition [276]. Several ongoing clinical
trials leverage MSCs potential for cancer therapy [71]. Also, Table 3
lists the key researches utilizing stem cells in various cancer treatments.
Nanotechnology and novel methods for TME manipulation
Nanotechnology brings several particular tools for manipulating the
function of the tumor component. For instance, photo-thermal therapy
(PTT) using nanoparticles (NPs) signicantly decreases the tumors
stiffness, which results in higher immune cell inltration [300]. As
illustrated in Fig. 8a, the carbon nanotube (CNT) content under laser
irradiation affects tumor collagen components dissociation. Marangon
et al. [300] demonstrate the role of thermal ablation or mild hyper-
thermia in the presence of CNT on the stiffness and tumor size. On the
other hand, the effect of hyperthermia on increasing the permeability of
tumor vessels is proven [301]. Also, increasing the distance between
collagen bers from 101 ±17 nm in the control group to 133 ±32 nm
in tumors treated with iron oxide under hyperthermia can indicate tu-
morsabnormal structure [302]. Irradiation effect on CNTs and tissue
function could be found in Fig. 8b and c.
Furthermore, photodynamic therapy (PDT) acts as a double-edged
sword on the collagen polymer in several tissues. This effect includes
crosslinking of the corneal tissue [303], and degradation upon
matrix-metalloproteinase (MMP) released upon the tumors broblasts
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
13
Table 3
Utilizing stem cells in various cancer treatment.
Cancer type In-vitro In-vivo Stem cell type Result Note Refs.
Ovarian cancer SK-OV-3 and OVCAR-3 Adipose derived MSCs
(AMSC)
Exosomes secreted from cancer
cells transforms MSCs to tumor
supportive myobroblast
Overexpression of TGF-β, SDF-1
and
α
-SMA
[277]
Ovarian cancer SK-OV-3 and OVCAR-3 -Bone marrow derived
MSC (BMMSC)-Tumor
associated MSC
MSC protects cancer cells against
hyperthermia
C-X-C motif chemokine 12 (CXCL
12) induces thermotolerance
[278]
Ovarian cancer SKOV3, CAOV3,
A2780 and ES2
SKOV3 AMSC Cancer cells secretes high levels of
MMP accompanied by
-Increased MMP2 and MMP9
expression in cancer
[279]
ES2
Ovarian cancer SKOV3 SKOV3 Adult BMMSC -Platelet activating factor secretion
by MSC increased the tumor
progression
[280]
Ovarian cancer SKOV3 SKOV3 Human endometrial
MSC
Enhanced cellular death and
chemosensitivity in-vivo
-Cell cycle arrest, mitochondrial
membrane potential alteration and
decreased AKT signaling
[281]
Ovarian cancer -SKOV3 Human AMSC Antiproliferation induced via
miRNA exosome
- BAX/Caspase-3/caspase-9 were
upregulated and BCL2
downregulated (Intrinsic
apoptosis)
[282]
Ovarian cancer -SKOV3 -SKOV3 AMSC -Increased migration, invasion and
proliferation
- Several proteins such as thymosin
beta 4 X-linked upregulated in
cancer cells
[283]
-HO8910
-ES2
Breast and ovarian
cancer
-TOV-112D -Human umbilical cord
whartons jelly MSC
(hUCMSC)
hUCMSC have not transformed to
tumor associated broblast
TAF (TGF-β etc.) were
overexpressed in BMMSC
[284]
-MDA-MB-231 -BMMSC
(Conditioned media)
Mammary cancer Mat B III Rat umbilical cord stem
cells from Whartons
jelly
Tumoral regression within 3438
days
- Apoptosis and antiproliferation
effect was evidenced
[285]
-No sign of recurrence in 100 days
Breast cancer MDA MB 231 MDA MB
231
hUCMSC Attenuated tumor growth and
colony formation
-Intravenous injected stem cells
were accumulated in tumor site
[286]
-Manipulated the cell cycles in G2
phase
Breast cancer MCF-7 Human MSC -Attenuated tumor growth -Wnt pathway inhibition in cancer
cells via MSC secreted dickkopf-1
(Dkk-1)
[102]
Colorectal cancer HT-29 HT-29 Human BM-MSC Increased tumor initiation and
sphere formation
IL-6 secretion by MSC stimulates
tumorigenesis
[287]
Colorectal cancer HCT-8/E11, HCT116,
SW480, HT29,LoVo
and T84
HCT 8/E11 BM-MSC Increased tumorigenesis, survival
and invasion
Neuregulin 1 secretion by MSC
activates PI3K/AKT
[288]
Colorectal cancer HCT116, LS180,
COLO205, HT29 and
SW480
LS180 BM-MSC -EMT induction -Cell-cell transactions through
surface TGF-β
[289]
-Excessive angiogenesis and tumor
growth
Colorectal cancer SW480, LS174T and
HT29
SW480,
LS174T and
HT29
MSC Increased angiogenesis IL-8 mediated increased HUVEC
proliferation and migration
[290]
Colorectal cancer SW480, SW620, HT29,
DLD1 and HCT116
HT29,
SW480, and
HCT116
MSC Somewhat strong CCR5 expression
dominates tumor growth
[291]
Colorectal cancer HT29, Lovo, SW1116,
and Caco2
HT29 MSC (TNF-
α
treated) -EMT induction CCL5 secretion by MSC targets
CCR1 receptors and activating
CCL5/β-catenin/Slug pathway
[292]
-Enhanced proliferation, migration
and invasion
Colorectal cancer SW48, SW480 and
SW620
SW48 Cancer derived MSC -EMT induction IL-6 secretion mediates IL-6/JAK2/
STAT3 signaling pathway
dominates
[293]
-Promoted invasiveness and
metastatic behavior
Colorectal cancer SW620 MSC TGF-β1 mediated invasiveness P53 downregulated [294]
Hepatocellular
carcinoma
MHCC97H, MHCC97L
and SMMC-7721 and
HEPG2
SMMC-721,
HEPG2
Cancer derived MSC Promoted proliferation and
invasion
Modulatory effect of S100A4
secretion is evidenced
[295]
Osteosarcoma SaOS-2, MG-63 and
HOS
BM-MSC -Inducing Mesenchymal to
amoeboid transition
MCP-1, GRO-
α
, IL-6 and IL-8
secretion induced GTPase Rhoa
activation
[296]
(continued on next page)
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
14
destruction [304]. Moreover, photochemical degradation and mechan-
ical properties alteration of hyaluronic acid (HA) is proved [305]. Due to
the incident laser and the utilized agent characteristics, PDT has various
vascular effects, from increasing permeability to occlusion [306,307].
Notably, there is a similar effect on TME using radio-sensitizers.
The hypoxic niche is the main hallmark of TME and plays a leading
role in cancer therapy resistance [308] and anaerobic
glycolysis-mediated acidication [309]. Routinely, hyperbaric oxygen
and ozone therapy were used to hamper the hypoxia-associated prob-
lems [310,311]. Oxygen carriers based on hemoglobin (Fig. 9a-c) [312]
and peruorocarbons (Fig. 9d-e) [313,314] are the emerging technolo-
gies to alleviate the tumors hypoxia. In addition, oxygen-generating
particles which produce oxygen through chemical reactions are quite
an of interest. Peroxide-based materials generate oxygen in a water
environment, catalyzed by catalase enzymes [315]. Most recently,
self-oxygen generators are synthesized, following the electrochemical
Table 3 (continued )
Cancer type In-vitro In-vivo Stem cell type Result Note Refs.
Esophageal
Squamous Cell
Carcinoma
Eca109 and TE-1 Eca109 and
TE-1
MSC β2-Microglobulin mediates EMT
induction
-E-cadherin downregulation [297]
Multiple myeloma U266 and LP-1 MSC Increased invasion and decreased
drug sensitivity
-CXCL13 secretion [298]
Melanoma B16-F10 B16-F10 MSC -Cell cycle arrest [299]
-Increases capsase-dependent
signaling
Fig. 8. (a) Effect of hyperthermia on the collagen component
of the tumor. Few intact collagen bers were observed in zones
devoid of CNTs while destructurated collagen bers were
visualized nearby CNTs (circled region). Cell damage evalua-
tion by TPEF and SHG microscopy in response to CNT-
mediated photothermal therapy in tumor slices. Circled re-
gions indicate the presence of CNTs. (b) Without laser irradi-
ation, the integrity of the tissue was observed both in the
presence or absence of CNTs for CNT-injected non-irradiated
control group. (c) Under laser exposure, tissue damage
correlated with the presence of CNTs while the tissue preserved
its integrity in the regions devoid of CNTs for CNT-injected and
irradiated group.
reprinted with permission from [300].
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
15
principles for oxygen generation [316318]. Therefore, using the oxy-
gen delivery method diminishes the hypoxic niche and diminishes the
acidosis and excessive angiogenesis in tumors.
Leveraging drugs for modifying the TME backs from the crosstalk
between cells and their residing niche. Drugs that act as angiogenesis
inhibitors, such as Axitinib, are the major subgroup that impedes tumor
progression and aggression via selectively inhibiting VEGF [319].
Quercetin, a plant-derived avonoid, remodels the TME to introduce
and retain nanoparticles [319,320]. Honokiol extracted from Magnolia
Ofcinalis acts as a vasculogenic mimicry channels inhibitor, encour-
aging their combination therapy with anticancer drugs [321,322].
Also, curcumin binds to the VEGF receptor and inhibits VEGF release
[323]. Moreover, conventional drugs, such as Aspirin, have recently
gained attention in modulating the tumors [324]. A drug delivery sys-
tem carries a drug to a specic tissue, minimizing the cytotoxicity while
increasing the efcacy [325]. Drugs are encapsulated within the reser-
voir and incorporated into the nanoparticles structure, physically and
chemically [326].
5. Conclusion
The concept of regenerative medicine highlights its potential in
treating various diseases through various facilities. The application of
regenerative medicine approaches in cancer treatment widens the hope
avenue of treating severely treatable diseases. Conventional cancer
treatments, such as surgery, chemotherapy, and radiotherapy, do not
entirely eradicate cancer since these methods cannot detect tumor cells
and, in addition to cancer cells, damage normal cells. Several regener-
ative medicine approaches, including cell therapy, gene therapy, and
tissue engineering, provide targeted cancer treatment. These methods
are specic, safer, and more effective than conventional methods. The
course of treatment is shortened, and the need for repeated doses is
eliminated, especially in the gene therapy approach. Scaffolds and tissue
engineering constructions can also be benecial in the targeted delivery
of antitumor drugs on tumor sites and overcome the short half-life of
chemotherapy drugs. Scaffolds can facilitate conventional immuno-
therapies and the proper homing and distribution of immune cells in
tumor sites by adjusting the tumor microenvironment. In addition, they
prevent the erosion of immune cells during the complex process of
immunotherapy due to activating factors loaded on the scaffold. Using
biomaterials to recruit immature antigen-presenting cells and further
direct the T-cell toward cancer cell death shows promising in cancer
therapy. Altering the TME by laser-assisted treatment, radiation ther-
apy, and drug delivery leveraged nanomedicines potential. These
methods guide the tumors toward regression by direct or indirect effects
on cancer cells and TME components.
Delivery and selection of appropriate vectors for transgenic expres-
sion in tissues, regulation of optimal therapeutic transgene expression,
and safe and efcient on-site delivery of the product are the most sig-
nicant challenges of gene therapy. Delivering the appropriate dose of
the therapeutic agent, including cell or gene therapy product to the
cancer tissue, trapping the therapeutic agents in organs such as the liver,
and their neutralizing by the treatment system are other signicant
challenges in this way. Furthermore, regulatory issues, proper imple-
mentation and monitoring of treatment, treatment durability, sophisti-
cated manufacturing methods, and the high cost of products are among
the most signicant challenges facing cancer treatment in regenerative
medicine. It is hoped that with the advancement of research in this eld,
the adverse effects of mentioned strategies can be controlled and make it
possible to perform personalized cancer treatments.
Funding
This study was not funded by any institute.
CRediT authorship contribution statement
Vahid Mansouri: Investigation, Conceptualization, Resources,
Visualization, Writing - original draft. Nima Beheshtizadeh: Concep-
tualization, Visualization, Writing - original draft, Writing - review &
editing, Supervision. Maliheh Gharibshahian: Investigation, Re-
sources, Writing - original draft. Leila Sabouri: Investigation, Re-
sources, Writing - original draft. Mohammad Varzandeh: Resources,
Writing - original draft. Nima Rezaei: Conceptualization, Supervision,
Validation, Writing - review & editing, Project administration.
Conict of interest statement
The authors declare that they have no conict of interest.
References
[1] A.S. Mao, D.J. Mooney, Regenerative medicine: current therapies and future
directions, Proc. Natl. Acad. Sci. USA 112 (47) (2015) 14452.
Fig. 9. (a) Hemoglobin-based oxygen carrier. Scanning electron micrographs of desolvated Hemoglobin nanoparticles show spherical particles of similar size and
shape for both (b) glutaraldehyde and (c) dextran cross-linked. The peruorocarbon-based nanoparticlesshape was recorded with (d) SEM, magnication 15,000
and (e) LSM magnication 8000.
(a-c) Reprinted with permission from [312]. (d-e) Reprinted with permission from [314].
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
16
[2] C. Pucci, C. Martinelli, G. Ciofani, Innovative approaches for cancer treatment:
current perspectives and new challenges, Ecancermedicalscience 13 (2019), 961.
[3] B. Mesure, P. Menu, J.K. Venkatesan, M. Cucchiarini, ´
E. Velot, Biomaterials and
gene therapy: a smart combination for MSC musculoskeletal engineering, Curr.
Stem Cell Res. Ther. 14 (4) (2019) 337343.
[4] V. Raeisdasteh Hokmabad, S. Davaran, A. Ramazani, R. Salehi, Design and
fabrication of porous biodegradable scaffolds: a strategy for tissue engineering,
J. Biomater. Sci. Polym. Ed. 28 (16) (2017) 17971825.
[5] L. Roseti, V. Parisi, M. Petretta, C. Cavallo, G. Desando, I. Bartolotti, B. Grigolo,
Scaffolds for bone tissue engineering: state of the art and new perspectives,
Mater. Sci. Eng. C. Mater. Biol. Appl. 78 (2017) 12461262.
[6] K. Xu, K. Ganapathy, T. Andl, Z. Wang, J.A. Copland, R. Chakrabarti, S.
J. Florczyk, 3D porous chitosan-alginate scaffold stiffness promotes differential
responses in prostate cancer cell lines, Biomaterials 217 (2019), 119311, 119311-
119311.
[7] J. Vanderburgh, J.A. Sterling, S.A. Guelcher, 3D printing of tissue engineered
constructs for in vitro modeling of disease progression and drug screening, Ann.
Biomed. Eng. 45 (1) (2017) 164179.
[8] F. Bray, J. Ferlay, I. Soerjomataram, R.L. Siegel, L.A. Torre, A. Jemal, Global
cancer statistics 2018: GLOBOCAN estimates of incidence and mortality
worldwide for 36 cancers in 185 countries, CA: Cancer J. Clin. 68 (6) (2018)
394424.
[9] R.L. Siegel, K.D. Miller, H.E. Fuchs, A. Jemal, Cancer statistics, 2021, CA: Cancer
J. Clin. 71 (1) (2021) 733.
[10] H. Sung, J. Ferlay, R.L. Siegel, M. Laversanne, I. Soerjomataram, A. Jemal,
F. Bray, Global Cancer Statistics 2020: GLOBOCAN estimates of incidence and
mortality worldwide for 36 cancers in 185 countries, CA: Cancer J. Clin. 71 (3)
(2021) 209249.
[11] C. Armengol, M.R. Sarrias, M. Sala, Hepatocellular carcinoma: present and future,
Med. Clín. 150 (10) (2018) 390397.
[12] A. Díaz De Lia˜
no, F. Oteiza Martínez, M.A. Ciga, M. Aizcorbe, F. Cobo, R. Trujillo,
Impact of surgical procedure for gastric cancer on quality of life, Br. J. Surg. 90
(1) (2003) 9194.
[13] J.-F. Liu, Q.-Z. Wang, J. Hou, Surgical treatment for cancer of the oesophagus and
gastric cardia in Hebei, China, Br. J. Surg. 91 (1) (2004) 9098.
[14] M. Loos, J. Kleeff, H. Friess, M.W. Büchler, Surgical treatment of pancreatic
cancer, Ann. N. Y. Acad. Sci. 1138 (1) (2008) 169180.
[15] M.L. James, M. Lehman, P.N. Hider, M. Jeffery, B.E. Hickey, D.P. Francis,
Fraction size in radiation treatment for breast conservation in early breast cancer,
Cochrane Database Syst. Rev. (11) (2010), 003860.
[16] M. McQuestion, M. Fitch, D. Howell, The changed meaning of food: physical,
social and emotional loss for patients having received radiation treatment for
head and neck cancer, Eur. J. Oncol. Nurs. 15 (2) (2011) 145151.
[17] Y. Suh, S.-J. Lee, Radiation treatment and cancer stem cells, Arch. Pharmacal Res.
38 (3) (2015) 408413.
[18] T.J. Whelan, J.-P. Pignol, M.N. Levine, J.A. Julian, R. MacKenzie, S. Parpia,
W. Shelley, L. Grimard, J. Bowen, H. Lukka, F. Perera, A. Fyles, K. Schneider,
S. Gulavita, C. Freeman, Long-term results of hypofractionated radiation therapy
for breast cancer, N. Engl. J. Med. 362 (6) (2010) 513520.
[19] E. De Mattia, E. Cecchin, M. Guardascione, L. Foltran, T. Di Raimo, F. Angelini,
M. DAndrea, G. Toffoli, Pharmacogenetics of the systemic treatment in advanced
hepatocellular carcinoma, World J. Gastroenterol. 25 (29) (2019) 38703896.
[20] R. Dutta, R.I. Mahato, Recent advances in hepatocellular carcinoma therapy,
Pharmacol. Ther. 173 (2017) 106117.
[21] L.A. Riba, R.A. Gruner, A. Fleishman, T.A. James, Surgical risk factors for the
delayed initiation of adjuvant chemotherapy in breast cancer, Ann. Surg. Oncol.
25 (7) (2018) 19041911.
[22] M. Schlumberger, B. Jarzab, M.E. Cabanillas, B. Robinson, F. Pacini, D.W. Ball,
J. McCaffrey, K. Newbold, R. Allison, R.G. Martins, L.F. Licitra, M.H. Shah,
D. Bodenner, R. Elisei, L. Burmeister, Y. Funahashi, M. Ren, J.P. OBrien, S.
I. Sherman, A phase II trial of the multitargeted tyrosine kinase inhibitor
lenvatinib (E7080) in advanced medullary thyroid cancer, Clinical Cancer Res. 22
(1) (2016) 4453.
[23] D. Strumberg, B. Schultheis, Regorafenib for cancer, Expert Opin. Investig. Drugs
21 (6) (2012) 879889.
[24] Y. Xiao, D. Yu, Tumor microenvironment as a therapeutic target in cancer,
Pharmacol. Ther. 221 (2021), 107753, 107753-107753.
[25] A.J. Sorrin, M. Kemal Ruhi, N.A. Ferlic, V. Karimnia, W.J. Polacheck, J.P. Celli, H.
C. Huang, I. Rizvi, Photodynamic therapy and the biophysics of the tumor
microenvironment, Photochem. Photobiol. 96 (2) (2020) 232259.
[26] O. Farc, V. Cristea, An overview of the tumor microenvironment, from cells to
complex networks (Review), Exp. Ther. Med. 21 (1) (2021) 96.
[27] L. Fozzatti, S.-Y. Cheng, Tumor cells and cancer-associated broblasts: a
synergistic crosstalk to promote thyroid cancer, Endocrinol. Metab. 35 (4) (2020)
673680.
[28] P.L. Bedard, A.R. Hansen, M.J. Ratain, L.L. Siu, Tumour heterogeneity in the
clinic, Nature 501 (7467) (2013) 355364.
[29] M.-Z. Jin, W.-L. Jin, The updated landscape of tumor microenvironment and drug
repurposing, Signal Transduct. Target. Ther. 5 (1) (2020), 166.
[30] L. Falzone, S. Salomone, M. Libra, Evolution of cancer pharmacological
treatments at the turn of the third millennium, Front. Pharmacol. 9 (2018) 1300.
[31] A.S. Cheung, D.J. Mooney, Engineered materials for cancer immunotherapy,
Nano Today 10 (4) (2015) 511531.
[32] J.I. Andorko, C.M. Jewell, Designing biomaterials with immunomodulatory
properties for tissue engineering and regenerative medicine, Bioeng. Transl. Med.
2 (2) (2017) 139155.
[33] T. Seya, Y. Takeda, K. Takashima, S. Yoshida, M. Azuma, M. Matsumoto,
Adjuvant immunotherapy for cancer: both dendritic cell-priming and check-point
inhibitor blockade are required for immunotherapy, Proc. Jpn. Acad. Ser. B Phys.
Biol. Sci. 94 (3) (2018) 153160.
[34] P. Huang, X. Wang, X. Liang, J. Yang, C. Zhang, D. Kong, W. Wang, Nano-, micro-,
and macroscale drug delivery systems for cancer immunotherapy, Acta Biomater.
85 (2019) 126.
[35] S.N. Truong, P. Van Pham, Stem cell technology and engineering for cancer
treatment, Biomed. Res. Ther. 2 (6) (2015) 13.
[36] K.B. Kaufmann, H. Büning, A. Galy, A. Schambach, M. Grez, Gene therapy on the
move, EMBO Mol. Med. 5 (11) (2013) 16421661.
[37] B. Arjmand, B. Larijani, M. Sheikh Hosseini, M. Payab, K. Gilany, et al., The
horizon of gene therapy in modern medicine: advances and challenges, in:
K. Turksen (Ed.), Cell Biology and Translational Medicine, Volume 8: Stem Cells
in Regenerative Medicine, Springer International Publishing, Cham, 2020,
pp. 3364.
[38] E. Hanna, C. R´
emuzat, P. Auquier, M. Toumi, Gene therapies development: slow
progress and promising prospect, J. Mark. Access Health Policy 5 (1) (2017),
1265293.
[39] S.L. Ginn, A.K. Amaya, I.E. Alexander, M. Edelstein, M.R. Abedi, Gene therapy
clinical trials worldwide to 2017: an update, J. Gene Med. 20 (5) (2018), 3015.
[40] X.M. Anguela, K.A. High, Entering the modern era of gene therapy, Annu. Rev.
Med. 70 (2019) 273288.
[41] D.B. Kohn, Gene therapy for blood diseases, Curr. Opin. Biotechnol. 60 (2019)
3945.
[42] V.A. Chow, M. Shadman, A.K. Gopal, Translating anti-CD19 CAR T-cell therapy
into clinical practice for relapsed/refractory diffuse large B-cell lymphoma, Blood
132 (8) (2018) 777781.
[43] M. Bowerman, C.G. Becker, R.J. Y´
a˜
nez-Mu˜
noz, K. Ning, M.J.A. Wood, T.
H. Gillingwater, K. Talbot, C. UK SMA Research, Therapeutic strategies for spinal
muscular atrophy: SMN and beyond, Dis. Models Mech. 10 (8) (2017) 943954.
[44] M.S. Maurer, J.H. Schwartz, B. Gundapaneni, P.M. Elliott, G. Merlini,
M. Waddington-Cruz, A.V. Kristen, M. Grogan, R. Witteles, T. Damy, B.
M. Drachman, S.J. Shah, M. Hanna, D.P. Judge, A.I. Barsdorf, P. Huber, T.
A. Patterson, S. Riley, J. Schumacher, M. Stewart, M.B. Sultan, C. Rapezzi,
I. ATTR-ACT Study, Tafamidis treatment for patients with transthyretin amyloid
cardiomyopathy, N. Engl. J. Med. 379 (11) (2018) 10071016.
[45] C. Zuccato, L. Breda, F. Salvatori, G. Breveglieri, S. Gardenghi, N. Bianchi,
E. Brognara, I. Lampronti, M. Borgatti, S. Rivella, R. Gambari, A combined
approach for β-thalassemia based on gene therapy-mediated adult hemoglobin
(HbA) production and fetal hemoglobin (HbF) induction, Ann. Hematol. 91 (8)
(2012) 12011213.
[46] S. Samakoglu, L. Lisowski, T. Budak-Alpdogan, Y. Usachenko, S. Acuto, R. Di
Marzo, A. Maggio, P. Zhu, J.F. Tisdale, I. Rivi`
ere, M. Sadelain, A genetic strategy
to treat sickle cell anemia by coregulating globin transgene expression and RNA
interference, Nat. Biotechnol. 24 (1) (2006) 8994.
[47] E.A. Stadtmauer, J.A. Fraietta, M.M. Davis, A.D. Cohen, K.L. Weber, E. Lancaster,
P.A. Mangan, I. Kulikovskaya, M. Gupta, F. Chen, L. Tian, V.E. Gonzalez, J. Xu, I.
Y. Jung, J.J. Melenhorst, G. Plesa, J. Shea, T. Matlawski, A. Cervini, A.L. Gaymon,
S. Desjardins, A. Lamontagne, J. Salas-Mckee, A. Fesnak, D.L. Siegel, B.L. Levine,
J.K. Jadlowsky, R.M. Young, A. Chew, W.T. Hwang, E.O. Hexner, B.M. Carreno,
C.L. Nobles, F.D. Bushman, K.R. Parker, Y. Qi, A.T. Satpathy, H.Y. Chang, Y. Zhao,
S.F. Lacey, C.H. June, CRISPR-engineered T cells in patients with refractory
cancer, Science 367 (2020) 6481.
[48] N. Nayerossadat, T. Maedeh, P.A. Ali, Viral and nonviral delivery systems for
gene delivery, Adv. Biomed. Res. 1 (2012) 27.
[49] R. Gardlík, R. P´
alffy, J. Hodosy, J. Luk´
acs, J. Turna, P. Celec, Vectors and delivery
systems in gene therapy, Med. Sci. Monit.: Int. Med. J. Exp. Clin. Res. 11 (4)
(2005) RA110RA121.
[50] R. Stripecke, N. Kasahara, Lentiviral and retroviral vector systems, in: K.K. Hunt,
S.A. Vorburger, S.G. Swisher (Eds.), Gene Therapy for Cancer, Humana Press,
Totowa, NJ, 2007, pp. 3971.
[51] N.S. Templeton, Gene and Cell Therapy: Therapeutic Mechanisms and Strategies,
CRC Press, 2008.
[52] C.L. Hardee, L.M. Ar´
evalo-Soliz, B.D. Hornstein, L. Zechiedrich, Advances in non-
viral DNA vectors for gene therapy, Genes 8 (2017) 2.
[53] R. Deev, I. Plaksa, I. Bozo, A. Isaev, Results of an international postmarketing
surveillance study of pl-VEGF165 Safety and Efcacy in 210 Patients with
peripheral arterial disease, Am. J. Cardiovasc. Drugs 17 (3) (2017) 235242.
[54] S.M. Selkirk, Gene therapy in clinical medicine, Postgrad. Med. J. 80 (948) (2004)
560570.
[55] S.R. Kumar, D.M. Markusic, M. Biswas, K.A. High, R.W. Herzog, Clinical
development of gene therapy: results and lessons from recent successes, Mol.
Ther. Methods Clin. Dev. 3 (2016) 16034.
[56] M. Jafarlou, B. Baradaran, T.A. Saedi, V. Jafarlou, D. Shanehbandi, M. Maralani,
F. Othman, An overview of the history, applications, advantages, disadvantages
and prospects of gene therapy, J. Biol. Regul. Homeost. Agents 30 (2) (2016)
315321.
[57] S. Razi Sooyani, B. Baradaran, F. Lotpour, T. Kazemi, L. Mohammadnejad,
Gene therapy, early promises, subsequent problems, and recent breakthroughs,
Adv. Pharm. Bull. 3 (2) (2013) 249255.
[58] X.M. Anguela, K.A. High, Entering the modern era of gene therapy, Annu. Rev.
Med. 70 (1) (2019) 273288.
[59] J.K. R¨
aty, J.T. Pikkarainen, T. Wirth, S. Yl¨
a-Herttuala, Gene therapy: the rst
approved gene-based medicines, molecular mechanisms and clinical indications,
Curr. Mol. Pharmacol. 1 (1) (2008) 1323.
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
17
[60] Z. Peng, Current status of gendicine in China: recombinant human Ad-p53 agent
for treatment of cancers, Hum. Gene Ther. 16 (9) (2005) 10161027.
[61] S.P. Chawla, H. Bruckner, M.A. Morse, N. Assudani, F.L. Hall, E.M. Gordon,
A phase I-II study using rexin-G tumor-targeted retrovector encoding a dominant-
negative Cyclin G1 inhibitor for advanced pancreatic cancer, Mol. Ther.
Oncolytics 12 (2018) 5667.
[62] S. Kim, N. Federman, E.M. Gordon, F.L. Hall, S.P. Chawla, Rexin-G(®), a tumor-
targeted retrovector for malignant peripheral nerve sheath tumor: a case report,
Mol. Clin. Oncol. 6 (6) (2017) 861865.
[63] R. Wang, P.S. Billone, W.M. Mullett, Nanomedicine in action: an overview of
cancer nanomedicine on the market and in clinical trials, J. Nanomater. 2013
(2013) 112.
[64] J. Pol, G. Kroemer, L. Galluzzi, First oncolytic virus approved for melanoma
immunotherapy, Oncoimmunology 5 (1) (2016), 1115641 e1115641-e1115641.
[65] O. Hamid, B. Hoffner, E. Gasal, J. Hong, R.D. Carvajal, Oncolytic immunotherapy:
unlocking the potential of viruses to help target cancer, Cancer
Immunol. Immunother. 66 (10) (2017) 12491264.
[66] M.L. Lim, P. Jungebluth, P. Macchiarini, Regenerative medicine for diseases of
the respiratory system. Translational Regenerative Medicine, Elsevier, 2015,
pp. 449456.
[67] P.L.F. Giangrande, The history of blood transfusion, Br. J. Haematol. 110 (4)
(2000) 758767.
[68] M. Liang, Oncorine, the world rst oncolytic virus medicine and its update in
China, Curr. Cancer Drug Targets 18 (2) (2018) 171176.
[69] S.A. Holstein, M.A. Lunning, CAR T-cell therapy in hematologic malignancies: a
voyage in progress, Clin. Pharmacol. Ther. 107 (1) (2020) 112122.
[70] M.L. Grant, C.M. Bollard, Cell therapies for hematological malignancies: dont
forget non-gene-modied t cells!, Blood Rev. 32 (3) (2018) 203224.
[71] D.-T. Chu, T.T. Nguyen, N. Tien, D.-K. Tran, J.-H. Jeong, P.G. Anh, V.V. Thanh, D.
T. Truong, T.C. Dinh, Recent progress of stem cell therapy in cancer treatment:
molecular mechanisms and potential applications, Cells 9 (2020) 3.
[72] H. Rafei, R.S. Mehta, K. Rezvani, Editorial: cellular therapies in cancer, Front.
Immunol. 10 (2019) 2788.
[73] C.-L. Zhang, T. Huang, B.-L. Wu, W.-X. He, D. Liu, Stem cells in cancer therapy:
opportunities and challenges, Oncotarget 8 (43) (2017) 7575675766.
[74] N. Chopra, S. Choudhury, S. Bhargava, S. Wajid, N.K. Ganguly, Potentials of
stem cell-therapyin pancreatic cancer: an update, Pancreatology 19 (8) (2019)
10341042.
[75] C. Tran, M.S. Damaser, Stem cells as drug delivery methods: application of stem
cell secretome for regeneration, Adv. Drug Deliv. Rev. 82 (2015) 111.
[76] S.-K. Kim, S.U. Kim, I.H. Park, J.H. Bang, K.S. Aboody, K.C. Wang, B.K. Cho,
M. Kim, L.G. Menon, P.M. Black, R.S. Carroll, Human neural stem cells target
experimental intracranial medulloblastoma and deliver a therapeutic gene
leading to tumor regression, Clin. Cancer Res. 12 (18) (2006) 55505556.
[77] J. Xiao, J. Mu, T. Liu, H. Xu, Dig the root of cancer: targeting cancer stem cells
therapy, J. Med. Discov. 2 (2) (2017) 16.
[78] M. Duebgen, J. Martinez-Quintanilla, K. Tamura, S. Hingtgen, N. Redjal,
H. Wakimoto, K. Shah, Stem cells loaded with multimechanistic oncolytic herpes
simplex virus variants for brain tumor therapy, J. Natl. Cancer Inst. 106 (2014)
90.
[79] Y. Reisner, H. Segall, Hematopoietic stem cell transplantation for cancer therapy,
Curr. Opin. Immunol. 7 (5) (1995) 687693.
[80] F. Vanessa, J. Christian, Mesenchymal stem cells: an emerging tool for cancer
targeting and therapy, Curr. Stem Cell Res. Ther. 3 (1) (2008) 3242.
[81] J.R. Bag´
o, K.T. Sheets, S.D. Hingtgen, Neural stem cell therapy for cancer,
Methods 99 (2016) 3743.
[82] H.-T. Lin, M. Otsu, H. Nakauchi, Stem cell therapy: an exercise in patience and
prudence, Philos. Trans. R. Soc. Lond. Ser. B Biol. Sci. 368 (1609) (2013),
20110334.
[83] K. Takahashi, S. Yamanaka, Induction of pluripotent stem cells from mouse
embryonic and adult broblast cultures by dened factors, Cell 126 (4) (2006)
663676.
[84] Y. Qiao, O.S. Agboola, X. Hu, Y. Wu, L. Lei, Tumorigenic and immunogenic
properties of induced pluripotent stem cells: a promising cancer vaccine, Stem
Cell Rev. Rep. 16 (6) (2020) 10491061.
[85] S.J. Sharkis, R.J. Jones, C. Civin, Y.Y. Jang, Pluripotent stem cell-based cancer
therapy: promise and challenges, Sci. Transl. Med. 4 (127) (2012) 127.
[86] H. Liu, Y. Kim, S. Sharkis, L. Marchionni, Y.-Y. Jang, In vivo liver regeneration
potential of human induced pluripotent stem cells from diverse origins, Sci.
Transl. Med. 3 (82) (2011) 82.
[87] M. Katsukawa, Y. Nakajima, A. Fukumoto, D. Doi, J. Takahashi, Fail-safe therapy
by gamma-ray irradiation against tumor formation by human-induced pluripotent
stem cell-derived neural progenitors, Stem Cells Dev. 25 (11) (2016) 815825.
[88] N.G. Kooreman, Y. Kim, P.E. de Almeida, V. Termglinchan, S. Diecke, N.Y. Shao,
T.T. Wei, H. Yi, D. Dey, R. Nelakanti, T.P. Brouwer, D.T. Paik, I. Sagiv-Bar,
A. Han, P. Quax, J.F. Hamming, R. Levy, M.M. Davis, J.C. Wu, Autologous iPSC-
based vaccines elicit anti-tumor responses in vivo, Cell Stem Cell 22 (4) (2018)
501513, e7.
[89] D.A. Knorr, Z. Ni, D. Hermanson, M.K. Hexum, L. Bendzick, L.J. Cooper, D.A. Lee,
D.S. Kaufman, Clinical-scale derivation of natural killer cells from human
pluripotent stem cells for cancer therapy, Stem Cells Transl. Med. 2 (4) (2013)
274283.
[90] F.H. Gage, S. Temple, Neural stem cells: generating and regenerating the brain,
Neuron 80 (3) (2013) 588601.
[91] P. Ellis, B.M. Fagan, S.T. Magness, S. Hutton, O. Taranova, S. Hayashi,
A. McMahon, M. Rao, L. Pevny, SOX2, a persistent marker for multipotential
neural stem cells derived from embryonic stem cells, the embryo or the adult,
Dev. Neurosci. 26 (24) (2004) 148165.
[92] D. Kanojia, I.V. Balyasnikova, R.A. Morshed, R.T. Frank, D. Yu, L. Zhang, D.
A. Spencer, J.W. Kim, Y. Han, D. Yu, A.U. Ahmed, K.S. Aboody, M.S. Lesniak,
Neural stem cells secreting anti-HER2 antibody improve survival in a preclinical
model of HER2 overexpressing breast cancer brain metastases, Stem Cells 33 (10)
(2015) 29852994.
[93] H.J. Lee, S.W. Doo, D.H. Kim, Y.J. Cha, J.H. Kim, Y.S. Song, S.U. Kim, Cytosine
deaminase-expressing human neural stem cells inhibit tumor growth in prostate
cancer-bearing mice, Cancer Lett. 335 (1) (2013) 5865.
[94] B.-R. Yi, S.U. Kim, K.-C. Choi, Co-treatment with therapeutic neural stem cells
expressing carboxyl esterase and CPT-11 inhibit growth of primary and
metastatic lung cancers in mice, Oncotarget 5 (24) (2014) 1283512848.
[95] R. Mooney, A.A. Majid, J. Batalla-Covello, D. Machado, X. Liu, J. Gonzaga,
R. Tirughana, M. Hammad, M.S. Lesniak, D.T. Curiel, K.S. Aboody, Enhanced
delivery of oncolytic adenovirus by neural stem cells for treatment of metastatic
ovarian cancer, Mol. Ther. Oncolytics 12 (2019) 7992.
[96] L. da Silva Meirelles, P.C. Chagastelles, N.B. Nardi, Mesenchymal stem cells reside
in virtually all post-natal organs and tissues, J. Cell Sci. 119 (Pt 11) (2006)
22042213.
[97] Z. Sun, S. Wang, R.C. Zhao, The roles of mesenchymal stem cells in tumor
inammatory microenvironment, J. Hematol. Oncol. 7 (1) (2014) 14.
[98] E. N Momin, G. Vela, H. A Zaidi, A. Qui˜
nones-Hinojosa, The oncogenic potential
of mesenchymal stem cells in the treatment of cancer: directions for future
research, Curr. Immunol. Rev. 6 (2) (2010) 137148.
[99] S.A. Park, C.H. Ryu, S.M. Kim, J.Y. Lim, S.I. Park, C.H. Jeong, J.A. Jun, J.H. Oh, S.
H. Park, W. Oh, S.S. Jeun, CXCR4-transfected human umbilical cord blood-
derived mesenchymal stem cells exhibit enhanced migratory capacity toward
gliomas, Int. J. Oncol. 38 (1) (2011) 97103.
[100] K.-W. Seo, H.-W. Lee, Y.-I. Oh, J.-O. Ahn, Y.-R. Koh, S.H. Oh, S.K. Kang, H.
Y. Youn, Anti-tumor effects of canine adipose tissue-derived mesenchymal
stromal cell-based interferon-β gene therapy and cisplatin in a mouse melanoma
model, Cytotherapy 13 (8) (2011) 944955.
[101] B.-R. Yi, S.N. O, N.H. Kang, K.A. Hwang, S.U. Kim, E.B. Jeung, Y.B. Kim, G.J. Heo,
K.C. Choi, Genetically engineered stem cells expressing cytosine deaminase and
interferon-β migrate to human lung cancer cells and have potentially therapeutic
anti-tumor effects, Int. J. Oncol. 39 (4) (2011) 833839.
[102] L. Qiao, Z.-l Xu, T.-j Zhao, L.-h Ye, X.-d Zhang, Dkk-1 secreted by mesenchymal
stem cells inhibits growth of breast cancer cells via depression of Wnt signalling,
Cancer Lett. 269 (1) (2008) 6777.
[103] P. Barcellos-de-Souza, V. Gori, F. Bambi, P. Chiarugi, Tumor microenvironment:
bone marrow-mesenchymal stem cells as key players, Biochim. Biophys. Acta
(BBA) - Rev. Cancer 1836 (2) (2013) 321335.
[104] A.H. Klopp, A. Gupta, E. Spaeth, M. Andreeff, F. Marini Iii, Concise review:
dissecting a discrepancy in the literature: do mesenchymal stem cells support or
suppress tumor growth? Stem Cells 29 (1) (2011) 1119.
[105] T. Sugiyama, T. Nagasawa, Bone marrow niches for hematopoietic stem cells and
immune cells, Inamm. Allergy Drug Targets 11 (3) (2012) 201206.
[106] E. Gschweng, S. De Oliveira, D.B. Kohn, Hematopoietic stem cells for cancer
immunotherapy, Immunol. Rev. 257 (1) (2014) 237249.
[107] F. Barriga, N. Rojas, A. Wietstruck, Alternative donor sources for hematopoietic
stem, Cell Transplant. Innov. Stem Cell Transplant. (2013) 349.
[108] Y. Abe, T. Ito, E. Baba, K. Nagafuji, K. Kawabe, I. Choi, Y. Arita, T. Miyamoto,
T. Teshima, S. Nakano, M. Harada, Nonmyeloablative allogeneic hematopoietic
stem cell transplantation as immunotherapy for pancreatic cancer, Pancreas 38
(7) (2009) 815819.
[109] S. Khurana, L. Margamuljana, C. Joseph, S. Schouteden, S.M. Buckley, C.
M. Verfaillie, Glypican-3-mediated inhibition of CD26 by TFPI: a novel
mechanism in hematopoietic stem cell homing and maintenance, Blood 121 (14)
(2013) 25872595.
[110] S.J. Dylla, L. Beviglia, I.K. Park, C. Chartier, J. Raval, L. Ngan, K. Pickell,
J. Aguilar, S. Lazetic, S. Smith-Berdan, M.F. Clarke, T. Hoey, J. Lewicki, A.
L. Gurney, Colorectal cancer stem cells are enriched in xenogeneic tumors
following chemotherapy, PLOS One 3 (6) (2008), 2428.
[111] A. Eramo, L. Ricci-Vitiani, A. Zeuner, R. Pallini, F. Lotti, G. Sette, E. Pilozzi, L.
M. Larocca, C. Peschle, R. De Maria, Chemotherapy resistance of glioblastoma
stem cells, Cell Death Differ. 13 (7) (2006) 12381241.
[112] M. Dean, T. Fojo, S. Bates, Tumour stem cells and drug resistance, Nat. Rev.
Cancer 5 (4) (2005) 275284.
[113] J.P. Medema, Cancer stem cells: the challenges ahead, Nat. Cell Biol. 15 (4)
(2013) 338344.
[114] D.R. Pattabiraman, R.A. Weinberg, Tackling the cancer stem cellswhat
challenges do they pose? Nat. Rev. Drug Discov. 13 (7) (2014) 497512.
[115] J.C. Chang, Cancer stem cells: role in tumor growth, recurrence, metastasis, and
treatment resistance, Medicine 95 (Suppl 1) (2016) S20S25.
[116] N. Haraguchi, H. Ishii, K. Mimori, F. Tanaka, M. Ohkuma, H.M. Kim, H. Akita,
D. Takiuchi, H. Hatano, H. Nagano, G.F. Barnard, Y. Doki, M. Mori, CD13 is a
therapeutic target in human liver cancer stem cells, J. Clin. Investig. 120 (9)
(2010) 33263339.
[117] L. Vermeulen, F. De Sousa E Melo, M. van der Heijden, K. Cameron, J.H. de Jong,
T. Borovski, J.B. Tuynman, M. Todaro, C. Merz, H. Rodermond, M.R. Sprick,
K. Kemper, D.J. Richel, G. Stassi, J.P. Medema, Wnt activity denes colon cancer
stem cells and is regulated by the microenvironment, Nat. Cell Biol. 12 (5) (2010)
468476.
[118] Z. Li, S. Bao, Q. Wu, H. Wang, C. Eyler, S. Sathornsumetee, Q. Shi, Y. Cao,
J. Lathia, R.E. McLendon, A.B. Hjelmeland, J.N. Rich, Hypoxia-inducible factors
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
18
regulate tumorigenic capacity of glioma stem cells, Cancer Cell 15 (6) (2009)
501513.
[119] K. Hansen, F.-J. Müller, M. Messing, F. Zeigler, J.F. Loring, K. Lamszus,
M. Westphal, N.O. Schmidt, A 3-dimensional extracellular matrix as a delivery
system for the transplantation of glioma-targeting neural stem/progenitor cells,
Neuro-Oncol. 12 (7) (2010) 645654.
[120] K.S. Panageas, A.S. Reiner, F.M. Iwamoto, T.F. Cloughesy, K.D. Aldape, A.
L. Rivera, A.F. Eichler, D.N. Louis, N.A. Paleologos, B.J. Fisher, L.S. Ashby, J.
G. Cairncross, G.B. Urgoiti, P.Y. Wen, K.L. Ligon, D. Schiff, H.I. Robins, B.
G. Rocque, M.C. Chamberlain, W.P. Mason, S.A. Weaver, R.M. Green, F.G. Kamar,
L.E. Abrey, L.M. DeAngelis, S.C. Jhanwar, M.K. Rosenblum, A.B. Lassman,
Recursive partitioning analysis of prognostic variables in newly diagnosed
anaplastic oligodendroglial tumors, Neuro-Oncology 16 (suppl_3) (2014)
15411546, iii50-iii50.
[121] V. Golubeva, J. Mikhalevich, J. Novikova, O. Tupizina, S. Tromova, Y. Zueva,
Novel cell population data from a haematology analyzer can predict timing and
efciency of stem cell transplantation, Transfus. Apher. Sci. 50 (1) (2014) 3945.
[122] A.L. Tobias, B. Thaci, B. Aufnger, E. Rinc´
on, I.V. Balyasnikova, C.K. Kim, Y. Han,
L. Zhang, K.S. Aboody, A.U. Ahmed, M.S. Lesniak, The timing of neural stem cell-
based virotherapy is critical for optimal therapeutic efcacy when applied with
radiation and chemotherapy for the treatment of glioblastoma, Stem Cells Transl.
Med. 2 (9) (2013) 655666.
[123] E.K. Sage, R.M. Thakrar, S.M. Janes, Genetically modied mesenchymal stromal
cells in cancer therapy, Cytotherapy 18 (11) (2016) 14351445.
[124] H.J. Jackson, S. Raq, R.J. Brentjens, Driving CAR T-cells forward, Nat. Rev. Clin.
Oncol. 13 (6) (2016) 370383.
[125] L. Labanieh, R.G. Majzner, C.L. Mackall, Programming CAR-T cells to kill cancer,
Nat. Biomed. Eng. 2 (6) (2018) 377391.
[126] P.-P. Zheng, J.M. Kros, J. Li, Approved CAR T cell therapies: ice bucket challenges
on glaring safety risks and long-term impacts, Drug Discov. Today 23 (6) (2018)
11751182.
[127] M. Alcantara, P. Du Rusquec, E. Romano, Current clinical evidence and potential
solutions to increase benet of CAR T-cell therapy for patients with solid tumors,
OncoImmunology 9 (1) (2020), 1777064.
[128] M. Martinez, E.K. Moon, CAR T cells for solid tumors: new strategies for nding,
inltrating, and surviving in the tumor microenvironment, Front. Immunol. 10
(2019) 128.
[129] E. Donnadieu, L. Dupr´
e, L.G. Pinho, V. Cotta-de-Almeida, Surmounting the
obstacles that impede effective CAR T cell trafcking to solid tumors, J. Leukoc.
Biol. 108 (4) (2020) 10671079.
[130] M. Cheng, Y. Chen, W. Xiao, R. Sun, Z. Tian, NK cell-based immunotherapy for
malignant diseases, Cell. Mol. Immunol. 10 (3) (2013) 230252.
[131] K. Choucair, J.R. Duff, C.S. Cassidy, M.T. Albrethsen, J.D. Kelso, A. Lenhard,
H. Staats, R. Patel, F.C. Brunicardi, L. Dworkin, J. Nemunaitis, Natural killer cells:
a review of biology, therapeutic potential and challenges in treatment of solid
tumors, Future Oncol. 15 (26) (2019) 30533069.
[132] E. Vivier, E. Tomasello, M. Baratin, T. Walzer, S. Ugolini, Functions of natural
killer cells, Nat. Immunol. 9 (5) (2008) 503510.
[133] A. Malhotra, A. Shanker, NK cells: immune cross-talk and therapeutic
implications, Immunotherapy 3 (10) (2011) 11431166.
[134] M.L. Saetersmoen, Q. Hammer, B. Valamehr, D.S. Kaufman, K.J. Malmberg, Off-
the-shelf cell therapy with induced pluripotent stem cell-derived natural killer
cells, Semin Immunopathol. 41 (1) (2019) 5968.
[135] K. Wang, Y. Han, W.C. Cho, H. Zhu, The rise of human stem cell-derived natural
killer cells for cancer immunotherapy, Expert Opin. Biol. Ther. 19 (2) (2019)
141148.
[136] K. Rezvani, R. Rouce, E. Liu, E. Shpall, Engineering natural killer cells for cancer
immunotherapy, Mol. Ther. 25 (8) (2017) 17691781.
[137] F. Morandi, M. Yazdanifar, C. Cocco, A. Bertaina, I. Airoldi, Engineering the
bridge between innate and adaptive immunity for cancer immunotherapy: focus
on γδ T and NK cells, Cells 9 (8) (2020), 1757.
[138] M.A. Morgan, H. Büning, M. Sauer, A. Schambach, Use of cell and genome
modication technologies to generate improved Off-the-ShelfCAR T and CAR
NK cells, Front. Immunol. 11 (2020) 1965.
[139] M. Katoh, Multi layered prevention and treatment of chronic inammation,
organ brosis and cancer associated with canonical WNT/βcatenin signaling
activation (Review), Int. J. Mol. Med. 42 (2) (2018) 713725.
[140] P.R. Buckley, K. Alden, M. Coccia, A. Chalon, C. Collignon, S.T. Temmerman, A.
M. Didierlaurent, R. van der Most, J. Timmis, C.A. Andersen, M.C. Coles,
Application of modeling approaches to explore vaccine adjuvant mode-of-action,
Front. Immunol. 10 (2019) 2150.
[141] M. Sadeghzadeh, S. Bornehdeli, H. Mohahammadrezakhani, M. Abolghasemi,
E. Poursaei, M. Asadi, V. Zafari, L. Aghebati-Maleki, D. Shanehbandi, Dendritic
cell therapy in cancer treatment; the state-of-the-art, Life Sci. 254 (2020),
117580, 117580-117580.
[142] R. Zhang, M.M. Billingsley, M.J. Mitchell, Biomaterials for vaccine-based cancer
immunotherapy, J. Control. Release: Off. J. Control. Release Soc. 292 (2018)
256276.
[143] H.-G. Hu, Y.-M. Li, Emerging adjuvants for cancer immunotherapy, Front. Chem.
8 (2020) 601.
[144] C.A. Vaine, M.K. Patel, J. Zhu, E. Lee, R.W. Finberg, R.C. Hayward, E.A. Kurt-
Jones, Tuning innate immune activation by surface texturing of polymer
microparticles: the role of shape in inammasome activation, J. Immunol. 190 (7)
(2013) 35253532.
[145] K. Niikura, T. Matsunaga, T. Suzuki, S. Kobayashi, H. Yamaguchi, Y. Orba,
A. Kawaguchi, H. Hasegawa, K. Kajino, T. Ninomiya, K. Ijiro, H. Sawa, Gold
nanoparticles as a vaccine platform: inuence of size and shape on immunological
responses in vitro and in vivo, ACS Nano 7 (5) (2013) 39263938.
[146] K.A. Fitzgerald, J. Guo, E.G. Tierney, C.M. Curtin, M. Malhotra, R. Darcy, F.
J. OBrien, C.M. ODriscoll, The use of collagen-based scaffolds to simulate
prostate cancer bone metastases with potential for evaluating delivery of
nanoparticulate gene therapeutics, Biomaterials 66 (2015) 5366.
[147] S.S. Rao, G.G. Bushnell, S.M. Azarin, G. Spicer, B.A. Aguado, J.R. Stoehr, E.
J. Jiang, V. Backman, L.D. Shea, J.S. Jeruss, Enhanced survival with implantable
scaffolds that capture metastatic breast cancer cells in vivo, Cancer Res. 76 (18)
(2016) 52095218.
[148] G. Doglioni, S. Parik, S.M. Fendt, Interactions in the (Pre)metastatic niche support
metastasis formation, Front. Oncol. 9 (2019) 219.
[149] R.N. Kaplan, S. Rai, D. Lyden, Preparing the soil: the premetastatic niche,
Cancer Res. 66 (23) (2006) 1108911093.
[150] B.A. Aguado, R.M. Harteld, G.G. Bushnell, J.T. Decker, S.M. Azarin, D. Nanavati,
M.J. Schipma, S.S. Rao, R.S. Oakes, Y. Zhang, J.S. Jeruss, L.D. Shea, Biomaterial
scaffolds as pre-metastatic niche mimics systemically alter the primary tumor and
tumor microenvironment, Adv. Healthc. Mater. 7 (10) (2018), 1700903
e1700903-e1700903.
[151] D.G. Leach, S. Young, J.D. Hartgerink, Advances in immunotherapy delivery from
implantable and injectable biomaterials, Acta Biomater. 88 (2019) 1531.
[152] G.G. Bushnell, T.P. Hardas, R.M. Harteld, Y. Zhang, R.S. Oakes, S. Ronquist,
H. Chen, I. Rajapakse, M.S. Wicha, J.S. Jeruss, L.D. Shea, Biomaterial scaffolds
recruit an aggressive population of metastatic tumor cells in vivo, Cancer Res. 79
(8) (2019) 20422053.
[153] Y. Wang, Z.G. Liu, H. Yuan, W. Deng, J. Li, Y. Huang, B. Kim, M.D. Story,
W. Jiang, The reciprocity between radiotherapy and cancer immunotherapy, Clin.
Cancer Res. 25 (6) (2019) 17091717.
[154] M.R. Egyud, J.F. Tseng, K. Suzuki, Multidisciplinary therapy of esophageal
cancer, Surg. Clin. N. Am. 99 (3) (2019) 419437.
[155] N. Chaicharoenaudomrung, P. Kunhorm, P. Noisa, Three-dimensional cell culture
systems as an in vitro platform for cancer and stem cell modeling, World J. Stem
Cells 11 (12) (2019) 10651083.
[156] C. Jensen, Y. Teng, Is it time to start transitioning from 2D to 3D cell culture?
Front. Mol. Biosci. 7 (2020) 33.
[157] Y. Zhang, F. Huang, C. Ren, J. Liu, L. Yang, S. Chen, J. Chang, C. Yang, W. Wang,
C. Zhang, Q. Liu, X.J. Liang, J. Liu, Enhanced radiosensitization by gold
nanoparticles with acid-triggered aggregation in cancer radiotherapy, Adv. Sci. 6
(8) (2019), 1801806.
[158] B.T. Emmer, D. Ginsburg, Genome editing and hematologic malignancy, Annu.
Rev. Med. 71 (1) (2020) 7183.
[159] A.L. David, S.N. Waddington, Candidate diseases for prenatal gene therapy, in:
C. Coutelle, S.N. Waddington (Eds.), Prenatal Gene Therapy: Concepts, Methods,
and Protocols, Humana Press, Totowa, NJ, 2012, pp. 939.
[160] C.C. Ma, Z.L. Wang, T. Xu, Z.Y. He, Y.Q. Wei, The approved gene therapy drugs
worldwide: from 1998 to 2019, Biotechnol. Adv. 40 (2020), 107502.
[161] B.S. Yilmaz, S. Gurung, D. Perocheau, J. Counsell, J. Baruteau, Gene therapy for
inherited metabolic diseases, J. Mother Child 24 (2) (2020) 5364.
[162] F. Arabi, M. Torabi-Rahvar, A. Shariati, N. Ahmadbeigi, M. Naderi, Antigenic
targets of CAR T cell therapy. A retrospective view on clinical trials, Exp. Cell Res.
369 (1) (2018) 110.
[163] C.A. Stein, D. Castanotto, FDA-approved oligonucleotide therapies in 2017, Mol.
Ther. 25 (5) (2017) 10691075.
[164] W. Guo, H. Song, Development of gene therapeutics for head and neck cancer in
china: from bench to bedside, Hum. Gene Ther. 29 (2) (2018) 180187.
[165] S. Pearson, H. Jia, K. Kandachi, China approves rst gene therapy, Nature
Publishing Group, 2004.
[166] E.M. Gordon, F.L. Hall, The ‘timelydevelopment of Rexin-G: rst targeted
injectable gene vector, Int. J. Oncol. 35 (2) (2009) 229238.
[167] C. Willyard, Limb-saving medicines sought to prevent amputations, Nature
Publishing Group, 2012.
[168] S. Yl¨
a-Herttuala, Endgame: glybera nally recommended for approval as the rst
gene therapy drug in the European union, Mol. Ther. 20 (10) (2012) 18311832.
[169] A. Poh, First oncolytic viral therapy for melanoma, Cancer Discov. 6 (1) (2016) 6.
[170] P.A. Ott, F.S. Hodi, Talimogene laherparepvec for the treatment of advanced
melanoma, Clin. Cancer Res. 22 (13) (2016) 31273131.
[171] A.M. Farkas, S. Mariz, V. Stoyanova-Beninska, P. Celis, S. Vamvakas, K. Larsson,
B. Sepodes, Advanced therapy medicinal products for rare diseases: state of play
of incentives supporting development in Europe, Front. Med. 4 (2017) 53.
[172] J. Hoggatt, Gene therapy for bubble boydisease, Cell 166 (2) (2016) 263.
[173] J. Schimmer, S. Breazzano, Investor outlook: rising from the ashes; GSKs
European Approval of Strimvelis for ADA-SCID, Hum. Gene Ther. Clin. Dev. 27
(2) (2016) 5761.
[174] H. Stirnadel-Farrant, M. Kudari, N. Garman, J. Imrie, B. Chopra, S. Giannelli,
M. Gabaldo, A. Corti, S. Zancan, A. Aiuti, M.P. Cicalese, R. Batta, J. Appleby,
M. Davinelli, P. Ng, Gene therapy in rare diseases: the benets and challenges of
developing a patient-centric registry for Strimvelis in ADA-SCID, Orphanet J. Rare
Dis. 13 (1) (2018) 49.
[175] E.W. Ottesen, ISS-N1 makes the rst FDA-approved drug for spinal muscular
atrophy, Transl. Neurosci. 8 (1) (2017) 16.
[176] A. Aartsma-Rus, N. Goemans, A sequel to the eteplirsen saga: eteplirsen is
approved in the United States but was not approved in Europe, Nucleic Acid.
Ther. 29 (1) (2019) 1315.
[177] R. Korinthenberg, A new era in the management of Duchenne muscular
dystrophy, Dev. Med. Child Neurol. 61 (3) (2019) 292297.
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
19
[178] S.L. Maude, T.W. Laetsch, J. Buechner, S. Rives, M. Boyer, H. Bittencourt,
P. Bader, M.R. Verneris, H.E. Stefanski, G.D. Myers, M. Qayed, B. De Moerloose,
H. Hiramatsu, K. Schlis, K.L. Davis, P.L. Martin, E.R. Nemecek, G.A. Yanik,
C. Peters, A. Baruchel, N. Boissel, F. Mechinaud, A. Balduzzi, J. Krueger, C.
H. June, B.L. Levine, P. Wood, T. Taran, M. Leung, K.T. Mueller, Y. Zhang, K. Sen,
D. Lebwohl, M.A. Pulsipher, S.A. Grupp, Tisagenlecleucel in children and young
adults with B-cell lymphoblastic leukemia, N. Engl. J. Med. 378 (5) (2018)
439448.
[179] S.S. Neelapu, F.L. Locke, N.L. Bartlett, L.J. Lekakis, D.B. Miklos, C.A. Jacobson,
I. Braunschweig, O.O. Oluwole, T. Siddiqi, Y. Lin, J.M. Timmerman, P.J. Stiff, J.
W. Friedberg, I.W. Flinn, A. Goy, B.T. Hill, M.R. Smith, A. Deol, U. Farooq,
P. McSweeney, J. Munoz, I. Avivi, J.E. Castro, J.R. Westin, J.C. Chavez,
A. Ghobadi, K.V. Komanduri, R. Levy, E.D. Jacobsen, T.E. Witzig, P. Reagan,
A. Bot, J. Rossi, L. Navale, Y. Jiang, J. Aycock, M. Elias, D. Chang, J. Wiezorek, W.
Y. Go, Axicabtagene ciloleucel CAR T-cell therapy in refractory large B-cell
lymphoma, N. Engl. J. Med. 377 (26) (2017) 25312544.
[180] S. Russell, J. Bennett, J.A. Wellman, D.C. Chung, Z.F. Yu, A. Tillman, J. Wittes,
J. Pappas, O. Elci, S. McCague, D. Cross, K.A. Marshall, J. Walshire, T.L. Kehoe,
H. Reichert, M. Davis, L. Rafni, L.A. George, F.P. Hudson, L. Dingeld, X. Zhu, J.
A. Haller, E.H. Sohn, V.B. Mahajan, W. Pfeifer, M. Weckmann, C. Johnson,
D. Gewaily, A. Drack, E. Stone, K. Wachtel, F. Simonelli, B.P. Leroy, J.F. Wright,
K.A. High, A.M. Maguire, Efcacy and safety of voretigene neparvovec (AAV2-
hRPE65v2) in patients with RPE65-mediated inherited retinal dystrophy: a
randomised, controlled, open-label, phase 3 trial, Lancet 390 (10097) (2017)
849860.
[181] C.H. Evans, S.C. Ghivizzani, P.D. Robbins, Arthritis gene therapy is becoming a
reality, Nat. Rev. Rheumatol. 14 (7) (2018) 381382.
[182] C.H. Evans, S.C. Ghivizzani, P.D. Robbins, Arthritis gene therapy approved in
Korea, J. Am. Acad. Orthop. Surg. 26 (2) (2018) e36e38.
[183] H. Suda, A. Murakami, T. Kaga, H. Tomioka, R. Morishita, Beperminogene
perplasmid for the treatment of critical limb ischemia, Expert Rev. Cardiovasc
Ther. 12 (10) (2014) 11451156.
[184] S.M. Hoy, Onasemnogene abeparvovec: rst global approval, Drugs 79 (11)
(2019) 12551262.
[185] J. Paik, S. Duggan, Volanesorsen: rst global approval, Drugs 79 (12) (2019)
13491354.
[186] Q. Jin, G. Liu, S. Li, H. Yuan, Z. Yun, W. Zhang, S. Zhang, Y. Dai, Y. Ma,
Decellularized breast matrix as bioactive microenvironment for in vitro three-
dimensional cancer culture, J. Cell. Physiol. 234 (4) (2019) 34253435.
[187] Y.A. Heo, Golodirsen: rst approval, Drugs 80 (3) (2020) 329333.
[188] https://www.fda.gov/vaccines-blood-biologics/cellular-gene-therapy-produc
ts/tecartus-brexucabtagene-autoleucel.
[189] https://www.fda.gov/vaccines-blood-biologics/abecma-idecabtagene-vicleucel.
[190] https://www.fda.gov/vaccines-blood-biologics/cellular-gene-therapy-products/b
reyanzi-lisocabtagene-maraleucel.
[191] L. Zaulyanov, R.S. Kirsner, A review of a bi-layered living cell treatment
(Apligraf) in the treatment of venous leg ulcers and diabetic foot ulcers, Clin.
Interv. Aging 2 (1) (2007) 9398.
[192] C. Schmidt, FDA approves rst cell therapy for wrinkle-free visage, Nat.
Biotechnol. 29 (8) (2011) 674675.
[193] J.L. Carey, A.E. Remmers, D.C. Flanigan, Use of MACI (autologous cultured
chondrocytes on porcine collagen membrane) in the United States: preliminary
experience, Orthop. J. Sports Med. 8 (8) (2020), 2325967120941816,
2325967120941816.
[194] C. Schmidt, Gintuit cell therapy approval signals shift at US regulator, Nat.
Biotechnol. 30 (6) (2012), 479.
[195] https://www.fda.gov/downloads/BiologicsBloodVaccines/CellularGeneTherapy
Products/ApprovedProducts/UCM519084.pdf.
[196] https://www.fda.gov/downloads/BiologicsBloodVaccines/CellularGeneTherapy
Products/ApprovedProducts/UCM279612.pdf.
[197] https://www.fda.gov/BiologicsBloodVaccines/CellularGeneTherapyProducts
/ApprovedProducts/ucm322728.htm.
[198] https://www.fda.gov/BiologicsBloodVaccines/CellularGeneTherapyProducts
/ApprovedProducts/ucm483851.htm.
[199] https://www.fda.gov/BiologicsBloodVaccines/CellularGeneTherapyProducts
/ApprovedProducts/ucm356850.htm.
[200] https://www.fda.gov/downloads/BiologicsBloodVaccines/CellularGeneTherapy
Products/ApprovedProducts/UCM305761.pdf.
[201] https://www.fda.gov/media/114119/download.
[202] https://www.fda.gov/downloads/BiologicsBloodVaccines/CellularGeneTherapy
Products/ApprovedProducts/UCM354696.pdf.
[203] https://www.fda.gov/BiologicsBloodVaccines/CellularGeneTherapyProducts
/ApprovedProducts/ucm210012.htm.
[204] G. Sampogna, S.Y. Guraya, A. Forgione, Regenerative medicine: historical roots
and potential strategies in modern medicine, J. Microsc. Ultrastruct. 3 (3) (2015)
101107.
[205] F. Berthiaume, T.J. Maguire, M.L. Yarmush, Tissue engineering and regenerative
medicine: history, progress, and challenges, Annu. Rev. Chem. Biomol. Eng. 2
(2011) 403430.
[206] M. Uchiyama, Regulatory science, PDA J. Pharm. Sci. Technol. 49 (4) (1995)
185187.
[207] T. Owaki, T. Shimizu, M. Yamato, T. Okano, Cell sheet engineering for
regenerative medicine: current challenges and strategies, Biotechnol. J. 9 (7)
(2014) 904914.
[208] N. Amariglio, A. Hirshberg, B.W. Scheithauer, Y. Cohen, R. Loewenthal,
L. Trakhtenbrot, N. Paz, M. Koren-Michowitz, D. Waldman, L. Leider-Trejo,
A. Toren, S. Constantini, G. Rechavi, Donor-derived brain tumor following neural
stem cell transplantation in an ataxia telangiectasia patient, PLOS Med. 6 (2)
(2009), 1000029.
[209] A.E. Karnoub, A.B. Dash, A.P. Vo, A. Sullivan, M.W. Brooks, G.W. Bell, A.
L. Richardson, K. Polyak, R. Tubo, R.A. Weinberg, Mesenchymal stem cells within
tumour stroma promote breast cancer metastasis, Nature 449 (7162) (2007)
557563.
[210] G.V. Røsland, A. Svendsen, A. Torsvik, E. Sobala, E. McCormack, H. Immervoll,
J. Mysliwietz, J.C. Tonn, R. Goldbrunner, P.E. Lønning, R. Bjerkvig, C. Schichor,
Long-term cultures of bone marrowderived human mesenchymal stem cells
frequently undergo spontaneous malignant transformation, Cancer Res. 69 (13)
(2009) 53315339.
[211] M.E. Bernardo, N. Zaffaroni, F. Novara, A.M. Cometa, M.A. Avanzini, A. Moretta,
D. Montagna, R. Maccario, R. Villa, M.G. Daidone, O. Zuffardi, F. Locatelli,
Human bone marrowderived mesenchymal stem cells do not undergo
transformation after long-term in vitro culture and do not exhibit telomere
maintenance mechanisms, Cancer Res. 67 (19) (2007) 91429149.
[212] C.-L. Zhang, T. Huang, B.-L. Wu, W.-X. He, D. Liu, Stem cells in cancer therapy:
opportunities and challenges, Oncotarget 8 (43) (2017) 7575675766.
[213] M. Malecki, ‘Above all, do no harm: safeguarding pluripotent stem cell therapy
against iatrogenic tumorigenesis, Stem Cell Res. Ther. 5 (3) (2014) 73.
[214] B. Arjmand, M. Sarvari, S. Alavi-Moghadam, M. Payab, P. Goodarzi, et al.,
Prospect of stem cell therapy and regenerative medicine in osteoporosis, Front.
Endocrinol. 11 (2020) 430.
[215] J.M. Robillard, D. Roskams-Edris, B. Kuzeljevic, J. Illes, Prevailing public
perceptions of the ethics of gene therapy, Hum. Gene Ther. 25 (8) (2014)
740746.
[216] S.E. Raper, N. Chirmule, F.S. Lee, N.A. Wivel, A. Bagg, G.P. Gao, J.M. Wilson, M.
L. Batshaw, Fatal systemic inammatory response syndrome in a ornithine
transcarbamylase decient patient following adenoviral gene transfer, Mol.
Genet. Metab. 80 (1) (2003) 148158.
[217] S. Hacein-Bey-Abina, A. Garrigue, G.P. Wang, J. Soulier, A. Lim, E. Morillon,
E. Clappier, L. Caccavelli, E. Delabesse, K. Beldjord, V. Asna, E. MacIntyre, L. Dal
Cortivo, I. Radford, N. Brousse, F. Sigaux, D. Moshous, J. Hauer, A. Borkhardt, B.
H. Belohradsky, U. Wintergerst, M.C. Velez, L. Leiva, R. Sorensen, N. Wulffraat,
S. Blanche, F.D. Bushman, A. Fischer, M. Cavazzana-Calvo, Insertional
oncogenesis in 4 patients after retrovirus-mediated gene therapy of SCID-X1,
J. Clin. Investig. 118 (9) (2008) 31323142.
[218] S.J. Howe, M.R. Mansour, K. Schwarzwaelder, C. Bartholomae, M. Hubank,
H. Kempski, M.H. Brugman, K. Pike-Overzet, S.J. Chatters, D. de Ridder, K.
C. Gilmour, S. Adams, S.I. Thornhill, K.L. Parsley, F.J. Staal, R.E. Gale, D.C. Linch,
J. Bayford, L. Brown, M. Quaye, C. Kinnon, P. Ancliff, D.K. Webb, M. Schmidt,
C. von Kalle, H.B. Gaspar, A.J. Thrasher, Insertional mutagenesis combined with
acquired somatic mutations causes leukemogenesis following gene therapy of
SCID-X1 patients, J. Clin. Investig. 118 (9) (2008) 31433150.
[219] A. Jaff´
e, S.A. Prasad, V. Larcher, S. Hart, Gene therapy for children with cystic
brosiswho has the right to choose? J. Med. Ethics 32 (6) (2006) 361364.
[220] J. Delhove, I. Osenk, I. Prichard, M. Donnelley, Public acceptability of gene
therapy and gene editing for human use: a systematic review, Hum. Gene Ther. 31
(12) (2019) 2046.
[221] J. Hudson, M. Orviska, European attitudes to gene therapy and
pharmacogenetics, Drug Discov. Today 16 (19) (2011) 843847.
[222] Regenerative, T. and Expert, M., Building on our own potential: a UK pathway for
regenerative medicine (A report from the Regenerative Medicine Expert Group).
[223] J. Gardner, A. Faulkner, A. Mahalatchimy, A. Webster, Are there specic
translational challenges in regenerative medicine? Lessons from other elds,
Regen. Med. 10 (7) (2015) 885895.
[224] B.-R. Yi, M.-A. Park, H.-R. Lee, N.-H. Kang, K.J. Choi, S.U. Kim, K.C. Choi,
Suppression of the growth of human colorectal cancer cells by therapeutic stem
cells expressing cytosine deaminase and interferon-β via their tumor-tropic effect
in cellular and xenograft mouse models, Mol. Oncol. 7 (3) (2013) 543554.
[225] S.P. Zielske, D.L. Livant, T.S. Lawrence, Radiation increases invasion of gene-
modied mesenchymal stem cells into tumors, Int. J. Radiat. Oncol. Biol. Phys. 75
(3) (2009) 843853.
[226] S.M. Kim, J.S. Woo, C.H. Jeong, C.H. Ryu, J.-D. Jang, S.S. Jeun, Potential
application of temozolomide in mesenchymal stem cell-based TRAIL gene therapy
against malignant glioma, Stem Cells Transl. Med. 3 (2) (2014) 172182.
[227] L.A. Couture, Scalable pluripotent stem cell culture, Nat. Biotechnol. 28 (6)
(2010) 562563.
[228] L. Foley, M. Whitaker, Concise review: cell therapies: the route to widespread
adoption, Stem Cells Transl. Med. 1 (5) (2012) 438447.
[229] J. Barthes, H. ¨
Ozçelik, M. Hindi´
e, A. Ndreu-Halili, A. Hasan, N.E. Vrana, Cell
microenvironment engineering and monitoring for tissue engineering and
regenerative medicine: the recent advances, BioMed Res. Int. 2014 (2014),
921905.
[230] J.G. Clohessy, P.P. Pandol, Mouse hospital and co-clinical trial projectfrom
bench to bedside, Nat. Rev. Clin. Oncol. 12 (8) (2015) 491498.
[231] D. Armstrong, R. Lilford, J. Ogden, S. Wessely, Health-related quality of life and
the transformation of symptoms, Sociol. Health Illn. 29 (4) (2007) 570583.
[232] H. Youse, J. Yuan, M. Keshavarz-Fathi, J.F. Murphy, N. Rezaei, Immunotherapy
of cancers comes of age, Expert Rev. Clin. Immunol. 13 (10) (2017) 10011015.
[233] O.A. Ali, E. Doherty, W.J. Bell, T. Fradet, J. Hudak, M.T. Laliberte, D.J. Mooney,
D.F. Emerich, Biomaterial-based vaccine induces regression of established
intracranial glioma in rats, Pharm. Res. 28 (5) (2011) 10741080.
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
20
[234] O.A. Ali, E. Doherty, W.J. Bell, T. Fradet, J. Hudak, M.T. Laliberte, D.J. Mooney,
D.F. Emerich, The efcacy of intracranial PLG-based vaccines is dependent on
direct implantation into brain tissue, J. Control. Release 154 (3) (2011) 249257.
[235] O.A. Ali, N. Huebsch, L. Cao, G. Dranoff, D.J. Mooney, Infection-mimicking
materials to program dendritic cells in situ, Nat. Mater. 8 (2) (2009) 151158.
[236] O.A. Ali, P. Tayalia, D. Shvartsman, S. Lewin, D.J. Mooney, Inammatory
cytokines presented from polymer matrices differentially generate and activate
DCs in situ, Adv. Funct. Mater. 23 (36) (2013) 46214628.
[237] O.A. Ali, E. Doherty, D.J. Mooney, D. Emerich, Relationship of vaccine efcacy to
the kinetics of DC and T-cell responses induced by PLG-based cancer vaccines,
Biomatter 1 (1) (2011) 6675.
[238] S.A. Bencherif, R. Warren Sands, O.A. Ali, W.A. Li, S.A. Lewin, T.M. Braschler, T.
Y. Shih, C.S. Verbeke, D. Bhatta, G. Dranoff, D.J. Mooney, Injectable cryogel-
based whole-cell cancer vaccines, Nat. Commun. 6 (1) (2015) 113.
[239] T.Y. Shih, S.O. Blacklow, A.W. Li, B.R. Freedman, S. Bencherif, S.T. Koshy, M.
C. Darnell, D.J. Mooney, Injectable, tough alginate cryogels as cancer vaccines,
Adv. Healthc. Mater. 7 (10) (2018), 1701469.
[240] O.A. Ali, C. Verbeke, C. Johnson, R.W. Sands, S.A. Lewin, D. White, E. Doherty,
G. Dranoff, D.J. Mooney, Identication of immune factors regulating antitumor
immunity using polymeric vaccines with multiple adjuvants, Cancer Res. 74 (6)
(2014) 16701681.
[241] O.A. Ali, S.A. Lewin, G. Dranoff, D.J. Mooney, Vaccines combined with immune
checkpoint antibodies promote cytotoxic T-cell activity and tumor eradication,
Cancer Immunol. Res. 4 (2) (2016) 95100.
[242] A.W. Li, M.C. Sobral, S. Badrinath, Y. Choi, A. Graveline, A.G. Stafford, J.
C. Weaver, M.O. Dellacherie, T.Y. Shih, O.A. Ali, J. Kim, K.W. Wucherpfennig, D.
J. Mooney, A facile approach to enhance antigen response for personalized cancer
vaccination, Nat. Mater. 17 (6) (2018) 528534.
[243] C. Palena, S.I. Abrams, J. Schlom, J.W. Hodge, Cancer vaccines: preclinical
studies and novel strategies, Adv. Cancer Res. 95 (2006) 115145.
[244] Y. Li, H. Zeng, R.H. Xu, B. Liu, Z. Li, Vaccination with human pluripotent stem
cells generates a broad spectrum of immunological and clinical responses against
colon cancer, Stem Cells 27 (12) (2009) 31033111.
[245] W. Dong, J. Du, H. Shen, D. Gao, Z. Li, G. Wang, X. Mu, Q. Liu, Administration of
embryonic stem cells generates effective antitumor immunity in mice with minor
and heavy tumor load, Cancer Immunol. Immunother. 59 (11) (2010)
16971705.
[246] Z. Zhang, X. Chen, X. Chang, X. Ye, Y. Li, H. Cui, Vaccination with embryonic
stem cells generates effective antitumor immunity against ovarian cancer, Int. J.
Mol. Med. 31 (1) (2012) 147153.
[247] K. Yaddanapudi, R.A. Mitchell, K. Putty, S. Willer, R.K. Sharma, J. Yan,
H. Bodduluri, J.W. Eaton, Vaccination with embryonic stem cells protects against
lung cancer: is a broad-spectrum prophylactic vaccine against cancer possible?
PLOS One 7 (7) (2012), 42289.
[248] K. Yaddanapudi, S. Meng, A.G. Whitt, N. Al Rayyan, J. Richie, A. Tu, J.W. Eaton,
C. Li, Exosomes from GM-CSF expressing embryonic stem cells are an effective
prophylactic vaccine for cancer prevention, Oncoimmunology 8 (3) (2019),
1561119.
[249] J. Li, J. Chen, X. Li, Y. Qian, Vaccination efcacy with marrow mesenchymal stem
cell against cancer was enhanced under simulated microgravity, Biochem.
Biophys. Res. Commun. 485 (3) (2017) 606613.
[250] H. Mashima, R. Zhang, T. Kobayashi, Y. Hagiya, H. Tsukamoto, T. Liu, T. Iwama,
M. Yamamoto, C. Lin, R. Nakatsuka, Y. Mishima, N. Watanabe, T. Yamada,
S. Senju, S. Kaneko, A. Idiris, T. Nakatsura, H. Ohdan, Y. Uemura, Generation of
GM-CSF-producing antigen-presenting cells that induce a cytotoxic T cell-
mediated antitumor response, OncoImmunology 9 (1) (2020), 1814620.
[251] J. Wang, L. Shao, L. Wu, W. Ma, Y. Zheng, C. Hu, F. Li, Expression levels of a gene
signature in hiPSC associated with lung adenocarcinoma stem cells and its
capability in eliciting specic antitumor immune-response in a humanized mice
model, Thorac. Cancer 11 (6) (2020) 16031612.
[252] A. Malandrino, M. Mak, R.D. Kamm, E. Moeendarbary, Complex mechanics of the
heterogeneous extracellular matrix in cancer, Extrem. Mech. Lett. 21 (2018)
2534.
[253] G. Ciasca, M. Papi, E. Minelli, V. Palmieri, M. De Spirito, Changes in cellular
mechanical properties during onset or progression of colorectal cancer, World J.
Gastroenterol. 22 (32) (2016) 72037214.
[254] S. Kawano, M. Kojima, Y. Higuchi, M. Sugimoto, K. Ikeda, N. Sakuyama,
S. Takahashi, R. Hayashi, A. Ochiai, N. Saito, Assessment of elasticity of colorectal
cancer tissue, clinical utility, pathological and phenotypical relevance, Cancer
Sci. 106 (9) (2015) 12321239.
[255] P. Costa, T.M. Scales, J. Ivaska, M. Parsons, Integrin-specic control of focal
adhesion kinase and RhoA regulates membrane protrusion and invasion, PLOS
One 8 (9) (2013), 74659.
[256] C.T. Mierke, The integrin alphav beta3 increases cellular stiffness and cytoskeletal
remodeling dynamics to facilitate cancer cell invasion, New J. Phys. 15 (1)
(2013), 015003.
[257] W. Tao, N. Kong, X. Ji, Y. Zhang, A. Sharma, J. Ouyang, B. Qi, J. Wang, N. Xie,
C. Kang, H. Zhang, O.C. Farokhzad, J.S. Kim, Emerging two-dimensional
monoelemental materials (Xenes) for biomedical applications, Chem. Soc. Rev. 48
(11) (2019) 28912912.
[258] R. Kalluri, R.A. Weinberg, The basics of epithelial-mesenchymal transition,
J. Clin. Investig. 119 (6) (2009) 14201428.
[259] A.J. Rice, E. Cortes, D. Lachowski, B. Cheung, S.A. Karim, J.P. Morton, A. Del Río
Hern´
andez, Matrix stiffness induces epithelialmesenchymal transition and
promotes chemoresistance in pancreatic cancer cells, Oncogenesis 6 (7) (2017),
352.
[260] L.G. Coffman, A.T. Pearson, L.G. Frisbie, Z. Freeman, E. Christie, D.D. Bowtell, R.
J. Buckanovich, Ovarian carcinoma-associated mesenchymal stem cells arise from
tissue-specic normal stroma, Stem Cells 37 (2) (2019) 257269.
[261] S. Liu, C. Ginestier, S.J. Ou, S.G. Clouthier, S.H. Patel, F. Monville, H. Korkaya,
A. Heath, J. Dutcher, C.G. Kleer, Y. Jung, G. Dontu, R. Taichman, M.S. Wicha,
Breast cancer stem cells are regulated by mesenchymal stem cells through
cytokine networks, Cancer Res. 71 (2) (2011) 614624.
[262] L. Lin, L. Du, The role of secreted factors in stem cells-mediated immune
regulation, Cell. Immunol. 326 (2018) 2432.
[263] S.A. Patel, J.R. Meyer, S.J. Greco, K.E. Corcoran, M. Bryan, P. Rameshwar,
Mesenchymal stem cells protect breast cancer cells through regulatory T cells:
role of mesenchymal stem cell-derived TGF-β, J. Immunol. 184 (10) (2010)
58855894.
[264] M. Ono, N. Kosaka, N. Tominaga, Y. Yoshioka, F. Takeshita, R. Takahashi,
M. Yoshida, H. Tsuda, K. Tamura, T. Ochiya, Exosomes from bone marrow
mesenchymal stem cells contain a microRNA that promotes dormancy in
metastatic breast cancer cells, Sci. Signal. 7 (332) (2014) ra63.
[265] S.A. Bliss, G. Sinha, O.A. Sandiford, L.M. Williams, D.J. Engelberth, K. Guiro, L.
L. Isenalumhe, S.J. Greco, S. Ayer, M. Bryan, R. Kumar, N.M. Ponzio,
P. Rameshwar, Mesenchymal stem cellderived exosomes stimulate cycling
quiescence and early breast cancer dormancy in bone marrow, Cancer Res. 76
(19) (2016) 58325844.
[266] J.-K. Lee, S.-R. Park, B.-K. Jung, Y.-K. Jeon, Y.-S. Lee, M.K. Kim, Y.G. Kim, J.
Y. Jang, C.W. Kim, Exosomes derived from mesenchymal stem cells suppress
angiogenesis by down-regulating VEGF expression in breast cancer cells, PLOS
One 8 (12) (2013), 84256.
[267] K. Pakravan, S. Babashah, M. Sadeghizadeh, S.J. Mowla, M. Mossahebi-
Mohammadi, F. Ataei, N. Dana, M. Javan, MicroRNA-100 shuttled by
mesenchymal stem cell-derived exosomes suppresses in vitro angiogenesis
through modulating the mTOR/HIF-1
α
/VEGF signaling axis in breast cancer cells,
Cell. Oncol. 40 (5) (2017) 457470.
[268] R. Lin, S. Wang, R.C. Zhao, Exosomes from human adipose-derived mesenchymal
stem cells promote migration through Wnt signaling pathway in a breast cancer
cell model, Mol. Cell. Biochem. 383 (12) (2013) 1320.
[269] J.A. Cho, H. Park, E.H. Lim, K.W. Lee, Exosomes from breast cancer cells can
convert adipose tissue-derived mesenchymal stem cells into myobroblast-like
cells, Int. J. Oncol. 40 (1) (2011) 130138.
[270] W. Wang, W. Zhong, J. Yuan, C. Yan, S. Hu, Y. Tong, Y. Mao, T. Hu, B. Zhang,
G. Song, Involvement of Wnt/β-catenin signaling in the mesenchymal stem cells
promote metastatic growth and chemoresistance of cholangiocarcinoma,
Oncotarget 6 (39) (2015) 4227642289.
[271] W. Zhong, Y. Tong, Y. Li, J. Yuan, S. Hu, T. Hu, G. Song, Mesenchymal stem cells
in inammatory microenvironment potently promote metastatic growth of
cholangiocarcinoma via activating Akt/NF-κB signaling by paracrine CCL5,
Oncotarget 8 (43) (2017) 7369373704.
[272] L.V. Rhodes, S.E. Muir, S. Elliott, L.M. Guillot, J.W. Antoon, P. Penfornis, S.
L. Tilghman, V.A. Salvo, J.P. Fonseca, M.R. Lacey, B.S. Beckman, J.A. McLachlan,
B.G. Rowan, R. Pochampally, M.E. Burow, Adult human mesenchymal stem cells
enhance breast tumorigenesis and promote hormone independence, Breast Cancer
Res. Treat. 121 (2) (2010) 293300.
[273] S.-C. Hung, W.-P. Deng, W.K. Yang, R.-S. Liu, C.-C. Lee, T.C. Su, R.J. Lin, D.
M. Yang, C.W. Chang, W.H. Chen, H.J. Wei, J.G. Gelovani, Mesenchymal stem cell
targeting of microscopic tumors and tumor stroma development monitored by
noninvasive in vivo positron emission tomography imaging, Clin. Cancer Res. 11
(21) (2005) 77497756.
[274] N. Serakinci, R. Christensen, U. Fahrioglu, F.B. Sorensen, F. Dagnæs-Hansen,
M. Hajek, T.H. Jensen, S. Kolvraa, N.W. Keith, Mesenchymal stem cells as
therapeutic delivery vehicles targeting tumor stroma, Cancer Biotherapy
Radiopharm. 26 (6) (2011) 767773.
[275] R.M. Dwyer, S.M. Potter-Beirne, K.A. Harrington, A.J. Lowery, E. Hennessy, J.
M. Murphy, F.P. Barry, T. OBrien, M.J. Kerin, Monocyte chemotactic protein-1
secreted by primary breast tumors stimulates migration of mesenchymal stem
cells, Clin. Cancer Res. 13 (17) (2007) 50205027.
[276] Y. Jung, J.K. Kim, Y. Shiozawa, J. Wang, A. Mishra, J. Joseph, J.E. Berry,
S. McGee, E. Lee, H. Sun, J. Wang, T. Jin, H. Zhang, J. Dai, P.H. Krebsbach, E.
T. Keller, K.J. Pienta, R.S. Taichman, Recruitment of mesenchymal stem cells into
prostate tumours promotes metastasis, Nat. Commun. 4 (1) (2013) 111.
[277] J.A. Cho, H. Park, E.H. Lim, K.H. Kim, J.S. Choi, J.H. Lee, J.W. Shin, K.W. Lee,
Exosomes from ovarian cancer cells induce adipose tissue-derived mesenchymal
stem cells to acquire the physical and functional characteristics of tumor-
supporting myobroblasts, Gynecol. Oncol. 123 (2) (2011) 379386.
[278] R. Lis, C. Touboul, P. Mirshahi, F. Ali, S. Mathew, D.J. Nolan, M. Maleki, S.
A. Abdalla, C.M. Raynaud, D. Querleu, E. Al-Azwani, J. Malek, M. Mirshahi,
A. Rai, Tumor associated mesenchymal stem cells protects ovarian cancer cells
from hyperthermia through CXCL12, Int. J. Cancer 128 (3) (2011) 715725.
[279] Y. Chu, H. Tang, Y. Guo, J. Guo, B. Huang, F. Fang, J. Cai, Z. Wang, Adipose-
derived mesenchymal stem cells promote cell proliferation and invasion of
epithelial ovarian cancer, Exp. Cell Res. 337 (1) (2015) 1627.
[280] T. Gao, Y. Yu, Q. Cong, Y. Wang, M. Sun, L. Yao, C. Xu, W. Jiang, Human
mesenchymal stem cells in the tumour microenvironment promote ovarian cancer
progression: the role of platelet-activating factor, BMC Cancer 18 (1) (2018)
110.
[281] S. Bu, Q. Wang, Q. Zhang, J. Sun, B. He, C. Xiang, Z. Liu, D. Lai, Human
endometrial mesenchymal stem cells exhibit intrinsic anti-tumor properties on
human epithelial ovarian cancer cells, Sci. Rep. 6 (1) (2016) 114.
V. Mansouri et al.
Biomedicine & Pharmacotherapy 141 (2021) 111875
21
[282] A.M.M.T. Reza, Y.-J. Choi, H. Yasuda, J.-H. Kim, Human adipose mesenchymal
stem cell-derived exosomal-miRNAs are critical factors for inducing anti-
proliferation signalling to A2780 and SKOV-3 ovarian cancer cells, Sci. Rep. 6 (1)
(2016) 115.
[283] Y. Chu, M. You, J. Zhang, G. Gao, R. Han, et al., Adipose-derived mesenchymal
stem cells enhance ovarian cancer growth and metastasis by increasing thymosin
beta 4X-linked expression, Stem Cells Int. (2019).
[284] A. Subramanian, G. Shu-Uin, N. Kae-Siang, K. Gauthaman, A. Biswas,
M. Choolani, A. Bongso, F. Chui-Yee, Human umbilical cord whartons jelly
mesenchymal stem cells do not transform to tumor-associated broblasts in the
presence of breast and ovarian cancer cells unlike bone marrow mesenchymal
stem cells, J. Cell. Biochem. 113 (6) (2012) 18861895.
[285] C. Ganta, D. Chiyo, R. Ayuzawa, R. Rachakatla, M. Pyle, G. Andrews, M. Weiss,
M. Tamura, D. Troyer, Rat umbilical cord stem cells completely abolish rat
mammary carcinomas with no evidence of metastasis or recurrence 100 days
posttumor cell inoculation, Cancer Res. 69 (5) (2009) 18151820.
[286] R. Ayuzawa, C. Doi, R.S. Rachakatla, M.M. Pyle, D.K. Maurya, D. Troyer,
M. Tamura, Naive human umbilical cord matrix derived stem cells signicantly
attenuate growth of human breast cancer cells in vitro and in vivo, Cancer Lett.
280 (1) (2009) 3137.
[287] K.S. Tsai, S.H. Yang, Y.P. Lei, C.C. Tsai, H.W. Chen, C.Y. Hsu, L.L. Chen, H.
W. Wang, S.A. Miller, S.H. Chiou, M.C. Hung, S.C. Hung, Mesenchymal stem cells
promote formation of colorectal tumors in mice, Gastroenterology 141 (3) (2011)
10461056.
[288] A. De Boeck, P. Pauwels, K. Hensen, J.-L. Rummens, W. Westbroek, A. Hendrix,
D. Maynard, H. Denys, K. Lambein, G. Braems, C. Gespach, M. Bracke, O. De
Wever, Bone marrow-derived mesenchymal stem cells promote colorectal cancer
progression through paracrine neuregulin 1/HER3 signalling, Gut 62 (4) (2013)
550560.
[289] V. Mele, M.G. Muraro, D. Calabrese, D. Pfaff, N. Amatruda, F. Amicarella,
B. Kvinlaug, C. Bocelli-Tyndall, I. Martin, T.J. Resink, M. Heberer, D. Oertli,
L. Terracciano, G.C. Spagnoli, G. Iezzi, Mesenchymal stromal cells induce
epithelial-to-mesenchymal transition in human colorectal cancer cells through the
expression of surface-bound TGF-β, Int. J. Cancer 134 (11) (2014) 25832594.
[290] J. Wang, Y. Wang, S. Wang, J. Cai, J. Shi, X. Sui, Y. Cao, W. Huang, X. Chen,
Z. Cai, H. Li, A.S. Bardeesi, B. Zhang, M. Liu, W. Song, M. Wang, A.P. Xiang, Bone
marrow-derived mesenchymal stem cell-secreted IL-8 promotes the angiogenesis
and growth of colorectal cancer, Oncotarget 6 (40) (2015) 4282542837.
[291] G. Nishikawa, K. Kawada, J. Nakagawa, K. Toda, R. Ogawa, S. Inamoto,
R. Mizuno, Y. Itatani, Y. Sakai, Bone marrow-derived mesenchymal stem cells
promote colorectal cancer progression via CCR5, Cell Death Dis. 10 (4) (2019)
113.
[292] K. Chen, Q. Liu, L.L. Tsang, Q. Ye, H.C. Chan, Y. Sun, X. Jiang, Human MSCs
promotes colorectal cancer epithelialmesenchymal transition and progression
via CCL5/β-catenin/Slug pathway, Cell Death Dis. 8 (5) (2017) 2819, e2819-
e2819.
[293] X. Zhang, F. Hu, G. Li, G. Li, X. Yang, L. Liu, R. Zhang, B. Zhang, Y. Feng, Human
colorectal cancer-derived mesenchymal stem cells promote colorectal cancer
progression through IL-6/JAK2/STAT3 signaling, Cell Death Dis. 9 (2) (2018)
113.
[294] I.-R. Oh, B. Raymundo, M. Kim, C.-W. Kim, Mesenchymal stem cells co-cultured
with colorectal cancer cells showed increased invasive and proliferative abilities
due to its altered p53/TGF-β1 levels, Biosci. Biotechnol. Biochem. 84 (2) (2020)
256267.
[295] X.L. Yan, Y.L. Jia, L. Chen, Q. Zeng, J.N. Zhou, C.J. Fu, H.X. Chen, H.F. Yuan, Z.
W. Li, L. Shi, Y.C. Xu, J.X. Wang, X.M. Zhang, L.J. He, C. Zhai, W. Yue, X.T. Pei,
Hepatocellular carcinoma-associated mesenchymal stem cells promote
hepatocarcinoma progression: role of the S100A4miR155SOCS1MMP9 axis,
Hepatology 57 (6) (2013) 22742286.
[296] L. Pietrovito, A. Leo, V. Gori, M. Lulli, M. Parri, V. Becherucci, L. Piccini,
F. Bambi, M.L. Taddei, P. Chiarugi, Bone marrow-derived mesenchymal stem cells
promote invasiveness and transendothelial migration of osteosarcoma cells via a
mesenchymal to amoeboid transition, Mol. Oncol. 12 (5) (2018) 659676.
[297] J. Wang, W. Yang, T. Wang, X. Chen, J. Wang, X. Zhang, C. Cai, B. Zhong, J. Wu,
Z. Chen, A.P. Xiang, W. Huang, Mesenchymal stromal cells-derived β2-
microglobulin promotes epithelialmesenchymal transition of esophageal
squamous cell carcinoma cells, Sci. Rep. 8 (1) (2018) 113.
[298] G. Zhang, F. Miao, J. Xu, R. Wang, Mesenchymal stem cells from bone marrow
regulate invasion and drug resistance of multiple myeloma cells by secreting
chemokine CXCL13, Bosn. J. Basic Med. Sci. 20 (2) (2020) 209217.
[299] F.C. de Menezes, L.G. de Sousa Cabral, M.C. Petrellis, C.F. Neto, D.A. Maria,
Antitumor effect of cell therapy with mesenchymal stem cells on murine
melanoma B16-F10, Biomed. Pharmacother. =Biomed. Pharmacother. 128
(2020), 110294.
[300] I. Marangon, A.A. Silva, T. Guilbert, J. Kolosnjaj-Tabi, C. Marchiol,
S. Natkhunarajah, F. Chammings, C. M´
enard-Moyon, A. Bianco, J.L. Gennisson,
G. Renault, F. Gazeau, Tumor stiffening, a key determinant of tumor progression,
is reversed by nanomaterial-induced photothermal therapy, Theranostics 7 (2)
(2017) 329343.
[301] D.A. Cope, M.W. Dewhirst, H.S. Friedman, D.D. Bigner, M.R. Zalutsky, Enhanced
delivery of a monoclonal antibody F (ab) 2 fragment to subcutaneous human
glioma xenografts using local hyperthermia, Cancer Res. 50 (6) (1990)
18031809.
[302] J. Kolosnjaj-Tabi, R. Di Corato, L. Lartigue, I. Marangon, P. Guardia, A.K. Silva,
N. Luciani, O. Cl´
ement, P. Flaud, J.V. Singh, P. Decuzzi, T. Pellegrino, C. Wilhelm,
F. Gazeau, Heat-generating iron oxide nanocubes: subtle destructuratorsof the
tumoral microenvironment, ACS Nano 8 (5) (2014) 42684283.
[303] S.A. Alageel, S.N. Arafat, B. Salvador-Culla, P.E. Kolovou, K. Jahanseir, A. Kozak,
G. Braithwaite, J.B. Ciolino, Corneal cross-linking with verteporn and
nonthermal laser therapy, Cornea 37 (3) (2018) 362368.
[304] H. Tanaka, T. Okada, H. Konishi, T. Tsuji, The effect of reactive oxygen species on
the biosynthesis of collagen and glycosaminoglycans in cultured human dermal
broblasts, Arch. Dermatol. Res. 285 (6) (1993) 352355.
[305] L. Lapˇ
cík, J. Schurz, Photochemical degradation of hyaluronic acid by singlet
oxygen, Colloid Polym. Sci. 269 (6) (1991) 633635.
[306] H.-W. Wang, M.E. Putt, M.J. Emanuele, D.B. Shin, E. Glatstein, A.G. Yodh, T.
M. Busch, Treatment-induced changes in tumor oxygenation predict
photodynamic therapy outcome, Cancer Res. 64 (20) (2004) 75537561.
[307] J.W. Snyder, W.R. Greco, D.A. Bellnier, L. Vaughan, B.W. Henderson,
Photodynamic therapy: a means to enhanced drug delivery to tumors, Cancer Res.
63 (23) (2003) 81268131.
[308] X. Jing, F. Yang, C. Shao, K. Wei, M. Xie, H. Shen, Y. Shu, Role of hypoxia in
cancer therapy by regulating the tumor microenvironment, Mol. Cancer 18 (1)
(2019) 157.
[309] D.C. Belisario, J. Kopecka, M. Pasino, M. Akman, E. De Smaele, M. Donadelli,
C. Riganti, Hypoxia dictates metabolic rewiring of tumors: implications for
chemoresistance, Cells 9 (12) (2020) 2598.
[310] E. Fern´
andez, V. Morillo, M. Salvador, A. Santaf´
e, I. Beato, M. Rodríguez,
C. Ferrer, Hyperbaric oxygen and radiation therapy: a review, Clin. Transl. Oncol.
23 (2020) 10471053.
[311] M. Luongo, A.L. Brigida, L. Mascolo, G. Gaudino, Possible therapeutic effects of
ozone mixture on hypoxia in tumor development, Anticancer Res. 37 (2) (2017)
425435.
[312] R. Hickey, A.F. Palmer, Synthesis of hemoglobin-based oxygen carrier
nanoparticles by desolvation precipitation, Langmuir (2020).
[313] M.P. Krafft, Alleviating tumor hypoxia with peruorocarbon-based oxygen
carriers, Curr. Opin. Pharmacol. 53 (2020) 117125.
[314] A. Wrobeln, J. Laudien, C. Groß-Heitfeld, J. Linders, C. Mayer, B. Wilde, T. Knoll,
D. Naglav, M. Kirsch, K.B. Ferenz, Albumin-derived peruorocarbon-based
articial oxygen carriers: a physico-chemical characterization and rst in vivo
evaluation of biocompatibility, Eur. J. Pharm. Biopharm. 115 (2017) 5264.
[315] B. Newland, M. Baeger, D. Eigel, H. Newland, C. Werner, Oxygen-producing
gellan gum hydrogels for dual delivery of either oxygen or peroxide with
doxorubicin, ACS Biomater. Sci. Eng. 3 (5) (2017) 787792.
[316] S. Xu, X. Zhu, C. Zhang, W. Huang, Y. Zhou, D. Yan, Oxygen and Pt (II) self-
generating conjugate for synergistic photo-chemo therapy of hypoxic tumor, Nat.
Commun. 9 (1) (2018) 19.
[317] D. Wang, B. Xue, T.Y. Ohulchanskyy, Y. Liu, A. Yakovliev, R. Ziniuk, M. Xu,
J. Song, J. Qu, Z. Yuan, Inhibiting tumor oxygen metabolism and simultaneously
generating oxygen by intelligent upconversion nanotherapeutics for enhanced
photodynamic therapy, Biomaterials 251 (2020), 120088.
[318] Z. Yang, J. Wang, S. Ai, J. Sun, X. Mai, W. Guan, Self-generating oxygen enhanced
mitochondrion-targeted photodynamic therapy for tumor treatment with hypoxia
scavenging, Theranostics 9 (23) (2019) 68096823.
[319] M. Li, X. Yue, Y. Wang, J. Zhang, L. Kan, Z. Jing, Remodeling the tumor
microenvironment to improve drug permeation and antitumor effects by co-
delivering quercetin and doxorubicin, J. Mater. Chem. B 7 (47) (2019)
76197626.
[320] K. Hu, L. Miao, T.J. Goodwin, J. Li, Q. Liu, L. Huang, Quercetin remodels the
tumor microenvironment to improve the permeation, retention, and antitumor
effects of nanoparticles, ACS Nano 11 (5) (2017) 49164925.
[321] X. Wang, L. Cheng, H.-J. Xie, R.-J. Ju, Y. Xiao, M. Fu, J.J. Liu, X.T. Li, Functional
paclitaxel plus honokiol micelles destroying tumour metastasis in treatment of
non-small-cell lung cancer, Artif. Cells Nanomed. Biotechnol. 46 (suppl 2)
(2018) 11541169.
[322] R.-J. Ju, L. Cheng, X. Qiu, S. Liu, X.-L. Song, X.M. Peng, T. Wang, C.Q. Li, X.T. Li,
Hyaluronic acid modied daunorubicin plus honokiol cationic liposomes for the
treatment of breast cancer along with the elimination vasculogenic mimicry
channels, J. Drug Target. 26 (9) (2018) 793805.
[323] M. Saberi-Karimian, N. Katsiki, M. Caraglia, M. Boccellino, M. Majeed,
A. Sahebkar, Vascular endothelial growth factor: an important molecular target of
curcumin, Crit. Rev. Food Sci. Nutr. 59 (2) (2019) 299312.
[324] M.-Z. Jin, W.-L. Jin, The updated landscape of tumor microenvironment and drug
repurposing, Signal Transduct. Target. Ther. 5 (1) (2020) 116.
[325] Z. Zhao, A. Ukidve, J. Kim, S. Mitragotri, Targeting strategies for tissue-specic
drug delivery, Cell 181 (1) (2020) 151167.
[326] J.K. Patra, G. Das, L.F. Fraceto, E.V.R. Campos, M. Rodriguez-Torres, L.S. Acosta-
Torres, L.A. Diaz-Torres, R. Grillo, M.K. Swamy, S. Sharma, S. Habtemariam, H.
S. Shin, Nano based drug delivery systems: recent developments and future
prospects, J. Nanobiotechnol. 16 (1) (2018) 71.
V. Mansouri et al.
... In such applications TTE scaffolds act supplementary by promoting cell-cell interactions between tumor cells and cells from the host immune system. CAR-T cells are responsible for the identification of tumor-associated antigens, thus in TTE scaffolds upon interaction with tumor cells CAR-T cells would occupy a co-stimulatory domain for activating host immune T cells [31]. NK cells are utilized in TTE scaffolds in combination with cytokines and transcription factors, to indirectly strengthen the host immune system by activating innate immune functions. ...
... The potential applications of regenerative medicine in cancer therapy include the scientific fields of cell and gene therapy, and tissue engineering [31,36]. Regenerative medicine aims to restore or replace the damaged tissues or organs, and tissue engineering plays a crucial role in achieving this goal. ...
... These challenges include minimizing the risk of infection, increasing host tolerance for reducing the risk of immune rejection, addressing ethical and moral issues for implementation, and evaluating the long term potential tumorigenesis risk of cell therapies. Other potential complications are associated with repeatability and scalability of the protocols, and cost issues related to tumor therapy, immunotherapy and tissue regeneration [31]. ...
Preprint
Full-text available
The evaluation of nanostructured biomaterials and medicines is associated with 2D cultures that provide insight into the biological mechanisms at the molecular level, while critical aspects of the tumor microenvironment (TME) are provided by the study of animal xenograft models. More realistic models that can histologically reproduce human tumors are provided by tissue engineering methodologies of co-culturing cells of varied phenotypes, to provide 3D tumor spheroids that recapitulate the dynamic TME in 3D matrices. The novel approaches of 3D tumor models are combined with tumor tissue engineering (TTE) scaffolds including hydrogels, bioprinted materials, decellularized tissues, fibrous and nanostructured matrices. This review focuses on the use of nanostructured materials in cancer therapy and regeneration, and the development of realistic models for studying TME molecular and immune characteristics. Tissue regeneration is an important aspect of TTE scaffolds for restoring the normal function of the tissues, while providing cancer treatment. Thus, this article reports recent advancements on the development of 3D TTE models for antitumor drug screening, studying of tumor metastasis, and tissue regeneration. Also, this review identifies the great opportunities of 3D TTE scaffolds in the evaluation of the immunological mechanisms and processes involved in the application of immunotherapies.
... Radiation therapy can damage normal cells, leading to side effects such as secondary cancers, blisters, and other skin problems [3][4][5]. Surgical treatment is both risky and traumatic for the patient, and it is also prone to complications [6]. However, due to the unclear etiology of certain tumors, it is difficult to cure completely. ...
Article
Full-text available
Cancer treatment is presently a significant challenge in the medical domain, wherein the primary modalities of intervention include chemotherapy, radiation therapy and surgery. However, these therapeutic modalities carry side effects. Photothermal therapy (PTT) and photodynamic therapy (PDT) have emerged as promising modalities for the treatment of tumors in recent years. Phototherapy is a therapeutic approach that involves the exposure of materials to specific wavelengths of light, which can subsequently be converted into either heat or Reactive Oxygen Species (ROS) to effectively eradicate cancer cells. Due to the hydrophobicity and lack of targeting of many photoresponsive materials, the use of nano-carriers for their transportation has been extensively explored. Among these nanocarriers, liposomes have been identified as an effective drug delivery system due to their controllability and availability in the biomedical field. By binding photoresponsive materials to li-posomes, it is possible to reduce the cytotoxicity of the material and regulate drug release and accumulation at the tumor site. This article provides a comprehensive review of the progress made in cancer therapy using photoresponsive materials loaded onto liposomes. Additionally, the article discusses the potential synergistic treatment through the combination of phototherapy with chemo/immuno/gene therapy using liposomes.
... Globally, cancer remains a leading cause of death, presenting significant obstacles in both its treatment and diagnosis (Bray et al., 2018, Ferlay et al., 2021, Mansouri et al., 2021. Immune-checkpoint blockade therapies, known for their effectiveness in treating various types of cancer, target proteins such as programmed death-1 (PD-1) and programmed death ligand-1 (PD-L1), which are utilized by cancer cells to escape immune detection (Mahoney et al., 2015, Patel andKurzrock, 2015). ...
Article
Full-text available
The ongoing global concern of cancer worldwide necessitates the development of advanced diagnostic and therapeutic strategies. The majority of recent detection strategies involve the employment of biomarkers. A critical biomarker for cancer immunotherapy efficacy and patient prognosis is Programmed Death Ligand 1 (PD-L1), which is a key immune checkpoint protein. PD-L1 can be particularly linked to cancer progression and therapy response. Current detection methods, such as enzyme-linked immunosorbent assay (ELISA), face limitations like high cost, time consumption, and complexity. This study introduces a microcantilever-based biosensor designed for the detection of soluble PD-L1 (sPD-L1), which has a specific association with PD-L1. The biosensor utilizes anti-PD-L1 as the sensing layer, capitalizing on the specific binding affinity between anti-PD-L1 and sPD-L1. The presence of the sensing layer was confirmed through Atomic Force Microscopy (AFM) and contact angle measurements. Binding between sPD-L1 and anti-PD-L1 induces a shift in the microcantilever's resonance frequency, which is proportional to the PD-L1 concentration. Notably, the resonance frequency shift demonstrates a robust linear relationship with the increasing biomarker concentration, ranging from 0.05 ng/ml to 500 ng/ml. The detection limit of the biosensor was determined to be approximately 10 pg/ml. The biosensor demonstrates excellent performance in detecting PD-L1 with high specificity even in complex biological matrices. This innovative approach not only provides a promising tool for early cancer diagnosis but also holds potential for monitoring immunotherapy efficacy, paving the way for personalized and effective cancer treatments.
Article
BACKGROUND Gastrointestinal tumor organoids serve as an effective model for simulating cancer in vitro and have been applied in basic biology and preclinical research. Despite over a decade of development and increasing research achievements in this field, a systematic and comprehensive analysis of the research hotspots and future trends is lacking. AIM To address this problem by employing bibliometric tools to explore the publication years, countries/regions, institutions, journals, authors, keywords, and references in this field. METHODS The literature was collected from Web of Science databases. CiteSpace-6.2R4, a widely used bibliometric analysis software package, was used for institutional analysis and reference burst analysis. VOSviewer 1.6.19 was used for journal co-citation analysis, author co-authorship and co-citation analysis. The ‘online platform for bibliometric analysis (https://bibliometric.com/app )’ was used to assess the total number of publications and the cooperation relationships between countries. Finally, we employed the bibliometric R software package (version R.4.3.1) in R-studio, for a comprehensive scientific analysis of the literature. RESULTS Our analysis included a total of 1466 publications, revealing a significant yearly increase in articles on the study of gastrointestinal tumor organoids. The United States (n = 393) and Helmholtz Association (n = 93) have emerged as the leading countries and institutions, respectively, in this field, with Hans Clevers and Toshiro Sato being the most contributing authors. The most influential journal in this field is Gastroenterology. The most impactful reference is "Long term expansion of epithelial organs from human colon, adenoma, adenocarcinoma, and Barrett's epithelium". Keywords analysis and citation burst analysis indicate that precision medicine, disease modeling, drug development and screening, and regenerative medicine are the most cutting-edge directions. These focal points were further detailed based on the literature. CONCLUSION This bibliometric study offers an objective and quantitative analysis of the research in this field, which can be considered as an important guide for next scientific research.
Chapter
Colorectal cancer is a global health concern with high incidence and mortality rates. Conventional treatments like surgery, chemotherapy, and radiation therapy have limitations in improving patient survival rates. Recent research highlights the role of gut microbiota and intestinal stem cells in maintaining intestinal health and their potential therapeutic applications in colorectal cancer treatment. The interaction between gut microbiota and stem cells influences epithelial self-renewal and overall intestinal homeostasis. Novel therapeutic approaches, including immunotherapy, targeted therapy, regenerative medicine using stem cells, and modulation of gut microbiota, are being explored to improve treatment outcomes. Accordingly, this chapter provides an overview of the potential therapeutic applications of gut microbiota and intestinal stem cells in treating colorectal cancer.
Article
Full-text available
The evaluation of nanostructured biomaterials and medicines is associated with 2D cultures that provide insight into biological mechanisms at the molecular level, while critical aspects of the tumor microenvironment (TME) are provided by the study of animal xenograft models. More realistic models that can histologically reproduce human tumors are provided by tissue engineering methods of co-culturing cells of varied phenotypes to provide 3D tumor spheroids that recapitulate the dynamic TME in 3D matrices. The novel approaches of creating 3D tumor models are combined with tumor tissue engineering (TTE) scaffolds including hydrogels, bioprinted materials, decellularized tissues, fibrous and nanostructured matrices. This review focuses on the use of nanostructured materials in cancer therapy and regeneration, and the development of realistic models for studying TME molecular and immune characteristics. Tissue regeneration is an important aspect of TTE scaffolds used for restoring the normal function of the tissues, while providing cancer treatment. Thus, this article reports recent advancements in the development of 3D TTE models for antitumor drug screening, studying tumor metastasis, and tissue regeneration. Also, this review identifies the significant opportunities of using 3D TTE scaffolds in the evaluation of the immunological mechanisms and processes involved in the application of immunotherapies.
Chapter
Software and Programming Tools in Pharmaceutical Research is a detailed primer on the use for computer programs in the design and development of new drugs. Chapters offer information about different programs and computational techniques in pharmacology. The book will help readers to harness computer technologies in pharmaceutical investigations. Readers will also appreciate the pivotal role that software applications and programming tools play in revolutionizing the pharmaceutical industry. The book includes nine structured chapters, each addressing a critical aspect of pharmaceutical research and software utilization. From an introduction to pharmaceutical informatics and computational chemistry to advanced topics like molecular modeling, data mining, and high-throughput screening, this book covers a wide range of topics. Key Features: - Practical Insights: Presents practical knowledge on how to effectively utilize software tools in pharmaceutical research. - Interdisciplinary Approach: Bridges the gap between pharmaceutical science and computer science - Cutting-Edge Topics: Covers the latest advancements in computational drug development, including data analysis and visualization techniques, drug repurposing, pharmacokinetic modelling and screening. - Recommendations for Tools: Includes informative tables for software tools - Referenced content: Includes scientific references for advanced readers The book is an ideal primer for students and educators in pharmaceutical science and computational biology, providing a comprehensive foundation for this rapidly evolving field. It is also an essential resource for pharmaceutical researchers, scientists, and professionals looking to enhance their understanding of software tools and programming in drug development.
Chapter
Exosomes are small vesicles that are released by cells, including medicinal signaling cells (MSCs), and play a role in intercellular communication. In regenerative medicine, exosomes are being explored as potential therapeutic agents due to their ability to transfer biomolecules between cells. For example, exosomes have been used to deliver small interfering ribonucleic acid (RNA) molecules to target cells, which can silence specific genes and modulate cell function. Additionally, exosomes can contain non-coding RNA, which plays a significant role in epigenetic control and can influence tumor metastasis. The goal of regenerative medicine is to take advantage of the epigenetic effect of exosomes while promoting repair and regeneration in targeted areas of the body. Exosomes have also been shown to improve symptoms of sarcopenia and have shown potential in applications such as knee osteoarthritis (OA), cartilage restoration, and tendon regeneration. Further research is needed to fully understand the mechanisms of exosomes in promoting tissue repair and to develop new strategies for the treatment of musculoskeletal injuries.
Article
This study aimed to synthesize magnetite nanoparticles proposing a new core with maltodextrin using the coprecipitation method in an alkaline medium, with the potential application in magnetic hyperthermia for cancer treatment. The syntheses followed a 3 2 factorial design for optimization, investigating the impact of two parameters: concentrations of Fe+2 and Fe+3 (0.02, 0.06, and 0.1 mol L-1 ) and maltodextrin mass (0.5, 1.5, and 2.5 g). Characterizations reveal that the synthesized nanoparticles show of the magnetite phase (Fe3O4) with an average crystallite size ranging from 6.3 to 12.5 nm, within the expected size range. The nanoparticles show superparamagnetic properties. Despite maltodextrin being present on the nanoparticles surfaces, it was observed to be ineffective in mitigating the aggregation and agglomeration phenomenon. In magnetic hyperthermia tests, the majority of samples reached and surpassed 43 °C during the analyses, corresponding to the therapeutic temperature. Furthermore, the ILP values ranged from 0.02 to 0.16 nH m2 kg-1 . Statistical evaluation concluded that optimal synthesis parameters involve a higher concentration of iron precursors and lower amounts of maltodextrin. The M-53 cells exposed to two distinct concentrations of nanoparticles exhibited non-cytotoxicity but displayed limited cellular internalization into the cells. These properties render them an as promising material for magnetic hyperthermia in oncological treatment. Further studies are necessary to determine the viability of the new coating route with maltodextrin.
Article
Full-text available
El cáncer gástrico es el quinto cáncer más común en todo el mundo y es responsable de un porcentaje significativo de todas las muertes que se dan debido a esta enfermedad. Actualmente la cirugía sigue siendo el estándar de oro en los tratamientos curativos vistos desde el campo de la oncología. Sin embargo, los tratamientos como la radioterapia, la quimioterapia pueden mejorar los resultados previos o posteriores a la intervención quirúrgica. En tal sentido, se lleva a cabo el tratamiento neoadyuvante antes de la gastrectomía o resección quirúrgica primaria seguida de quimioterapia. Este enfoque, combinado con un descubrimiento temprano y una mejor detección, ha llevado a una disminución de la incidencia general del cáncer gástrico. Lamentablemente, los tumores malignos del estómago se diagnostican en etapas avanzadas o metastásicas cuando la mediana de supervivencia general sigue siendo pobre. La atención quirúrgica en estos casos debe ser brindada por un equipo especializado y de gran envergadura. Los aspectos quirúrgicos son la pieza clave para los resultados, tomando en consideración que la estandarización del tratamiento quirúrgico del cáncer gástrico de acuerdo con el perfil del paciente es de importancia decisiva para un mejor resultado. Esta revisión tiene como objetivo resumir los estándares actuales en el tratamiento quirúrgico del cáncer gástrico.
Article
Full-text available
This article provides an update on the global cancer burden using the GLOBOCAN 2020 estimates of cancer incidence and mortality produced by the International Agency for Research on Cancer. Worldwide, an estimated 19.3 million new cancer cases (18.1 million excluding nonmelanoma skin cancer) and almost 10.0 million cancer deaths (9.9 million excluding nonmelanoma skin cancer) occurred in 2020. Female breast cancer has surpassed lung cancer as the most commonly diagnosed cancer, with an estimated 2.3 million new cases (11.7%), followed by lung (11.4%), colorectal (10.0 %), prostate (7.3%), and stomach (5.6%) cancers. Lung cancer remained the leading cause of cancer death, with an estimated 1.8 million deaths (18%), followed by colorectal (9.4%), liver (8.3%), stomach (7.7%), and female breast (6.9%) cancers. Overall incidence was from 2‐fold to 3‐fold higher in transitioned versus transitioning countries for both sexes, whereas mortality varied <2‐fold for men and little for women. Death rates for female breast and cervical cancers, however, were considerably higher in transitioning versus transitioned countries (15.0 vs 12.8 per 100,000 and 12.4 vs 5.2 per 100,000, respectively). The global cancer burden is expected to be 28.4 million cases in 2040, a 47% rise from 2020, with a larger increase in transitioning (64% to 95%) versus transitioned (32% to 56%) countries due to demographic changes, although this may be further exacerbated by increasing risk factors associated with globalization and a growing economy. Efforts to build a sustainable infrastructure for the dissemination of cancer prevention measures and provision of cancer care in transitioning countries is critical for global cancer control.
Article
Full-text available
Each year, the American Cancer Society estimates the numbers of new cancer cases and deaths in the United States and compiles the most recent data on population‐based cancer occurrence. Incidence data (through 2017) were collected by the Surveillance, Epidemiology, and End Results Program; the National Program of Cancer Registries; and the North American Association of Central Cancer Registries. Mortality data (through 2018) were collected by the National Center for Health Statistics. In 2021, 1,898,160 new cancer cases and 608,570 cancer deaths are projected to occur in the United States. After increasing for most of the 20th century, the cancer death rate has fallen continuously from its peak in 1991 through 2018, for a total decline of 31%, because of reductions in smoking and improvements in early detection and treatment. This translates to 3.2 million fewer cancer deaths than would have occurred if peak rates had persisted. Long‐term declines in mortality for the 4 leading cancers have halted for prostate cancer and slowed for breast and colorectal cancers, but accelerated for lung cancer, which accounted for almost one‐half of the total mortality decline from 2014 to 2018. The pace of the annual decline in lung cancer mortality doubled from 3.1% during 2009 through 2013 to 5.5% during 2014 through 2018 in men, from 1.8% to 4.4% in women, and from 2.4% to 5% overall. This trend coincides with steady declines in incidence (2.2%‐2.3%) but rapid gains in survival specifically for nonsmall cell lung cancer (NSCLC). For example, NSCLC 2‐year relative survival increased from 34% for persons diagnosed during 2009 through 2010 to 42% during 2015 through 2016, including absolute increases of 5% to 6% for every stage of diagnosis; survival for small cell lung cancer remained at 14% to 15%. Improved treatment accelerated progress against lung cancer and drove a record drop in overall cancer mortality, despite slowing momentum for other common cancers.
Article
Full-text available
Hypoxia is a condition commonly observed in the core of solid tumors. The hypoxia-inducible factors (HIF) act as hypoxia sensors that orchestrate a coordinated response increasing the pro-survival and pro-invasive phenotype of cancer cells, and determine a broad metabolic rewiring. These events favor tumor progression and chemoresistance. The increase in glucose and amino acid uptake, glycolytic flux, and lactate production; the alterations in glutamine metabolism, tricarboxylic acid cycle, and oxidative phosphorylation; the high levels of mitochondrial reactive oxygen species; the modulation of both fatty acid synthesis and oxidation are hallmarks of the metabolic rewiring induced by hypoxia. This review discusses how metabolic-dependent factors (e.g., increased acidification of tumor microenvironment coupled with intracellular alkalinization, and reduced mitochondrial metabolism), and metabolic-independent factors (e.g., increased expression of drug efflux transporters, stemness maintenance, and epithelial-mesenchymal transition) cooperate in determining chemoresistance in hypoxia. Specific metabolic modifiers, however, can reverse the metabolic phenotype of hypoxic tumor areas that are more chemoresistant into the phenotype typical of chemosensitive cells. We propose these metabolic modifiers, able to reverse the hypoxia-induced metabolic rewiring, as potential chemosensitizer agents against hypoxic and refractory tumor cells.
Article
Full-text available
Tumor microenvironment denotes the non-cancerous cells and components presented in the tumor, including molecules produced and released by them. The constant interactions between tumor cells and the tumor microenvironment play decisive roles in tumor initiation, progression, metastasis, and response to therapies. The tumor microenvironment as a therapeutic target in cancer has attracted great research and clinical interest. Here we summarize the current progress in targeting the tumor microenvironment in both drug development and clinical trials; highlight challenges in targeting the tumor microenvironment to achieve therapeutic efficacy; explore new technologies and approaches to better decipher the tumor microenvironment; and discuss strategies to intervene in the pro-tumor microenvironment and maximize therapeutic benefits.
Article
Full-text available
For a long period, cancer has been believed to be a gene disease, in which oncogenic and suppressor mutations accumulate gradually, finally leading to the malignant transformation of cells. This vision has changed in the last few years, the involvement of the tumor microenvironment, the non-malignant part of the tumors, as an important contributor to the malignant growth being now largely recognized. There is a consensus according to which the understanding of the tumor microenvironment is important as a means to develop new approaches in the therapy of cancer. In this context, the present study is a review of the different types of non-malignant cells that can be found in tumors, with their pro or antitumoral actions, presence in tumors and therapeutic targeting. These cells establish complex relations between them, through cytokines, exosomes, cell adhesion, co-stimulation and co-inhibition; these relations will also be examined in the present work.
Article
Full-text available
Thyroid cancer is the most common endocrine malignancy. Although most thyroid cancer patients are successfully treated and have an excellent prognosis, a percentage of these patients will develop aggressive disease and, eventually, progress to anaplastic thyroid cancer. Since most patients with this type of aggressive thyroid carcinoma will die from the disease, new treatment strategies are urgently needed. Tumor cells live in a complex and dynamic tumor microenvironment composed of different types of stromal cells. Cancer-associated fibroblasts (CAFs) are one of the most important cell components in the tumor microenvironment of most solid tumors, including thyroid cancer. CAFs originate mainly from mesenchymal cells and resident fibroblasts that are activated and reprogrammed in response to paracrine factors and cytokines produced and released by tumor cells. Upon reprogramming, which is distinguished by the expression of different marker proteins, CAFs synthesize and secret soluble factors. The secretome of CAFs directly impacts different functions of tumor cells. This bi-directional interplay between CAFs and tumor cells within the tumor microenvironment ends up fostering tumor cancer progression. CAFs are therefore key regulators of tumor progression and represent an under-explored therapeutic target in thyroid cancer.
Article
Full-text available
Immunotherapy using dendritic cells (DCs) is a promising treatment modality for cancer. However, the limited number of functional DCs from peripheral blood has been linked to the unsatisfactory clinical efficacies of current DC-based cancer immunotherapies. We previously generated proliferating antigen-presenting cells (APCs) by genetically engineering myeloid cells derived from induced pluripotent stem cells (iPSC-pMCs), which offer infinite functional APCs for broad applications in cancer therapy. Herein, we aimed to further enhance the antitumor effect of these cells by genetic modification. GM-CSF gene transfer did not affect the morphology, or surface phenotype of the original iPSC-pMCs, however, it did impart good viability to iPSC-pMCs. The resultant cells induced GM-CSF-dependent CD8⁺ T cell homeostatic proliferation, thereby enhancing antigen-specific T cell priming in vitro. Administration of the tumor antigen-loaded GM-CSF-producing iPSC-pMCs (GM-pMCs) efficiently stimulated antigen-specific T cells and promoted effector cell infiltration of the tumor tissues, leading to an augmented antitumor effect. To address the potential tumorigenicity of iPSC-derived products, irradiation was applied and found to restrict the proliferation of GM-pMCs, while retaining their T cell-stimulatory capacity. Furthermore, the irradiated cells exerted an antitumor effect equivalent to that of bone marrow-derived DCs obtained from immunocompetent mice. Additionally, combination with immune checkpoint inhibitors increased the infiltration of CD8⁺ or NK1.1⁺ effector cells and decreased CD11b⁺/Gr-1⁺ cells without causing adverse effects. Hence, although GM-pMCs have certain characteristics that differ from endogenous DCs, our findings suggest the applicability of these cells for broad clinical use and will provide an unlimited source of APCs with uniform quality.
Article
Hemoglobin (Hb)-based oxygen carriers (HBOCs) present an alternative to red blood cells (RBCs) when blood is not available. However, the most widely used synthesis techniques have fundamental flaws, which may have contributed toward disappointing clinical application. Polymerized Hb contains a heterogeneous distribution of particle size and shape, while Hb encapsulation inside liposomes results in high lipid burden and low Hb content. Meanwhile, there are a variety of other nanoparticle synthetic techniques which, having found success as drug delivery vehicles, may be well suited to function as an HBOC. We synthesized desolvated Hb nanoparticles (Hb-dNPs) with diameters of approximately 250 nm by the controlled precipitation of Hb with ethanol. Oxidized dextran was found to be an effective surface stabilizing agent that maintained particle integrity. In vitro biophysical characterization showed a high-affinity oxygen delivery profile (P50 = 7.72 mm Hg), suggesting a potential for therapeutic use and opening a new avenue for HBOC research.
Article
About 5% of cancer patients treated with radiotherapy will have severe late-onset toxicity. Hyperbaric oxygen therapy (HBOT) has been used as a treatment for radiation injuries for decades, with many publications presenting data from small series or individual cases. Moreover, we know that the hypoxic areas of tumours are more resistant to radiation. HBOT increases the oxygen tension in tissues and, theoretically, it should enhance the efficiency of radiotherapy. To better understand how HBOT works, we carried out this bibliographic review. We found Grade B and C evidence that at pressures exceeding 2 absolute atmospheres (ata), HBOT reduced late-onset radiation injuries to the head and neck, bone, prostate and bladder. It also appeared to prevent osteoradionecrosis after exodontia in irradiated areas. Finally, HBOT at 2 ata increased the effectiveness of radiation in head and neck tumours and achieved promising results in the local control of high-grade gliomas.