Next Article in Journal
Association between Dysfunction of the Nucleolar Stress Response and Multidrug Resistance in Pediatric Acute Lymphoblastic Leukemia
Previous Article in Journal
Enhanced Epithelial-to-Mesenchymal Transition and Chemoresistance in Advanced Retinoblastoma Tumors Is Driven by miR-181a
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

TRAIL in the Treatment of Cancer: From Soluble Cytokine to Nanosystems

by
Hojjat Alizadeh Zeinabad
1 and
Eva Szegezdi
1,2,*
1
Apoptosis Research Centre, Biomedical Sciences Building, School of Biological and Chemical Sciences, University of Galway, H91 W2TY Galway, Ireland
2
Science Foundation Ireland (SFI) Centre for Research in Medical Devices (CÚRAM), Biomedical Sciences Building, University of Galway, H91 W2TY Galway, Ireland
*
Author to whom correspondence should be addressed.
Cancers 2022, 14(20), 5125; https://doi.org/10.3390/cancers14205125
Submission received: 21 September 2022 / Revised: 14 October 2022 / Accepted: 15 October 2022 / Published: 19 October 2022
(This article belongs to the Section Cancer Therapy)

Abstract

:

Simple Summary

TRAIL is a death ligand cytokine, predominantly used by effector immune cells to kill malignantly transformed cells. Since its discovery, TRAIL has attracted a lot of attention as a promising anticancer drug due to its selective action against cancer cells, promising a safe, low-toxicity treatment. Despite its promising characteristics, clinical trials have not delivered on this promise, due to issues with the poor in vivo biological activity of recombinant TRAIL formulations. Nanoparticles have the potential to overcome these limitations, and an increasing number of studies have reported very promising preclinical results. Here, we summarize the potential of TRAIL for cancer therapy, and provide a critical assessment of the challenges and the potential of various formulations of nanovehicles designed to date for TRAIL-based cancer therapy.

Abstract

The death ligand tumor necrosis factor (TNF)-related apoptosis-inducing ligand (TRAIL), a member of the TNF cytokine superfamily, has long been recognized for its potential as a cancer therapeutic due to its low toxicity against normal cells. However, its translation into a therapeutic molecule has not been successful to date, due to its short in vivo half-life associated with insufficient tumor accumulation and resistance of tumor cells to TRAIL-induced killing. Nanotechnology has the capacity to offer solutions to these limitations. This review provides a perspective and a critical assessment of the most promising approaches to realize TRAIL’s potential as an anticancer therapeutic, including the development of fusion constructs, encapsulation, nanoparticle functionalization and tumor-targeting, and discusses the current challenges and future perspectives.

1. Introduction

The immune system plays an essential role in eliminating malignant cells by identifying and killing such cells. Death ligands (DL), expressed on the surface of effector immune cells, play a central role in this process. DLs are members of the tumor necrosis factor (TNF) cytokine family. They bind and activate death receptors (DR), thus inducing a regulated cell death program in the receptor-carrying cell.
The three best-known DLs are TNF itself, with its receptor, TNF receptor 1 (TNFR1), TNF-related apoptosis-inducing ligand (TRAIL), inducing cell death via the receptors DR4 and DR5, and Fas ligand (FasL) binding to the Fas receptor. Of the DLs, TRAIL received a lot of attention due to its capacity to induce apoptosis in many types of cancer cells without affecting healthy, non-transformed cells [1,2,3,4,5,6]. A variety of immune cells, including eosinophil granulocytes [7], macrophages [8,9], neutrophil granulocytes [10,11], dendritic cells [12,13], monocytes [9,14], natural killer (NK) cells [15], T and B cells [16,17], express TRAIL, and it is well established that TRAIL plays an important role in tumor immune surveillance, making TRAIL a promising candidate for an anticancer therapeutic [18,19,20].
Dulanermin, a soluble, recombinant human TRAIL (rhTRAIL) formulation comprising of the extracellular portion of TRAIL (amino acids 114–281) was the first TRAIL variant tested for the treatment of cancer [21,22]. The efficacy of Dulanermin either as a single agent or in combination with chemotherapy has been tested in several phase II and III clinical trials, but the tests have had disappointing results, showing very limited anti-tumor efficacy [4,23,24]. Although rhTRAIL had a very good safety profile and very low toxicity, it showed poor efficacy. The poor anti-tumor activity was linked to its short in vivo half-life, insufficient accumulation in tumor tissues and resistance of tumor cells to TRAIL-induced cell death [25,26,27].
To overcome these limitations, in the past ten years, novel, nanomedicine-based approaches have been considered. This review presents the most successful approaches used to improve the efficacy of TRAIL-based therapeutics highlighting how nanotechnology can help to realize the expected therapeutic potential of TRAIL and outlines the remaining challenges.

2. The TRAIL-Induced Apoptotic Pathway

The physiological role of apoptosis is to eliminate superfluous and abnormal cells, such as damaged or stressed cells including tumor cells [28]. It follows two major pathways; the intrinsic and the extrinsic apoptotic pathway (Figure 1) [29]. The intrinsic pathway, also known as the mitochondrial pathway, is initiated by events on the surface of mitochondria and regulated by the B cell leukemia 2 (Bcl-2) protein family [30,31]. The Bcl-2 family proteins are characterized by the presence of at least one of the four Bcl-2 homology (BH1–4) domains and divided into three sub-families [31,32]. The first sub-family consists of the founding member of the family, Bcl-2, BCL-2-like protein 1 (BCL2L1, Bcl-extra large (Bcl-xL), BCL-2-like-2 (Bcl-W), Myeloid cell leukemia-1 (Mcl-1), BCL-2-like-10 (Bcl-B) and BCL-2-related protein A1 (A1)). These members contain all four BH domains and they inhibit apoptosis, while the other two sub-families are pro-apoptotic. BCL2-related ovarian killer (Bok), Bcl-2 associated X protein (Bax) and Bcl-2 homologous antagonist/killer (Bak) contain the BH1-BH3 domains but lack the BH4 domain and they form the multidomain, pro-apoptotic sub-group of the family. Finally, the so-called BH3-only proteins form the third group and, as their name shows, they only contain the BH3 domain. Members of the BH3-only group include for example BCL-2 antagonist of cell death (Bad), BH3-interacting domain death agonist (Bid), BCL-2-like-11 (Bim), BCL-2-modifying factor (Bmf), BCL-2-interacting killer (Bik), Harakiri (Hrk, also known as death protein-5), Phorbol-12-myristate-13-acetate-induced protein 1 (Noxa) and BCL-2-binding component-3 (Puma). Upon stress, such as DNA damage, oxidative stress, endoplasmic reticulum stress, etc., BH3-only proteins are induced and/or activated and they initiate the apoptotic program by activating the multidomain pro-apoptotic Bcl-2 proteins, Bax and Bak [32,33,34,35,36], which then oligomerize and form pores on the mitochondrial outer membrane (MOM), resulting in its permeabilization [37,38]. From the permeabilized mitochondria, pro-apoptotic factors such as cytochrome c (Cyt c) [39], endonuclease G [40], apoptosis-inducing factor (AIF) [41] and second mitochondria-derived activator of caspase (Smac) [42] are released [43]. In the cytosol, Cyt c interacts with apoptotic protease-activating factor (Apaf-1) and pro-caspase-9 to form the apoptosome complex, which cleaves and thus activates pro-caspase-9, which in turn cleaves and activates the effector caspases, caspase-3/-6/-7, thus committing the cell to apoptosis [44,45]. The other pro-apoptotic factors released from the mitochondria facilitate this pathway. For example, Smac can neutralize X-linked inhibitor of apoptosis protein (XIAP), thus relieving XIAP-mediated caspase-3/-7/-9 inhibition [46,47].
The extrinsic apoptotic pathway is initiated by DRs (Figure 1). Binding of a DL to its DR results in the formation of a protein complex, called the death-inducing signaling complex (DISC), starting with the recruitment of the adaptor protein, Fas-associated death domain (FADD) and pro-caspase-8 and/or -10 to the death receptor. Pro-caspase-8 comprises of two N-terminal death effector domains (DEDs) followed by two protease catalytic domains (p18 and p10). One DED of pro-caspase-8 interacts with the same (DED) domain of FADD (homotypic interaction), while its second DED domain enables recruitment of an additional pro-caspase-8 molecule. Via the DED–DED interactions, multiple pro-caspase-8 molecules are recruited to the DISC, forming either a chain or a filament in which pro-caspase-8 activation takes place [48,49,50,51]. Once active, caspase-8/-10 are released from the DISC into the cytoplasm where similar to caspase-9, they activate downstream effector caspases, resulting in apoptosis (for review: Sessler et al., 2013; [52]).
In addition to effector caspases, caspase-8 can also cleave and activate the BH3-only protein, Bid, thus interconnecting the extrinsic and the intrinsic apoptotic pathways (Figure 1) [32,53]. Because in some cells DR-generated caspase-8 activity is not sufficient to activate the downstream caspase cascade, in these cells amplification of the apoptotic signal through the mitochondrial pathway (Bid activation) is essential to commit the cell to apoptosis (so-called type II cells) [54,55,56].
There are numerous mechanisms that control and override apoptotic signaling, thus allowing cells to survive. For example, the intrinsic apoptotic pathway is inhibited by the anti-apoptotic members of the Bcl-2 family which can block Bax/Bak activation and consequent Cyt c release [19]. This is a common mechanism of DL-resistance in type II cells, and often employed by cancer cells [57,58,59,60].
More specific to the extrinsic apoptotic pathway, activation of DRs themselves is controlled by decoy receptors, such as decoy receptor (DcR)1, DcR2 and osteoprotegerin (OPG) for TRAIL, or DcR3 in the case of FasL [61,62,63]. Downstream of the DRs, pro-caspase-8 activation can be inhibited by cellular FLICE inhibitory protein (cFLIP), a pseudo-caspase that interacts with and prevents pro-caspase-8 activation in the DISC [48,49,64]. Further downstream, the apoptotic pathway is controlled by inhibitor of apoptosis (IAP) proteins, including XIAP (X-linked IAP), cIAP1 (cellular IAP1) and cIAP2 [46,65,66]. XIAP can directly bind and inhibit caspase-3, caspase-7 and caspase-9 [66,67,68], while cIAP1 and cIAP2 inhibit apoptosis through activation of the nuclear factor-κB (NF-κB) signaling pathway upon DR-activation leading to the induction of cFLIP [69], Bcl-2 and Bcl-xL [69,70,71,72,73].

3. TRAIL Formulations to Increase Its Serum Half-Life Time

Soluble TRAIL has a short in vivo half-life time of 0.56–1.02 h [21] that limits its efficacy [74]. The main reason for it is its relatively small size of approximately 60 kDa, in its biologically active, trimeric form. As kidney filtration cut-off is around 70 kDa [75], many proteins smaller than this size, including rhTRAIL, are quickly cleared from the blood. To overcome this issue, numerous recombinant TRAIL derivatives have been designed to increase its size by fusing it to a peptide or protein [76,77,78,79,80,81,82,83,84,85,86,87]. For example, TRAIL has been conjugated to serum albumin, either directly [88] or indirectly [89], which increases its circulatory half-life substantially. Additionally, because albumin can interact with albondin (gp60 receptor) on endothelial cells, the fusion also facilitated the transport of TRAIL across the endothelium into the tissues via caveolae-mediated transcytosis [90] leading to enhanced tumor-suppressing potential [88].
A similar strategy has been used by Brin and colleagues [91]. They fused arginine deiminase to rhTRAIL, as arginine deaminase was known to up-regulate DR5 expression. They showed that the fusion construct was more effective against a colorectal cancer tumor xenograft than rhTRAIL.
All these studies could achieve substantial increase in the circulatory half-life of TRAIL; however, it would be important to determine whether these modifications alter or influence receptor binding, as the fusion partners were quite large and thus, they could alter the conformation of TRAIL’s receptor-binding motifs.

4. TRAIL Formulations to Mimic Membrane-Bound TRAIL

Although the soluble form of TRAIL, comprising of its extracellular domain, is biologically active, it has been shown that the full-length, membrane-bound form of TRAIL is a 100–1000-fold more potent inducer of cell death ascribed to an oligomeric or clustered conformation of membrane-bound TRAIL as opposed to the trimeric structure of soluble TRAIL [92,93]. To address this issue, a variety of nanocarriers, including organic and inorganic nanoparticles (NPs), viral NPs and cells or cellular components have been assessed to replicate the oligomeric TRAIL complexes believed to be formed in the membrane upon receptor binding.

4.1. Cell- and Virus-Based Delivery Tools

One approach to mimic the structure and organization of membrane-bound TRAIL was the use of viral particles. For example, TRAIL has been conjugated to capsid protein IX of the Ad5 oncolytic adenovirus via its intracellular, N-terminus using a heterodimeric zipper domain thus enabling an orientation of TRAIL on the virus surface similar to that of on the plasma membrane [94]. The study found that this TRAIL formulation had increased cytotoxic potential both in vitro and in vivo. In a recent work by the same lab [95] this design was further improved by combining the oncolytic virus treatment with ginsenoside-Rh2 treatment. Ginsenoside-Rh2 is a derivative of the small molecule drug, dammarane found in ginseng species and has been shown to up-regulate TRAIL receptor expression [96] through which the study could achieve higher toxicity in tumor cells with low baseline TRAIL receptor expression, such as primary AML cells.
The rod-shaped flexuous filamentous potato virus X (PVX) is another virus tested for TRAIL delivery [97]. PVX is a unique, multifunctional vehicle increasingly considered for nanomedicine, because the surface of PVX is rich in cysteine and lysine side chains, which can be easily used as adaptors for functionalization [98,99,100]. His-tagged TRAIL could be immobilized on the virus surface via His-tag-nickel-nitrilotriacetic acid (Ni-NTA) interaction. Placing the His-tag on the N-terminus of TRAIL enabled conjugation of TRAIL to the surface of the virus in its native orientation. This formulation led to a 3–10-fold higher pro-apoptotic potential in vitro compared to rh-His-TRAIL. Furthermore, it also suppressed in vivo tumor growth in a human triple-negative breast cancer mouse model [97].
Several studies explored whether human cells can be used as TRAIL-delivery tools. Using cells as a vehicle for TRAIL therapy has several advantages, such as low immunogenicity [101], ease of genetic-engineering [102,103] and tumor-targeting using migratory properties of the cell vehicles (e.g., mesenchymal stem cells (MSCs) [103] or via other genetic modifications) [104]. In a smart approach [105], genetically engineered platelets expressing surface-bound TRAIL have been developed using a lentiviral transgene expression. Hematopoietic stem and progenitor cells (HSPCs) were transduced with a TRAIL gene whose expression was under the megakaryocyte-specific integrin αIIβ promoter. By transplanting the TRAIL-transduced HSPCs into the bone marrow of mice, TRAIL-expressing platelets were produced in vivo, during megakaryocyte maturation. This design reduced prostate cancer metastasis, indicating that TRAIL-expressing platelets could eliminate circulating tumor cells [105]. In a similar study, Shah and colleagues [106] engineered neural stem cells to secrete soluble TRAIL. This approach markedly decreased tumor burden and prolonged survival in mouse models of glioblastoma [106].
Another important cell type tested as TRAIL-delivery vehicles are mesenchymal stem cells (MSCs) due to their ability to migrate and home to sites of injury [107,108]. Based on this hypothesis, several studies have shown that TRAIL-expressing MSCs significantly inhibited tumor growth and prolonged survival in various mouse cancer models, including lung cancer [109], lung metastasis of breast cancer [110] and renal cell carcinoma [111], glioblastoma [102,112,113,114], hepatocellular carcinoma [115], cervical carcinoma [116], colorectal carcinoma [117] and pancreatic carcinoma [118].
Despite the advantages of cell-based TRAIL therapy, cell therapy also has considerable limitations, including the limited replicative potential of MSCs [119], the requirement of GMP (good manufacturing practice) facilities for generation of cellular therapeutics and the cost and time required to obtain sufficient number of cells for therapy and potential immunogenicity [120,121]. To overcome these problems, nanoparticle carriers are being developed that can not only mimic cell-based TRAIL therapy [122], but also provide a mechanism to combine several advantageous properties of TRAIL delivery from increased half-life, correct conformation, combination with other drugs to targeted delivery to tumor sites [123,124,125,126,127].

4.2. Nanoparticle-Based TRAIL Delivery

NPs have a great advantage over biological carriers, similar to cells and EVs, due to their high flexibility, which allows selective targeting of the tumor site facilitating cargo accumulation in the tumor microenvironment and thus enhancing efficacy, while also reducing potential toxic effects on normal cells. An additional benefit of nanoparticles is the possibility to encapsulate chemotherapeutic agents and co-deliver them with TRAIL, therefore overcoming the frequent issue of tumor TRAIL resistance.
The delivery of TRAIL via NPs has been extensively explored, and several successful nanoparticle-based TRAIL-delivery systems have been developed (Figure 2). These studies also confirmed that NPs can enable TRAIL oligomerization, increase serum half-life and overcome tumor TRAIL resistance, particularly when the TRAIL-NPs were combined with sensitizing agents [128,129,130,131,132].
A variety of nanocarriers, including metallic NPs [134,135,136], polymeric NPs [137,138,139,140], lipid-based NPs [141,142,143], protein NPs [128,144,145,146] and carbon-based NPs [132,147], have been tested as TRAIL carriers. Key findings of these studies are summarized in Figure 3.
In the following, we focus on discussing the most successful nano-based approaches for TRAIL delivery.

4.2.1. Cell Component-Based Nanoparticles for TRAIL Delivery

Since 2011, when particle-coating with cell membrane was first reported [148], a variety of cell types have been used as membrane sources to coat nanoparticles [149] and thus combine the advantages of designed nanoparticles with membrane properties of a specific cell type to enable tissue targeting, NP immune escape or prolonged blood circulation [150,151,152]. For example, monodispersed silica particles were coated with membranes of activated platelets followed by functionalization with TRAIL with the aim to target and kill circulating tumor cells (CTCs). In the circulation, tumor cells are vulnerable and to protect themselves from immune recognition, they hide from effector immune cells by binding and covering themselves with platelets (cloaking/microthrombi). The study showed that the propensity of cloaking can be turned against CTCs as CTCs bound the platelet membrane-coated NPs instead of platelets, which exposed them to TRAIL and killed them [153].
Exosomes are another cell component being explored as drug delivery tools for cancer therapy [154,155,156]. Exosomes were first described in 1981 [157] as small extracellular vehicles (EVs) of 30–100 nm diameter surrounded by a lipid bilayer, naturally released by various cell types [158]. Due to their native origin, they have low immunogenicity matched with an excellent ability to carry a wide spectrum of cargoes ranging from hydrophilic as well as hydrophobic drugs to genetic materials, making EVs promising nanocarriers [154,159].
Protein-based delivery tools: TRAIL protein conjugated to the surface of various NPs (e.g., liposomes, viruses, carbon nanotubes, metallic NPs) interacts with DR4/DR5 on the tumor cells to induce apoptosis. Some TRAIL protein-based delivery tools are functionalized with targeting ligands, e.g., antibodies, or molecules (i.e., hyaluronic acid) that preferentially target tumor cells and enable the induction of cell death via TRAIL. The unique structure of liposomes makes them an excellent delivery tool for combination therapy that can co-deliver chemotherapeutic reagents and TRAIL protein resulting in a synergistic cell death signal. Metallic NPs functionalized with TRAIL protein not only induce apoptosis via TRAIL, but they can also generate reactive oxygen species (ROS) as they can be internalized into tumor cells. The generated ROS leads to the release of pro-apoptotic proteins from mitochondria and activation of the intrinsic apoptosis pathway, thus amplifying the TRAIL death signal.
Tumors are also known to release EVs [160] containing various cytokines [161,162], immunosuppressive and inhibitory proteins such as transforming growth factor beta 1 (TGFβ1) [163], programmed death ligand 1 (PD-L1) [164,165,166], cyclooxygenase-2 (COX2), cytotoxic T-lymphocyte-associated protein 4 (CTLA4), TRAIL [167] and FasL [168,169]. Using the ability of cells to produce and release EVs, it was demonstrated that cells genetically modified to ectopically express TRAIL released exosomes decorated with biologically active TRAIL [170]. As an example, the Huber laboratory transduced K562 leukemic cells with TRAIL to produce TRAIL-expressing exosomes. These exosomes induced apoptosis of melanoma and lymphoma cells in vitro and reduced tumor growth in vivo in melanoma and lymphoma mouse models [170]. Similarly, Shamili and colleagues [171] reported that exosomes isolated from cultures of MSCs engineered to ectopically express TRAIL could induce apoptosis in vitro as well in vivo in a melanoma mouse model [171]. Another study [172] demonstrated that TRAIL-containing MSC-derived exosomes could induce cell death in a range of cancer cell lines including cell lines resistant to soluble TRAIL [172].
Despite all the above-mentioned unique advantages, EVs are highly heterogeneous [173] and their isolation and purification is challenging, time-consuming and expensive in comparison to artificially generated vesicles, such as liposomes [174,175,176], which drove the development of liposome-based TRAIL nanovehicles.

4.2.2. Liposome-Based TRAIL Delivery

So far, liposomes are one of the most commonly used NPs for cancer therapy with the first formulations in clinic use already, such as the sustained-release formulation of cytarabine (DepoCyt®) [177] or pegylated liposomal doxorubicin (DOXIL®/CAELYX®) [178].
Since 1961, liposomes have been the most widely investigated cargo delivery system due to good biocompatibility [179,180,181], biodegradability, ability to encapsulate both hydrophilic and hydrophobic cargos, a wide range of sizes and monodispersed composition. Easy surface modification and high liposome permeability offered solutions to cargo targeting and controlled release, which earned liposomes the title of “smart drug carriers” [176,182]. Furthermore, the high dynamic mobility of liposome membranes makes their structure similar to biological membranes and a good choice for artificial cell generation [183]. Although liposomes are not as stable as exosomes, they can be easily synthesized with the desired size on a large scale. Moreover, surface modification and functionalization with peptides or proteins, including antibodies, is very easy [176,184]. A final notable advantage of liposomes that makes them unique is the possibility to externally control the release of the encapsulated cargo at the tumor site [185,186].
One of the first studies developing TRAIL-functionalized liposomes was the Martinz-Lostao laboratory [143], who used large unilamellar vesicles (LUV) in which Ni2+ cations in the LUVs bound His-tagged TRAIL. This noncovalent interaction bound high amounts of TRAIL without changing the structure of the liposome and maintaining TRAIL bioactivity. The study proved that immobilization of TRAIL to liposomes increased TRAIL’s local concentration leading to increased pro-apoptotic potential compared to rhTRAIL. Additionally, LUV-TRAIL was also efficient against rhTRAIL-resistant leukemia cells [143], likely due to the arrangement of TRAIL into highly ordered oligomers on the LUVs which could induce a high level of DR4/DR5 clustering on the leukemic cells leading to increased pro-caspase-8 activation able to drive a type I extrinsic cell death pathway [129,187].
Due to the hyperpermeability of the vasculature in tumor tissues, the incorporation of TRAIL in nanoparticles can facilitate its preferential accumulation in tumor sites, a process called passive targeting [188] (Table 1). On the contrary, as a soluble protein, rhTRAIL can also pass through normal vessels, leading to a non-selective distribution of TRAIL between tumor sites and healthy tissues, thus lowering its concentration in the tumor.
Although the majority of FDA-approved nanoparticles for cancer therapy have used passive targeting [192,193], it has been shown that the decoration of nanoparticles with recognition molecules, such as an antibody to target tumor-specific antigens, enhances efficacy [194,195,196,197]. This is also true for TRAIL, with the first studies showing that active targeting of TRAIL to tumor sites can further increase its in vivo efficacy (Table 2).

4.2.3. Active Targeting

One of the most significant advantages of nanoparticles in cancer therapy is the ability to functionalize their surface with proteins and other biomolecules to selectively target tumor cells [198,199]. The most common strategies for NP tumor-targeting are (1) targeting NPs to immune cells and using the immune cells as NP carriers to indirectly target tumor cells, (2) functionalization with monoclonal antibodies to directly target unique or overexpressed antigens on the surface of tumor cells, and (3) functionalization with ligands/receptors against tumor-expressed receptors/ligands, such as hyaluronic acid (HA) to target CD44, often highly expressed on cancer cells [200,201].
Of these three strategies, indirect targeting of tumor cells is probably the most widely used. In this method, NPs are linked to immune cells, and the immune cells deliver the NPs to the tumor site or to CTCs. An appealing example of indirect targeting was the study by Mitchell and colleagues, who developed liposomes functionalized with E-selectin and TRAIL via Ni-NTA binding [202]. E-selectin enabled the liposomes to bind to E-selectin ligands widely expressed by leukocytes [203,204] as well as CTCs and to eliminate the CTCs via TRAIL-induced apoptosis thus preventing prostate cancer metastasis [205]. In a similar study, tethering a very low dose of TRAIL and E-selectin to liposomes was sufficient to induce TRAIL-mediated cell death in circulating tumor cells, inhibit metastasis and prolong survival in the 4T1 breast carcinoma mouse model [206].
NK cells have also been used as a carrier of TRAIL-functionalized liposomes. Functionalizing liposomes with TRAIL and anti-NK1.1 antibodies allowed the liposomes to bind to NK cells ubiquitously expressing NK1.1 to form “super-NK cells”. The super-NK cells displayed high cytotoxicity both in vitro and in vivo compared to liposomes functionalized with TRAIL alone or to unmodified NK cells [207,208].
Targeting NPs to tumor cells via immune cell-mediated delivery relies heavily on the ability of the immune cells to find the tumor cells. As tumors evolve, they acquire abilities to inhibit immune cells or to hide from them by removing antigens immune cells would recognize (the process of immune escape) [209,210], which is a major limitation of indirect, immune cell-based NP targeting and nanosystems with the ability to target the tumor directly may be more effective. For example, Seifert and colleagues [211] engineered liposomes able to target epidermal growth factor receptor (EGFR, overexpressed on a variety of cancers) [212,213] by functionalizing the liposomes with an EGFR-specific single-chain Fv molecule (EGFR scFv) and TRAIL (immune-LipoTRAIL) [211]. The immune-LipoTRAIL NP did not only increase TRAIL’s serum half-life, but it could target and inhibit the growth of an EGFR-positive colorectal tumor xenograft. Although, in this study, EGFR targeting did not enhance the efficacy of the TRAIL-liposome, direct NP targeting of tumors warrants further research [211].
Table 2. Active targeting approaches for TRAIL delivery.
Table 2. Active targeting approaches for TRAIL delivery.
FormulationTRAIL Form/LocationTargeting StrategyTumor TypeRef.
Inhalable HSA NPsrhTRAIL/SurfaceHSA to target gp60 transcytosis pathwayLung cancer, mouse model[214]
HSA NPsrhTRAIL/SurfaceTransferrin on HSA NPs to target transferrin receptor on tumor cells and HSA to target gp60 transcytosis pathwayColorectal cancer, mouse model[215]
LiposomepTRAIL/SurfaceAngiopeptide-2 to target the low-density LRP on BBB and glioma cellsGlioblastoma, mouse model[216]
Polymeric NPs coated with platelet membranerhTRAIL/SurfacePlatelet membrane to target CD44 on tumor cells via p-selectinMetastatic breast cancer, mouse model[217]
Polymeric NPspTRAIL/InsideReconstituted HDL on polymeric NPs to target scavenger receptor class B type I on MSCsPulmonary melanoma, mouse metastasis model[218]
Polymeric NPspTRAIL/SurfaceEGFR-specific peptide to target EGFR on laryngeal cancer cellsHep-2 laryngeal squamous cell carcinoma, mouse model[219]
PEI-coated gold nanocompositepTRAIL/SurfaceDexamethasone to target nucleusHep3B cell, mouse model[220]
Polymeric NPsrhTRAIL conjugated to PEG/InsideHA to target CD44 on tumor cellsCollagen-induced arthritis, mouse model[221]
Liposome inside a hyaluronic acid crosslinked-gel shellrhTRAIL/
B/w liposome and gel shell
HA to target CD44 on tumor cellsMDA-MB-231, breast cancer xenograft model[222]
ZnFe2O4 magnetic NPs w/mesoporous silica shell with PEITRAIL DNA/SurfaceAdipose tissue-derived MSCs treated with NPs to target tumor cellsOvarian cancer, mouse model[223]
Carbon dot coated polyethyleneiminepTRAIL/
Surface
MSCs for targeting tumor cellsLung cancer cell line (A549 cells)[224]
PEG-crosslinked albumin hydrogelrhTRAIL/InsideHSA to target gp60 transcytosis pathwayPancreatic cancer, mouse model[225]
Magnetic ternary nanohybrids (iron oxide NPs coated with HA)TRAIL DNA/InsideMSCs for targeting tumor cellsGlioma, mouse model[226]
Polymeric NPspTRAIL/InsideCholine-derivate to target choline transporters on BBB and glioma cellsGlioma, mouse model[227]
Abbreviations: HSA, human serum albumin; NPs, nanoparticles; rhTRAIL, recombinant human TRAIL; pTRAIL, plasmid TRAIL; LRP, lipoprotein receptor-related protein; BBB, blood-brain barrier; HDL, high-density lipoprotein; MSCs, mesenchymal stem cells; EGFR, epidermal growth factor receptor; PEI, polyethylenimine; PEG, polyethylene glycol.

4.2.4. Nanoparticles for TRAIL-Based Drug Combination Therapies

Although tethering TRAIL to cells or NPs significantly improved its therapeutic efficacy, primary tumors tend to be resistant to TRAIL. TRAIL resistance develops during tumor pathogenesis, as a mechanism to escape from immune recognition and elimination [74,228,229,230,231,232,233,234]. Accordingly, approximately two thirds of cancer cell lines have been found to be TRAIL-resistant, emphasizing that TRAIL resistance is a major limitation of TRAIL-based therapeutics [235].
Combination therapies have become the primary strategy to address tumor drug resistance, for which nanotechnology offers an excellent tool (Table 3) due to its capacity to simultaneously deliver multiple therapeutic reagents [236,237].

4.2.5. Combination of Chemotherapeutic Reagent Encapsulating-Nanoparticles with TRAIL

Since numerous studies revealed the synergistic cytotoxic effect of TRAIL with chemotherapeutic drugs [244], the unique potential of nanoparticles to encapsulate drugs may provide an opportunity to overcome TRAIL resistance. In this regard, liposomes and polymeric NPs received the most attention because of their advantageous properties. Incorporation of peptides and drugs in polymers has been known since the 1950s, with the first polymeric nanoparticles reported in 1976 [188]. Since then, various synthetic and biological polymers such as polyethylene glycol (PEG), polyglutamic acid, N-(2-hydroxypropyl) methacrylamide (HPMA), polylactic acid (PLA) and poly(lactic-co-glycolic acid) (PLGA) have been developed and been used in the clinic [245]. The key advantages offered by polymeric nanoparticles are their ability (1) to encapsulate a variety of therapeutic molecules such as RNA, DNA, proteins and chemotherapeutic drugs, (2) for controlled cargo-release, and (3) for surface functionalization with ligands and antibodies [246,247,248].
The first studies aimed to co-deliver a sensitizing agent used already approved chemotherapeutics known to induce DR4/DR5 expression by triggering cellular stress thus sensitizing cancer cells to TRAIL, such as derivatives of the microtubule inhibitor, taxol and DNA damaging agents, such as doxorubicin and mitoxantrone [240,249]. These NP formulations were more effective than the agents alone, paving the way for the development of more complex TRAIL-drug combination nanoformulations incorporating active tumor cell targeting and enhancing anti-tumor immune response.
Dendrimer, a polymer NP type, was a choice of carrier for many of these studies. Dendrimers are highly branched, globular polymers with internal cavities. This structure allows them to easily conjugate and encapsulate various drugs, proteins, and other molecules. In addition, they are biocompatible [250,251]. Liu and colleagues engineered a pH-sensitive dendrimer to co-deliver DOX and a TRAIL-encoding plasmid (pORF-hTRAIL) for glioma treatment [252]. They conjugated DOX and a T7 peptide that targets brain tumor cells by binding to the transferrin receptor (TfR), often overexpressed in glioblastoma [253] and brain capillary endothelial cells [254,255] to a poly-l-lysine dendrimer loaded with the pORF-hTRAIL plasmid. When the T7 peptide bound to the TfR on the tumor cells, the NP became endocytosed. Acidic pH in the endosomes then induced the release DOX and pORF-hTRAIL from the pH-sensitive dendrimers, thus simultaneously delivering DOX that triggered DNA damage and the TRAIL gene, which, after production of TRAIL protein, induced apoptotic signaling in glioma cells [252]. Several other dendrimer formulations, including PAMAM [256], diaminotriazine-modified PAMAM [189] and fluorinated dendrimers [257], have been used to increase the cytotoxicity of TRAIL by co-delivering a TRAIL-sensitizing drug and a TRAIL-expression cassette.
Another approach targeting NP delivery to tumor cells used an acrylamide-based doxorubicin nanogel (DOX-NG) fabricated by the single emulsion method [217]. For targeting, the DOX-NGs were coated with platelet membrane since platelets express p-selectin [258], a molecule that binds to CD44, a receptor often up-regulated on cancer cells [259]. Finally, the coated DOX-NPs were functionalized with TRAIL using cysteine thiol-conjugation. This nanoplatform decreased the number of circulating tumor cells and lung metastases in vivo and significantly reduced the number of tumor metastases [217]. In a similar approach, Suryaprakash and colleagues designed a hybrid TRAIL-MSC/nanocomposite spheroid system to target glioblastoma (GBM) cells [260]. The nanocomposite encapsulated mitoxantrone (MTX, a genotoxic drug/topoisomerase II inhibitor) and its surface was functionalized with a peptide to target interleukin-13 receptor alpha 2 (IL13Rα2), which is frequently overexpressed in GBM cells [261]. Finally, a spheroid hybrid TRAIL-MSC/nanocomposite was also shown to be effective [260]. These were generated by passing TRAIL-secreting MSCs and modified nanocomposites through a microfluidic system. This hybrid spheroid had the advantage of expressing TRAIL in a true, membrane-bound conformation.
Feng and colleagues addressed the question of whether NP-based TRAIL delivery could instigate an anti-tumor immune response, thus facilitating long-term disease control using another polymer NP, periodic mesoporous organosilica (PMO) [262]. DOX was encapsulated into a PMO NP and the NP surface was modified with TRAIL. This nanoplatform induced an immune response by activating dendritic cells, CD4+ and CD8+ T cells and inhibited tumor growth in a breast cancer mouse model [262].
More recent studies used liposomes to deliver TRAIL-chemotherapeutic drug combinations. This nanoplatform resulted in remarkable cell death compared to single treatments with DOX or liposome-presented TRAIL-NPs in the H-1080 melanoma mouse model [263]. In a similar study, TRAIL-functionalized liposomes loaded with polymeric micelles modified with piperlongumine, a natural alkaloid that triggers oxidative stress also showed enhanced TRAIL-induced apoptosis in TRAIL-resistant prostate cancer cell lines [264].
Overall, these data show that nanotools that can co-deliver TRAIL and chemotherapeutic agents significantly enhance the efficacy of TRAIL, making these nanoplatforms a promising therapeutic tool as they can overcome tumor TRAIL resistance by simultaneously activating both the intrinsic and the extrinsic apoptosis pathways.

4.2.6. Combination of ROS Producing-Nanoparticles with TRAIL

Metal NPs have been reported to induce cell death through inducing oxidative stress [265,266] owing to the fact that they could generate reactive oxygen species, such as singlet oxygens (1O2) [267,268,269,270,271] or hydrogen peroxide (H2O2) [272]. In the presence of ferrous or cuprous ions, H2O2 can be converted into OH through the Fenton reaction [269,273]. OH then reacts with proteins, lipids and DNA that damages the cell and leads to cell death [274]. Production of OH• from H2O2 is not limited to the ferrous or cuprous ions, other ions such as Fe, Mn, Cu, Co, etc., also can generate OH through Fenton-like reactions [272,275,276]. Since oxidative stress has been shown to sensitize some cancer cells to TRAIL, numerous studies have demonstrated that co-delivering TRAIL with metal NPs could achieve the same effect [184,277]. The challenge in using metal NPs is their limited flexibility for oriented conjugation of proteins, such as TRAIL and accordingly, studies to date have predominantly relied on electrostatic- and amide-based conjugation on iron and silver NPs (Figure 3) [278,279].
Although the delivery of TRAIL via these NPs is a promising approach, it can be challenging to link proteins to metal NPs in an oriented and stable manner. This can lead to the unspecific distribution of the metal NPs throughout the body causing widespread oxidative stress linked to kidney toxicity [280]. ROS can also disrupt cell membranes, including that of healthy cells, emphasizing that the safe application of metal NPs would require their selective targeting to cancer cells [281].

4.2.7. Combination Phototherapy with TRAIL

Recently, NP-based phototherapies such as photothermal therapy (PTT) and photodynamic therapy (PDT) have received considerable attention for cancer treatment. NP-based phototherapies can directly induce cell death through generating ROS (in PDT) and/or converting light energy into heat (in PTT) [282,283,284,285] and the first studies assessing this methodology to overcome TRAIL resistance have produced promising results. They encapsulated iron oxide nanoparticles and photothermal metallo-aromatic complexes (Ph556) into a hybrid micellar NP to generate a nanocomposite, and the surface of the nanocomposites was decorated with TRAIL via electrostatic interactions (TRAIL nanocomposites). They found that TRAIL nanocomposites significantly improved efficacy against TRAIL-resistant lung tumor xenografts (A549 cells) and the sensitization was associated with DR4/DR5 induction [286]. In another approach, outer membrane vesicles (OMV) from E. coli ectopically expressing human TRAIL were functionalized with an ανβ3 integrin-targeting ligand and the photothermal agent, indocyanine green [287]. Upon NIR irritation, indocyanine green produced singlet oxygen that led to oxidative stress and consequent TRAIL sensitization of resistant tumors [287]. Based on these results, TRAIL can be incorporated into PTT nanosystems without a problem of losing its biological activity. These systems enable not only the active targeting of tumor cells based on tumor antigens, but also site-specific activation of multiple pro-apoptotic mechanisms to effectively eradicate tumor cells while posing minimal risk to non-malignant tissues. To date, the design and generation of these nano-approaches is however labor-intensive and needs further optimization.

5. Conclusions and Future Perspectives

One of the most promising mechanisms to selectively eliminate malignant cells is by inducing their suicide program using death ligands. Regarding the death ligands, TRAIL has received the most attention as a potential cancer therapeutic molecule as it induces apoptosis selectively in cancer cells without affecting healthy, non-malignant cells. However, the low biological activity of TRAIL as a soluble protein, its short in vivo half-life time, insufficient accumulation in tumor tissues and acquired resistance of tumor cells to TRAIL-induced cell death showed poor efficacy in clinical trials.
In the last two decades, nanotechnology has opened a new avenue for cancer therapy, including that for TRAIL. Conjugating TRAIL to nanoparticles has addressed the poor pro-apoptotic activity of soluble rhTRAIL and prolonged its circulatory half-life. Furthermore, co-encapsulating drugs and conjugating TRAIL to NPs with particular properties, such as metal-based NPs, has led to overcoming TRAIL resistance in specific cancer models. Additionally, the co-functionalization of NPs with TRAIL and monoclonal antibodies or other molecules such as hyaluronic acid, has allowed specific tumor-targeting, thus enhancing TRAIL accumulation in tumor sites.
Overall, TRAIL-based NPs could address most of the limitations of soluble rhTRAIL therapy, and these nanoformulations have the potential to become a potent and safe treatment for cancer. However, all these formulations are still in preclinical development stage, and none of them have been translated to the clinic yet. The following considerations may pave the way to the design of an effective TRAIL-based nano-system that can potentially translate into the clinic.
-
Most TRAIL-NP formulations have only been tested in immunodeficient mouse models engrafted with cancer cell lines. These models cannot replicate crucial properties of the human body to assess targeting efficiency, the heterogeneity of primary tumors or the role of the immune system. Testing in patient-derived tumor xenografts or genetically engineered mouse models is crucial to progress TRAIL-NPs towards clinical development.
-
With the advance of personalized therapy, it is becoming possible to target the exact molecular mechanisms of TRAIL resistance in individual tumors. Cells become sensitive to TRAIL during the first stages of malignant transformation [288], but this sensitivity is lost as the tumor evolves by the up-regulation of one (but not multiple) inhibitory pathways (e.g., increased Bcl-2 expression, XIAP expression) [289]. On the contrary, healthy cells have redundant (multiple) mechanisms to maintain their TRAIL resistance [289]. Targeting the specific apoptosis inhibitor, instead of combining TRAIL with a broad-spectrum, and often toxic cell stressor (e.g., doxorubicin), is likely to achieve higher efficiency and/or lower systemic toxicity.
-
In most of the current formulations for co-delivery of TRAIL and TRAIL sensitizers, TRAIL is conjugated to the surface of the NPs, and its sensitizers are loaded inside the NPs. This means that first, TRAIL engages with DRs on the target cell, and after that, or in parallel, when the cell endocytoses the NPs, the sensitizing agent can exert its effect, when it might be too late. Formulations for time-dependent release nanoparticles, or development of sequential treatment regimes, might solve this issue.
-
Selective tumor-targeting with higher-specificity tumor markers have a great potential to maximize safety as well as potency of TRAIL-NPs and enable controlled activation and/or cargo-release. Thus, identifying and targeting more specific cancer biomarkers, such as CLL-1 (targeting leukemic stem cells) is required in order to minimize off-target toxicity [290].
Overall, nanoparticles have opened a vast array of opportunities and mechanisms to overcome the limitations of the traditional, soluble protein-based TRAIL therapy, and they also offer a spectrum of solutions to personalize TRAIL-based cancer therapy. The results of multiple preclinical models prove that in the right formulation TRAIL has a strong therapeutic potential which warrants progression into more advanced preclinical models and towards clinical trials.

Author Contributions

The concept, content and structure were formulated by H.A.Z. and E.S.; H.A.Z. and E.S. contributed to literature research, interpretation, writing, editing and proof-reading of manuscript; Figures were generated by H.A.Z. using BioRender. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Blood Cancer Network Ireland (BCNI) program funded by Science Foundation Ireland and the Irish Cancer Society (14-ICS-B3042), DISCOVER-RISE program (H2020-RISE-777995), the NUI, Galway College of Science Scholarship, and the NUI, Galway Thomas Crawford Hayes fund to H.A.Z.

Conflicts of Interest

The authors have no competing interests to declare.

References

  1. Ashkenazi, A.; Shahrokh, Z.; Schwall, R.H.; Ashkenazi, A.; Pai, R.C.; Fong, S.; Leung, S.; Lawrence, D.A.; Marsters, S.A.; Blackie, C.; et al. Schwall, Safety and antitumor activity of recombinant soluble Apo2 ligand. J. Clin. Investig. 1999, 104, 155–162. [Google Scholar] [CrossRef] [PubMed]
  2. Voss, O.H.; Arango, D.; Tossey, J.C.; Calero, M.A.V.; Doseff, A.I. Splicing reprogramming of TRAIL/DISC-components sensitizes lung cancer cells to TRAIL-mediated apoptosis. Cell Death Dis. 2021, 12, 287. [Google Scholar] [CrossRef] [PubMed]
  3. Chen, J.; Knudsen, S.; Mazin, W.; Dahlgaard, J.; Zhang, B. A 71-Gene Signature of TRAIL Sensitivity in Cancer CellsBiomarker of TRAIL Response. Mol. Cancer Ther. 2021, 11, 34–45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Lemke, J.; von Karstedt, S.; Zinngrebe, J.; Walczak, H. Getting TRAIL back on track for cancer therapy. Cell Death Differ. 2014, 21, 1350–1364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Walczak, H.; Miller, R.E.; Ariail, K.; Gliniak, B.; Griffith, T.S.; Kubin, M.; Chin, W.; Jones, J.; Woodward, A.; Le, T.; et al. Tumoricidal activity of tumor necrosis factor–related apoptosis–inducing ligand in vivo. Nat. Med. 1999, 5, 157–163. [Google Scholar] [CrossRef]
  6. Yagita, H.; Takeda, K.; Hayakawa, Y.; Smyth, M.J.; Okumura, K. TRAIL and its receptors as targets for cancer therapy. Cancer Sci. 2004, 95, 777–783. [Google Scholar] [CrossRef] [Green Version]
  7. Robertson, N.M.; Zangrilli, J.G.; Hastie, A.; Lindemeyer, R.G.; Planeta, A.; Smith, M.K.; Innocent, N.; Musani, A.; Pascual, R.; Peters, S.; et al. Differential Expression of TRAIL and TRAIL Receptors in Allergic Asthmatics Following Segmental Antigen Challenge: Evidence for a Role of TRAIL in Eosinophil Survival. J. Immunol. 2002, 169, 5986–5996. [Google Scholar] [CrossRef] [Green Version]
  8. Herold, S.; Steinmueller, M.; von Wulff, W.; Cakarova, L.; Pinto, R.; Pleschka, S.; Mack, M.; Kuziel, W.A.; Corazza, N.; Brunner, T.; et al. Lung epithelial apoptosis in infl uenza virus pneumonia: The role of macrophage- expressed TNF-related apoptosis- inducing ligand. J. Exp. Med. 2008, 205, 3065–3077. [Google Scholar] [CrossRef]
  9. Vik, R.; Espevik, T. Lipopolysaccharide induces expression of APO2 ligand/TRAIL in human monocytes and macrophages. Scand. J. Immunol. 2000, 51, 244–250. [Google Scholar]
  10. Tecchio, C.; Huber, V.; Scapini, P.; Calzetti, F.; Margotto, D.; Todeschini, G.; Pilla, L.; Martinelli, G.; Pizzolo, G.; Rivoltini, L. IFNα-stimulated neutrophils and monocytes release a soluble form of TNF-related apoptosis-inducing ligand (TRAIL/Apo-2 ligand) displaying apoptotic activity on leukemic cells. Blood 2004, 103, 3837–3844. [Google Scholar] [CrossRef] [Green Version]
  11. Kamohara, H.; Matsuyama, W.; Shimozato, O.; Abe, K.; Galligan, C. Regulation of tumour necrosis factor-related apoptosis-inducing ligand (TRAIL) and TRAIL receptor expression in human neutrophils. Immunology 2004, 111, 186–194. [Google Scholar] [CrossRef]
  12. Zahn, S.; Rehkämper, C.; Ferring-Schmitt, S.; Bieber, T.; Tüting, T.; Wenzel, J. Interferon-α stimulates TRAIL expression in human keratinocytes and peripheral blood mononuclear cells: Implications for the pathogenesis of cutaneous lupus erythematosus. Br. J. Dermatol. 2011, 165, 1118–1123. [Google Scholar] [CrossRef]
  13. Fanger, N.A.; Maliszewski, C.R.; Schooley, K.; Griffith, T.S. Human Dendritic Cells Mediate Cellular Apoptosis via Tumor Necrosis Factor–related Apoptosis-inducing Ligand (TRAIL). J. Exp. Med. 1999, 190, 1155–1164. [Google Scholar] [CrossRef]
  14. Griffith, B.T.S.; Wiley, S.R.; Kubin, M.Z.; Sedger, L.M.; Maliszewski, C.R.; Fanger, N.A. Monocyte-mediated tumoricidal activity via the tumor necrosis factor–related cytokine, TRAIL. J. Exp. Med. 1999, 189, 1343–1354. [Google Scholar] [CrossRef]
  15. Shimasaki, N.; Jain, A.; Campana, D. NK cells for cancer immunotherapy. Nat. Rev. Drug Discov. 2020, 19, 200–218. [Google Scholar] [CrossRef]
  16. Mariani, S.M.; Krammer, P.H. Surface expression of TRAIL/Apo-2 ligand in activated mouse T and B cells. Eur. J. Immunol. 1998, 28, 1492–1498. [Google Scholar] [CrossRef]
  17. Mariani, S.M.; Krammer, P.H. Differential regulation of TRAIL and CD95 ligand in transformed cells of the T and B lymphocyte lineage. Eur. J. Immunol. 1998, 28, 973–982. [Google Scholar] [CrossRef]
  18. Wajant, H. CD95L/FasL and TRAIL in tumour surveillance and cancer therapy. In The Link Between Inflammation and Cancer; Part of the book series: Cancer Treatment and Research (CTAR, volume 130); Rosen, S.T., Ed.; Springer: Berlin/Heidelberg, Germany, 2006; pp. 141–165. [Google Scholar]
  19. Kelley, S.K.; Ashkenazi, A. Targeting death receptors in cancer with Apo2L/TRAIL. Curr. Opin. Pharmacol. 2004, 4, 333–339. [Google Scholar] [CrossRef]
  20. Smyth, M.J.; Takeda, K.; Hayakawa, Y.; Peschon, J.J.; Van Den Brink, M.R.; Yagita, H. Nature’s TRAIL—On a path to cancer immunotherapy. Immunity 2003, 18, 1–6. [Google Scholar] [CrossRef] [Green Version]
  21. Herbst, R.S.; Eckhardt, S.G.; Kurzrock, R.; Ebbinghaus, S.; O’Dwyer, P.J.; Gordon, M.S.; Novotny, W.; Goldwasser, M.A.; Tohnya, T.M.; Lum, B.L. Phase I dose-escalation study of recombinant human Apo2L/TRAIL, a dual proapoptotic receptor agonist, in patients with advanced cancer. J. Clin. Oncol. 2010, 28, 2839–2846. [Google Scholar] [CrossRef]
  22. Singh, D.; Tewari, M.; Singh, S.; Narayan, G. Revisiting the role of TRAIL/TRAIL-R in cancer biology and therapy. Future Oncol. 2021, 17, 581–596. [Google Scholar] [CrossRef]
  23. Naka, T.; Sugamura, K.; Hylander, B.L.; Widmer, M.B.; Rustum, Y.M.; Repasky, E.A. Effects of tumor necrosis factor-related apoptosis-inducing ligand alone and in combination with chemotherapeutic agents on patients’ colon tumors grown in SCID mice. Cancer Res. 2002, 62, 5800–5806. [Google Scholar]
  24. Ashkenazi, A.; Holland, P.; Eckhardt, S.G. Ligand-based targeting of apoptosis in cancer: The potential of recombinant human apoptosis ligand 2/tumor necrosis factor–related apoptosis-inducing ligand (rhApo2L/TRAIL). J. Clin. Oncol. 2008, 26, 3621–3630. [Google Scholar] [CrossRef]
  25. Stöhr, D.; Jeltsch, A.; Rehm, M. TRAIL receptor signaling: From the basics of canonical signal transduction toward its entanglement with ER stress and the unfolded protein response. Int. Rev. Cell Mol. Biol. 2020, 351, 57–99. [Google Scholar]
  26. Danish, L.; Stöhr, D.; Scheurich, P.; Pollak, N. TRAIL-R3/R4 and Inhibition of TRAIL Signalling in Cancer. In TRAIL, Fas Ligand, TNF and TLR3 in Cancer; Micheau, O., Ed.; Springer: Berlin/Heidelberg, Germany, 2017; pp. 27–57. [Google Scholar]
  27. Ashkenazi, A. Targeting death and decoy receptors of the tumour-necrosis factor superfamily. Nat. Rev. Cancer 2002, 2, 420–430. [Google Scholar] [CrossRef]
  28. Kerr, J.F.; Winterford, C.M.; Harmon, B.V. Apoptosis. Its significance in cancer and cancer therapy. Cancer 1994, 73, 2013–2026. [Google Scholar] [CrossRef]
  29. Sun, S.Y.; Hail, N., Jr.; Lotan, R. Apoptosis as a novel target for cancer chemoprevention. J. Natl. Cancer Inst. 2004, 96, 662–672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Johnstone, R.W.; Ruefli, A.A.; Lowe, S.W. Apoptosis: A link between cancer genetics and chemotherapy. Cell 2002, 108, 153–164. [Google Scholar] [CrossRef] [Green Version]
  31. Cory, S.; Adams, J.M. The Bcl2 family: Regulators of the cellular life-or-death switch. Nat. Rev. Cancer 2002, 2, 647–656. [Google Scholar] [CrossRef] [PubMed]
  32. Zhou, F.; Yang, Y.; Xing, D. Bcl-2 and Bcl-xL play important roles in the crosstalk between autophagy and apoptosis. FEBS J. 2011, 278, 403–413. [Google Scholar] [CrossRef] [PubMed]
  33. Macdonald, D.C.; Ní Chonghaile, T.; Gupta, S.; Samali, A. Bcl-2 family on guard at the ER. Am. J. Physiol. Cell Physiol. 2009, 296, C941–C953. [Google Scholar]
  34. Adams, J.M.; Cory, S. The Bcl-2 apoptotic switch in cancer development and therapy. Oncogene 2007, 26, 1324–1337. [Google Scholar] [CrossRef] [Green Version]
  35. Cory, S.; Huang, D.; Adams, J.M. The Bcl-2 family: Roles in cell survival and oncogenesis. Oncogene 2003, 22, 8590–8607. [Google Scholar] [CrossRef] [Green Version]
  36. Denicourt, C.; Dowdy, S.F. Targeting apoptotic pathways in cancer cells. Science 2004, 305, 1411–1413. [Google Scholar] [CrossRef]
  37. Wei, M.C.; Zong, W.X.; Cheng, E.H.Y.; Lindsten, T.; Panoutsakopoulou, V.; Ross, A.J.; Roth, K.A.; MacGregor, G.R.; Thompson, C.B.; Korsmeyer, S.J. Proapoptotic BAX and BAK: A requisite gateway to mitochondrial dysfunction and death. Science 2001, 292, 727–730. [Google Scholar] [CrossRef] [Green Version]
  38. Lopez, J.; Tait, S.W.G. Mitochondrial apoptosis: Killing cancer using the enemy within. Br. J. Cancer 2015, 112, 957–962. [Google Scholar] [CrossRef] [Green Version]
  39. Kluck, R.M.; Bossy-Wetzel, E.; Green, D.R.; Newmeyer, D.D. The release of cytochrome c from mitochondria: A primary site for Bcl-2 regulation of apoptosis. Science 1997, 275, 1132–1136. [Google Scholar] [CrossRef] [Green Version]
  40. Li, L.Y.; Luo, X.; Wang, X. Endonuclease G is an apoptotic DNase when released from mitochondria. Nature 2001, 412, 95–99. [Google Scholar] [CrossRef]
  41. Strauss, G.; Westhoff, M.A.; Fischer-Posovszky, P.; Fulda, S.; Schanbacher, M.; Eckhoff, S.M.; Stahnke, K.; Vahsen, N.; Kroemer, G.; Debatin, K.-M. 4-hydroperoxy-cyclophosphamide mediates caspase-independent T-cell apoptosis involving oxidative stress-induced nuclear relocation of mitochondrial apoptogenic factors AIF and EndoG. Cell Death Differ. 2008, 15, 332–343. [Google Scholar] [CrossRef] [Green Version]
  42. Du, C.; Fang, M.; Li, Y.; Li, L.; Wang, X. Smac, a mitochondrial protein that promotes cytochrome c–dependent caspase activation by eliminating IAP inhibition. Cell 2000, 102, 33–42. [Google Scholar] [CrossRef] [Green Version]
  43. Elmallah, M.I.; Micheau, O. Marine drugs regulating apoptosis induced by tumor necrosis factor-related apoptosis-inducing ligand (TRAIL). Mar. Drugs 2015, 13, 6884–6909. [Google Scholar] [CrossRef] [Green Version]
  44. Green, D.R.; Kroemer, G. The pathophysiology of mitochondrial cell death. Science 2004, 305, 626–630. [Google Scholar] [CrossRef]
  45. Ashkenazi, A.; Dixit, V.M. Death receptors: Signaling and modulation. Science 1998, 281, 1305–1309. [Google Scholar] [CrossRef] [Green Version]
  46. Reed, J.C. Apoptosis-targeted therapies for cancer. Cancer Cell 2003, 3, 17–22. [Google Scholar] [CrossRef]
  47. Hunter, A.M.; LaCasse, E.C.; Korneluk, R.G. The inhibitors of apoptosis (IAPs) as cancer targets. Apoptosis 2007, 12, 1543–1568. [Google Scholar]
  48. Fu, T.M.; Li, Y.; Lu, A.; Li, Z.; Vajjhala, P.R.; Cruz, A.C.; Srivastava, D.B.; DiMaio, F.; Penczek, P.A.; Siegel, R.M.; et al. Cryo-EM structure of caspase-8 tandem DED filament reveals assembly and regulation mechanisms of the death-inducing signaling complex. Mol. Cell 2016, 64, 236–250. [Google Scholar] [CrossRef] [Green Version]
  49. Schleich, K.; Buchbinder, J.H.; Pietkiewicz, S.; Kähne, T.; Warnken, U.; Öztürk, S.; Schnölzer, M.; Naumann, M.; Krammer, P.H.; Lavrik, I.N. Molecular architecture of the DED chains at the DISC: Regulation of procaspase-8 activation by short DED proteins c-FLIP and procaspase-8 prodomain. Cell Death Differ. 2016, 23, 681–694. [Google Scholar] [CrossRef] [Green Version]
  50. Dickens, L.S.; Boyd, R.S.; Jukes-Jones, R.; Hughes, M.A.; Robinson, G.L.; Fairall, L.; Schwabe, J.W.R.; Cain, K.; MacFarlane, M. A death effector domain chain DISC model reveals a crucial role for caspase-8 chain assembly in mediating apoptotic cell death. Mol. Cell 2012, 47, 291–305. [Google Scholar] [CrossRef] [Green Version]
  51. Humphreys, L.M.; Fox, J.P.; Higgins, C.A.; Majkut, J.; Sessler, T.; McLaughlin, K.; McCann, C.; Roberts, J.Z.; Crawford, N.T.; McDade, S.S.; et al. A revised model of TRAIL-R2 DISC assembly explains how FLIP (L) can inhibit or promote apoptosis. EMBO Rep. 2020, 21, e49254. [Google Scholar] [CrossRef]
  52. Sessler, T.; Healy, S.; Samali, A.; Szegezdi, E. Structural determinants of DISC function: New insights into death receptor-mediated apoptosis signalling. Pharmacol. Ther. 2013, 140, 186–199. [Google Scholar] [CrossRef] [PubMed]
  53. Eskes, R.; Desagher, S.; Antonsson, B.; Martinou, J.C. Bid induces the oligomerization and insertion of Bax into the outer mitochondrial membrane. Mol. Cell. Biol. 2000, 20, 929–935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Özören, N.; El-Deiry, W.S. Defining characteristics of Types I and II apoptotic cells in response to TRAIL. Neoplasia 2002, 4, 551–557. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Barnhart, B.C.; Alappat, E.C.; Peter, M.E. The CD95 type I/type II model. Semin. Immunol. 2003, 15, 185–193. [Google Scholar] [CrossRef]
  56. Algeciras-Schimnich, A.; Pietras, E.M.; Barnhart, B.C.; Legembre, P.; Vijayan, S.; Holbeck, S.L.; Peter, M.E. Two CD95 tumor classes with different sensitivities to antitumor drugs. Proc. Natl. Acad. Sci. USA 2003, 100, 11445–11450. [Google Scholar] [CrossRef] [Green Version]
  57. Lee, E.F.; Harris, T.J.; Tran, S.; Evangelista, M.; Arulananda, S.; John, T.; Ramnac, C.; Hobbs, C.; Zhu, H.; Gunasingh, G. BCL-XL and MCL-1 are the key BCL-2 family proteins in melanoma cell survival. Cell Death Dis. 2019, 10, 342. [Google Scholar] [CrossRef]
  58. Mukherjee, N.; Skees, J.; Todd, K.J.; West, D.A.; Lambert, K.A.; Robinson, W.A.; Amato, C.M.; Couts, K.L.; van Gulick, R.; MacBeth, M.; et al. MCL1 inhibitors S63845/MIK665 plus Navitoclax synergistically kill difficult-to-treat melanoma cells. Cell Death Dis. 2020, 11, 443. [Google Scholar] [CrossRef]
  59. Mukherjee, N.; Almeida, A.; Partyka, K.A.; Lu, Y.; Schwan, J.V.; Lambert, K.; Rogers, M.; Robinson, W.A.; Robinson, S.E.; Applegate, A.J.; et al. Combining a GSI and BCL-2 inhibitor to overcome melanoma’s resistance to current treatments. Oncotarget 2016, 7, 84594. [Google Scholar] [CrossRef] [Green Version]
  60. Hata, A.N.; Engelman, J.A.; Faber, A.C. The BCL2 Family: Key Mediators of the Apoptotic Response to Targeted Anticancer TherapeuticsBCL2 Family and Targeted Therapies. Cancer Discov. 2015, 5, 475–487. [Google Scholar] [CrossRef] [Green Version]
  61. Li, W.; Zhang, C.; Chen, C.; Zhuang, G. Correlation between expression of DcR3 on tumor cells and sensitivity to FasL. Cell. Mol. Immunol. 2007, 4, 455–460. [Google Scholar]
  62. Mocellin, S.; Rossi, C.R.; Pilati, P.; Nitti, D. Tumor necrosis factor, cancer and anticancer therapy. Cytokine Growth Factor Rev. 2005, 16, 35–53. [Google Scholar] [CrossRef]
  63. Eberle, J. Countering TRAIL resistance in melanoma. Cancers 2019, 11, 656. [Google Scholar] [CrossRef] [Green Version]
  64. Panaitiu, A.E.; Basiashvili, T.; Mierke, D.F.; Pellegrini, M. An engineered construct of cFLIP provides insight into DED1 structure and interactions. Structure 2022, 30, 229–239. [Google Scholar] [CrossRef]
  65. Hajra, K.M.; Liu, J.R. Apoptosome dysfunction in human cancer. Apoptosis 2004, 9, 691–704. [Google Scholar] [CrossRef]
  66. Danial, N.N.; Korsmeyer, S.J. Cell death: Critical control points. Cell 2004, 116, 205–219. [Google Scholar] [CrossRef] [Green Version]
  67. Riedl, S.J.; Renatus, M.; Schwarzenbacher, R.; Zhou, Q.; Sun, C.; Fesik, S.W.; Liddington, R.C.; Salvesen, G.S. Structural basis for the inhibition of caspase-3 by XIAP. Cell 2001, 104, 791–800. [Google Scholar] [CrossRef]
  68. Chai, J.; Shiozaki, E.; Srinivasula, S.M.; Wu, Q.; Dataa, P.; Alnemri, E.S.; Shi, Y. Structural basis of caspase-7 inhibition by XIAP. Cell 2001, 104, 769–780. [Google Scholar] [CrossRef]
  69. Wang, C.Y.; Mayo, M.W.; Korneluk, R.G.; Goeddel, D.V.; Baldwin, A.S. NF-κB antiapoptosis: Induction of TRAF1 and TRAF2 and c-IAP1 and c-IAP2 to suppress caspase-8 activation. Science 1998, 281, 1680–1683. [Google Scholar] [CrossRef]
  70. Chen, C.; Edelstein, L.C.; Gélinas, C. The Rel/NF-κB family directly activates expression of the apoptosis inhibitor Bcl-xL. Mol. Cell. Biol. 2000, 20, 2687–2695. [Google Scholar] [CrossRef]
  71. Zarnegar, B.J.; Wang, Y.; Mahoney, D.J.; Dempsey, P.W.; Cheung, H.H.; He, J.; Shiba, T.; Yang, X.; Yeh, W.; Mak, T.W.; et al. Noncanonical NF-κB activation requires coordinated assembly of a regulatory complex of the adaptors cIAP1, cIAP2, TRAF2 and TRAF3 and the kinase NIK. Nat. Immunol. 2008, 9, 1371–1378. [Google Scholar] [CrossRef] [Green Version]
  72. Chu, Z.L.; McKinsey, T.A.; Liu, L.; Gentry, J.J.; Malim, M.H.; Ballard, D.W. Suppression of tumor necrosis factor-induced cell death by inhibitor of apoptosis c-IAP2 is under NF-κB control. Proc. Natl. Acad. Sci. USA 1997, 94, 10057–10062. [Google Scholar] [CrossRef] [Green Version]
  73. Karin, M.; Lin, A. NF-κB at the crossroads of life and death. Nat. Immunol. 2002, 3, 221–227. [Google Scholar] [CrossRef]
  74. Yuan, X.; Gajan, A.; Chu, Q.; Xiong, H.; Wu, K.; Wu, G.S. Developing TRAIL/TRAIL death receptor-based cancer therapies. Cancer Metastasis Rev. 2018, 37, 733–748. [Google Scholar] [CrossRef]
  75. Caliceti, P.; Veronese, F.M. Pharmacokinetic and biodistribution properties of poly (ethylene glycol)–protein conjugates. Adv. Drug Deliv. Rev. 2003, 55, 1261–1277. [Google Scholar] [CrossRef]
  76. Seifert, O.; Plappert, A.; Fellermeier, S.; Siegemund, M.; Pfizenmaier, K.; Kontermann, R.E. Tetravalent Antibody–scTRAIL Fusion Proteins with Improved PropertiesTetravalent Fusion Proteins. Mol. Cancer Ther. 2014, 13, 101–111. [Google Scholar] [CrossRef] [Green Version]
  77. Rozga, P.; Kloska, D.; Pawlak, S.; Teska-Kaminska, M.; Galazka, M.; Bukato, K.; Pieczykolan, A.; Jaworski, A.; Molga-Kaczmarska, A.; Kopacz, A.; et al. Novel engineered TRAIL-based chimeric protein strongly inhibits tumor growth and bypasses TRAIL resistance. Int. J. Cancer 2020, 147, 1117–1130. [Google Scholar] [CrossRef]
  78. Yan, C.; Li, S.; Li, Z.; Peng, H.; Yuan, X.; Jiang, L.; Zhang, Y.; Fan, D.; Hu, X.; Yang, M.; et al. Human umbilical cord mesenchymal stem cells as vehicles of CD20-specific TRAIL fusion protein delivery: A double-target therapy against non-Hodgkin’s lymphoma. Mol. Pharm. 2013, 10, 142–151. [Google Scholar] [CrossRef]
  79. Trebing, J.; El-Mesery, M.; Schäfer, V.; Weisenberger, D.; Siegmund, D.; Silence, K.; Wajant, H. CD70-restricted specific activation of TRAILR1 or TRAILR2 using scFv-targeted TRAIL mutants. Cell Death Dis. 2014, 5, e1035. [Google Scholar] [CrossRef] [Green Version]
  80. Müller, N.; Schneider, B.; Pfizenmaier, K.; Wajant, H. Superior serum half life of albumin tagged TNF ligands. Biochem. Biophys. Res. Commun. 2010, 396, 793–799. [Google Scholar] [CrossRef]
  81. Hendriks, D.; Choi, G.; de Bruyn, M.; Wiersma, V.R.; Bremer, E. Antibody-based cancer therapy: Successful agents and novel approaches. Int. Rev. Cell Mol. Biol. 2017, 331, 289–383. [Google Scholar]
  82. Kholodenko, R.V.; Kalinovsky, D.V.; Doronin, I.I.; Ponomarev, E.D.; Kholodenko, I.V. Antibody fragments as potential biopharmaceuticals for cancer therapy: Success and limitations. Curr. Med. Chem. 2019, 26, 396–426. [Google Scholar] [CrossRef]
  83. Madhumathi, J.; Sridevi, S.; Verma, R.S. Novel TNF-related apoptotic-inducing ligand-based immunotoxin for therapeutic targeting of CD25 positive leukemia. Target. Oncol. 2016, 11, 535–547. [Google Scholar] [CrossRef] [PubMed]
  84. Wajant, H.; Gerspach, J.; Pfizenmaier, K. Engineering death receptor ligands for cancer therapy. Cancer Lett. 2013, 332, 163–174. [Google Scholar] [CrossRef] [PubMed]
  85. Assohou-Luty, C.; Gerspach, J.; Siegmund, D.; Müller, N.; Huard, B.; Tiegs, G.; Pfizenmaier, K.; Wajant, H. A CD40–CD95L fusion protein interferes with CD40L-induced prosurvival signaling and allows membrane CD40L-restricted activation of CD95. J. Mol. Med. 2006, 84, 785–797. [Google Scholar] [CrossRef] [PubMed]
  86. Bremer, E.; van Dam, G.M.; de Bruyn, M.; van Riezen, M.; Dijkstra, M.; Kamps, G.; Helfrich, W.; Haisma, H. Potent systemic anticancer activity of adenovirally expressed EGFR-selective TRAIL fusion protein. Mol. Ther. 2008, 16, 1919–1926. [Google Scholar] [CrossRef] [Green Version]
  87. Stieglmaier, J.; Bremer, E.; Kellner, C.; Liebig, T.M.; Cate, B.T.; Peipp, M.; Schulze-Koops, H.; Pfeiffer, M.; Bühring, H.-J.; Greil, J.; et al. Selective induction of apoptosis in leukemic B-lymphoid cells by a CD19-specific TRAIL fusion protein. Cancer Immunol. Immunother. 2008, 57, 233–246. [Google Scholar] [CrossRef]
  88. Li, R.; Yang, H.; Jia, D.; Nie, Q.; Cai, H.; Fan, Q.; Wan, L.; Li, L.; Lu, X. Fusion to an albumin-binding domain with a high affinity for albumin extends the circulatory half-life and enhances the in vivo antitumor effects of human TRAIL. J. Control. Release 2016, 228, 96–106. [Google Scholar] [CrossRef]
  89. Byeon, H.J.; Min, S.Y.; Kim, I.; Lee, E.S.; Oh, K.T.; Shin, B.S.; Lee, K.C.; Youn, Y.S. Human serum albumin-TRAIL conjugate for the treatment of rheumatoid arthritis. Bioconjug. Chem. 2014, 25, 2212–2221. [Google Scholar] [CrossRef]
  90. Galley, H.F.; Webster, N.R. Physiology of the endothelium. Br. J. Anaesth. 2004, 93, 105–113. [Google Scholar] [CrossRef] [Green Version]
  91. Brin, E.; Wu, K.; Dagostino, E.; Kuo, M.M.C.; He, Y.; Shia, W.J.; Chen, L.C.; Stempniak, M.; Hickey, R.; Almassy, R.; et al. TRAIL stabilization and cancer cell sensitization to its pro-apoptotic activity achieved through genetic fusion with arginine deiminase. Oncotarget 2018, 9, 36914. [Google Scholar] [CrossRef]
  92. Von Karstedt, S.; Montinaro, A.; Walczak, H. Exploring the TRAILs less travelled: TRAIL in cancer biology and therapy. Nat. Rev. Cancer 2017, 17, 352–366. [Google Scholar] [CrossRef]
  93. Naval, J.; de Miguel, D.; Gallego-Lleyda, A.; Anel, A.; Martinez-Lostao, L. Importance of TRAIL molecular anatomy in receptor oligomerization and signaling. Implications for cancer therapy. Cancers 2019, 11, 444. [Google Scholar] [CrossRef] [Green Version]
  94. Wang, Z.; Yu, B.; Wang, B.; Yan, J.; Feng, X.; Wang, Z.; Wang, L.; Zhang, H.; Wu, H.; Wu, J.; et al. A novel capsid-modified oncolytic recombinant adenovirus type 5 for tumor-targeting gene therapy by intravenous route. Oncotarget 2016, 7, 47287. [Google Scholar] [CrossRef] [Green Version]
  95. Wang, Z.; Liu, W.; Wang, L.; Gao, P.; Li, Z.; Wu, J.; Zhang, H.; Wu, H.; Kong, W.; Yu, B.; et al. Enhancing the antitumor activity of an engineered TRAIL-coated oncolytic adenovirus for treating acute myeloid leukemia. Signal Transduct. Target. Ther. 2020, 5, 40. [Google Scholar] [CrossRef] [Green Version]
  96. Wang, C.; He, H.; Dou, G.; Li, J.; Zhang, X.; Jiang, M.; Li, P.; Huang, X.; Chen, H.; Li, L.; et al. Ginsenoside 20 (S)-Rh2 induces apoptosis and differentiation of acute myeloid leukemia cells: Role of orphan nuclear receptor Nur77. J. Agric. Food Chem. 2017, 65, 7687–7697. [Google Scholar] [CrossRef]
  97. Le, D.H.; Commandeur, U.; Steinmetz, N.F. Presentation and delivery of tumor necrosis factor-related apoptosis-inducing ligand via elongated plant viral nanoparticle enhances antitumor efficacy. ACS Nano 2019, 13, 2501–2510. [Google Scholar] [CrossRef]
  98. Lee, K.L.; Murray, A.A.; Le, D.H.T.; Sheen, M.R.; Shukla, S.; Commandeur, U.; Fiering, S.; Steinmetz, N.F. Combination of plant virus nanoparticle-based in situ vaccination with chemotherapy potentiates antitumor response. Nano Lett. 2017, 17, 4019–4028. [Google Scholar] [CrossRef] [Green Version]
  99. Le, D.H.; Lee, K.L.; Shukla, S.; Commandeur, U.; Steinmetz, N.F. Potato virus X, a filamentous plant viral nanoparticle for doxorubicin delivery in cancer therapy. Nanoscale 2017, 9, 2348–2357. [Google Scholar] [CrossRef]
  100. Le, D.H.; Hu, H.; Commandeur, U.; Steinmetz, N.F. Chemical addressability of potato virus X for its applications in bio/nanotechnology. J. Struct. Biol. 2017, 200, 360–368. [Google Scholar] [CrossRef]
  101. Zhang, J.; Huang, X.; Wang, H.; Liu, X.; Zhang, T.; Wang, Y.; Hu, D. The challenges and promises of allogeneic mesenchymal stem cells for use as a cell-based therapy. Stem Cell Res. Ther. 2015, 6, 234. [Google Scholar] [CrossRef] [Green Version]
  102. Menon, L.G.; Kelly, K.; Yang, H.W.; Kim, S.; Black, P.M.; Carroll, R.S. Human bone marrow-derived mesenchymal stromal cells expressing S-TRAIL as a cellular delivery vehicle for human glioma therapy. Stem Cells 2009, 27, 2320–2330. [Google Scholar] [CrossRef]
  103. Shah, K. Mesenchymal stem cells engineered for cancer therapy. Adv. Drug Deliv. Rev. 2012, 64, 739–748. [Google Scholar] [CrossRef] [Green Version]
  104. Golinelli, G.; Grisendi, G.; Prapa, M.; Bestagno, M.; Spano, C.; Rossignoli, F.; Bambi, F.; Sardi, I.; Cellini, M.; Horwitz, E.M.; et al. Targeting GD2-positive glioblastoma by chimeric antigen receptor empowered mesenchymal progenitors. Cancer Gene Ther. 2020, 27, 558–570. [Google Scholar] [CrossRef] [Green Version]
  105. Li, J.; Sharkey, C.C.; Wun, B.; Liesveld, J.L.; King, M.R. Genetic engineering of platelets to neutralize circulating tumor cells. J. Control. Release 2016, 228, 38–47. [Google Scholar] [CrossRef] [Green Version]
  106. Kauer, T.M.; Figueiredo, J.L.; Hingtgen, S.; Shah, K. Encapsulated therapeutic stem cells implanted in the tumor resection cavity induce cell death in gliomas. Nat. Neurosci. 2012, 15, 197–204. [Google Scholar] [CrossRef] [Green Version]
  107. Labusca, L.; Herea, D.D.; Mashayekhi, K. Stem cells as delivery vehicles for regenerative medicine-challenges and perspectives. World J. Stem Cells 2018, 10, 43. [Google Scholar] [CrossRef]
  108. Kawada, H.; Fujita, J.; Kinjo, K.; Matsuzaki, Y.; Tsuma, M.; Miyatake, H.; Muguruma, Y.; Tsuboi, K.; Itabashi, Y.; Ikeda, Y.; et al. Nonhematopoietic mesenchymal stem cells can be mobilized and differentiate into cardiomyocytes after myocardial infarction. Blood 2004, 104, 3581–3587. [Google Scholar] [CrossRef]
  109. Choi, Y.U.; Yoon, Y.; Jung, P.Y.; Hwang, S.; Hong, J.E.; Kim, W.-S.; Sohn, J.H.; Rhee, K.-J.; Bae, K.S.; Eom, Y.W. TRAIL-Overexpressing Adipose Tissue-Derived Mesenchymal Stem Cells Efficiently Inhibit Tumor Growth in an H460 Xenograft Model. Cancer Genom. Proteom. 2021, 18, 569–578. [Google Scholar] [CrossRef]
  110. Yoon, N.; Park, M.S.; Peltier, G.C.; Lee, R.H. Pre-activated human mesenchymal stromal cells in combination with doxorubicin synergistically enhance tumor-suppressive activity in mice. Cytotherapy 2015, 17, 1332–1341. [Google Scholar] [CrossRef]
  111. Kim, S.W.; Kim, S.J.; Park, S.H.; Yang, H.G.; Kang, M.C.; Choi, Y.W.; Kim, S.M.; Jeun, S.-S.; Sung, Y.C. Complete Regression of Metastatic Renal Cell Carcinoma by Multiple Injections of Engineered Mesenchymal Stem Cells Expressing Dodecameric TRAIL and HSV-TKComplete Cure of Metastatic Tumor by Engineered MSCs. Clin. Cancer Res. 2013, 19, 415–427. [Google Scholar] [CrossRef] [Green Version]
  112. Redjal, N.; Zhu, Y.; Shah, K. Combination of Systemic Chemotherapy with Local Stem Cell Delivered S-TRAIL in Resected Brain Tumors. Stem Cells 2015, 33, 101–110. [Google Scholar] [CrossRef] [Green Version]
  113. Kim, S.M.; Woo, J.S.; Jeong, C.H.; Ryu, C.H.; Jang, J.D.; Jeun, S.S. Potential application of temozolomide in mesenchymal stem cell-based TRAIL gene therapy against malignant glioma. Stem Cells Transl. Med. 2014, 3, 172–182. [Google Scholar] [CrossRef] [PubMed]
  114. Kim, S.M.; Lim, J.Y.; Park, S.I.; Jeong, C.H.; Oh, J.H.; Jeong, M.; Oh, W.; Park, S.H.; Sung, Y.C.; Jeun, S.-S. Gene therapy using TRAIL-secreting human umbilical cord blood–derived mesenchymal stem cells against intracranial glioma. Cancer Res. 2008, 68, 9614–9623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Zhang, B.; Shan, H.; Li, D.; Li, Z.R.; Zhu, K.S.; Jiang, Z.B. The inhibitory effect of MSCs expressing TRAIL as a cellular delivery vehicle in combination with cisplatin on hepatocellular carcinoma. Cancer Biol. Ther. 2012, 13, 1175–1184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Grisendi, G.; Bussolari, R.; Cafarelli, L.; Petak, I.; Rasini, V.; Veronesi, E.; de Santis, G.; Spano, C.; Tagliazzucchi, M.; Barti-Juhasz, H.; et al. Adipose-derived mesenchymal stem cells as stable source of tumor necrosis factor–related apoptosis-inducing ligand delivery for cancer therapy. Cancer Res. 2010, 70, 3718–3729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Mueller, L.P.; Luetzkendorf, J.; Widder, M.; Nerger, K.; Caysa, H.; Mueller, T. TRAIL-transduced multipotent mesenchymal stromal cells (TRAIL-MSC) overcome TRAIL resistance in selected CRC cell lines in vitro and in vivo. Cancer Gene Ther. 2011, 18, 229–239. [Google Scholar] [CrossRef] [PubMed]
  118. Mohr, A.; Albarenque, S.M.; Deedigan, L.; Yu, R.; Reidy, M.; Fulda, S.; Zwacka, R.M. Targeting of XIAP combined with systemic mesenchymal stem cell-mediated delivery of sTRAIL ligand inhibits metastatic growth of pancreatic carcinoma cells. Stem Cells 2010, 28, 2109–2120. [Google Scholar] [CrossRef] [PubMed]
  119. Kim, H.J.; Park, J.S. Usage of human mesenchymal stem cells in cell-based therapy: Advantages and disadvantages. Dev. Reprod. 2017, 21, 1. [Google Scholar] [CrossRef]
  120. Nishikawa, S.; Goldstein, R.A.; Nierras, C.R. The promise of human induced pluripotent stem cells for research and therapy. Nat. Rev. Mol. Cell Biol. 2008, 9, 725–729. [Google Scholar] [CrossRef]
  121. Kimbrel, E.A.; Lanza, R. Next-generation stem cells—Ushering in a new era of cell-based therapies. Nat. Rev. Drug Discov. 2020, 19, 463–479. [Google Scholar] [CrossRef]
  122. Anselmo, A.C.; Mitragotri, S. Cell-mediated delivery of nanoparticles: Taking advantage of circulatory cells to target nanoparticles. J. Control. Release 2014, 190, 531–541. [Google Scholar] [CrossRef] [Green Version]
  123. Mitchell, M.J.; Jain, R.K.; Langer, R. Engineering and physical sciences in oncology: Challenges and opportunities. Nat. Rev. Cancer 2017, 17, 659–675. [Google Scholar] [CrossRef]
  124. Wong, C.; Stylianopoulos, T.; Cui, J.; Martin, J.; Chauhan, V.P.; Jiang, W.; Popović, Z.; Jain, R.K.; Bawendi, M.G.; Fukumura, D. Multistage nanoparticle delivery system for deep penetration into tumor tissue. Proc. Natl. Acad. Sci. USA 2011, 108, 2426–2431. [Google Scholar] [CrossRef] [Green Version]
  125. Chauhan, V.P.; Jain, R.K. Strategies for advancing cancer nanomedicine. Nat. Mater. 2013, 12, 958–962. [Google Scholar] [CrossRef] [Green Version]
  126. Sindhwani, S.; Syed, A.M.; Ngai, J.; Kingston, B.R.; Maiorino, L.; Rothschild, J.; MacMillan, P.; Zhang, Y.; Rajesh, N.U.; Hoang, T.; et al. The entry of nanoparticles into solid tumours. Nat. Mater. 2020, 19, 566–575. [Google Scholar] [CrossRef]
  127. Nam, J.; Son, S.; Park, K.S.; Zou, W.; Shea, L.D.; Moon, J.J. Cancer nanomedicine for combination cancer immunotherapy. Nat. Rev. Mater. 2019, 4, 398–414. [Google Scholar] [CrossRef]
  128. Kih, M.; Lee, E.J.; Lee, N.K.; Kim, Y.K.; Lee, K.E.; Jeong, C.; Yang, Y.; Kim, D.H.; Kim, I.S. Designed trimer-mimetic TNF superfamily ligands on self-assembling nanocages. Biomaterials 2018, 180, 67–77. [Google Scholar] [CrossRef]
  129. De Miguel, D.; Gallego-Lleyda, A.; Ayuso, J.M.; Pejenaute-Ochoa, D.; Jarauta, V.; Marzo, I.; Fernández, L.J.; Ochoa, I.; Conde, B.; Anel, A.; et al. High-order TRAIL oligomer formation in TRAIL-coated lipid nanoparticles enhances DR5 cross-linking and increases antitumour effect against colon cancer. Cancer Lett. 2016, 383, 250–260. [Google Scholar] [CrossRef]
  130. Jiang, X.; Fitch, S.; Wang, C.; Wilson, C.; Li, J.; Grant, G.A.; Yang, F. Nanoparticle engineered TRAIL-overexpressing adipose-derived stem cells target and eradicate glioblastoma via intracranial delivery. Proc. Natl. Acad. Sci. USA 2016, 113, 13857–13862. [Google Scholar] [CrossRef] [Green Version]
  131. De Miguel, D.; Gallego-Lleyda, A.; Ayuso, J.M.; Erviti-Ardanaz, S.; Pazo-Cid, R.; del Agua, C.; Fernández, L.J.; Ochoa, I.; Anel, A.; Martinez-Lostao, L. TRAIL-coated lipid-nanoparticles overcome resistance to soluble recombinant TRAIL in non-small cell lung cancer cells. Nanotechnology 2016, 27, 185101. [Google Scholar] [CrossRef]
  132. De Miguel, D.; Lemke, J.; Anel, A.; Walczak, H.; Martinez-Lostao, L. Onto better TRAILs for cancer treatment. Cell Death Differ. 2016, 23, 733–747. [Google Scholar] [CrossRef] [Green Version]
  133. Rozanov, D.V.; Savinov, A.Y.; Golubkov, V.S.; Rozanova, O.L.; Postnova, T.I.; Sergienko, E.A.; Vasile, S.; Aleshin, A.E.; Rega, M.F.; Pellecchia, M.; et al. Engineering a leucine zipper-TRAIL homotrimer with improved cytotoxicity in tumor cells. Mol. Cancer Ther. 2009, 8, 1515–1525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Huang, Y.J.; Hsu, S. TRAIL-functionalized gold nanoparticles selectively trigger apoptosis in polarized macrophages. Nanotheranostics 2017, 1, 326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Belkahla, H.; Haque, A.; Revzin, A.; Gharbi, T.; Constantinescu, A.A.; Micheau, O.; Hémadi, M.; Ammar, S. Coupling tumor necrosis factor-related apoptosis-inducing ligand to iron oxide nanoparticles increases its apoptotic activity on HCT116 and HepG2 malignant cells: Effect of magnetic core size. J. Interdiscip. Nanomed. 2019, 4, 34–50. [Google Scholar] [CrossRef]
  136. Ke, S.; Zhou, T.; Yang, P.; Wang, Y.; Zhang, P.; Chen, K.; Ren, L.; Ye, S. Gold nanoparticles enhance TRAIL sensitivity through Drp1-mediated apoptotic and autophagic mitochondrial fission in NSCLC cells. Int. J. Nanomed. 2017, 12, 2531. [Google Scholar] [CrossRef] [Green Version]
  137. Kim, H.; Jeong, D.; Kang, H.E.; Lee, K.C.; Na, K. A sulfate polysaccharide/TNF-related apoptosis-inducing ligand (TRAIL) complex for the long-term delivery of TRAIL in poly (lactic-co-glycolic acid)(PLGA) microspheres. J. Pharm. Pharmacol. 2013, 65, 11–21. [Google Scholar] [CrossRef]
  138. Chen, Z.; Penet, M.F.; Krishnamachary, B.; Banerjee, S.R.; Pomper, M.G.; Bhujwalla, Z.M. PSMA-specific theranostic nanoplex for combination of TRAIL gene and 5-FC prodrug therapy of prostate cancer. Biomaterials 2016, 80, 57–67. [Google Scholar] [CrossRef] [Green Version]
  139. Huang, S.; Li, J.; Han, L.; Liu, S.; Ma, H.; Huang, R.; Jiang, C. Dual targeting effect of Angiopep-2-modified, DNA-loaded nanoparticles for glioma. Biomaterials 2011, 32, 6832–6838. [Google Scholar] [CrossRef]
  140. Hu, C.; Gu, F.; Tai, Z.; Yao, C.; Gong, C.; Xia, Q.; Gao, Y.; Gao, S. Synergistic effect of reduced polypeptide micelle for co-delivery of doxorubicin and TRAIL against drug-resistance in breast cancer. Oncotarget 2016, 7, 61832. [Google Scholar] [CrossRef] [Green Version]
  141. Nair, P.M.; Flores, H.; Gogineni, A.; Marsters, S.; Lawrence, D.A.; Kelley, R.F.; Ngu, H.; Sagolla, M.; Komuves, L.; Bourgon, R.; et al. Enhancing the antitumor efficacy of a cell-surface death ligand by covalent membrane display. Proc. Natl. Acad. Sci. USA 2015, 112, 5679–5684. [Google Scholar] [CrossRef] [Green Version]
  142. Gallego-Lleyda, A.; de Miguel, D.; Anel, A.; Martinez-Lostao, L. Lipid nanoparticles decorated with TNF-related aptosis-inducing ligand (TRAIL) are more cytotoxic than soluble recombinant TRAIL in sarcoma. Int. J. Mol. Sci. 2018, 19, 1449. [Google Scholar] [CrossRef] [Green Version]
  143. De Miguel, D.; Basáñez, G.; Sánchez, D.; Malo, P.G.; Marzo, I.; Larrad, L.; Naval, J.; Pardo, J.; Anel, A.; Martinez-Lostao, L. Liposomes decorated with Apo2L/TRAIL overcome chemoresistance of human hematologic tumor cells. Mol. Pharm. 2013, 10, 893–904. [Google Scholar] [CrossRef]
  144. Kim, T.H.; Jiang, H.-H.; Youn, Y.S.; Park, C.W.; Lim, S.M.; Jin, C.-H.; Tak, K.K.; Lee, H.S.; Lee, K.C. Preparation and characterization of Apo2L/TNF-related apoptosis-inducing ligand–loaded human serum albumin nanoparticles with improved stability and tumor distribution. J. Pharm. Sci. 2011, 100, 482–491. [Google Scholar] [CrossRef]
  145. Li, H.; Zhao, J.; Wang, A.; Li, Q.; Cui, W. Supramolecular assembly of protein-based nanoparticles based on tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) for cancer therapy. Colloids Surf. A Physicochem. Eng. Asp. 2020, 590, 124486. [Google Scholar] [CrossRef]
  146. Chen, Y.F.; Chen, G.Y.; Chang, C.H.; Su, Y.C.; Chen, Y.C.; Jiang, Y.; Jan, J.S. TRAIL encapsulated to polypeptide-crosslinked nanogel exhibits increased anti-inflammatory activities in Klebsiella pneumoniae-induced sepsis treatment. Mater. Sci. Eng. C 2019, 102, 85–95. [Google Scholar] [CrossRef]
  147. Arroyo, N.; Herlem, G.; Picaud, F. Ligand nanovectorization using graphene to target cellular death receptors of cancer cell. Proteins Struct. Funct. Bioinform. 2020, 88, 94–105. [Google Scholar] [CrossRef]
  148. Hu, C.M.J.; Zhang, L.; Aryal, S.; Cheung, C.; Fang, R.H.; Zhang, L. Erythrocyte membrane-camouflaged polymeric nanoparticles as a biomimetic delivery platform. Proc. Natl. Acad. Sci. USA 2011, 108, 10980–10985. [Google Scholar] [CrossRef] [Green Version]
  149. He, Z.; Zhang, Y.; Feng, N. Cell membrane-coated nanosized active targeted drug delivery systems homing to tumor cells: A review. Mater. Sci. Eng. C 2020, 106, 110298. [Google Scholar] [CrossRef]
  150. Luo, G.F.; Chen, W.H.; Zeng, X.; Zhang, X.Z. Cell primitive-based biomimetic functional materials for enhanced cancer therapy. Chem. Soc. Rev. 2021, 50, 945–985. [Google Scholar] [CrossRef]
  151. Dash, P.; Piras, A.M.; Dash, M. Cell membrane coated nanocarriers-an efficient biomimetic platform for targeted therapy. J. Control. Release 2020, 327, 546–570. [Google Scholar] [CrossRef]
  152. Liu, W.; Yan, Q.; Xia, C.; Wang, X.; Kumar, A.; Wang, Y.; Liu, Y.; Pan, Y.; Liu, J. Recent advances in cell membrane coated metal–organic frameworks (MOFs) for tumor therapy. J. Mater. Chem. B 2021, 9, 4459–4474. [Google Scholar] [CrossRef]
  153. Li, J.; Ai, Y.; Wang, L.; Bu, P.; Sharkey, C.C.; Wu, Q.; Wun, B.; Roy, S.; Shen, X.; King, M.R. Targeted drug delivery to circulating tumor cells via platelet membrane-functionalized particles. Biomaterials 2016, 76, 52–65. [Google Scholar] [CrossRef] [Green Version]
  154. Xu, R.; Rai, A.; Chen, M.; Suwakulsiri, W.; Greening, D.W.; Simpson, R.J. Extracellular vesicles in cancer—Implications for future improvements in cancer care. Nat. Rev. Clin. Oncol. 2018, 15, 617–638. [Google Scholar] [CrossRef]
  155. Chang, C.H.; Pauklin, S. Extracellular vesicles in pancreatic cancer progression and therapies. Cell Death Dis. 2021, 12, 973. [Google Scholar] [CrossRef]
  156. Van Niel, G.; d’Angelo, G.; Raposo, G. Shedding light on the cell biology of extracellular vesicles. Nat. Rev. Mol. Cell Biol. 2018, 19, 213–228. [Google Scholar] [CrossRef]
  157. Trams, E.G.; Lauter, C.J.; Salem, J.N.; Heine, U. Exfoliation of membrane ecto-enzymes in the form of micro-vesicles. Biochim. Biophys. Acta (BBA)-Biomembr. 1981, 645, 63–70. [Google Scholar] [CrossRef]
  158. Li, Q.; Huang, Q.; Huyan, T.; Wang, Y.; Huang, Q.; Shi, J. Bifacial effects of engineering tumour cell-derived exosomes on human natural killer cells. Exp. Cell Res. 2018, 363, 141–150. [Google Scholar] [CrossRef]
  159. Xu, R.; Greening, D.W.; Zhu, H.J.; Takahashi, N.; Simpson, R.J. Extracellular vesicle isolation and characterization: Toward clinical application. J. Clin. Investig. 2016, 126, 1152–1162. [Google Scholar] [CrossRef]
  160. Kalluri, R.; LeBleu, V.S. The biology, function, and biomedical applications of exosomes. Science 2020, 367, eaau6977. [Google Scholar] [CrossRef]
  161. Robbins, P.D.; Morelli, A.E. Regulation of immune responses by extracellular vesicles. Nat. Rev. Immunol. 2014, 14, 195–208. [Google Scholar] [CrossRef] [Green Version]
  162. Fitzgerald, W.; Freeman, M.L.; Lederman, M.M.; Vasilieva, E.; Romero, R.; Margolis, L. A system of cytokines encapsulated in extracellular vesicles. Sci. Rep. 2018, 8, 8973. [Google Scholar] [CrossRef] [Green Version]
  163. Webber, J.; Steadman, R.; Mason, M.D.; Tabi, Z.; Clayton, A. Cancer exosomes trigger fibroblast to myofibroblast differentiation. Cancer Res. 2010, 70, 9621–9630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Chen, G.; Huang, A.C.; Zhang, W.; Zhang, G.; Wu, M.; Xu, W.; Yu, Z.; Yang, J.; Wang, B.; Sun, H.; et al. Exosomal PD-L1 contributes to immunosuppression and is associated with anti-PD-1 response. Nature 2018, 560, 382–386. [Google Scholar] [CrossRef] [PubMed]
  165. Poggio, M.; Hu, T.; Pai, C.-C.; Chu, B.; Belair, C.D.; Chang, A.; Montabana, E.; Lang, U.E.; Fu, Q.; Fong, L.; et al. Suppression of exosomal PD-L1 induces systemic anti-tumor immunity and memory. Cell 2019, 177, 414–427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Theodoraki, M.N.; Yerneni, S.S.; Hoffmann, T.K.; Gooding, W.E.; Whiteside, T.L. Clinical significance of PD-L1+ exosomes in plasma of head and neck cancer patients. Clin. Cancer Res. 2018, 24, 896–905. [Google Scholar] [CrossRef] [Green Version]
  167. Ludwig, S.; Floros, T.; Theodoraki, M.N.; Hong, C.S.; Jackson, E.K.; Lang, S.; Whiteside, T.L. Suppression of lymphocyte functions by plasma exosomes correlates with disease activity in patients with head and neck cancer. Clin. Cancer Res. 2017, 23, 4843–4854. [Google Scholar] [CrossRef] [Green Version]
  168. Andreola, G.; Rivoltini, L.; Castelli, C.; Huber, V.; Perego, P.; Deho, P.; Squarcina, P.; Accornero, P.; Lozupone, F.; Lugini, L.; et al. Induction of lymphocyte apoptosis by tumor cell secretion of FasL-bearing microvesicles. J. Exp. Med. 2002, 195, 1303–1316. [Google Scholar] [CrossRef] [Green Version]
  169. Abusamra, A.J.; Zhong, Z.; Zheng, X.; Li, M.; Ichim, T.E.; Chin, J.L.; Min, W.P. Tumor exosomes expressing Fas ligand mediate CD8+ T-cell apoptosis. Blood Cells Mol. Dis. 2005, 35, 169–173. [Google Scholar] [CrossRef]
  170. Rivoltini, L.; Chiodoni, C.; Squarcina, P.; Tortoreto, M.; Villa, A.; Vergani, B.; Bürdek, M.; Botti, L.; Arioli, I.; Cova, A.; et al. TNF-related apoptosis-inducing ligand (TRAIL)–armed exosomes deliver proapoptotic signals to tumor site. Clin. Cancer Res. 2016, 22, 3499–3512. [Google Scholar] [CrossRef] [Green Version]
  171. Shamili, F.H.; Bayegi, H.R.; Salmasi, Z.; Sadri, K.; Mahmoudi, M.; Kalantari, M.; Ramezani, M.; Abnous, K. Exosomes derived from TRAIL-engineered mesenchymal stem cells with effective anti-tumor activity in a mouse melanoma model. Int. J. Pharm. 2018, 549, 218. [Google Scholar] [CrossRef]
  172. Yuan, Z.; Kolluri, K.K.; Gowers, K.H.C.; Janes, S.M. TRAIL delivery by MSC-derived extracellular vesicles is an effective anticancer therapy. J. Extracell. Vesicles 2017, 6, 1265291. [Google Scholar] [CrossRef] [Green Version]
  173. Setroikromo, R.; Zhang, B.; Reis, C.R.; Mistry, R.H.; Quax, W.J. Death Receptor 5 Displayed on Extracellular Vesicles Decreases TRAIL Sensitivity of Colon Cancer Cells. Front. Cell Dev. Biol. 2020, 8, 318. [Google Scholar] [CrossRef]
  174. Wang, J.; Chen, D.; Ho, E.A. Challenges in the development and establishment of exosome-based drug delivery systems. J. Control. Release 2020, 329, 894–906. [Google Scholar] [CrossRef]
  175. Gutierrez-Millan, C.; Díaz, C.C.; Lanao, J.M.; Colino, C.I. Advances in Exosomes-Based Drug Delivery Systems. Macromol. Biosci. 2020, 21, 2000269. [Google Scholar] [CrossRef]
  176. Leggio, L.; Arrabito, G.; Ferrara, V.; Vivarelli, S.; Paternò, G.; Marchetti, B.; Pignataro, B.; Iraci, N. Mastering the Tools: Natural versus Artificial Vesicles in Nanomedicine. Adv. Healthc. Mater. 2020, 9, 2000731. [Google Scholar] [CrossRef]
  177. Glantz, M.J.; Jaeckle, K.A.; Chamberlain, M.C.; Phuphanich, S.; Recht, L.; Swinnen, L.J.; Maria, B.; LaFollette, S.; Schumann, G.B.; Cole, B.F.; et al. A randomized controlled trial comparing intrathecal sustained-release cytarabine (DepoCyt) to intrathecal methotrexate in patients with neoplastic meningitis from solid tumors. Clin. Cancer Res. 1999, 5, 3394–3402. [Google Scholar]
  178. Judson, I.; Radford, J.A.; Harris, M.; Blay, J.Y.; van Hoesel, Q.; le Cesne, A.; van Oosterom, A.T.; Clemons, M.J.; Kamby, C.; Hermans, C.; et al. Randomised phase II trial of pegylated liposomal doxorubicin (DOXIL®/CAELYX®) versus doxorubicin in the treatment of advanced or metastatic soft tissue sarcoma: A study by the EORTC Soft Tissue and Bone Sarcoma Group. Eur. J. Cancer 2001, 37, 870–877. [Google Scholar] [CrossRef]
  179. Van Tran, V.; Moon, J.Y.; Lee, Y.C. Liposomes for delivery of antioxidants in cosmeceuticals: Challenges and development strategies. J. Control. Release 2019, 300, 114–140. [Google Scholar] [CrossRef]
  180. Bangham, A.D.; Standish, M.M.; Watkins, J.C. Diffusion of univalent ions across the lamellae of swollen phospholipids. J. Mol. Biol. 1965, 13, 238-IN27. [Google Scholar] [CrossRef]
  181. Bangham, A.D.; Horne, R.W. Negative staining of phospholipids and their structural modification by surface-active agents as observed in the electron microscope. J. Mol. Biol. 1964, 8, 660-IN10. [Google Scholar] [CrossRef]
  182. Pattni, B.S.; Chupin, V.V.; Torchilin, V.P. New developments in liposomal drug delivery. Chem. Rev. 2015, 115, 10938–10966. [Google Scholar] [CrossRef]
  183. Rideau, E.; Dimova, R.; Schwille, P.; Wurm, F.R.; Landfester, K. Liposomes and polymersomes: A comparative review towards cell mimicking. Chem. Soc. Rev. 2018, 47, 8572–8610. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Guimarães, P.P.G.; Gaglione, S.; Sewastianik, T.; Carrasco, R.D.; Langer, R.; Mitchell, M.J. Nanoparticles for immune cytokine TRAIL-based cancer therapy. ACS Nano 2018, 12, 912–931. [Google Scholar] [CrossRef] [PubMed]
  185. Wang, Y.; Kohane, D.S. External triggering and triggered targeting strategies for drug delivery. Nat. Rev. Mater. 2017, 2, 17020. [Google Scholar] [CrossRef]
  186. Lovell, J.F.; Jin, C.S.; Huynh, E.; Jin, H.; Kim, C.; Rubinstein, J.L.; Chan, W.C.W.; Cao, W.; Wang, L.V.; Zheng, G. Porphysome nanovesicles generated by porphyrin bilayers for use as multimodal biophotonic contrast agents. Nat. Mater. 2011, 10, 324–332. [Google Scholar] [CrossRef] [Green Version]
  187. De Miguel, D.; Gallego-Lleyda, A.; Anel, A.; Martinez-Lostao, L. Liposome-bound TRAIL induces superior DR5 clustering and enhanced DISC recruitment in histiocytic lymphoma U937 cells. Leuk. Res. 2015, 39, 657–666. [Google Scholar] [CrossRef]
  188. Petros, R.A.; DeSimone, J.M. Strategies in the design of nanoparticles for therapeutic applications. Nat. Rev. Drug Discov. 2010, 9, 615–627. [Google Scholar] [CrossRef]
  189. Wang, Y.; Li, L.; Shao, N.; Hu, Z.; Chen, H.; Xu, L.; Wang, C.; Cheng, Y.; Xiao, J. Triazine-modified dendrimer for efficient TRAIL gene therapy in osteosarcoma. Acta Biomater. 2015, 17, 115–124. [Google Scholar] [CrossRef]
  190. Kim, T.H.; Jiang, H.H.; Park, C.W.; Youn, Y.S.; Lee, S.; Chen, X.; Lee, K.C. PEGylated TNF-related apoptosis-inducing ligand (TRAIL)-loaded sustained release PLGA microspheres for enhanced stability and antitumor activity. J. Control. Release 2011, 150, 63–69. [Google Scholar] [CrossRef]
  191. Zakaria, A.B.; Picaud, F.; Rattier, T.; Pudlo, M.; Saviot, L.; Chassagnon, R.; Lherminier, J.; Gharbi, T.; Micheau, O.; Herlem, G. Nanovectorization of TRAIL with single wall carbon nanotubes enhances tumor cell killing. Nano Lett. 2015, 15, 891–895. [Google Scholar] [CrossRef]
  192. Ventola, C.L. Progress in nanomedicine: Approved and investigational nanodrugs. Pharm. Ther. 2017, 42, 742. [Google Scholar]
  193. Bobo, D.; Robinson, K.J.; Islam, J.; Thurecht, K.J.; Corrie, S.R. Nanoparticle-based medicines: A review of FDA-approved materials and clinical trials to date. Pharm. Res. 2016, 33, 2373–2387. [Google Scholar] [CrossRef]
  194. Mitchell, M.J.; Billingsley, M.M.; Haley, R.M.; Wechsler, M.E.; Peppas, N.A.; Langer, R. Engineering precision nanoparticles for drug delivery. Nat. Rev. Drug Discov. 2021, 20, 101–124. [Google Scholar] [CrossRef]
  195. Chen, L.; Hong, W.; Ren, W.; Xu, T.; Qian, Z.; He, Z. Recent progress in targeted delivery vectors based on biomimetic nanoparticles. Signal Transduct. Target. Ther. 2021, 6, 225. [Google Scholar] [CrossRef]
  196. Cheng, C.J.; Tietjen, G.T.; Saucier-Sawyer, J.K.; Saltzman, W.M. A holistic approach to targeting disease with polymeric nanoparticles. Nat. Rev. Drug Discov. 2015, 14, 239–247. [Google Scholar] [CrossRef] [Green Version]
  197. Marques, A.C.; Costa, P.J.; Velho, S.; Amaral, M.H. Functionalizing nanoparticles with cancer-targeting antibodies: A comparison of strategies. J. Control. Release 2020, 320, 180–200. [Google Scholar] [CrossRef]
  198. Patra, J.K.; Das, G.; Fraceto, L.F.; Campos, E.V.R.; Rodriguez-Torres, M.d.; Acosta-Torres, L.S.; Diaz-Torres, L.A.; Grillo, R.; Swamy, M.K.; Sharma, S.; et al. Nano based drug delivery systems: Recent developments and future prospects. J. Nanobiotechnol. 2018, 16, 71. [Google Scholar] [CrossRef] [Green Version]
  199. Misra, S.; Heldin, P.; Hascall, V.C.; Karamanos, N.K.; Skandalis, S.S.; Markwald, R.R.; Ghatak, S. Hyaluronan–CD44 interactions as potential targets for cancer therapy. FEBS J. 2011, 278, 1429–1443. [Google Scholar] [CrossRef] [Green Version]
  200. Senbanjo, L.T.; Chellaiah, M.A. CD44: A multifunctional cell surface adhesion receptor is a regulator of progression and metastasis of cancer cells. Front. Cell Dev. Biol. 2017, 5, 18. [Google Scholar] [CrossRef] [Green Version]
  201. Platt, V.M.; Szoka, F.C., Jr. Anticancer therapeutics: Targeting macromolecules and nanocarriers to hyaluronan or CD44, a hyaluronan receptor. Mol. Pharm. 2008, 5, 474–486. [Google Scholar] [CrossRef]
  202. Mitchell, M.J.; Wayne, E.; Rana, K.; Schaffer, C.B.; King, M.R. TRAIL-coated leukocytes that kill cancer cells in the circulation. Proc. Natl. Acad. Sci. USA 2014, 111, 930–935. [Google Scholar] [CrossRef] [Green Version]
  203. Nimrichter, L.; Burdick, M.M.; Aoki, K.; Laroy, W.; Fierro, M.A.; Hudson, S.A.; von Seggern, C.E.; Cotter, R.J.; Bochner, B.S.; Tiemeyer, M.; et al. E-selectin receptors on human leukocytes, Blood. J. Am. Soc. Hematol. 2008, 112, 3744–3752. [Google Scholar]
  204. Larsen, G.R.; Sako, D.; Ahern, T.J.; Shaffer, M.; Erban, J.; Sajer, S.A.; Gibson, R.M.; Wagner, D.D.; Furie, B.C.; Furie, B. P-selectin and E-selectin. Distinct but overlapping leukocyte ligand specificities. J. Biol. Chem. 1992, 267, 11104–11110. [Google Scholar] [CrossRef]
  205. Wayne, E.C.; Chandrasekaran, S.; Mitchell, M.J.; Chan, M.F.; Lee, R.E.; Schaffer, C.B.; King, M.R. TRAIL-coated leukocytes that prevent the bloodborne metastasis of prostate cancer. J. Control. Release 2016, 223, 215–223. [Google Scholar]
  206. Jyotsana, N.; Zhang, Z.; Himmel, L.E.; Yu, F.; King, M.R. Minimal dosing of leukocyte targeting TRAIL decreases triple-negative breast cancer metastasis following tumor resection. Sci. Adv. 2019, 5, eaaw4197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Chandrasekaran, S.; McGuire, M.J.; King, M.R. Sweeping lymph node micrometastases off their feet: An engineered model to evaluate natural killer cell mediated therapeutic intervention of circulating tumor cells that disseminate to the lymph nodes. Lab Chip 2014, 14, 118–127. [Google Scholar] [CrossRef] [PubMed]
  208. Chandrasekaran, S.; Chan, M.F.; Li, J.; King, M.R. Super natural killer cells that target metastases in the tumor draining lymph nodes. Biomaterials 2016, 77, 66–76. [Google Scholar] [PubMed] [Green Version]
  209. O’Reilly, E.; Tirincsi, A.; Logue, S.E.; Szegezdi, E. The Janus face of death receptor signaling during tumor immunoediting. Front. Immunol. 2016, 7, 446. [Google Scholar]
  210. Schreiber, R.D.; Old, L.J.; Smyth, M.J. Cancer immunoediting: Integrating immunity’s roles in cancer suppression and promotion. Science 2011, 331, 1565–1570. [Google Scholar]
  211. Seifert, O.; Pollak, N.; Nusser, A.; Steiniger, F.; Rüger, R.; Pfizenmaier, K.; Kontermann, R.E. Immuno-LipoTRAIL: Targeted delivery of TRAIL-functionalized liposomal nanoparticles. Bioconjug. Chem. 2014, 25, 879–887. [Google Scholar]
  212. Sigismund, S.; Avanzato, D.; Lanzetti, L. Emerging functions of the EGFR in cancer. Mol. Oncol. 2018, 12, 3–20. [Google Scholar] [CrossRef] [Green Version]
  213. Tebbutt, N.; Pedersen, M.W.; Johns, T.G. Targeting the ERBB family in cancer: Couples therapy. Nat. Rev. Cancer 2013, 13, 663–673. [Google Scholar] [CrossRef]
  214. Choi, S.H.; Byeon, H.J.; Choi, J.S.; Thao, L.; Kim, I.; Lee, E.S.; Shin, B.S.; Lee, K.C.; Youn, Y.S. Inhalable self-assembled albumin nanoparticles for treating drug-resistant lung cancer. J. Control. Release 2015, 197, 199–207. [Google Scholar] [CrossRef]
  215. Bae, S.; Ma, K.; Kim, T.H.; Lee, E.S.; Oh, K.T.; Park, E.S.; Lee, K.C.; Youn, Y.S. Doxorubicin-loaded human serum albumin nanoparticles surface-modified with TNF-related apoptosis-inducing ligand and transferrin for targeting multiple tumor types. Biomaterials 2012, 33, 1536–1546. [Google Scholar] [CrossRef]
  216. Sun, X.; Pang, Z.; Ye, H.; Qiu, B.; Guo, L.; Li, J.; Ren, J.; Qian, Y.; Zhang, Q.; Chen, J.; et al. Co-delivery of pEGFP-hTRAIL and paclitaxel to brain glioma mediated by an angiopep-conjugated liposome. Biomaterials 2012, 33, 916–924. [Google Scholar] [CrossRef]
  217. Hu, Q.; Sun, W.; Qian, C.; Wang, C.; Bomba, H.N.; Gu, Z. Anticancer platelet-mimicking nanovehicles. Adv. Mater. 2015, 27, 7043–7050. [Google Scholar] [CrossRef]
  218. Chen, K.; Cao, X.; Li, M.; Su, Y.; Li, H.; Xie, M.; Zhang, Z.; Gao, H.; Xu, X.; Han, Y.; et al. A TRAIL-delivered lipoprotein-bioinspired nanovector engineering stem cell-based platform for inhibition of lung metastasis of melanoma. Theranostics 2019, 9, 2984. [Google Scholar] [CrossRef]
  219. Ren, H.; Zhou, L.; Liu, M.; Lu, W.; Gao, C. Peptide GE11–polyethylene glycol–polyethylenimine for targeted gene delivery in laryngeal cancer. Med. Oncol. 2015, 32, 185. [Google Scholar] [CrossRef]
  220. Chen, Z.; Zhang, L.; He, Y.; Li, Y. Sandwich-type Au-PEI/DNA/PEI-Dexa nanocomplex for nucleus-targeted gene delivery in vitro and in vivo. ACS Appl. Mater. Interfaces 2014, 6, 14196–14206. [Google Scholar] [CrossRef]
  221. Kim, Y.J.; Chae, S.Y.; Jin, C.H.; Sivasubramanian, M.; Son, S.; Choi, K.Y.; Jo, D.G.; Kim, K.; Kwon, I.C.; Lee, K.C.; et al. Ionic complex systems based on hyaluronic acid and PEGylated TNF-related apoptosis-inducing ligand for treatment of rheumatoid arthritis. Biomaterials 2010, 31, 9057–9064. [Google Scholar] [CrossRef]
  222. Jiang, T.; Mo, R.; Bellotti, A.; Zhou, J.; Gu, Z. Gel–liposome-mediated co-delivery of anticancer membrane-associated proteins and small-molecule drugs for enhanced therapeutic efficacy. Adv. Funct. Mater. 2014, 24, 2295–2304. [Google Scholar] [CrossRef]
  223. Yin, P.T.; Shah, S.; Pasquale, N.J.; Garbuzenko, O.B.; Minko, T.; Lee, K.B. Stem cell-based gene therapy activated using magnetic hyperthermia to enhance the treatment of cancer. Biomaterials 2016, 81, 46–57. [Google Scholar] [CrossRef] [Green Version]
  224. Han, J.; Na, K. Transfection of the TRAIL gene into human mesenchymal stem cells using biocompatible polyethyleneimine carbon dots for cancer gene therapy. J. Ind. Eng. Chem. 2019, 80, 722–728. [Google Scholar] [CrossRef]
  225. Kim, I.; Choi, J.S.; Lee, S.; Byeon, H.J.; Lee, E.S.; Shin, B.S.; Choi, H.-G.; Lee, K.C.; Youn, Y.S. In situ facile-forming PEG cross-linked albumin hydrogels loaded with an apoptotic TRAIL protein. J. Control. Release 2015, 214, 30–39. [Google Scholar] [CrossRef]
  226. Huang, R.Y.; Lin, Y.H.; Lin, S.Y.; Li, Y.N.; Chiang, C.S.; Chang, C.W. Magnetic ternary nanohybrids for nonviral gene delivery of stem cells and applications on cancer therapy. Theranostics 2019, 9, 2411. [Google Scholar] [CrossRef]
  227. Li, J.; Guo, Y.; Kuang, Y.; An, S.; Ma, H.; Jiang, C. Choline transporter-targeting and co-delivery system for glioma therapy. Biomaterials 2013, 34, 9142–9148. [Google Scholar] [CrossRef]
  228. Carneiro, B.A.; El-Deiry, W.S. Targeting apoptosis in cancer therapy. Nat. Rev. Clin. Oncol. 2020, 17, 395–417. [Google Scholar] [CrossRef]
  229. Huang, Y.; Yang, X.; Xu, T.; Kong, Q.; Zhang, Y.; Shen, Y.; Wei, Y.; Wang, G.; Chang, K.J. Overcoming resistance to TRAIL-induced apoptosis in solid tumor cells by simultaneously targeting death receptors, c-FLIP and IAPs. Int. J. Oncol. 2016, 49, 153–163. [Google Scholar] [CrossRef] [Green Version]
  230. Kretz, A.-L.; Trauzold, A.; Hillenbrand, A.; Knippschild, U.; Henne-Bruns, D.; von Karstedt, S.; Lemke, J. TRAILblazing strategies for cancer treatment. Cancers 2019, 11, 456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Deng, D.; Shah, K. TRAIL of hope meeting resistance in cancer. Trends Cancer 2020, 6, 989–1001. [Google Scholar] [CrossRef]
  232. Igney, F.H.; Krammer, P.H. Immune escape of tumors: Apoptosis resistance and tumor counterattack. J. Leukoc. Biol. 2002, 71, 907–920. [Google Scholar] [CrossRef]
  233. Zhang, L.; Fang, B. Mechanisms of resistance to TRAIL-induced apoptosis in cancer. Cancer Gene Ther. 2005, 12, 228–237. [Google Scholar] [CrossRef] [PubMed]
  234. He, W.; Liu, Q.; Wang, L.; Chen, W.; Li, N.; Cao, X. TLR4 signaling promotes immune escape of human lung cancer cells by inducing immunosuppressive cytokines and apoptosis resistance. Mol. Immunol. 2007, 44, 2850–2859. [Google Scholar] [CrossRef] [PubMed]
  235. Wagner, K.W.; Punnoose, E.A.; Januario, T.; Lawrence, D.A.; Pitti, R.M.; Lancaster, K.; Lee, D.; von Goetz, M.; Yee, S.F.; Totpal, K.; et al. Death-receptor O-glycosylation controls tumor-cell sensitivity to the proapoptotic ligand Apo2L/TRAIL. Nat. Med. 2007, 13, 1070–1077. [Google Scholar] [CrossRef] [PubMed]
  236. Fan, W.; Yung, B.; Huang, P.; Chen, X. Nanotechnology for multimodal synergistic cancer therapy. Chem. Rev. 2017, 117, 13566–13638. [Google Scholar] [CrossRef]
  237. Dai, W.; Wang, X.; Song, G.; Liu, T.; He, B.; Zhang, H.; Wang, X.; Zhang, Q. Combination antitumor therapy with targeted dual-nanomedicines. Adv. Drug Deliv. Rev. 2017, 115, 23–45. [Google Scholar] [CrossRef]
  238. Guo, L.; Fan, L.; Ren, J.; Pang, Z.; Ren, Y.; Li, J.; Wen, Z.; Jiang, X. A novel combination of TRAIL and doxorubicin enhances antitumor effect based on passive tumor-targeting of liposomes. Nanotechnology 2011, 22, 265105. [Google Scholar] [CrossRef]
  239. Jiang, H.H.; Kim, T.H.; Lee, S.; Chen, X.; Youn, Y.S.; Lee, K.C. PEGylated TNF-related apoptosis-inducing ligand (TRAIL) for effective tumor combination therapy. Biomaterials 2011, 32, 8529–8537. [Google Scholar] [CrossRef]
  240. Kim, I.; Byeon, H.J.; Kim, T.H.; Lee, E.S.; Oh, K.T.; Shin, B.S.; Lee, K.C.; Youn, Y.S. Doxorubicin-loaded porous PLGA microparticles with surface attached TRAIL for the inhalation treatment of metastatic lung cancer. Biomaterials 2013, 34, 6444–6453. [Google Scholar] [CrossRef]
  241. Alvizo-Baez, C.A.; Luna-Cruz, I.E.; Vilches-Cisneros, N.; Rodríguez-Padilla, C.; Alcocer-González, J.M. Systemic delivery and activation of the TRAIL gene in lungs, with magnetic nanoparticles of chitosan controlled by an external magnetic field. Int. J. Nanomed. 2016, 11, 6449. [Google Scholar] [CrossRef] [Green Version]
  242. Cui, W.; Cui, Y.; Zhao, J.; Li, J. Fabrication of tumor necrosis factor-related apoptosis inducing ligand (TRAIL)/ALG modified CaCO3 as drug carriers with the function of tumor selective recognition. J. Mater. Chem. B 2013, 1, 1326–1332. [Google Scholar] [CrossRef]
  243. Perlstein, B.; Finniss, S.A.; Miller, C.; Okhrimenko, H.; Kazimirsky, G.; Cazacu, S.; Lee, H.K.; Lemke, N.; Brodie, S.; Umansky, F.; et al. TRAIL conjugated to nanoparticles exhibits increased anti-tumor activities in glioma cells and glioma stem cells in vitro and in vivo. Neuro-Oncol. 2013, 15, 29–40. [Google Scholar] [CrossRef] [Green Version]
  244. Johnstone, R.W.; Frew, A.J.; Smyth, M.J. The TRAIL apoptotic pathway in cancer onset, progression and therapy. Nat. Rev. Cancer 2008, 8, 782–798. [Google Scholar] [CrossRef]
  245. Ediriwickrema, A.; Saltzman, W.M. Nanotherapy for cancer: Targeting and multifunctionality in the future of cancer therapies. ACS Biomater. Sci. Eng. 2015, 1, 64–78. [Google Scholar] [CrossRef]
  246. Kamaly, N.; Xiao, Z.; Valencia, P.M.; Radovic-Moreno, A.F.; Farokhzad, O.C. Targeted polymeric therapeutic nanoparticles: Design, development and clinical translation. Chem. Soc. Rev. 2012, 41, 2971–3010. [Google Scholar] [CrossRef]
  247. Kamaly, N.; Yameen, B.; Wu, J.; Farokhzad, O.C. Degradable controlled-release polymers and polymeric nanoparticles: Mechanisms of controlling drug release. Chem. Rev. 2016, 116, 2602–2663. [Google Scholar] [CrossRef] [Green Version]
  248. George, A.; Shah, P.A.; Shrivastav, P.S. Natural biodegradable polymers based nano-formulations for drug delivery: A review. Int. J. Pharm. 2019, 561, 244–264. [Google Scholar] [CrossRef]
  249. Feng, C.; Han, X.; Chi, L.; Sun, J.; Gong, F.; Shen, Y. Synthesis, characterization, and in vitro evaluation of TRAIL-modified, cabazitaxel-loaded polymeric micelles for achieving synergistic anticancer therapy. J. Biomater. Sci. Polym. Ed. 2018, 29, 1729–1744. [Google Scholar] [CrossRef]
  250. Sharma, A.K.; Gothwal, A.; Kesharwani, P.; Alsaab, H.; Iyer, A.K.; Gupta, U. Dendrimer nanoarchitectures for cancer diagnosis and anticancer drug delivery. Drug Discov. Today 2017, 22, 314–326. [Google Scholar] [CrossRef]
  251. Chis, A.A.; Dobrea, C.; Morgovan, C.; Arseniu, A.M.; Rus, L.L.; Butuca, A.; Juncan, A.M.; Totan, M.; Vonica-Tincu, A.L.; Cormos, G.; et al. Applications and limitations of dendrimers in biomedicine. Molecules 2020, 25, 3982. [Google Scholar] [CrossRef]
  252. Liu, S.; Guo, Y.; Huang, R.; Li, J.; Huang, S.; Kuang, Y.; Han, L.; Jiang, C. Gene and doxorubicin co-delivery system for targeting therapy of glioma. Biomaterials 2012, 33, 4907–4916. [Google Scholar] [CrossRef]
  253. Jhaveri, A.; Deshpande, P.; Pattni, B.; Torchilin, V. Transferrin-targeted, resveratrol-loaded liposomes for the treatment of glioblastoma. J. Control. Release 2018, 277, 89–101. [Google Scholar] [CrossRef]
  254. Johnsen, K.B.; Moos, T. Revisiting nanoparticle technology for blood–brain barrier transport: Unfolding at the endothelial gate improves the fate of transferrin receptor-targeted liposomes. J. Control. Release 2016, 222, 32–46. [Google Scholar] [CrossRef]
  255. Pang, Z.; Gao, H.; Yu, Y.; Chen, J.; Guo, L.; Ren, J.; Wen, Z.; Su, J.; Jiang, X. Brain delivery and cellular internalization mechanisms for transferrin conjugated biodegradable polymersomes. Int. J. Pharm. 2011, 415, 284–292. [Google Scholar] [CrossRef]
  256. Pishavar, E.; Ramezani, M.; Hashemi, M. Co-delivery of doxorubicin and TRAIL plasmid by modified PAMAM dendrimer in colon cancer cells, in vitro and in vivo evaluation. Drug Dev. Ind. Pharm. 2019, 45, 1931–1939. [Google Scholar] [CrossRef]
  257. Wang, Y.; Wang, M.; Chen, H.; Liu, H.; Zhang, Q.; Cheng, Y. Fluorinated dendrimer for TRAIL gene therapy in cancer treatment. J. Mater. Chem. B 2016, 4, 1354–1360. [Google Scholar] [CrossRef]
  258. Merten, M.; Thiagarajan, P. P-selectin in arterial thrombosis. Z. Kardiol. 2004, 93, 855–863. [Google Scholar] [CrossRef]
  259. Li, L.; Ding, Q.; Zhou, J.; Wu, Y.; Zhang, M.; Guo, X.; Long, M.; Lü, S. Distinct binding kinetics of E-, P-and L-selectins to CD44. FEBS J. 2021, 289, 2877–2894. [Google Scholar] [CrossRef]
  260. Suryaprakash, S.; Lao, Y.-H.; Cho, H.-Y.; Li, M.; Ji, H.Y.; Shao, D.; Hu, H.; Quek, C.H.; Huang, D.; Mintz, R.L.; et al. Engineered mesenchymal stem cell/nanomedicine spheroid as an active drug delivery platform for combinational glioblastoma therapy. Nano Lett. 2019, 19, 1701–1705. [Google Scholar]
  261. Brown, C.E.; Badie, B.; Barish, M.E.; Weng, L.; Ostberg, J.R.; Chang, W.-C.; Naranjo, A.; Starr, R.; Wagner, J.; Wright, C.; et al. Bioactivity and safety of IL13Rα2-redirected chimeric antigen receptor CD8+ T cells in patients with recurrent glioblastoma. Clin. Cancer Res. 2015, 21, 4062–4072. [Google Scholar] [CrossRef] [Green Version]
  262. Feng, X.; Li, F.; Zhang, L.; Liu, W.; Wang, X.; Zhu, R.; Qiao, Z.-A.; Yu, B.; Yu, X. TRAIL-modified, doxorubicin-embedded periodic mesoporous organosilica nanoparticles for targeted drug delivery and efficient antitumor immunotherapy. Acta Biomater. 2022, 143, 392–405. [Google Scholar]
  263. De Miguel, D.; Gallego-Lleyda, A.; Martinez-Ara, M.; Plou, J.; Anel, A.; Martinez-Lostao, L. Double-edged lipid nanoparticles combining liposome-bound TRAIL and encapsulated doxorubicin showing an extraordinary synergistic pro-apoptotic potential. Cancers 2019, 11, 1948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Zhang, Z.; Patel, S.B.; King, M.R. Micelle-in-Liposomes for Sustained Delivery of Anticancer Agents That Promote Potent TRAIL-Induced Cancer Cell Apoptosis. Molecules 2021, 26, 157. [Google Scholar] [CrossRef] [PubMed]
  265. Zhang, X.; Zhang, H.; Liang, X.; Zhang, J.; Tao, W.; Zhu, X.; Chang, D.; Zeng, X.; Liu, G.; Mei, L. Iron oxide nanoparticles induce autophagosome accumulation through multiple mechanisms: Lysosome impairment, mitochondrial damage, and ER stress. Mol. Pharm. 2016, 13, 2578–2587. [Google Scholar] [CrossRef]
  266. Wu, Y.N.; Yang, L.X.; Shi, X.Y.; Li, I.C.; Biazik, J.M.; Ratinac, K.R.; Chen, D.H.; Thordarson, P.; Shieh, D.B.; Braet, F. The selective growth inhibition of oral cancer by iron core-gold shell nanoparticles through mitochondria-mediated autophagy. Biomaterials 2011, 32, 4565–4573. [Google Scholar] [CrossRef] [PubMed]
  267. Azizi, M.; Ghourchian, H.; Yazdian, F.; Dashtestani, F.; AlizadehZeinabad, H. Cytotoxic effect of albumin coated copper nanoparticle on human breast cancer cells of MDA-MB 231. PLoS ONE 2017, 12, e0188639. [Google Scholar]
  268. Azizi, M.; Ghourchian, H.; Yazdian, F.; Alizadehzeinabad, H. Albumin coated cadmium nanoparticles as chemotherapeutic agent against MDA-MB 231 human breast cancer cell line. Artif. Cells Nanomed. Biotechnol. 2018, 46, 787–797. [Google Scholar] [CrossRef] [Green Version]
  269. Yang, B.; Chen, Y.; Shi, J. Reactive oxygen species (ROS)-based nanomedicine. Chem. Rev. 2019, 119, 4881–4985. [Google Scholar]
  270. Chen, Z.; Yin, J.-J.; Zhou, Y.-T.; Zhang, Y.; Song, L.; Song, M.; Hu, S.; Gu, N. Dual enzyme-like activities of iron oxide nanoparticles and their implication for diminishing cytotoxicity. ACS Nano 2012, 6, 4001–4012. [Google Scholar] [CrossRef]
  271. Khan, M.I.; Mohammad, A.; Patil, G.; Naqvi, S.A.H.; Chauhan, L.K.S.; Ahmad, I. Induction of ROS, mitochondrial damage and autophagy in lung epithelial cancer cells by iron oxide nanoparticles. Biomaterials 2012, 33, 1477–1488. [Google Scholar]
  272. Ranji-Burachaloo, H.; Gurr, P.A.; Dunstan, D.E.; Qiao, G.G. Cancer treatment through nanoparticle-facilitated fenton reaction. ACS Nano 2018, 12, 11819–11837. [Google Scholar] [CrossRef]
  273. DeBerardinis, R.J.; Chandel, N.S. Fundamentals of cancer metabolism. Sci. Adv. 2016, 2, e1600200. [Google Scholar] [CrossRef] [Green Version]
  274. Qian, X.; Zhang, J.; Gu, Z.; Chen, Y. Nanocatalysts-augmented Fenton chemical reaction for nanocatalytic tumor therapy. Biomaterials 2019, 211, 1–13. [Google Scholar] [CrossRef]
  275. Gao, S.; Jin, Y.; Ge, K.; Li, Z.; Liu, H.; Dai, X.; Zhang, Y.; Chen, S.; Liang, X.; Zhang, J. Self-supply of O2 and H2O2 by a Nanocatalytic medicine to enhance combined chemo/Chemodynamic therapy. Adv. Sci. 2019, 6, 1902137. [Google Scholar] [CrossRef] [Green Version]
  276. Bokare, A.D.; Choi, W. Review of iron-free Fenton-like systems for activating H2O2 in advanced oxidation processes. J. Hazard. Mater. 2014, 275, 121–135. [Google Scholar] [CrossRef]
  277. Dianat-Moghadam, H.; Heidarifard, M.; Mahari, A.; Shahgolzari, M.; Keshavarz, M.; Nouri, M.; Amoozgar, Z. TRAIL in oncology: From recombinant TRAIL to nano-and self-targeted TRAIL-based therapies. Pharmacol. Res. 2020, 155, 104716. [Google Scholar] [CrossRef]
  278. Shi, Y.; Wang, J.; Liu, J.; Lin, G.; Xie, F.; Pang, X.; Pei, Y.; Cheng, Y.; Zhang, Y.; Lin, Z.; et al. Oxidative stress-driven DR5 upregulation restores TRAIL/Apo2L sensitivity induced by iron oxide nanoparticles in colorectal cancer. Biomaterials 2020, 233, 119753. [Google Scholar] [CrossRef]
  279. Sur-Erdem, I.; Muslu, K.; Pınarbası, N.; Altunbek, M.; Seker-Polat, F.; Cingöz, A.; Aydın, S.O.; Kahraman, M.; Culha, M.; Solaroglu, I.; et al. TRAIL-conjugated silver nanoparticles sensitize glioblastoma cells to TRAIL by regulating CHK1 in the DNA repair pathway. Neurol. Res. 2020, 42, 1061–1069. [Google Scholar] [CrossRef]
  280. Makhdoumi, P.; Karimi, H.; Khazaei, M. Review on metal-based nanoparticles: Role of reactive oxygen species in renal toxicity. Chem. Res. Toxicol. 2020, 33, 2503–2514. [Google Scholar] [CrossRef]
  281. Attarilar, S.; Yang, J.; Ebrahimi, M.; Wang, Q.; Liu, J.; Tang, Y.; Yang, J. The toxicity phenomenon and the related occurrence in metal and metal oxide nanoparticles: A brief review from the biomedical perspective. Front. Bioeng. Biotechnol. 2020, 8, 822. [Google Scholar] [CrossRef]
  282. Khan, S.; Hasan, A.; Attar, F.; Babadaei, M.M.N.; Zeinabad, H.A.; Salehi, M.; Alizadeh, M.; Hassan, M.; Derakhshankhah, H.; Hamblin, M.R. Diagnostic and drug release systems based on microneedle arrays in breast cancer therapy. J. Control. Release 2021, 338, 341–357. [Google Scholar] [CrossRef]
  283. Liu, Y.; Bhattarai, P.; Dai, Z.; Chen, X. Photothermal therapy and photoacoustic imaging via nanotheranostics in fighting cancer. Chem. Soc. Rev. 2019, 48, 2053–2108. [Google Scholar] [CrossRef] [PubMed]
  284. Pham, T.C.; Nguyen, V.-N.; Choi, Y.; Lee, S.; Yoon, J. Recent strategies to develop innovative photosensitizers for enhanced photodynamic therapy. Chem. Rev. 2021, 121, 13454–13619. [Google Scholar] [CrossRef] [PubMed]
  285. Zhi, D.; Yang, T.; O’hagan, J.; Zhang, S.; Donnelly, R.F. Photothermal therapy. J. Control. Release 2020, 325, 52–71. [Google Scholar] [CrossRef] [PubMed]
  286. Lin, G.; Zhang, Y.; Zhu, C.; Chu, C.; Shi, Y.; Pang, X.; Ren, E.; Wu, Y.; Mi, P.; Xia, H.; et al. Photo-excitable hybrid nanocomposites for image-guided photo/TRAIL synergistic cancer therapy. Biomaterials 2018, 176, 60–70. [Google Scholar] [CrossRef]
  287. Peng, L.-H.; Wang, M.-Z.; Chu, Y.; Zhang, L.; Niu, J.; Shao, H.-T.; Yuan, T.-J.; Jiang, Z.-H.; Gao, J.-Q.; Ning, X.-H. Engineering bacterial outer membrane vesicles as transdermal nanoplatforms for photo-TRAIL–programmed therapy against melanoma. Sci. Adv. 2020, 6, eaba2735. [Google Scholar] [CrossRef]
  288. Nesterov, A.; Nikrad, M.; Johnson, T.; Kraft, A.S. Oncogenic Ras sensitizes normal human cells to tumor necrosis factor-α-related apoptosis-inducing ligand-induced apoptosis. Cancer Res. 2004, 64, 3922–3927. [Google Scholar] [CrossRef] [Green Version]
  289. Van Dijk, M.; Halpin-McCormick, A.; Sessler, T.; Samali, A.; Szegezdi, E. Resistance to TRAIL in non-transformed cells is due to multiple redundant pathways. Cell Death Dis. 2013, 4, e702. [Google Scholar] [CrossRef] [Green Version]
  290. Gurney, M.; O’Reilly, E.; Corcoran, S.; Brophy, S.; Hardwicke, D.; Krawczyk, J.; Hermanson, D.; Childs, R.W.; Szegezdi, E.; O’Dwyer, M.E. Tc Buster Transposon Engineered CLL-1 CAR-NK Cells Efficiently Target Acute Myeloid Leukemia. Blood 2021, 138, 1725. [Google Scholar] [CrossRef]
Figure 1. The TRAIL-induced cell death pathway and its regulation. Interaction of TRAIL with DR4 and DR5 leads to the recruitment of the adaptor protein, FADD, which then binds pro-caspase-8 and/or -10 leading to their activation. Activated caspase-8/-10 are released from the DISC into the cytoplasm, where they activate the downstream, effector caspases, which by cleaving vital proteins commit the cell to die. Caspase-8 can also cleave the BH3-only protein, Bid thus activating it. Truncated Bid (tBid) then binds and activates Bax and Bak, leading to mitochondrial membrane permeabilization and release of apoptotic factors such as Cyt c and Smac into the cytosol. Cyt c binds to APAF-1 thus initiating its oligomerization and recruitment of pro-caspase-9. The formed protein complex is called the apoptosome, serving as an activation platform for pro-caspase-9. Activated caspase-9 cleaves and activates the effector caspases, caspase-3/-6/-7, thus amplifying the DR-initiated apoptotic signal. Apoptosis signaling is kept under control at all stages of the process. First, decoy receptors (DcR1, DcR2) and osteoprotegerin (OPG) can sequester TRAIL thus preventing DR4/5 activation. cFLIP can interact and inhibit pro-caspase-8 in the DISC. At the mitochondrion, antiapoptotic Bcl-2 proteins, e.g., Bcl-2 itself and Bcl-xL inhibit Smac and Cyt c release. XIAP can inhibit caspase-3/-7/-9 activation; while Smac can block XIAP, thus allowing the apoptotic program to progress. TRAIL, tumor necrosis factor (TNF)-related apoptosis-inducing ligand; DR, death receptor; OPG, osteoprotegerin; DcR, decoy receptor; DD, death domain, FADD, Fas-associated death domain; Active/Pro-C, pro-caspase; DISC, death-inducing signaling complex; cFLIP, Cellular FLICE inhibitory protein; Cyt c, cytochrome c; APAF-1, apoptotic protease-activating factor-1; Smac, XIAP, X-linked inhibitor of apoptosis protein; Bak, BCL-2 antagonist/killer; Bax, Bcl-2 associated X protein. Figure was generated with BioRender.
Figure 1. The TRAIL-induced cell death pathway and its regulation. Interaction of TRAIL with DR4 and DR5 leads to the recruitment of the adaptor protein, FADD, which then binds pro-caspase-8 and/or -10 leading to their activation. Activated caspase-8/-10 are released from the DISC into the cytoplasm, where they activate the downstream, effector caspases, which by cleaving vital proteins commit the cell to die. Caspase-8 can also cleave the BH3-only protein, Bid thus activating it. Truncated Bid (tBid) then binds and activates Bax and Bak, leading to mitochondrial membrane permeabilization and release of apoptotic factors such as Cyt c and Smac into the cytosol. Cyt c binds to APAF-1 thus initiating its oligomerization and recruitment of pro-caspase-9. The formed protein complex is called the apoptosome, serving as an activation platform for pro-caspase-9. Activated caspase-9 cleaves and activates the effector caspases, caspase-3/-6/-7, thus amplifying the DR-initiated apoptotic signal. Apoptosis signaling is kept under control at all stages of the process. First, decoy receptors (DcR1, DcR2) and osteoprotegerin (OPG) can sequester TRAIL thus preventing DR4/5 activation. cFLIP can interact and inhibit pro-caspase-8 in the DISC. At the mitochondrion, antiapoptotic Bcl-2 proteins, e.g., Bcl-2 itself and Bcl-xL inhibit Smac and Cyt c release. XIAP can inhibit caspase-3/-7/-9 activation; while Smac can block XIAP, thus allowing the apoptotic program to progress. TRAIL, tumor necrosis factor (TNF)-related apoptosis-inducing ligand; DR, death receptor; OPG, osteoprotegerin; DcR, decoy receptor; DD, death domain, FADD, Fas-associated death domain; Active/Pro-C, pro-caspase; DISC, death-inducing signaling complex; cFLIP, Cellular FLICE inhibitory protein; Cyt c, cytochrome c; APAF-1, apoptotic protease-activating factor-1; Smac, XIAP, X-linked inhibitor of apoptosis protein; Bak, BCL-2 antagonist/killer; Bax, Bcl-2 associated X protein. Figure was generated with BioRender.
Cancers 14 05125 g001
Figure 2. Correlation between TRAIL formulations and anticancer therapeutic potential. (a) RhTRAIL-based therapy, (b) crosslinking TRAIL monomers through an anti-tag antibody or (c) protein fusions facilitating TRAIL oligomerization (i.e., leucine zipper-TRAIL) [5,133]. (d) Conjugation of antibody fragment to TRAIL using a linker (i.e., scFv425) [86] to target tumor cells. (e) Conjugation of TRAIL to nanoparticles (NPs) to mimic membrane-bound TRAIL, (f) additional functionalization of the NPs with antibodies or other molecules to target cancer cells or tumor sites. (g) Encapsulation of chemotherapeutic agents into TRAIL-functionalized NPs and (h) further modification of NPs surface with molecules or antibodies for active targeting. Figure was generated with BioRender.
Figure 2. Correlation between TRAIL formulations and anticancer therapeutic potential. (a) RhTRAIL-based therapy, (b) crosslinking TRAIL monomers through an anti-tag antibody or (c) protein fusions facilitating TRAIL oligomerization (i.e., leucine zipper-TRAIL) [5,133]. (d) Conjugation of antibody fragment to TRAIL using a linker (i.e., scFv425) [86] to target tumor cells. (e) Conjugation of TRAIL to nanoparticles (NPs) to mimic membrane-bound TRAIL, (f) additional functionalization of the NPs with antibodies or other molecules to target cancer cells or tumor sites. (g) Encapsulation of chemotherapeutic agents into TRAIL-functionalized NPs and (h) further modification of NPs surface with molecules or antibodies for active targeting. Figure was generated with BioRender.
Cancers 14 05125 g002
Figure 3. Schematic representation of using nanoparticles to improve TRAIL efficacy. TRAIL DNA-based delivery tools: Uptake by tumor cells via either passive- or active (receptor-mediated) endocytosis. The endocytosed NPs escape from the endosome and release the encapsulated TRAIL gene into the cytosol, which then translocates to the nucleus. After transcription and translation, soluble TRAIL is released into the tumor microenvironment, where it can interact with its receptors (death receptor (DR)4 and 5 to induce apoptosis. Alternatively, TRAIL gene-encapsulating liposomes can fuse with the cell membrane and directly release the TRAIL gene into the cytosol. Dendrimers and polymeric NPs can co-deliver TRAIL gene and chemotherapeutic reagents, resulting in DNA damage thus amplifying the TRAIL-induced cell death signals. The figure was generated with BioRender.
Figure 3. Schematic representation of using nanoparticles to improve TRAIL efficacy. TRAIL DNA-based delivery tools: Uptake by tumor cells via either passive- or active (receptor-mediated) endocytosis. The endocytosed NPs escape from the endosome and release the encapsulated TRAIL gene into the cytosol, which then translocates to the nucleus. After transcription and translation, soluble TRAIL is released into the tumor microenvironment, where it can interact with its receptors (death receptor (DR)4 and 5 to induce apoptosis. Alternatively, TRAIL gene-encapsulating liposomes can fuse with the cell membrane and directly release the TRAIL gene into the cytosol. Dendrimers and polymeric NPs can co-deliver TRAIL gene and chemotherapeutic reagents, resulting in DNA damage thus amplifying the TRAIL-induced cell death signals. The figure was generated with BioRender.
Cancers 14 05125 g003
Table 1. Nanoformulations used for passive TRAIL delivery.
Table 1. Nanoformulations used for passive TRAIL delivery.
FormulationSize (nm)TRAIL Form/LocalizationMain FindingsTumor TypeRef.
Liposome100–140rhTRAIL/SurfaceMimicking the membrane properties of natural TRAIL increased receptor clustering and improved cytotoxic potential of TRAIL thus overcoming resistance to soluble TRAIL.Colorectal cancer, mouse model[141]
Liposome100rhTRAIL/SurfaceEnhanced clustering of DR5 overcoming resistance to soluble TRAIL.Colorectal cancer, mouse model[129]
Triazine modified dendrimer200pTRAIL/SurfaceEnhancement in transfection efficacy of TRAIL gene. Improved tumor growth inhibition.Osteosarcoma mouse model[189]
PEG-TRAIL microencapsulated into PLGA11,000–15,500TRAIL/InsideIncreased biological half-life time and sustained TRAIL delivery up to 18 days.Colorectal cancer, mouse model[190]
Single-walled carbon nanotubesNDrhTRAIL/Surface20-fold increase in TRAIL cytotoxicity against cancer cells without toxicity against normal cells.Colorectal cancer, NSCLC, hepatocellular carcinoma[191]
Abbreviations: nm, nanometer; rhTRAIL, recombinant human TRAIL; pTRAIL, plasmid TRAIL; DR, death receptor; PEG, polyethylene glycol; PLGA, poly (lactic-co-glycolic acid); ND, not determined; NSCLC: non-small cell lung carcinoma.
Table 3. Nanoparticles for combinatorial approaches for TRAIL delivery.
Table 3. Nanoparticles for combinatorial approaches for TRAIL delivery.
FormulationTRAIL Form/LocationStrategy to Overcome TRAIL ResistanceTumor TypeRef.
Inhalable HSA NPs; loaded w/DoxrhTRAIL/SurfaceDNA damage caused by DOX.Lung cancer, mouse model[214]
HSA NPs; loaded w/DOXrhTRAIL/SurfaceDNA damage caused by DOX.Colorectal cancer, mouse model[215]
Liposome, loaded w/DOXrhTRAIL/SurfaceDNA damage caused by DOX.NSCLC, mouse model[238]
Liposome, loaded w/PTXpTRAIL/SurfacePTX induced M-phase cell cycle arrest.Glioblastoma, mouse model[216]
Polymeric NPs coated with platelet membrane, loaded w/DOXrhTRAIL/SurfaceDNA damage caused by DOX.Metastatic breast cancer, mouse model[217]
Polymeric NPs, loaded w/DOXrhTRAIL/InsideDNA damage caused by DOX.Prostate and colon cancer, mouse model[239]
PEI-coated gold nanocompositepTRAIL/SurfaceROS generation by iron oxide NPs inducing DNA damage.Hep3B cell xenograft mouse model[220]
Inhalable highly porous PLGA microparticles, loaded w/DOXrhTRAIL/SurfaceDNA damage caused by DOX.Mouse model of H226 cell metastasis[240]
Liposome inside a hyaluronic acid crosslinked-gel shell, loaded w/DOXrhTRAIL/Between liposome and the gel shellDNA damage caused by DOX.MDA-MB-231 breast cancer cell, xenograft mouse model[222]
Chitosan modified magnetic nanoparticlespTRAIL/InsideHyperthermia induced by a magnetic field.Pulmonary metastatic mouse model[241]
Alginate modified CaCO3 NPs, loaded w/DOXrhTRAIL/SurfaceDNA damage caused by DOX.Cervical cancer cell line (HeLa cells)[242]
Magnetic ferric oxide NPrhTRAIL/SurfaceROS generation by iron oxide causing DNA damage.Glioma mouse model[243]
Polymeric NPs, DOX intercalatedpTRAIL/InsideDNA damage caused by DOX.Glioma mouse model[227]
Abbreviations: HSA, human serum albumin; NPs, nanoparticles; rhTRAIL, recombinant human TRAIL; DOX, doxorubicin; DNA, deoxyribonucleic acid; NSCLC, non-small cell lung cancer; pTRAIL, plasmid encoding TRAIL; PTX, paclitaxel; PEI, polyethylenimine; ROS, reactive oxygen species; PLGA, poly (lactic-co-glycolic acid).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Alizadeh Zeinabad, H.; Szegezdi, E. TRAIL in the Treatment of Cancer: From Soluble Cytokine to Nanosystems. Cancers 2022, 14, 5125. https://doi.org/10.3390/cancers14205125

AMA Style

Alizadeh Zeinabad H, Szegezdi E. TRAIL in the Treatment of Cancer: From Soluble Cytokine to Nanosystems. Cancers. 2022; 14(20):5125. https://doi.org/10.3390/cancers14205125

Chicago/Turabian Style

Alizadeh Zeinabad, Hojjat, and Eva Szegezdi. 2022. "TRAIL in the Treatment of Cancer: From Soluble Cytokine to Nanosystems" Cancers 14, no. 20: 5125. https://doi.org/10.3390/cancers14205125

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop