Next Article in Journal
Protection against Marburg Virus and Sudan Virus in NHP by an Adenovector-Based Trivalent Vaccine Regimen Is Correlated to Humoral Immune Response Levels
Next Article in Special Issue
Assessment of Postvaccination Neutralizing Antibodies Response against SARS-CoV-2 in Cancer Patients under Treatment with Targeted Agents
Previous Article in Journal
Use of a Practitioner-Friendly Behavior Model to Identify Factors Associated with COVID-19 Vaccination and Other Behaviors
Previous Article in Special Issue
The Big Potential of Small Particles: Lipid-Based Nanoparticles and Exosomes in Vaccination
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Application of mRNA Technology in Cancer Therapeutics

Research Institute of Senology, Acıbadem University, Istanbul 34457, Turkey
Vaccines 2022, 10(8), 1262; https://doi.org/10.3390/vaccines10081262
Submission received: 11 July 2022 / Revised: 31 July 2022 / Accepted: 1 August 2022 / Published: 5 August 2022
(This article belongs to the Special Issue Oncology in the Era of SARS-CoV-2)

Abstract

:
mRNA-based therapeutics pose as promising treatment strategies for cancer immunotherapy. Improvements in materials and technology of delivery systems have helped to overcome major obstacles in generating a sufficient immune response required to fight a specific type of cancer. Several in vivo models and early clinical studies have suggested that various mRNA treatment platforms can induce cancer-specific cytolytic activity, leading to numerous clinical trials to determine the optimal method of combinations and sequencing with already established agents in cancer treatment. Nevertheless, further research is required to optimize RNA stabilization, delivery platforms, and improve clinical efficacy by interacting with the tumor microenvironment to induce a long-term antitumor response. This review provides a comprehensive summary of the available evidence on the recent advances and efforts to overcome existing challenges of mRNA-based treatment strategies, and how these efforts play key roles in offering perceptive insights into future considerations for clinical application.

1. Introduction

During the past decade, major technological advances have enabled the use of mRNA-based therapeutics as a promising innovative approach in cancer therapy. Initial cancer vaccine trials date back to the 19th century with reports of tumor regression in patients injected with samples of erysipelas [1] and have continued since then with insufficient success. A dendritic cell-based vaccine directed against PSA, namely Sipuleucel-T, remains the only approved therapeutic cancer vaccine with limited clinical applicability [2].
Navigating the host immune system through induction of tumor-specific immunity by activating cytotoxic T cells to eliminate cancer cells is the mainstay of cancer immunotherapy. In order to induce an adaptive immune response, T cells have to recognize cancer-specific neoantigens processed by antigen presenting cells (APCs) such as dendritic cells or tissue macrophages, which delineate the trenches of immune-related cell killing. Thus, identification of tumor neoantigens has played a key role to establish the development of cancer vaccines [3,4].
mRNA vaccine is a recent innovative approach to cancer immunotherapy by encoding tumor-specific antigens to be introduced into APCs to synthesize the required antigens by the intracellular machinery. Early preclinical studies have shown generation of robust antitumor immune responses to mRNA-based cancer vaccines that are capable of cytotoxic activity [5,6]. Furthermore, mRNA vaccines are more advantageous in terms of production cost and feasibility as compared to DNA-based vaccines. They are less expensive, require less comprehensive and time-consuming manufacturing processes, and enable enhanced precision in tumor-directed genomic targeting as well as inability for DNA integration, which is a major drawback in alternative nucleic acid vaccines [7].
Despite the enthusiasm generated in the field of mRNA-based immune-oncology, many challenges remain to be addressed before this strategy can be adopted in routine practice. Numerous efforts are being placed to improve translational ability and stability through molecular engineering. Furthermore, much research is being carried out to optimize delivery systems in order to facilitate intracellular uptake and mitigate inherent immunogenicity ensued through degradation by extracellular ribonucleases [8]. In this review, data on current mRNA-based vaccines are discussed in the context of available evidence from preclinical and clinical studies, highlighting future prospects of incorporating this novel therapeutic strategy in cancer immunotherapy.

2. Cancer Immunology and Immunotherapy

The immune system comprises of elements of myeloid and lymphoid lineage such as lymphocytes and macrophages, which are specialized to generate an immune response against foreign-appearing structures in the host, including cancerous cells. When a tumor cell encounters the innate immune system, an inflammatory signaling cascade is initiated, which stimulates induction of dendritic cell maturation, immunostimulatory cytokine secretion, and natural killer cell activity. Once tumor cells are internalized by dendritic cells, these APCs interact with the microenvironment to present the neoantigen to cytotoxic T cells and B cells, which are subsequently activated to develop an antigen-specific immune response. The generation of an active adaptive immunity requires dendritic cell maturation and induction of danger signals from both apoptotic or necrotic cells and the TME during antigen processing [9,10]. Nevertheless, as tumors progress, the anticancer immunologic activity may be hampered by the upregulation of various checkpoints and immune-suppressive elements of the host immune system, which are naturally programmed to balance any excessive immune activity against the host. Recently identified inhibitory immunoreceptors, such as lymphocyte-activation gene 3 (LAG-3) and T cell immunoreceptor with immunoglobulin and ITIM domain (TIGIT), indoleamine 2,3-dioxygenase (IDO) controls immune activity by suppressing tumor-specific T and B lymphocytes, resulting in a shift towards generation of an immune-suppressive stroma comprising of exhausted T cells, which lack the ability to generate an anti-tumor immune response; as well as myeloid and lymphoid elements of the immune system associated with immune escape, namely regulatory T cells (Treg), immature dendritic cells, M2 macrophages, and myeloid derived suppressor cells (MDSC) [11]. It has been shown that cancer cells tap into the host immune system to overcome immune-mediated cell killing by switching the tumor microenvironment to an immunosuppressive or immune-cold phenotype mediated by several cytokines and molecules [11,12,13].
Immunotherapy refers to all types of treatment strategies aiming to restore immune dysregulation or to modulate the host immune system to destroy cancer cells. The current immunotherapy approaches involve strategies that aim to release the brakes on the host immune regulatory systems by inhibiting checkpoint upregulation by programmed cell death protein-1 (PD-1), programmed cell death ligand-1 (PD-L1), or cytotoxic T lymphocyte antigen-4 (CTLA) inhibitors, or to stimulate the immune system to generate a cancer-specific response by cancer vaccines or to administer ex vivo activated autologous or allogeneic immune cells that target cancer cells, such as chimeric antigen receptor-T (CAR-T) cells or engineered natural killer (NK) cells, otherwise referred to as adoptive cell therapy [14,15,16,17,18,19].

3. The Evolving Role of mRNA Technology in Cancer Immunotherapy

Nucleic acids in the form of RNA or DNA, whether exogenous from bacterial or viral causes, or endogenous, shed from cancer cells are capable of inducing variable degrees of immune response [20]. Cancer antigens, whether as whole-cell lysates, peptides, or as nucleic acids intended to translate into the structural protein of the antigen itself, can be delivered into the host in order to generate a cancer-specific immune response, otherwise referred to as a “cancer vaccine”. In contrast to peptide-based or whole cell vaccines, nucleic acid vaccines are more advantageous since they enable delivery of multiple or full-length tumor antigens, leading to a broader immune response.
In vitro transcribed RNAs, which are proven to have a wide applicability in SARS-CoV-2 vaccinations, have recently gained interest as cancer vaccines due to their versatility to encode chimeric peptide structures, allowing for targeting cancer cells with diverse and complex mutational structures. Furthermore, mRNA vaccines have emerged as an appealing alternative to DNA vaccines not only due to their ability to be translated in both dividing and nondividing cells, but also due to their safety since they cannot integrate into the host genome [21].
As synthetic mRNA enters the host cells through the cell membrane or by endocytosis, translation to the peptide of interest occurs within the cytosol. This protein structure, which is undistinguishable from the product of endogenous mRNA, undergoes post-transcriptional modifications eventually leading to degradation by intracellular compartments. These peptides are then presented on major histocompatibility complexes (MHC) of the antigen-presenting cells to be introduced to the effector cells of the host immune system to induce cancer-specific killer T cells along with activated helper T lymphocytes and NK cells [22] (Figure 1). In addition to the generation of a cancer-specific immune response, exogenous mRNA helps to maintain an immune-friendly tumor microenvironment (TME) by triggering secretion of type I Interferon (IFN) and other inflammatory cytokines through activation of toll-like receptors (TLR) and retinoic acid-inducible gene I (RIG-I) [23]. Furthermore, mRNA constructs can be engineered to express proinflammatory cytokines including, but not limited to Interleukin 2 (IL-2) IL-7, IL-12 and IL-15, which act synergistically to enhance generation of antigen-specific CD 8 + cytotoxic T cells, increase the ratio of active CD8 cells to immune suppressive Tregs, and induce memory T cells for a long-lasting immune response [24,25,26,27,28]. In addition, mRNAs are also being developed to encode monoclonal antibodies (mAbs), which have an established role as a passive targeted immunotherapy approach for various cancer types such as Her-2 positive breast cancer and lymphomas. There is in vivo evidence that suggests mRNA encoded mAbs are indeed able to induce a more sustained antitumor effect as compared to their recombinant equivalents in murine models [29,30]. These constructs can be modified to encode bispecific mAbs comprising an anti-CD3 Fv and a tumor-specific Fv, which are able to redirect T cells to the TME to elicit a stronger immune-mediated tumor cell killing [31].

4. The Immunogenicity and Molecular Biology of mRNA-Based Immunotherapy

Therapeutic mRNAs are produced through in vitro transcription (IVT) catalyzed by DNA-dependent RNA polymerase, which selectively recognize the promoter region of DNA templates [32]. The end product is a naked mRNA strand, which should be modified to optimize the stability and translational ability. These modifications include capping the 5′ end, optimizing the sequence of the untranslated translating regions, and adding a poly-A tail [33]. Nevertheless, these alterations and byproducts generated during the IVT process may impede the desired antitumor response through activation of the innate immune system, leading to the recognition of the modified mRNA molecules as nonself, as well as interference with the transcriptional capacity by cellular stress mechanisms [34].
As the first line of defense against external and internal pathogens, the innate immune system initiates a cascade of events subsequently triggering adaptive cancer-specific immunity. Endogenous and exogenous non-self-nucleic acids are recognized by intracellular pattern recognition receptor family (PRR) comprising TLRs, RIG-I-like receptors, nucleotide-binding and oligomerization domain (NOD)-like receptors, C-typelectin receptors, absent-in-melanoma 2 (AIM2)-like receptors, and the cyclic GMP-AMP synthase. Activation of PRRs localized in the cytosol and endosomal compartment in turn lead to transcriptional activity of several proinflammatory cytokines and chemokines, and stimulation of transcription-independent intracellular pathways such as autophagy, apoptosis, and phagocytosis. Studies with synthetic nucleic acids to manipulate the immune system have shown that different sequences of dsRNA (siRNA) varying in length induce distinct immune responses, which may be in opposite directions [20,35,36]. Therefore, purity and nucleotide composition of therapeutic mRNAs play a significant role in generating an optimal immune response.

4.1. mRNA Vaccine Structure

Two types of mRNA-based vaccines are available: nonreplicating (NRM) and self-replicating mRNA (SRM) vaccines, which are composed of a universal 5′ cap, 3′ and 5′ noncoding regions, an open reading frame, and a 3′ poly-A tail. While the cap structure protects the mRNA from quick degradation and induces IFN-mediated immune responses, the untranslated regions regulate the translational efficiency of mRNA. The poly A tail plays a significant role in the translation by regulating the stability of mRNA. Enrichment of G:C content and utilizing modified codons in the ORF constructs and optimizing the length of the poly-A sequence are some of the structural modifications that promote a translational process [37,38,39,40] (Figure 2). The NRM, though technically less demanding to produce, has the disadvantage of limited activity and stability, which can be overcome to a certain extent by structural optimization. SRM vaccines differ from NRMs by including an extra construct that encodes a replicase component. Generally, these vaccines are produced through engineering of single-strand RNA viral structural genes, which have been substituted by the gene of interest (i.e., cancer antigen), while keeping the nonstructural genes (i.e., replicase), leading to a high level of antigen expression within a delivery system. Picornaviruses, alphaviruses, and flaviviruses are the most common RNA viral systems employed to generate SRM vaccines [5,41,42,43,44].

4.2. mRNA Delivery Platforms

Although mRNA technology is a promising tool for cancer immunotherapy, a number of challenges have to be faced to facilitate an effective immune response. First, the large and negatively charged RNA molecule has to cross the cell membrane, which is a significant barrier to intracellular delivery due to its negative charge. Once mRNA enters the cell, there is a high risk for degradation through ribonucleases, which are abundant throughout the skin and systemic circulation. Although delivery of naked mRNA is possible through intradermal, subcutaneous, and intramuscular routes, the efficacy of such approaches is hindered by a short half-life, rapid degradation and inadequate immune response due to ineffective access to intracellular compartments. Therefore, an efficient delivery is crucial to achieve favorable therapeutic potential. Therapeutic advances in mRNA technology have been linked to the development of various nanotechnological delivery systems that have been engineered to ensure optimal translational capability [5,21,45].

4.2.1. Synthetic Systems

  • Lipid-based Delivery Systems
Lipid-based materials are the most extensively investigated delivery systems for RNA-based therapeutics. Referred to as lipid nanoparticles (LNP), these structures consist of a cationic or more recently a pH-dependent ionizable lipid layer; a polyethylene glycol (PEG) component; phospholipids and cholesterol [45,46]. The ionizable amino lipid layer is designed to obtain a positive charge as pH drops, facilitating endosomal uptake of the liposome, while retaining encapsulation of the negatively charged mRNA molecule. The PEG molecule plays a significant role in preventing macrophage-mediated degradation, together with providing stability along with cholesterol [47,48,49]. The structure of the amino lipid component plays a key role in delivery efficacy, tolerance, and tissue clearance [50]. Efforts to optimize LNP delivery of mRNA vaccines have yielded efficient RNA delivery in cell lines, and strong, long-lasting humoral immune responses against several viral pathogens in murine models [51,52,53]. Clinical studies with two LNP-based mRNA vaccines against the SARS-CoV2 virus during the pandemic have confirmed the favorable results, showing a strong efficacy across different populations, leading to regulatory approval despite the short-term follow-up period [54,55].
Nevertheless, as part of a cancer vaccine, LNP design should further be developed to deliver the mRNA cargo specifically to antigen-presenting cells, while preventing degradation and retain effective translational capacity. Moreover, the amino lipid structure should be biodegradable to prevent toxicity and allow for multiple dosing at the same time. Emerging evidence from preclinical studies suggest that LNP mRNA vaccines provide robust antigen-specific antitumor with memory T cell responses by specifically targeting dendritic cells, leading to prevention of tumor growth in murine models [56,57,58,59].
2.
Polymer-based Delivery Systems
Polymeric materials and dendrimers, modified with nanotechnologic fatty side chains to reduce toxicity and avoid enzymatic degradation in vivo, have gained popularity to deliver mRNA as vaccines against fatal viral pathogens such as HIV, Zika, Ebola, and H1N1 Influenza [60,61,62,63,64]. Polymeric structures surrounded by a PEG outer shell have been used in murine models to deliver an antiangiogenic RNA sequence, which was shown to inhibit growth in a pancreatic cancer model [65]. Similarly designed mRNA vaccines have been shown to effectively translate into tumor-associated antigens in vivo [66]. Furthermore, a polymer-based RNA vaccine encoding PTEN has successfully been introduced into several castration-resistant prostate cancer models and has been shown to inhibit tumor growth by restoring PTEN function [67].
3.
Peptide-based Delivery
Cell-penetrating peptides (CPPs) are cationic peptides that can translocate through the cell membrane independent of receptors and can transport proteins, small organic molecules, nanoparticles, and oligonucleotides. Because of a favorable safety profile and efficient transfection capability, CPPs represent a promising class of nonviral delivery vectors [68,69,70]. Nevertheless, low cell and tissue selectivity, and impaired internalization of the cargo by conjugation through different cellular layers limit efficient clinical application [71]. Recent efforts have been focused on identifying the most optimal CPP for enhanced immune activity. Protamine is a cationic peptide that can prevent lysosomal degradation during delivery of RNA. Protamine-based deliveries have been shown to induce a strong immune response through toll-like receptor 7 activation [72,73]. More recently, advances in biotechnology have led to promising developments in peptide-based mRNA delivery. For example, a pegylated cationic KL4 peptide complex in powder form has been successfully used as an aerosolized delivery system for pulmonary delivery [74]. Furthermore, an optimized GALA-peptide conjugated mRNA encoding the Ova peptide exhibited a strong APC uptake and an efficient endosomal escape, leading to enhanced antigen-specific T cell activity and dendritic cell maturation compared to naked RNA or different peptide complexes [75].

4.2.2. Biological Systems

  • Ex Vivo Transfected Cellular Systems
Immunotherapy against cancer requires transfection of APC with specific antigens or nucleic acids such as mRNA, which translate into tumor-specific antigens. Although in vivo transfection via the intramuscular, intravenous, or subcutaneous routes are possible, the immune response generated is usually weak and unsustained. Therefore, ex vivo transfected engineered dendritic cells or chimeric antigen receptor T cells have been developed as cancer vaccines or adoptive cell therapy strategies to target cancer cells once introduced in the host [76].
  • Dendritic Cells
Dendritic cells play a crucial role in reprogramming the immune system by their ability to uptake and present the tumor antigens, leading to generation of potent effector cell activity directed against cancer cells. Additionally, mature dendritic cells are capable of modulating chemokine- and cytokine-induced lymphoid activation, which are strictly relevant for a systemic and sustainable anticancer immune response. As autologous cancer vaccines, dendritic cells are harvested from the host by apheresis, isolated from mononuclear cells or progenitor stem cells, subsequently stimulated by various cytokines to achieve maturity, followed by transfection with specific antigens as nucleic acids or peptides. Numerous efforts have been focused on methods to achieve a stronger immune response through more efficient antigen presentation, migration to required lymphatic tissues, and induction of a stronger cytokine production through generation of Notch differentiated dendritic cells with engineered receptor expression capability using clustered regularly interspaced short palindromic repeats (CRISP-R) gene editing and RNA interference, as well as the use of optimized maturation cocktails [77,78,79]. Ex vivo transfection of mRNA-loaded dendritic cell vaccines against a variety of tumor specific antigens such as telomerase reverse transcriptase (TERT) and the melanoma cell line B16F10 have led to generation of a strong antitumor immune response in murine melanoma models [80,81].
b.
CAR-T Cells
Chimeric antigen receptor (CAR)-modified T cells represent a novel adoptive cell therapy approach that has been shown to effectively target tumor cells leading to a potent immune-mediated cancer cell killing [82]. CAR-Ts confer several advantages over natural host immunity by MHC independent tumor antigen presentation, more potent cell receptor binding, and ability to bypass escape mechanisms such as HLA downregulation [83]. Direct transfection by electroporation or viral systems has been utilized to deliver CAR-encoding mRNA to generate cancer-specific CAR-T cells [84]. More recently, RNA optimization by nanoparticles and gene editing through CRISP-R technology has been utilized to engineer CAR-Ts that have improved stability and transfection ability [85,86]. Preclinical studies investigating ex vivo transfection of patient-derived T cells by retroviral constructs to deliver mRNA encoding bi-CARs targeting tumor-specific epitopes have shown that the engineered CAR-Ts are capable of recognizing target antigens and overcoming escape variants, eventually leading to improved survival in a glioblastoma (GBM) murine model [87]. Furthermore, profound cytotoxic cell lysis has been demonstrated with RNA transfected CAR-T constructs expressing CD19 in xenograft models with leukemia [88,89], extensively reviewed elsewhere by Rajan et al. [90].
2.
Viral Constructs
Viral constructs generated from RNA viruses have been evaluated extensively as self-replicating RNA (SRM) vaccines against several cancer types. Single-strand RNA viruses including alphaviruses, flaviviruses, and rhabdoviruses can be engineered to form naked RNA replicons and recombinant viral-like particles (VLP), which are capable of producing a high level of tumor antigen expression in APCs, in turn leading to a strong immune response [91,92]. An SRM vaccine comprises a replicon carrying the gene of interest in conjunction with the replicase gene and a defective virus encoding structural genes, forming VLP in a packaging cell construct. The VLP, taken up by APCs when introduced into the host, deliver self-replicating RNA constructs to the cytosol by receptor-mediated endocytosis, leading to a high level of RNA production and tumor–antigen expression through translation [44,93,94]. Preclinical studies evaluating the role of replicon-based SRM vaccines have shown the success in eliciting strong humoral and cellular immune responses against several cancer types in xenograft models harboring Her-2 neu breast cancer, prostate cancer, GBM, and human papilloma virus (HPV)-induced tumors [95,96,97,98].

5. Clinical Applications

5.1. Personalized mRNA Vaccines

5.1.1. Naked mRNA Vaccines

Direct intradermal injection of naked mRNA sequences was shown to effectively produce the encoded protein leading to generation of tumor-specific functional immune response in various cancer types including lung cancer and prostate cancer, hence were introduced as an alternative vaccination method [99,100]. Early clinical studies of naked mRNA fortified with protamine stabilization showed a favorable immune response in 50% of the cohort comprising seven patients with melanoma, and a complete clinical response in one patient [101]. Similarly, another multiplex mRNA vaccine encoding carcinoembryonic antigen (CEA), mucin 1 (MUC1), human epidermal growth factor receptor 2 (Her-2), melanoma-associated antigen (MAGE), survivin and telomerase as tumor-associated antigens was evaluated in 30 patients with advanced renal cell carcinoma (RCC). Vaccinations were reported to be feasible with specific CD4 and CD8 immune responses and no serious toxicity. Median survival was 29 months, with approximately one-third of the cohort alive at 4 years [102]. Long-term follow up extending to 10 years revealed a strong correlation of immune responses with prolonged survival [103]. A sequence-optimized RNA vaccine encoding five non-small cell-associated tumor antigens, BI1361849 was investigated as part of a combination strategy with radiotherapy in stage 4 lung cancer patients responding to platin-based chemotherapy. Immunologic analysis revealed 40% of patients reaching the prespecified threshold of two-fold increase in functional tumor-specific CD4 and CD8 cell generation. The best overall response rate (ORR) was disease stabilization in 46% of patients [104].

5.1.2. LNP mRNA

The only available clinical data have been reported by Sahin et al. [105], who have investigated the efficacy of a liposomal RNA vaccine (FixVac), with and without PD-1 inhibition in the dose-escalation phase I Lipo-MERIT trial. FixVac, which is specifically designed to target dendritic cells, encodes four tumor antigens associated with malignant melanoma. All patients enrolled in this trial had previously received immune checkpoint inhibitors. In the interim analysis, 16% objective responses were observed in the monotherapy arm of 25 patients. The vaccine showed synergistic activity with PD-1 inhibition showing an ORR of 35%, which increased to 50% with increasing dose of the FixVac. Translational work on some patients with long-term immune monitoring has indicated generation of memory T cells, along with helper and cytotoxic T cell response in some responders. Data from this trial and others are awaited to provide further insight on the clinical applications of LNP mRNA vaccines and optimal combinations.

5.1.3. Dendritic Cell-Based Vaccines

Early clinical trials with RNA transfected dendritic cells have shown that vaccination with this strategy is feasible and is able to stimulate tumor-specific T cell responses in vivo. Based on encouraging preclinical data, a phase Ib study evaluated the role of PSA encoding mRNA dendritic cell vaccine in 16 patients with metastatic castration-resistant prostate cancer patients [106]. There was a significant increase in CTL response seen in all patients after completion of therapy and an ongoing prostate-specific antigen (PSA) response in six of seven evaluable patients who did not have subsequent therapies. An autologous RNA transfected DC vaccine was investigated in a pioneering trial in 10 patients with metastatic RCC. Confirmatory to the previous trial, there were strong antitumor T cell responses against three RCC antigens generated in five out of six patients who were evaluable [107]. Carcinoembryonic antigen is a frequently expressed tumor antigen in gastrointestinal tumors. In a phase Ib/II trial, 24 patients with resected hepatic metastases of colorectal carcinoma were vaccinated with CEA-encoding DC. In addition to the immune responses, there was one complete response and two minor responses with a clinical benefit ratio of 25%. The median RFS in the phase II cohort was reported as 122 days [108]. A different approach from the same group evaluated CEA transfected mRNA vaccination in three patients with localized pancreatic carcinoma, who were operated after neoadjuvant chemoradiation. All but one patient received the planned 6-month treatment, and no side effects were noted. Patients were reported as disease-free at approximately 4 years from diagnosis [109]. A DC-based mRNA vaccine encoding CD40Ligand, CD70, and TLR4 (TriMixDC), and transfected with melanoma associated genes (MAGE, Tyrosinase, gp100-TriMix-MEL) was evaluated in a cohort of advanced melanoma patients. There were two partial responses (13.3%), and the median progression-free survival (PFS) and overall survival (OS) were reported as 5 months and 14 months, respectively [110]. Based on the encouraging early data, the vaccine was used in an earlier disease setting following resection for stage III/IV melanoma. In the investigational arm, 21 patients received four vaccinations with a booster over 6 months, whereas the control group had no adjuvant therapy. Although vaccination was deemed to be feasible, the study was closed early due to futility. In evaluable patients, the median time to progression was 8 months, with more early relapses in the vaccine group and 13 months in the control group. Nevertheless, the 1-year disease-free survival (DFS) appeared to be higher in the vaccine arm: 71% vs. 35%, respectively [111]. TriMix-MEL was also evaluated as part of a combination with ipilimumab in pretreated patients with melanoma who had not received prior immune checkpoint blockade. Out of 39 patients, there were 8 patients with a complete response (CR) and an ORR of 38%. Reaching the primary endpoint of 6-month DCR, the median PFS and OS were 27 and 58 weeks, respectively [112]. A personalized mRNA transfected DC vaccine used in combination with PD-1 inhibition and low dose cyclophosphamide also confirmed generation of a high level of tumor-specific immune response and favorable survival outcomes in a cohort comprising 10 patients with lung cancer and GBM [113]. Both trials provide strong insight into the combined use of vaccines with immune checkpoint blockade, which deserve further investigation.

5.1.4. Viral-Based Self-Replicating mRNA Vaccines

Self-replicating viral-based constructs from RNA delivery have been shown to be safe and feasible in early clinical trials for both infectious diseases and cancer immunotherapy [114]. Early clinical studies with viral constructs have utilized alphavirus-based viral replicon particles (VRP) more frequently, which were designed to efficiently express the desired tumor antigen in high amounts following targeted uptake in DCs. In fact, a first- in-human study with a replication incompetent Semliki Forest replicon encoding HPV-associated antigens E6 and E7 showed strong antigen-specific interferon-gamma responses in three cervical in situ neoplasia patients [115]. Furthermore, one of the initial studies evaluated the feasibility of a CEA-encoding alphavirus particle in 30 patients with advanced solid tumors who received four doses with 3 weekly intervals. As a main endpoint, investigators were able to deduce that multiple injections of the VPR-CEA (Tricom-CEA) were feasible and could generate specific CD8 and CD4 T cell responses. However, the level of immune responses remained stable during the booster period in some patients who were able to receive them, suggesting that more than four injections would probably be unnecessary. Yet, there was one responder in the cohort and two patients with stabilization of disease in this heterogeneous cohort with some correlation between long-term efficacy and the level of immune response [116]. Long-term follow up results of a separate cohort comprising stage III colorectal cancer patients treated with the same vaccine were recently reported. In parallel with the phase I–II data, a specific immune response was induced in all patients, which was higher compared to stage IV patients from the previous trial. After a median follow up of 5 years, all patients were reported to be alive, with a 25% recurrence rate [117]. The same viral construct was used to produce a VPR encoding the transmembrane and extracellular domains of the Her-2 receptor, which was evaluated in a phase IB-II trial including patients with advanced Her-2 (+) breast cancer. All patients had received prior her-2 blockade, and one of the two cohorts included in the trial received combined anti-Her2 therapy with the vaccine, while cohort one received vaccine as monotherapy, given as three injections every 2 weeks. Investigators reported detectable levels of anti-Her2 immune responses, which unfortunately did not translate into relevant clinical responses, with median PFS of 1.8 and 3.6 months in cohorts 1 and 2, respectively [118]. Limited data from small studies including prostate, lung, and colorectal cancer suggest that VPR-based mRNA vaccines are able to generate antigen-specific immune responses, the level of which may be correlated with the outcome [119]. Still, this area of mRNA technology requires further investigation before finding a place in the immunotherapeutic landscape for cancer patients.

5.2. mRNA-Engineered Cellular-Based Immunotherapy and Gene Editing

Engineered CAR-T cells have revolutionized adoptive cellular therapy in hematologic cancers. The application of this technology to other immune cells has led to the development of viral transfected CAR-Ms (CAR-macrophages) that could potentially be used against solid tumors. Despite disadvantages in the manufacturing process, in vitro transcribed, mRNA-based CAR-encoding immune cells represent a safe and effective alternative to the first- and second-generation CAR-T cells that are currently approved for clinical use. Lipid nanoparticle delivery systems, as described previously, have been successfully utilized to deliver anti-CD19 coding mRNA in M1 macrophages and cytotoxic T cells [86]. RNA CAR-T cells have the advantages of rapid production and multiple administrations leading to enhanced efficacy, which has overcome several limitations pertaining to the routine use of CAR-T cells in the clinic. Furthermore, advances in genome editing have also refined selective targeting by immune cells. The CRISP/Cas9 system is a novel editing tool that can be utilized to select and delete the desired genome sites or result in knockdown of several genes through epigenetic silencing in order to enhance CAR-T efficacy [120,121]. Nevertheless, early phase I trials in various solid tumors comprising pancreatic and breast cancers have yielded unsatisfactory outcomes with transient responses despite generation of immune activity [84].
A list of ongoing trials on different RNA-based therapeutics is provided in Table 1.

6. Boosting Immune Response

6.1. Modulation of the Tumor Microenvironment

The tumor microenvironment (TME) plays a major role in controlling the cancer-immunity cycle through interaction with the signaling pathways leading to generation of a tumor-specific immune response. In some circumstances, mRNA can act as a tumor promoter, whereas, in other instances it can be modified to help generate an immune-friendly environment. In fact, tumor-derived mRNA has been implicated in activating angiogenesis under hypoxic conditions to promote tumor growth [122]. Furthermore, mRNA modification by N6-methyladenosine (m6-A), a redundant modification of eucaryotic mRNA, has been shown to be involved in generation of stemness property of cancer cells, promoting tumor growth and resistance to immunotherapy. Accumulating data suggests that targeting distinct m6-A regulators may reprogram the TME through secretion of immune-activating cytokines, upregulating costimulatory receptors and skewing the immune cell population toward an activated state by increasing the ratio of mature dendritic cells, M1 macrophages and Th1 cells to Tregs and MDSC [123]. In addition, distinct formulations with encoding mRNA have been shown to enhance immune-mediated cell killing through TME modulation. Research on the immunomodulatory use of RNA has been mainly focused on intratumoral or systemic delivery of mRNA engineered to produce costimulatory cytokines and receptors. Interleukin-2, an inflammatory cytokine playing a key role in differentiation and generation of effector T cell responses, has been used for the treatment of various tumor types since the turn of the century. Nevertheless, the unfavorable response–toxicity ratio with potentially fatal side effects has hampered its widespread use in routine practice and led to the adoption of alternative treatment strategies. Preclinical studies focusing on engineered mRNA encoding IL-2 in conjunction with other immune stimulatory cytokines such as IL-7, IL-15, and IL-12 used alone or as part of a combined approach with cancer vaccines, and adoptive T cell infusions, have been shown to generate synergistic immune activity in murine models with the advantage of an improved toxicity profile [24,26,124,125]. mRNA can also be used to transfect immune cells with costimulatory ligands including CD-40L and CD-70, enhancing immune-mediated cellular cytotoxicity in experimental models bearing several cancer types [126,127,128]. Advances in technology have also led to the production of fusion constructs, which have the ability of targeting tumor cells and the microenvironment simultaneously. Engineered mRNA encoding bispecific monoclonal antibodies expressing antitumor proteins, with receptors targeting effector T cells and immune checkpoints such as PD-1 and CTLA-4, and dendritic cells transfected with mRNA encoding agonistic ligands such as glucorticoid-induced TNFR-related protein (GITR), have been shown to elicit strong and sustained immune responses in experimental models [31,129,130]. Data from early clinical trials evaluating these novel strategies are awaited with enthusiasm.

6.2. Potential Combinations

Though mRNA fusion transcripts have provided an unprecedented opportunity to target multiple pathways in the cancer-immunity network, further combination strategies are required to ensure enhanced cancer cell killing. Energetic efforts are being placed to target several inhibitory elements of the TME, comprising angiogenesis, immune checkpoints, desmoplastic reaction, and fibrosis ensued by focal adhesion kinase (FAK) phosphorylation, epithelial–mesenchymal transition, and inhibitory signal transduction through phosphatase and tensin homolog (PTEN) loss or Myc activation [11,131]. There are numerous early-phase clinical trials evaluating the feasibility of combination approaches utilizing cancer vaccines and adoptive cell treatment with checkpoint inhibitors and antiangiogenic agents. Cytotoxic chemotherapy at various dose levels is an integral part in many of these studies. In fact, an ongoing clinical trial (NCT04503278) with an autologous CAR-T combined with an RNA-based cancer vaccine targeting claudin 6, has yielded encouraging activity in patients with refractory testicular or ovarian cancer [132]. Despite the theoretical advantage, an optimal combination has yet to be proven, rendering this as an active area of ongoing research. A comprehensive overview of ongoing and completed clinical trials has been provided elsewhere [133].

7. Conclusions and Future Perspectives

The advent of nanoparticle technology and genome editing tools has led to the generation of novel methods for RNA-based treatment. RNA can not only be modified to be used as a drug in itself, but can also serve as an efficient platform to deliver genomic information. During the last two decades, energetic efforts have been placed to optimize methods of mRNA-based gene therapies. Advances in nanotechnology have led to developments in delivery platforms, yielding encouraging results in preclinical and clinical studies for the treatment of a wide spectrum of diseases, including viral to bacterial pathogens, together with rare conditions related to genetic disorders and cancer.
Nevertheless, instability, impaired translational capacity, and lack of effective delivery methods have remained as major challenges, which require further research before this approach gains wide clinical applicability. The success of mRNA vaccines against SARS-CoV-2 during the recent pandemic has generated renewed enthusiasm to exploit this technology further. Despite the evident potential in the prevention of infectious diseases, the evolution of mRNA-based therapeutic strategies in oncologic care have lagged behind antiviral indications due to inadequate clinical responses. However, IVT RNAs delivered in various platforms as cancer vaccines or adoptive cell therapies have proved to be attractive and versatile tools to elicit cancer-specific immune responses, not only through effective antigen presentation, but also induction of an immune-friendly TME. Furthermore, there is early evidence suggesting that the clinical efficacy of mRNA-based systems can be enhanced through novel combinations with different anticancer strategies, including immune checkpoint inhibitors and chemotherapy.
Looking beyond cancer vaccines, advances in mRNA technology and gene editing have unraveled distinct innovative strategies for future development across several cancer types. Accumulating data from well-designed preclinical studies suggest that exogenously produced short, noncoding RNA fragments comprising antisense oligonucleotides (ASO), short interfering RNA (siRNA), and microRNA (miRNA) can be modified to target mRNAs for a wide array of genomic functions ranging from epigenetic silencing to restoration of inactive tumor suppressant genes such as TP53 or PTEN [134,135]. Another exciting advance in gene therapies with LNP-based mRNA platforms edited for aberrant genome targeting or protein replacement therapy has been the modification for specific organ targeting, a process called SORT for diverse cellular origins [136].
The oncology community is eagerly awaiting validated novel mRNA-based combinations for enhanced anticancer activity. The versatility of mRNA platforms and rapid production capacity of clinical grade products underscores the potential role of various mRNA therapeutic approaches in the future of cancer treatment.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created in this study. Data sharing is not applicable.

Acknowledgments

The author would like to thank Ipek G. Eralp for proofreading this article.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Cooley, W.B. The treatment of malignant tumors by repeated inoculations of erysipelas: With a report of ten original cases. Am. J. Med. Sci. 1893, 105, 487–511. [Google Scholar] [CrossRef]
  2. Small, E.J.; Schellhammer, P.F.; Higano, C.S.; Redfern, C.H.; Nemunaitis, J.J.; Valone, F.H.; Verjee, S.S.; Jones, L.A.; Hershberg, R.M. Placebo-controlled phase III trial of immunologic therapy with sipuleucel-T (APC8015) in patients with metastatic, asymptomatic hormone refractory prostate cancer. J. Clin. Oncol. 2006, 24, 3089–3094. [Google Scholar] [CrossRef] [PubMed]
  3. Ebrahimi, N.; Akbari, M.; Ghanaatian, M.; Roozbahani, M.P.; Adelian, S.; Borjian, B.M.; Yazdani, E.; Ahmadi, A.; Hamblin, M.R. Development of neoantigens: From identification in cancer cells to application in cancer vaccines. Expert. Rev. Vaccines 2021, 19, 941–955. [Google Scholar] [CrossRef]
  4. Fritsch, E.F.; Burkhardt, U.; Hacohen, N.; Wu, C.J. Personal Neoantigen Cancer Vaccines: A Road Not Fully Paved. Cancer Immunol. Res. 2020, 8, 1465–1469. [Google Scholar] [CrossRef]
  5. Pardi, N.; Hogan, M.J.; Porter, F.W.; Weissman, D. mRNA vaccines—A new era in vaccinology. Nat. Rev. Drug Discov. 2018, 17, 261–279. [Google Scholar] [CrossRef]
  6. Sahin, U.; Karikó, K.; Türeci, Ö. mRNA-based therapeutics—Developing a new class of drugs. Nat. Rev. Drug Discov. 2014, 13, 759–780. [Google Scholar] [CrossRef] [PubMed]
  7. Sahin, U.; Derhovanessian, E.; Miller, M.; Kloke, B.P.; Simon, P.; Löwer, M.; Bukur, V.; Tadmor, A.D.; Luxemburger, U.; Schrörs, B.; et al. Personalized RNA mutanome vaccines mobilize poly-specific therapeutic immunity against cancer. Nature 2017, 547, 222–226. [Google Scholar] [CrossRef]
  8. Xu, S.; Yang, K.; Li, R.; Zhang, L. mRNA Vaccine Era—Mechanisms, Drug Platform and Clinical Prospection. Int. J. Mol. Sci. 2020, 21, 6582. [Google Scholar] [CrossRef]
  9. Upadhyay, S.; Sharma, N.; Gupta, K.B.; Dhiman, M. Role of immune system in tumor progression and carcinogenesis. J. Cell Biochem. 2018, 119, 5028–5042. [Google Scholar] [CrossRef]
  10. Anari, F.; Ramamurthy, C.; Zibelman, M. Impact of tumor microenvironment composition on therapeutic responses and clinical outcomes in cancer. Future Oncol. 2018, 14, 1409–1421. [Google Scholar] [CrossRef]
  11. Khalaf, K.; Hana, D.; Chou, J.T.; Singh, C.; Mackiewicz, A.; Kaczmarek, M. Aspects of the Tumor Microenvironment Involved in Immune Resistance and Drug Resistance. Front. Immunol. 2021, 12, 656364. [Google Scholar] [CrossRef]
  12. Li, J.; Byrne, K.T.; Yan, F.; Yamazoe, T.; Chen, Z.; Baslan, T.; Richman, L.P.; Lin, J.H.; Sun, Y.H.; Rech, A.J.; et al. Tumor Cell-Intrinsic Factors Underlie Heterogeneity of Immune Cell Infiltration and Response to Immunotherapy. Immunity 2018, 49, 178–193. [Google Scholar] [CrossRef]
  13. Gerard, C.L.; Delyon, J.; Wicky, A.; Homicsko, K.; Cuendet, M.A.; Michielin, O. Turning tumors from cold to inflamed to improve immunotherapy response. Cancer Treat. Rev. 2021, 101, 102227. [Google Scholar] [CrossRef] [PubMed]
  14. Ito, F.; Chang, A.E. Cancer immunotherapy: Current status and future directions. Surg. Oncol. Clin. N. Am. 2013, 22, 765–783. [Google Scholar] [CrossRef] [PubMed]
  15. Ayana, R.; Kumar, A.R.; Devan, A.R.; Nair, B.; Vinod, B.S.; Nath, L.R. Harnessing the immune system against cancer: Current immunotherapy approaches and therapeutic targets. Mol. Biol. Rep. 2021, 48, 8075–8095. [Google Scholar]
  16. Stephan, K.; Matthias, I.; Sebastian, K. Advances in cancer immunotherapy 2019—Latest trends. Exp. Clin. Cancer Res. 2019, 38, 268. [Google Scholar]
  17. Marshall, H.T.; Djamgoz, M.B.A. Immuno-Oncology: Emerging targets and combination therapies. Front. Oncol. 2018, 8, 315. [Google Scholar] [CrossRef]
  18. Khalil, D.N.; Smith, E.L.; Brentjens, R.J.; Wolchok, J.D. The future of cancer treatment: Immunomodulation, CARs and combination immunotherapy. Nat. Rev. Clin. Oncol. 2016, 13, 273–290. [Google Scholar] [CrossRef]
  19. Wang, Z.; Cao, Y.J. Adoptive Cell Therapy Targeting Neoantigens: A Frontier for Cancer Research. Front. Immunol. 2020, 11, 176. [Google Scholar] [CrossRef] [PubMed]
  20. Bishani, A.; Chernolovskaya, E.L. Activation of innate immunity by therapeutic nucleic acids. Int. J. Mol. Sci. 2021, 22, 13360. [Google Scholar] [CrossRef]
  21. Miao, L.; Zhang, Y.; Huang, L. mRNA vaccine for cancer immunotherapy. Mol. Cancer 2021, 20, 41. [Google Scholar] [CrossRef]
  22. Beck, J.D.; Reidenbach, D.; Salomon, N.; Sahin, U.; Türeci, Ö.; Vormehr, M.; Kranz, L.M. mRNA therapeutics in cancer immunotherapy. Mol. Cancer 2021, 20, 69. [Google Scholar] [CrossRef] [PubMed]
  23. Pastor, F.; Berraondo, P.; Etxeberria, I.; Frederick, J.; Sahin, U.; Gilboa, E.; Melero, I. An RNA toolbox for cancer immunotherapy. Nat. Rev. Drug Discov. 2018, 17, 751–767. [Google Scholar] [CrossRef] [PubMed]
  24. Vormehr, M. Substantial improvement of cancer immunotherapy by an RNA encoded extended half-life Interleukin-2 variant. Abstract P626. In Proceedings of the 34th Annual Meeting & Pre-Conference Programs (SITC 2019), National Harbor, MD, USA, 6–10 November 2019. [Google Scholar]
  25. Waldmann, T.A. The shared and contrasting roles of IL2 and IL15 in the life and death of normal and neoplastic lymphocytes: Implications for cancer therapy. Cancer Immunol. Res. 2015, 3, 219–227. [Google Scholar] [CrossRef] [PubMed]
  26. Etxeberria, I.; Bolaños, E.; Quetglas, J.I.; Gros, A.; Villanueva, A.; Palomero, J.; Sánchez-Paulete, A.R.; Piulats, J.M.; Matias-Guiu, X.; Olivera, I.; et al. Intratumor adoptive transfer of IL-12 mRNA transiently engineered antitumor CD8+ T cells. Cancer Cell 2019, 36, 613–629. [Google Scholar] [CrossRef]
  27. Hewitt, S.L.; Bailey, D.; Zielinski, J.; Apte, A.; Musenge, F.; Karp, R.; Burke, S.; Garcon, F.; Mishra, A.; Gurumurthy, S.; et al. Intratumoral IL12 mRNA therapy promotes TH1 transformation of the tumor microenvironment. Clin. Cancer Res. 2020, 26, 6284–6298. [Google Scholar] [CrossRef]
  28. Lai, I.; Swaminathan, S.; Baylot, V.; Mosley, A.; Dhanasekaran, R.; Gabay, M.; Felsher, D.W. Lipid nanoparticles that deliver IL-12 messenger RNA suppress tumorigenesis in MYC oncogene-driven hepatocellular carcinoma. J. Immunother. Cancer 2018, 6, 125. [Google Scholar] [CrossRef]
  29. Thran, M.; Mukherjee, J.; Ponisch, M.; Fiedler, K.; Thess, A.; Mui, B.L.; Hope, M.J.; Tam, Y.K.; Horscroft, N.; Heidenreich, R.; et al. mRNA mediates passive vaccination against infectious agents, toxins, and tumors. EMBO Mol. Med. 2017, 9, 1434–1447. [Google Scholar] [CrossRef]
  30. Rybakova, Y.; Kowalski, P.S.; Huang, Y.; Gonzalez, J.T.; Heartlein, M.W.; DeRosa, F.; Delcassian, D.; Anderson, D.G. mRNA delivery for therapeutic anti-HER2 antibody expression in vivo. Mol. Ther. 2019, 27, 1415–1423. [Google Scholar] [CrossRef]
  31. Stadler, C.R.; Bähr-Mahmud, H.; Celik, L.; Hebich, B.; Roth, A.S.; Roth, R.P.; Karikó, K.; Türeci, Ö.; Sahin, U. Elimination of large tumors in mice by mRNA-encoded bispecific antibodies. Nat. Med. 2017, 23, 815–817. [Google Scholar] [CrossRef]
  32. Mu, X.; Hur, S. Immunogenicity of In vitro transcribed RNA. Acc. Chem. Res. 2021, 54, 4012–4023. [Google Scholar] [CrossRef] [PubMed]
  33. Linares-Fernandez, S.; Lacroix, C.; Exposito, J.Y.; Verrier, B. Tailoring MRNA Vaccine to Balance Innate/Adaptive Immune Response. Trends Mol. Med. 2020, 26, 311–323. [Google Scholar] [CrossRef] [PubMed]
  34. De Beuckelaer, A.; Pollard, C.; Van Lint, S.; Roose, K.; Van Hoecke, L.; Naessens, T.; Udhayakumar, V.K.; Smet, M.; Sanders, N.; Lienenklaus, S.; et al. Type I Interferons Interfere with the Capacity of MRNA Lipoplex Vaccines to Elicit Cytolytic T Cell Responses. Mol. Ther. 2016, 24, 2012–2020. [Google Scholar] [CrossRef] [PubMed]
  35. Chen, N.; Xia, P.; Li, S.; Zhang, T.; Wang, T.T.; Zhu, J. RNA sensors of the innate immune system and their detection of pathogens. IUBMB Life 2017, 69, 297–304. [Google Scholar] [CrossRef]
  36. Mian, M.F.; Ahmed, A.N.; Rad, M.; Babaian, A.; Bowdish, D.; Ashkar, A.A. Length of dsRNA (poly I:C) drives distinct innate immune responses, depending on the cell type. J. Leukoc. Biol. 2013, 94, 1025–1036. [Google Scholar] [CrossRef]
  37. Ramanathan, A.; Robb, G.B.; Chan, S.H. mRNA Capping: Biological Functions and Applications. Nucleic Acids Res. 2016, 44, 7511–7526. [Google Scholar] [CrossRef]
  38. Devarkar, S.C.; Wang, C.; Miller, M.T.; Ramanathan, A.; Jiang, F.; Khan, A.G.; Patel, S.S.; Marcotrigiano, J. Structural Basis for M7g Recognition and 2′-O-Methyl Discrimination in Capped RNAs by the Innate Immune Receptor RIG-I. Proc. Natl. Acad. Sci. USA 2016, 113, 596–601. [Google Scholar] [CrossRef]
  39. Mauro, V.P.; Chappell, S.A. A Critical Analysis of Codon Optimization in Human Therapeutics. Trends Mol. Med. 2014, 20, 604–613. [Google Scholar] [CrossRef]
  40. Gallie, D.R. The Cap and Poly (A) Tail Function Synergistically to Regulate mRNA Translational Efficiency. Genes Dev. 1991, 5, 2108–2116. [Google Scholar] [CrossRef]
  41. Jackson, N.A.; Kester, K.E.; Casimiro, D.; Gurunathan, S.; DeRosa, F. The Promise of mRNA Vaccines: A Biotech and Industrial Perspective. NPJ Vaccines 2020, 5, 11. [Google Scholar] [CrossRef]
  42. Bloom, K.; van den Berg, F.; Arbuthnot, P. Self-Amplifying RNA Vaccines for Infectious Diseases. Gene Ther. 2021, 28, 117–129. [Google Scholar] [CrossRef] [PubMed]
  43. Lundstrom, K. Replicon RNA Viral Vectors as Vaccines. Vaccines 2016, 4, 39–61. [Google Scholar] [CrossRef]
  44. Singh, A.; Koutsoumpli, G.; van de Wall, S.; Daemen, T. An alphavirus-based therapeutic cancer vaccine: From design to clinical trial. Cancer Immunol. Immunother. 2019, 68, 849–859. [Google Scholar] [CrossRef]
  45. Kowalski, P.S.; Rudra, A.; Miao, L.; Anderson, D.G. Delivering the Messenger: Advances in Technologies for Therapeutic mRNA Delivery. Mol. Ther. 2019, 27, 710–728. [Google Scholar] [CrossRef] [PubMed]
  46. Fobian, S.F.; Cheng, Z.; Ten Hagen, T.L.M. Smart Lipid-Based Nanosystems for Therapeutic Immune Induction against Cancers: Perspectives and Outlooks. Pharmaceutics 2021, 14, 26. [Google Scholar] [CrossRef]
  47. Semple, S.C.; Akinc, A.; Chen, J.; Sandhu, A.P.; Mui, B.L.; Cho, C.K.; Sah, D.W.; Stebbing, D.; Crosley, E.J.; Yaworski, E.; et al. Rational design of cationic lipids for siRNA delivery. Nat. Biotechnol. 2010, 28, 172–176. [Google Scholar] [CrossRef] [PubMed]
  48. Love, K.T.; Mahon, K.P.; Levins, C.G.; Whitehead, K.A.; Querbes, W.; Dorkin, J.R.; Qin, J.; Cantley, W.; Qin, L.L.; Racie, T.; et al. Lipid-like materials for low-dose, in vivo gene silencing. Proc. Natl. Acad. Sci. USA 2010, 107, 1864–1869. [Google Scholar] [CrossRef] [PubMed]
  49. Whitehead, K.A.; Sahay, G.; Li, G.Z.; Love, K.T.; Alabi, C.A.; Ma, M.; Zurenko, C.; Querbes, W.; Langer, R.S.; Anderson, D.G. Synergistic silencing: Combinations of lipid-like materials for efficacious siRNA delivery. Mol. Ther. 2011, 19, 1688–1694. [Google Scholar] [CrossRef]
  50. Sabnis, S.; Kumarasinghe, E.S.; Salerno, T.; Mihai, C.; Ketova, T.; Senn, J.J.; Lynn, A.; Bulychev, A.; McFadyen, I.; Chan, J.; et al. A Novel Amino Lipid Series for mRNA Delivery: Improved Endosomal Escape and Sustained Pharmacology and Safety in Non-human Primates. Mol. Ther. 2018, 26, 1509–1519. [Google Scholar] [CrossRef]
  51. Tanaka, H.; Miyama, R.; Sakurai, Y.; Tamagawa, S.; Nakai, Y.; Tange, K.; Yoshioka, H.; Akita, H. Improvement of mRNA Delivery Efficiency to a T Cell Line by Modulating PEG-Lipid Content and Phospholipid Components of Lipid Nanoparticles. Pharmaceutics 2021, 13, 2097. [Google Scholar] [CrossRef]
  52. Yanez Arteta, M.; Kjellman, T.; Bartesaghi, S.; Wallin, S.; Wu, X.; Kvist, A.J.; Dabkowska, A.; Székely, N.; Radulescu, A.; Bergenholtz, J.; et al. Successful reprogramming of cellular protein production through mRNA delivered by functionalized lipid nanoparticles. Proc. Natl. Acad. Sci. USA 2018, 115, E3351–E3360. [Google Scholar] [CrossRef]
  53. Pardi, N.; Hogan, M.J.; Naradikian, M.S.; Parkhouse, K.; Cain, D.W.; Jones, L.; Moody, M.A.; Verkerke, H.P.; Myles, A.; Willis, E.; et al. Nucleoside-modified mRNA vaccines induce potent T follicular helper and germinal center B cell responses. J. Exp. Med. 2018, 215, 1571–1588. [Google Scholar] [CrossRef] [PubMed]
  54. Baden, L.R.; El Sahly, H.M.; Essink, B.; Kotloff, K.; Frey, S.; Novak, R.; Diemert, D.; Spector, S.A.; Rouphael, N.; Creech, C.B.; et al. Efficacy and Safety of the mRNA-1273 SARS-CoV-2 Vaccine. N. Engl. J. Med. 2021, 384, 403–416. [Google Scholar] [CrossRef] [PubMed]
  55. Thomas, S.J.; Moreira, E.D., Jr.; Kitchin, N.; Absalon, J.; Gurtman, A.; Lockhart, S.; Perez, J.L.; Pérez, M.G.; Polack, F.P.; Zerbini, C.; et al. Safety and Efficacy of the BNT162b2 mRNA Covid-19 Vaccine through 6 Months. N. Engl. J. Med. 2021, 385, 1761–1773. [Google Scholar] [CrossRef]
  56. Oberli, M.A.; Reichmuth, A.M.; Dorkin, J.R.; Mitchell, M.J.; Fenton, O.S.; Jaklenec, A.; Anderson, D.G.; Langer, R.; Blankschtein, D. Lipid Nanoparticle Assisted mRNA Delivery for Potent Cancer Immunotherapy. Nano Lett. 2017, 17, 1326–1335. [Google Scholar] [CrossRef] [PubMed]
  57. Fan, Y.N.; Li, M.; Luo, Y.L.; Chen, Q.; Wang, L.; Zhang, H.B.; Shen, S.; Gu, Z.; Wang, J. Cationic lipid-assisted nanoparticles for delivery of mRNA cancer vaccine. Biomater. Sci. 2018, 6, 3009–3018. [Google Scholar] [CrossRef]
  58. Zhang, Y.; Xie, F.; Yin, Y.; Zhang, Q.; Jin, H.; Wu, Y.; Pang, L.; Li, J.; Gao, J. Immunotherapy of Tumor RNA-Loaded Lipid Nanoparticles Against Hepatocellular Carcinoma. Int. J. Nanomed. 2021, 16, 1553–1564. [Google Scholar] [CrossRef] [PubMed]
  59. Kranz, L.M.; Diken, M.; Haas, H.; Kreiter, S.; Loquai, C.; Reuter, K.C.; Meng, M.; Fritz, D.; Vascotto, F.; Hefesha, H.; et al. Systemic RNA delivery to dendritic cells exploits antiviral defence for cancer immunotherapy. Nature 2016, 534, 396–401. [Google Scholar] [CrossRef]
  60. Dahlman, J.E.; Barnes, C.; Khan, O.; Thiriot, A.; Jhunjunwala, S.; Shaw, T.E.; Xing, Y.; Sager, H.B.; Sahay, G.; Speciner, L.; et al. In vivo endothelial siRNA delivery using polymeric nanoparticles with low molecular weight. Nat. Nanotechnol. 2014, 9, 648–655. [Google Scholar] [CrossRef]
  61. Dong, Y.; Dorkin, J.R.; Wang, W.; Chang, P.H.; Webber, M.J.; Tang, B.C.; Yang, J.; Abutbul-Ionita, I.; Danino, D.; DeRosa, F.; et al. Poly(glycoamidoamine) Brushes Formulated Nanomaterials for Systemic siRNA and mRNA Delivery In Vivo. Nano Lett. 2016, 16, 842–848. [Google Scholar] [CrossRef]
  62. Kaczmarek, J.C.; Kauffman, K.J.; Fenton, O.S.; Sadtler, K.; Patel, A.K.; Heartlein, M.W.; DeRosa, F.; Anderson, D.G. Optimization of a Degradable Polymer-Lipid Nanoparticle for Potent Systemic Delivery of mRNA to the Lung Endothelium and Immune Cells. Nano Lett. 2018, 18, 6449–6454. [Google Scholar] [CrossRef] [PubMed]
  63. Chahal, J.S.; Khan, O.F.; Cooper, C.L.; McPartlan, J.S.; Tsosie, J.K.; Tilley, L.D.; Sidik, S.M.; Lourido, S.; Langer, R.; Bavari, S.; et al. Dendrimer-RNA nanoparticles generate protective immunity against lethal Ebola, H1N1 influenza, and Toxoplasma gondii challenges with a single dose. Proc. Natl. Acad. Sci. USA 2016, 113, E4133–E4142. [Google Scholar] [CrossRef] [PubMed]
  64. Chahal, J.S.; Fang, T.; Woodham, A.W.; Khan, O.F.; Ling, J.; Anderson, D.G.; Ploegh, H.L. An RNA nanoparticle vaccine against Zika virus elicits antibody and CD8+ T cell responses in a mouse model. Sci. Rep. 2017, 7, 252. [Google Scholar] [CrossRef] [PubMed]
  65. Uchida, S.; Kinoh, H.; Ishii, T.; Matsui, A.; Tockary, T.A.; Takeda, K.M.; Uchida, H.; Osada, K.; Itaka, K.; Kataoka, K. Systemic delivery of messenger RNA for the treatment of pancreatic cancer using polyplex nanomicelles with a cholesterol moiety. Biomaterials 2016, 82, 221–228. [Google Scholar] [CrossRef]
  66. Chen, Q.; Qi, R.; Chen, X.; Yang, X.; Wu, S.; Xiao, H.; Dong, W. A Targeted and Stable Polymeric Nanoformulation Enhances Systemic Delivery of mRNA to Tumors. Mol. Ther. 2017, 25, 92–101. [Google Scholar] [CrossRef]
  67. Islam, M.A.; Xu, Y.; Tao, W.; Ubellacker, J.M.; Lim, M.; Aum, D.; Lee, G.Y.; Zhou, K.; Zope, H.; Yu, M.; et al. Restoration of tumour-growth suppression in vivo via systemic nanoparticle-mediated delivery of PTEN mRNA. Nat. Biomed. Eng. 2018, 2, 850–864. [Google Scholar] [CrossRef]
  68. Hoerr, I.; Obst, R.; Rammensee, H.G.; Jung, G. In vivo application of RNA leads to induction of specific cytotoxic T lymphocytes and antibodies. Eur. J. Immunol. 2000, 30, 1–7. [Google Scholar] [CrossRef]
  69. Hoyer, J.; Neundorf, I. Peptide vectors for the nonviral delivery of nucleic acids. Acc. Chem. Res. 2012, 45, 1048–1056. [Google Scholar] [CrossRef] [PubMed]
  70. Nakase, I.; Akita, H.; Kogure, K.; Gräslund, A.; Langel, U.; Harashima, H.; Futaki, S. Efficient intracellular delivery of nucleic acid pharmaceuticals using cell-penetrating peptides. Acc. Chem. Res. 2012, 45, 1132–1139. [Google Scholar] [CrossRef]
  71. Reissmann, S. Cell penetration: Scope and limitations by the application of cell-penetrating peptides. J. Pept. Sci. 2014, 20, 760–784. [Google Scholar] [CrossRef]
  72. Li, H.; Tsui, T.Y.; Ma, W. Intracellular Delivery of Molecular Cargo Using Cell-Penetrating Peptides and the Combination Strategies. Int. J. Mol. Sci. 2015, 16, 19518–19536. [Google Scholar] [CrossRef] [PubMed]
  73. Kallen, K.J.; Heidenreich, R.; Schnee, M.; Petsch, B.; Schlake, T.; Thess, A.; Baumhof, P.; Scheel, B.; Koch, S.D.; Fotin-Mleczek, M. A novel, disruptive vaccination technology: Self-adjuvanted RNActive(®) vaccines. Hum. Vaccines Immunother. 2013, 9, 2263–2276. [Google Scholar] [CrossRef] [PubMed]
  74. Qiu, Y.; Man, R.C.H.; Liao, Q.; Kung, K.L.K.; Chow, M.Y.T.; Lam, J.K.W. Effective mRNA pulmonary delivery by dry powder formulation of PEGylated synthetic KL4 peptide. J. Control. Release 2019, 314, 102–115. [Google Scholar] [CrossRef]
  75. Lou, B.; De Koker, S.; Lau, C.Y.J.; Hennink, W.E.; Mastrobattista, E. mRNA Polyplexes with Post-Conjugated GALA Peptides Efficiently Target, Transfect, and Activate Antigen Presenting Cells. Bioconjug. Chem. 2019, 30, 461–475. [Google Scholar] [CrossRef]
  76. Hajj, K.; Whitehead, K. Tools for translation: Non-viral materials for therapeutic mRNA delivery. Nat. Rev. Mater. 2017, 2, 17056. [Google Scholar] [CrossRef]
  77. Mastelic-Gavillet, B.; Balint, K.; Boudousquie, C.; Gannon, P.O.; Kandalaft, L.E. Personalized dendritic cell vaccines—Recent breakthroughs and encouraging clinical results. Front. Immunol. 2019, 10, 766. [Google Scholar] [CrossRef]
  78. Saxena, M.; Balan, S.; Roudko, V.; Bhardwaj, N. Towards superior dendritic-cell vaccines for cancer therapy. Nat. Biomed. Eng. 2018, 2, 341–346. [Google Scholar] [CrossRef]
  79. Perez, C.R.; De Palma, M. Engineering dendritic cell vaccines to improve cancer immunotherapy. Nat. Commun. 2019, 10, 5408–5417. [Google Scholar] [CrossRef]
  80. Nair, S.K.; Heiser, A.; Boczkowski, D.; Majumdar, A.; Naoe, M.; Lebkowski, J.S.; Vieweg, J.; Gilboa, E. Induction of cytotoxic T cell responses and tumor immunity against unrelated tumors using telomerase reverse transcriptase RNA transfected dendritic cells. Nat. Med. 2000, 6, 1011–1017. [Google Scholar] [CrossRef]
  81. Perche, F.; Benvegnu, T.; Berchel, M.; Lebegue, L.; Pichon, C.; Jaffrès, P.A.; Midoux, P. Enhancement of dendritic cells transfection in vivo and of vaccination against B16F10 melanoma with mannosylated histidylated lipopolyplexes loaded with tumor antigen messenger RNA. Nanomedicine 2011, 7, 445–453. [Google Scholar] [CrossRef]
  82. Eshhar, Z.; Waks, T.; Bendavid, A.; Schindler, D.G. Functional expression of chimeric receptor genes in human T cells. J. Immunol. Methods 2001, 248, 67–76. [Google Scholar] [CrossRef]
  83. Kalos, M.; June, C.H. Adoptive T cell transfer for cancer immunotherapy in the era of synthetic biology. Immunity 2013, 39, 49–60. [Google Scholar] [CrossRef] [PubMed]
  84. Irving, M.; Lanitis, E.; Migliorini, D.; Ivics, Z.; Guedan, S. Choosing the Right Tool for Genetic Engineering: Clinical Lessons from Chimeric Antigen Receptor-T Cells. Hum. Gene Ther. 2021, 32, 1044–1058. [Google Scholar] [CrossRef] [PubMed]
  85. Liu, J.; Zhou, G.; Zhang, L.; Zhao, Q. Building Potent Chimeric Antigen Receptor T Cells with CRISPR Genome Editing. Front. Immunol. 2019, 10, 456. [Google Scholar] [CrossRef] [PubMed]
  86. Ye, Z.; Chen, J.; Zhao, X.; Li, Y.; Harmon, J.; Huang, C.; Chen, J.; Xu, Q. In Vitro Engineering Chimeric Antigen Receptor Macrophages and T Cells by Lipid Nanoparticle-Mediated mRNA Delivery. ACS Biomater. Sci. Eng. 2022, 8, 722–733. [Google Scholar] [CrossRef]
  87. Hegde, M.; Corder, A.; Chow, K.K.; Mukherjee, M.; Ashoori, A.; Kew, Y.; Zhang, Y.J.; Baskin, D.S.; Merchant, F.A.; Brawley, V.S.; et al. Combinational targeting offsets antigen escape and enhances effector functions of adoptively transferred T cells in glioblastoma. Mol. Ther. 2013, 21, 2087–2101. [Google Scholar] [CrossRef]
  88. Barrett, D.M.; Zhao, Y.; Liu, X.; Jiang, S.; Carpenito, C.; Kalos, M.; Carroll, R.G.; June, C.H.; Grupp, S.A. Treatment of advanced leukemia in mice with mRNA engineered T cells. Hum. Gene Ther. 2011, 22, 1575–1586. [Google Scholar] [CrossRef]
  89. Barrett, D.M.; Liu, X.; Jiang, S.; June, C.H.; Grupp, S.A.; Zhao, Y. Regimen-specific effects of RNA-modified chimeric antigen receptor T cells in mice with advanced leukemia. Hum. Gene Ther. 2013, 24, 717–727. [Google Scholar] [CrossRef]
  90. Rajan, S.T.; Gugliandolo, A.; Bramanti, P.; Mazzon, E. In Vitro-Transcribed mRNA Chimeric Antigen Receptor T Cell (IVT mRNA CAR T) Therapy in Hematologic and Solid Tumor Management: A Preclinical Update. Int. J. Mol. Sci. 2020, 21, 6514. [Google Scholar] [CrossRef]
  91. Ljungberg, K.; Liljeström, P. Self-replicating alphavirus RNA vaccines. Expert Rev. Vaccines 2015, 14, 177–194. [Google Scholar] [CrossRef]
  92. Smerdou, C.; Liljeström, P. Two-helper RNA system for production of recombinant Semliki forest virus particles. J. Virol. 1999, 73, 1092–1098. [Google Scholar] [CrossRef]
  93. Beissert, T.; Perkovic, M.; Vogel, A.; Erbar, S.; Walzer, K.C.; Hempel, T.; Brill, S.; Haefner, E.; Becker, R.; Türeci, Ö.; et al. A Trans-amplifying RNA Vaccine Strategy for Induction of Potent Protective Immunity. Mol. Ther. 2020, 28, 119–128. [Google Scholar] [CrossRef] [PubMed]
  94. Biddlecome, A.; Habte, H.H.; McGrath, K.M.; Sambanthamoorthy, S.; Wurm, M.; Sykora, M.M.; Knobler, C.M.; Lorenz, I.C.; Lasaro, M.; Elbers, K.; et al. Delivery of self-amplifying RNA vaccines in in vitro reconstituted virus-like particles. PLoS ONE 2019, 14, e0215031. [Google Scholar] [CrossRef] [PubMed]
  95. Nelson, E.L.; Prieto, D.; Alexander, T.G.; Pushko, P.; Lofts, L.A.; Rayner, J.O.; Kamrud, K.I.; Fralish, B.; Smith, J.F. Venezuelan equine encephalitis replicon immunization overcomes intrinsic tolerance and elicits effective anti-tumor immunity to the ‘self’ tumor-associated antigen, neu in a rat mammary tumor model. Breast Cancer Res. Treat. 2003, 82, 169–183. [Google Scholar] [CrossRef]
  96. Riabov, V.; Tretyakova, I.; Alexander, R.B.; Pushko, P.; Klyushnenkova, E.N. Anti-tumor effect of the alphavirus-based virus-like particle vector expressing prostate-specific antigen in a HLA-DR transgenic mouse model of prostate cancer. Vaccine 2015, 33, 5386–5395. [Google Scholar] [CrossRef]
  97. Zhang, X.; Mao, G.; van den Pol, A.N. Chikungunya-vesicular stomatitis chimeric virus targets and eliminates brain tumors. Virology 2018, 522, 244–259. [Google Scholar] [CrossRef]
  98. Velders, M.P.; McElhiney, S.; Cassetti, M.C.; Eiben, G.L.; Higgins, T.; Kovacs, G.R.; Elmishad, A.G.; Kast, W.M.; Smith, L.R. Eradication of established tumors by vaccination with Venezuelan equine encephalitis virus replicon particles delivering human papillomavirus 16 E7 RNA. Cancer Res. 2001, 61, 7861–7867. [Google Scholar] [PubMed]
  99. Sebastian, M.; Schröder, A.; Scheel, B.; Hong, H.S.; Muth, A.; von Boehmer, L.; Zippelius, A.; Mayer, F.; Reck, M.; Atanackovic, D.; et al. A phase I/IIa study of the mRNA-based cancer immunotherapy CV9201 in patients with stage IIIB/IV non-small cell lung cancer. Cancer Immunol. Immunother. 2019, 68, 799–812. [Google Scholar] [CrossRef]
  100. Kübler, H.; Scheel, B.; Gnad-Vogt, U.; Miller, K.; Schultze-Seemann, W.; Vom Dorp, F.; Parmiani, G.; Hampel, C.; Wedel, S.; Trojan, L.; et al. Self-adjuvanted mRNA vaccination in advanced prostate cancer patients: A first-in-man phase I/IIa study. J. Immunother. Cancer 2015, 3, 26. [Google Scholar] [CrossRef]
  101. Weide, B.; Pascolo, S.; Scheel, B.; Derhovanessian, E.; Pflugfelder, A.; Eigentler, T.K.; Pawelec, G.; Hoerr, I.; Rammensee, H.G.; Garbe, C. Direct injection of protamine-protected mRNA: Results of a phase 1/2 vaccination trial in metastatic melanoma patients. J. Immunother. 2009, 32, 498–507. [Google Scholar] [CrossRef]
  102. Rittig, S.M.; Haentschel, M.; Weimer, K.J.; Heine, A.; Muller, M.R.; Brugger, W.; Horger, M.S.; Maksimovic, O.; Stenzl, A.; Hoerr, I.; et al. Intradermal vaccinations with RNA coding for TAA generate CD8+ and CD4+ immune responses and induce clinical benefit in vaccinated patients. Mol. Ther. 2011, 19, 990–999. [Google Scholar] [CrossRef]
  103. Rittig, S.M.; Haentschel, M.; Weimer, K.J.; Heine, A.; Müller, M.R.; Brugger, W.; Horger, M.S.; Maksimovic, O.; Stenzl, A.; Hoerr, I.; et al. Long-term survival correlates with immunological responses in renal cell carcinoma patients treated with mRNA-based immunotherapy. Oncoimmunology 2015, 5, e1108511. [Google Scholar] [CrossRef] [PubMed]
  104. Papachristofilou, A.; Hipp, M.M.; Klinkhardt, U.; Früh, M.; Sebastian, M.; Weiss, C.; Pless, M.; Cathomas, R.; Hilbe, W.; Pall, G.; et al. Phase Ib evaluation of a self-adjuvanted protamine formulated mRNA-based active cancer immunotherapy, BI1361849 (CV9202), combined with local radiation treatment in patients with stage IV non-small cell lung cancer. J. Immunother. Cancer 2019, 7, 38. [Google Scholar] [CrossRef] [PubMed]
  105. Sahin, U.; Oehm, P.; Derhovanessian, E.; Jabulowsky, R.A.; Vormehr, M.; Gold, M.; Maurus, D.; Schwarck-Kokarakis, D.; Kuhn, A.N.; Omokoko, T.; et al. An RNA vaccine drives immunity in checkpoint-inhibitor-treated melanoma. Nature 2020, 585, 107–112. [Google Scholar] [CrossRef] [PubMed]
  106. Heiser, A.; Coleman, D.; Dannull, J.; Yancey, D.; Maurice, M.A.; Lallas, C.D.; Dahm, P.; Niedzwiecki, D.; Gilboa, E.; Vieweg, J. Autologous dendritic cells transfected with prostate-specific antigen RNA stimulate CTL responses against metastatic prostate tumors. J. Clin. Investig. 2002, 109, 409–417. [Google Scholar] [CrossRef]
  107. Su, Z.; Dannull, J.; Heiser, A.; Yancey, D.; Pruitt, S.; Madden, J.; Coleman, D.; Niedzwiecki, D.; Gilboa, E.; Vieweg, J. Immunological and clinical responses in metastatic renal cancer patients vaccinated with tumor RNA-transfected dendritic cells. Cancer Res. 2003, 63, 2127–2133. [Google Scholar]
  108. Morse, M.A.; Nair, S.K.; Mosca, P.J.; Hobeika, A.C.; Clay, T.M.; Deng, Y.; Boczkowski, D.; Proia, A.; Neidzwiecki, D.; Clavien, P.A.; et al. Immunotherapy with autologous, human dendritic cells transfected with carcinoembryonic antigen mRNA. Cancer Investig. 2003, 21, 341–349. [Google Scholar] [CrossRef]
  109. Morse, M.A.; Nair, S.K.; Boczkowski, D.; Tyler, D.; Hurwitz, H.I.; Proia, A.; Clay, T.M.; Schlom, J.; Gilboa, E.; Lyerly, H.K. The feasibility and safety of immunotherapy with dendritic cells loaded with CEA mRNA following neoadjuvant chemoradiotherapy and resection of pancreatic cancer. Int. J. Gastrointest. Cancer 2002, 32, 1–6. [Google Scholar] [CrossRef]
  110. Wilgenhof, S.; Van Nuffel, A.M.T.; Benteyn, D.; Corthals, J.; Aerts, C.; Heirman, C.; Van Riet, I.; Bonehill, A.; Thielemans, K.; Neyns, B. A phase IB study on intravenous synthetic mRNA electroporated dendritic cell immunotherapy in pretreated advanced melanoma patients. Ann. Oncol. 2013, 24, 2686–2693. [Google Scholar] [CrossRef] [PubMed]
  111. Jansen, Y.; Kruse, V.; Corthals, J.; Schats, K.; van Dam, P.J.; Seremet, T.; Heirman, C.; Brochez, L.; Kockx, M.; Thielemans, K.; et al. A randomized controlled phase II clinical trial on mRNA electroporated autologous monocyte-derived dendritic cells (TriMixDC-MEL) as adjuvant treatment for stage III/IV melanoma patients who are disease-free following the resection of macrometastases. Cancer Immunol. Immunother. 2020, 69, 2589–2598. [Google Scholar] [CrossRef]
  112. Wilgenhof, S.; Corthals, J.; Heirman, C.; van Baren, N.; Lucas, S.; Kvistborg, P.; Thielemans, K.; Neyns, B. Phase II Study of Autologous Monocyte-Derived mRNA Electroporated Dendritic Cells (TriMixDC-MEL) Plus Ipilimumab in Patients with Pretreated Advanced Melanoma. J. Clin. Oncol. 2016, 34, 1330–1338. [Google Scholar] [CrossRef] [PubMed]
  113. Wang, Q.T.; Nie, Y.; Sun, S.N.; Lin, T.; Han, R.J.; Jiang, J.; Li, Z.; Li, J.Q.; Xiao, Y.P.; Fan, Y.Y.; et al. Tumor-associated antigen-based personalized dendritic cell vaccine in solid tumor patients. Cancer Immunol. Immunother. 2020, 69, 1375–1387. [Google Scholar] [CrossRef] [PubMed]
  114. Lundstrom, K. Self-Replicating RNA Viruses for Vaccine Development against Infectious Diseases and Cancer. Vaccines 2021, 9, 1187. [Google Scholar] [CrossRef] [PubMed]
  115. Komdeur, F.L.; Singh, A.; van de Wall, S.; Meulenberg, J.J.M.; Boerma, A.; Hoogeboom, B.N.; Paijens, S.T.; Oyarce, C.; de Bruyn, M.; Schuuring, E.; et al. First-in-Human Phase I Clinical Trial of an SFV-Based RNA Replicon Cancer Vaccine against HPV-Induced Cancers. Mol. Ther. 2021, 29, 611–625. [Google Scholar] [CrossRef] [PubMed]
  116. Morse, M.A.; Hobeika, A.C.; Osada, T.; Berglund, P.; Hubby, B.; Negri, S.; Niedzwiecki, D.; Devi, G.R.; Burnett, B.K.; Clay, T.M.; et al. An alphavirus vector overcomes the presence of neutralizing antibodies and elevated numbers of Tregs to induce immune responses in humans with advanced cancer. J. Clin. Investig. 2010, 120, 3234–3241. [Google Scholar] [CrossRef] [PubMed]
  117. Crosby, E.J.; Hobeika, A.C.; Niedzwiecki, D.; Rushing, C.; Hsu, D.; Berglund, P.; Smith, J.; Osada, T.; Gwin, W.R., III; Hartman, Z.C.; et al. Long-term survival of patients with stage III colon cancer treated with VRP-CEA(6D), an alphavirus vector that increases the CD8+ effector memory T cell to Treg ratio. J. Immunother. Cancer 2020, 8, e001662. [Google Scholar] [CrossRef]
  118. Crosby, E.J.; Gwin, W.; Blackwell, K.; Marcom, P.K.; Chang, S.; Maecker, H.T.; Broadwater, G.; Hyslop, T.; Kim, S.; Rogatko, A.; et al. Vaccine-Induced Memory CD8+T Cells Provide Clinical Benefit in HER2 Expressing Breast Cancer: A Mouse to Human Translational Study. Clin. Cancer Res. 2019, 25, 2725–2736. [Google Scholar] [CrossRef]
  119. Aliahmad, P.; Miyake-Stoner, S.J.; Geall, A.J.; Wang, N.S. Next generation self-replicating RNA vectors for vaccines and immunotherapies. Cancer Gene Ther. 2022, 22, 1–9. [Google Scholar] [CrossRef]
  120. Dimitri, A.; Herbst, F.; Fraietta, J.A. Engineering the next-generation of CAR T-cells with CRISPR-Cas9 gene editing. Mol. Cancer 2022, 21, 78. [Google Scholar] [CrossRef]
  121. Giuffrida, L.; Sek, K.; Henderson, M.A.; Lai, J.; Chen, A.X.Y.; Meyran, D.; Todd, K.; Petley, E.V.; Mardiana, S.; Mølck, C.; et al. CRISPR/Cas9 mediated deletion of the adenosine A2A receptor enhances CAR T cell efficacy. Nat. Commun. 2021, 12, 3236. [Google Scholar] [CrossRef]
  122. Zhang, P.; Lim, S.B.; Jiang, K.; Chew, T.W.; Low, B.C.; Lim, C.T. Distinct mRNAs in Cancer Extracellular Vesicles Activate Angiogenesis and Alter Transcriptome of Vascular Endothelial Cells. Cancers 2021, 13, 2009. [Google Scholar] [CrossRef]
  123. Zhang, Z.; Zhang, C.; Luo, Y.; Zhang, G.; Wu, P.; Sun, N.; He, J. RNA N6-methyladenosine modification in the lethal teamwork of cancer stem cells and the tumor immune microenvironment: Current landscape and therapeutic potential. Clin. Transl. Med. 2021, 11, e525. [Google Scholar] [CrossRef] [PubMed]
  124. Hewitt, S.L.; Bai, A.; Bailey, D.; Ichikawa, K.; Zielinski, J.; Karp, R.; Apte, A.; Arnold, K.; Zacharek, S.J.; Iliou, M.S.; et al. Durable anticancer immunity from intratumoral administration of IL-23, IL-36γ, and OX40L mRNAs. Sci. Transl. Med. 2019, 11, eaat9143. [Google Scholar] [CrossRef] [PubMed]
  125. Patel, M.R.; Todd, M.B.; Jimeno, A.; Wang, D.; LoRusso, P.; Do, K.T.; Stemmer, S.M.; Maurice-Dror, C.; Geva, R.; Zacharek, S.; et al. A phase I study of mRNA-2752, a lipid nanoparticle encapsulating mRNAs encoding human OX40L, IL-23, and IL-36γ, for intratumoral (iTu) injection alone and in combination with durvalumab. J. Clin. Oncol. 2020, 38 (Suppl. 15), 3092. [Google Scholar] [CrossRef]
  126. Van Lint, S.; Renmans, D.; Broos, K.; Goethals, L.; Maenhout, S.; Benteyn, D.; Goyvaerts, C.; du Four, S.; van der Jeught, K.; Bialkowski, L.; et al. Intratumoral delivery of TriMix mRNA results in T-cell activation by cross-presenting dendritic cells. Cancer Immunol Res. 2016, 4, 146–156. [Google Scholar] [CrossRef] [PubMed]
  127. Jimeno, A. A phase 1/2, open-label, multicenter, dose escalation and efficacy study of mRNA-2416, a lipid nanoparticle encapsulated mRNA encoding human OX40L, for intratumoral injection alone or in combination with durvalumab for patients with advanced malignancies. Abstract CT032. In Proceedings of the American Association for Cancer Research Annual Meeting, Philadelphia, PA, USA, 27–28 April 2020; 22–24 June 2020. [Google Scholar]
  128. Haabeth, O.A.W.; Blake, T.R.; McKinlay, C.J.; Tveita, A.A.; Sallets, A.; Waymouth, R.M.; Wender, P.A.; Levy, R. Local Delivery of Ox40l, Cd80, and Cd86mRNA Kindles Global Anticancer Immunity. Cancer Res. 2019, 79, 1624–1634. [Google Scholar] [CrossRef] [PubMed]
  129. Spiess, C.; Zhai, Q.; Carter, P.J. Alternative molecular formats and therapeutic applications for bispecific antibodies. Mol. Immunol. 2015, 67 Pt A, 95–106. [Google Scholar] [CrossRef]
  130. Pruitt, S.K.; Boczkowski, D.; de Rosa, N.; Haley, N.R.; Morse, M.A.; Tyler, D.S.; Dannull, J.; Nair, S. Enhancement of anti-tumor immunity through local modulation of CTLA-4 and GITR by dendritic cells. Eur. J. Immunol. 2011, 41, 3553–3563. [Google Scholar] [CrossRef]
  131. Nguyen, K.B.; Spranger, S. Modulation of the immune microenvironment by tumor-intrinsic oncogenic signaling. J. Cell Biol. 2020, 219, e201908224. [Google Scholar] [CrossRef]
  132. Haanen, J.B.; Mackensen, A.; Koenecke, C.; Alsdorf, W.; Desuki, A.; Wagner-Drouet, E.; Heudobler, D.; Borchmann, P.; Wiegert, E.; Schulz, C.; et al. BNT211: A Phase I trial to evaluate safety and efficacy of CLDN6 CAR-T cells and CARVac-mediated in vivo expansion in patients with CLDN6-positive advanced solid tumors. Abstract CT002. In Proceedings of the American Association for Cancer Research Annual Meeting, New Orleans, LA, USA, 8–13 April 2022. [Google Scholar]
  133. Barbier, A.J.; Jiang, A.Y.; Zhang, P.; Wooster, R.; Anderson, D.G. The clinical progress of mRNA vaccines and immunotherapies. Nat. Biotechnol. 2022, 40, 840–854. [Google Scholar] [CrossRef] [PubMed]
  134. Gupta, A.; Andresen, J.L.; Manan, R.S.; Langer, R. Nucleic acid delivery for therapeutic applications. Adv. Drug Deliv. Rev. 2021, 178, 113834. [Google Scholar] [CrossRef] [PubMed]
  135. Mendes, B.B.; Conniot, J.; Avital, A.; Yao, D.; Jiang, X.; Zhou, X.; Sharf-Pauker, N.; Xiao, Y.; Adir, O.; Liang, H.; et al. Nanodelivery of nucleic acids. Nat. Rev. Methods Primers 2022, 2, 24. [Google Scholar] [CrossRef] [PubMed]
  136. Cheng, Q.; Wei, T.; Farbiak, L.; Johnson, L.T.; Dilliard, S.A.; Siegwart, D.J. Selective organ targeting (SORT) nanoparticles for tissue-specific mRNA delivery and CRISPR-Cas gene editing. Nat. Nanotechnol. 2020, 15, 313–320. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic diagram of intracellular events in mRNA processing by the antigen presenting cells. mRNA enters the cell through the cytosol and translated by the ribosome into the encoded antigen. The antigen is then: (A) degraded into small protein fragments and epitope by the proteosome, which combine with the MHC-I complex at the rough endoplasmic reticulum and traffics to the cell membrane for presentation to naive CD8 (+) cytotoxic T cells; (B) either exocytosed to re-enter the APC through endocytosis or enters the autophagic pathway. Then, the antigen is split into fragments and its epitopes by lysosomal degradation. These epitopes bind with the MHC-II complex and are transferred to the cell membrane to activate naive CD4 (+) T lymphocytes. (MHC: major histocompatibility complex, RER: rough endoplasmic reticulum).
Figure 1. Schematic diagram of intracellular events in mRNA processing by the antigen presenting cells. mRNA enters the cell through the cytosol and translated by the ribosome into the encoded antigen. The antigen is then: (A) degraded into small protein fragments and epitope by the proteosome, which combine with the MHC-I complex at the rough endoplasmic reticulum and traffics to the cell membrane for presentation to naive CD8 (+) cytotoxic T cells; (B) either exocytosed to re-enter the APC through endocytosis or enters the autophagic pathway. Then, the antigen is split into fragments and its epitopes by lysosomal degradation. These epitopes bind with the MHC-II complex and are transferred to the cell membrane to activate naive CD4 (+) T lymphocytes. (MHC: major histocompatibility complex, RER: rough endoplasmic reticulum).
Vaccines 10 01262 g001
Figure 2. Schematic structure of an mRNA vaccine construct: (A) non-replicating and (B) self-replicating, 5′ cap—all eukaryotic mRNA has a cap that contains an m7GpppN structure, preserved throughout evolution. The cap structure not only prevents degradation, but also assists in binding with the eIF to activate translation. Untranslated regions regulate the translational efficiency, whereas the coding sequence contains codons that encode the gene of interest. The poly-A tail acts to maintain the stability of the RNA molecule [37,38,39,40].
Figure 2. Schematic structure of an mRNA vaccine construct: (A) non-replicating and (B) self-replicating, 5′ cap—all eukaryotic mRNA has a cap that contains an m7GpppN structure, preserved throughout evolution. The cap structure not only prevents degradation, but also assists in binding with the eIF to activate translation. Untranslated regions regulate the translational efficiency, whereas the coding sequence contains codons that encode the gene of interest. The poly-A tail acts to maintain the stability of the RNA molecule [37,38,39,40].
Vaccines 10 01262 g002
Table 1. Ongoing trials with mRNA-based therapeutics.
Table 1. Ongoing trials with mRNA-based therapeutics.
Trial IDTrial DesignTarget Patient Population (n)Cancer TypeInvestigational TreatmentPrimary OutcomesTrial Responsible Party/Collaborators
Cancer vaccines:
NCT05192460Phase I;
Dose escalation and expansion with mRNA vaccine (PGV002)
Adult patients
(n: 36)
Advanced gastric cancer, esophageal cancer, and liver cancerDose expansion: vaccine + PD-1/L1 inhibitorSafety
Tolerability
Feasibility
Affiliated Hospital of the Chinese Academy of Military Medical Sciences, China
NCT05202561Phase I;
Open label; 2 arms;
Arm I: mRNA cancer vaccine
Adult patients with HLA-A11:01 or C08:02 subtype
(n: 10)
Refractory advanced solid tumors with KRAS mutArm II: vaccine + PD-1 inhibitor (Navuilumab)Safety
Tolerability
Feasibility
Bengbu Medical College, China
NCT04534205
AHEAD-MERIT
Phase II;
Open label nonrandomized 2 arm with run-in dose evaluation;
mRNA vaccine + PD-1 inh vs. PD-1 inh monotherapy
Adult patients
(n: 285)
Unresectable recurrent or metastatic HPV16+ HNSCC expressing PD-L1 with CPS ≥ 1BNT113 (HPV 16 B6/7 mRNA vaccine) + PembrolizumabRun-in: Safety
Phase II: OS and ORR; QoL
BionTech SE
NCT03313778
KEYNOTE 603
Phase I, Open label, dose escalation mRNA-4157 vaccine monotherapy (Part A); combined with PD-1 inhibitor (Part B, C, D);Adult patients
(n: 142)
Part A: clinically disease-free after early cancer diagnosis
Part B, C: unresectable (locally advanced or metastatic) solid malignancies
NSCLC, SCLC, HPV (-) HNSCC; Bladder urothelial; melanoma; MSI-H; high TMB
Part D: resected melanoma
Part B, C, D: mRNA-4157 vaccine (lipid encapsulated mRNA vaccine encoding 20 tumor neoantigens) + PembrolizumabSafetyModerna TX, Inc.
NCT01686334
WIDEA
Phase II randomized; Open label
mRNA dendritic vaccine vs. surveillance
Adult patients
(n: 130)
AML with minimal residual disease following front-line chemotherapy (morphological CR or CRi)Autologous dendritic cells loaded by mRNA electroporation with the Wilms’ tumor antigen (WT1)OSAntwerp University Hospital; Belgium
NCT04526899Phase II randomized; Open label
BNT111 and Cemiplimab Combination vs. single agents
Adult patients
(n: 180)
Anti-PD-1-refractory/Relapsed, Unresectable Stage III or IV Melanoma; ≥1–5 prior lines treatment including nivolumab/pembrolizumab or BRAFinhBNT111 and Cemiplimab Combination vs. BNT111 (mRNA vaccine encoding 4 melanoma tumor antigents- NY-ESO-1, MAGE-A3, tyrosinase, and TPTE) vs. CemiplimabORRBionTech SE
NCT04573140
PNOC020
Phase I; dose escalation;
Autologous LP-mRNA tumor vaccine
Pediatric and adult Patients
(n: 28)
Newly diagnosed
adult MGMT unmethylated glioblastoma and Pediatric High-Grade Gliomas (pHGG); <3 cm residual tumor following surgery and completed chemoradiation
Autologous total tumor mRNA and pp65 lysosomal associated membrane protein (LAMP) loaded lipid particles (liposomal vaccine)Feasibility, Safety, Dose findingUniversity of Florida
NCT04911621
ADDICT-PedGLIO
Phase I–II
mRNA loaded autologous mRNA dendritic cell vaccine
Pediatric Patients (Aged ≥ 12 months and < 18 years)
(n: 10)
Adjuvant Dendritic Cell Immunotherapy complementing standard therapy in High-grade Glioma and Diffuse Intrinsic Pontine GliomaWT1 mRNA-loaded autologous monocyte derived DC: Phase I newly diagnosed: combined with first line chemoradiation treatment Phase II prior therapy: Dendritic cell vaccination plus optional conventional antiglioma treatmentFeasibility, SafetyUniversity Hospital, Antwerp,
Belgium
NCT02465268
ATTAC-II
Phase II Randomized, Blinded, and Placebo-controlled; Autologous LP-mRNA dendritic cell vaccine with chemotherapyAdult patients
(n: 175)
Adjuvant CMV RNA-Pulsed Dendritic Cells with Tetanus–Diphtheria Toxoid Vaccine; Newly Diagnosed Glioblastoma with < 3 cm residual tumor following surgery and completed chemoradiationmRNA DCs encoding the pp65 neoantigen and LAMP (lysosomal associated membrane protein) with GM-CSF vs. placebo and unpulsed PBMC combined with adjuvant TMZOSImmunomic Therapeutics, Inc.; University of Florida; NCI
NCT03688178
DERIVe
Phase II Randomized, Blinded; Autologous LP-mRNA dendritic cell vaccine alone or combined with CD27 mabAdult patients
(n: 80)
Adjuvant CMV pp65-LAMP mRNA-pulsed autologous DCs ± Varlilumab; Newly Diagnosed Glioblastoma with < 1 cm residual tumor following surgery and completed chemoradiationAdjuvant CMV RNA-Pulsed Dendritic Cells with pp65-lysosomal-associated membrane protein DCs ± anti CD27 mAb (Varlilumab) and Td preconditioning during adjuvant TMZ
Group 1 and 2 (blinded)
Group 3 (nonblinded)
OS
Safety
Change in Treg Depletion
Duke University
Celldex Therapeutics
NCT05357898Phase I/II first in human, open labelEngineered vaccine alone and combined with chemotherapyAdult patients
(n: 60)
Recurrent, locally advanced, or metastatic HPV16+ solid tumors (head and neck, cervical, anal, vulvar, or penile cancer)SQZ-eAPC-HPV vaccine (mRNA engineered APC-targeting multiple tumor antigens and encoding cytokines) as monotherapy and in combination with pembrolizumab Safety, Dose-findingSQZ Biotechnologies
NCT03548571
DEN-STEM
Phase II–III; Open, randomized study
mRNA pulsed dendritic cell therapy vs. standard therapy
Adult patients
(n: 60)
Newly diagnosed IDH wild-type, MGMT-methylated glioblastoma with <1 mm3 residual tumor following surgery and completed chemoradiationAdjuvant autologous trivalent dendritic cells transfected with tm stem cells, survivin, and hTERT combined with TMZ
compared to TMZ after surgery and RT
PFSOslo University Hospital
NCT04382898
PRO-MERIT
Phase I–II; Open label
Dose expansion of W_pro1 vaccine alone and combined with PD-1 inhibitor
Adult patients (n: 130)Metastatic castration-resistant prostate cancer (mCRPC) progressing after 2–3 prior lines of treatment; localized high risk prostate cancer (LPC)W_pro1 liposomal mRNA vaccine encoding 5 tumor antigens
Part 1, Part 2-1B (mCRPC): dose finding; Part 2-1A (mCRPC): vaccine + Cemiplimab
Part 2-2 (LPC): vaccine; Part 2-3 (LPC): vaccine + Cemiplimab
Safety, ORRBionTech SE
NCT03739931Phase I
Open label, dose escalation study of mRNA-2752 alone and combined with PD-L1 inhibition
Adult patients (n: 264)Advanced or metastatic solid tumor malignancies (TNBC, HNSCC, NSCLC, urothelial cancer, melanoma) or lymphoma progressing after standard 1 line of prior therapyArm A: mRNA 2752 alone
Arm B: mRNA 2752 + Durvalumab
Safety, ORRModernaTX, Inc.
AstraZeneca
NCT03788083
TMBA
Phase I
Open label, intratumoral TriMix injection compared with placebo
Adult patients (n: 36)Newly diagnosed stage 1–2 breast cancer; intratumoral administration before surgeryDose escalation of TriMix (naked mRNA vaccine encoding CD70, CD40 ligand, and constitutively active TLR4 that activate
dendritic cells)
Safety; Immune-modulatory EffectUniversitair Ziekenhuis, Brussels
Nonvaccine therapies
NCT04981691
(Amaretto)
Phase I, mRNA-engineered anti-Mesothelin CAR-T cells therapyAdult patients
(n: 12)
Unresectable or metastatic mesothelin expression-positive, advanced solid tumorsDose-escalation of mRNA transduced mesothelin expressing CAR-T cells SafetyRuijin Hospital
UTC Therapeutics Inc
NCT04683939Phase I/IIa dose escalation; Open label; BNT 141 alone and combined with chemotherapyAdult patients
(n: 96)
Unresectable or metastatic Claudin 18.2 (CLDN18.2)-positive GI, hepatobiliary or ovarian cancerPart 1a: Dose-escalation monotherapy with BNT 141 (mRNA-encoded mAb targeting claudin 18.2)
Part 1b: Dose escalation with Nab-Pac and gemcitabine
Safety, Dose findingBionTech SE
NCT04995536Phase I
CpG-STAT3 siRNA combined with RT
Adult patients (n: 18)Recurrent/Refractory B-cell NHL; ≥2 prior lines treatmentDose escalation of siRNA targeting TLR9 and STAT3 with local RTSafety, Dose findingCity of Hope Medical Center
NCI
NCT05392699Phase I
ABOD2011 hsc IL-12 mRNA
Adult patients (n: 60)Recurrent/Refractory solid tumors progressing after standard therapyABOD2011
(Humanized Single chain mRNA encoding IL-12)
Safety, Dose findingCancer Institute and Hospital, Chinese Academy of Medical Sciences
Abbreviations: HNSCC: head and neck squamous cell cancer; KRAS: Kirsten rat sarcoma virus; NSCLC: nonsmall cell lung cancer; SCLC: small cell lung cancer; MSI-H: microsatellite instability—high; TMB: tumor mutation burden; AML: acute myeloid leukemia; BRAF: murine sarcoma viral oncogene homolog B; NY-ESO-1: New York esophageal squamous cell carcinoma-1; MAGE-A3: melanoma-associated antigen 3; TPTE: tyrosine-protein phosphatase; MGMT: O-6-methylguanine-DNA methyltransferase; TMZ: temozolamide; RT: radiotherapy; GI: gastrointestinal; Nab-Pac: nao-bound paclitaxel; NHL: non-Hodgkin lymphoma; STAT: signal transduction and activator of transcription 1.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Eralp, Y. Application of mRNA Technology in Cancer Therapeutics. Vaccines 2022, 10, 1262. https://doi.org/10.3390/vaccines10081262

AMA Style

Eralp Y. Application of mRNA Technology in Cancer Therapeutics. Vaccines. 2022; 10(8):1262. https://doi.org/10.3390/vaccines10081262

Chicago/Turabian Style

Eralp, Yesim. 2022. "Application of mRNA Technology in Cancer Therapeutics" Vaccines 10, no. 8: 1262. https://doi.org/10.3390/vaccines10081262

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