Engineering

Multi-Modal RNA

Given the inherent heterogeneity of cancer in its causation and progression, CAR_Ma was designed to be multi-modal in attempts to fully address the many facets of the complex disease, whilst at the same time optimizing the resource utilization and function of transfected macrophages against cancer cells. This year’s CAR_Ma was designed with elements that require three modal functions: transience, replicative, and release. The VEE replication complex designed into CAR_Ma facilitates the functioning of the different modalities, allowing for different sections of the RNA construct to be transient, repeated, or released.


Cancer is a Complex Problem Cancer in itself is unilaterally complex. A multifactorial disease capable of both sporadic and familial origins (Fridovich et al., 1998). Cancer has and will eventually evolve both within the body and alongside humanity. Throughout a tumor’s development from initiation to late metastasis, the tumor itself evolves to sustain itself and avoid detection and destruction. Tumors modulate and protect themselves by maintaining complex relationships with immune cells enabling tumor evasion and shielding itself within its own tumor microenvironment (Baghban et al., 2020; De Visser & Joyce, 2023; Kim & Cho, 2022). To tackle such a multifaceted and ever-evolving enemy, any solution needs to be numerous in its methods and modalities as well.


Enabling Multimodality Multimodality in RNA depends on the ability of different genes and elements to be active or translated whilst on the same, singular RNA strand - requiring polycistronic behavior. However, mammalian mRNA molecules are typically monocistronic (Karginov et al., 2017). To enable multimodality, CAR_Ma was designed with the addition of 26S subgenomic promoters (26S), as well as internal ribosome entry sites (IRES), which effectively convert monocistronic RNA to polycistronic RNA (Karginov et al., 2017). hat as many people globally will be able to receive the most advanced therapeutics.

26S is a subgenomic promoter that directs transcription for the subsequent downstream gene, resulting in the formation of mRNA for the single gene (Kamrud et al., 2007). IRES elements are responsible for assembling functional ribosomes at a start codon, allowing for 5’ cap-independent translation initiation, thus allowing for genes downstream of IRES to be translated (Kamrud et al., 2007). Our design adds several 26S and IRES elements in our multimodal RNA, allowing for multiple genes to be present, co-transcribable, and co-translatable.


Transience The transient modality of CAR_Ma focuses on improving therapeutic efficiency while retaining clinical applicability and safety. CAR_Ma itself is a transient system. By relying on transience, we are able to avoid safety risks and risks of malignancies with DNA or viral-based treatments, which may lead to random insertional mutagenesis. Such issues have plagued even the latest CAR immunotherapies, which caused the CAR-T cells to become malignant and are required to come with boxed warnings (U.S. Food and Drug Administration, 2024). Further uses of the transient system facilitate the elimination of immune evasion by tumor cells using CRISPR/Cas systems, or RNAi pipelines. Many cancers hijack systems such as SIRPα-CD47 or PD-1/PD-L1 to evade detection and subsequent targeting by the immune system, preventing clearance of cancerous cells. Knockout of SIRPα expression on macrophages at the genomic level by the CRISPR/Cas systems using Cas9-GFP or Cas12 prevents the SIRPα-CD47 interaction, inhibiting immune evasion via this pathway. By knocking out these evasion pathways, immune cells are able to better recognize their immune targets and destroy the tumors (Ray et al., 2018).

Given knockout by a CRISPR/Cas system occurs at the genomic level, the expression of Cas is not required to be continuously maintained, as a genomic knockout only needs to occur once to result in permanent changes in the expression. Thus, the permanence of genomic changes caused by CRISPR/Cas systems implies the need for a transient modality to avoid unnecessary expenditure of cellular resources. Alternatively, this action can also be carried out via RNAi instead of CRISPR using the release cycle.


Replicative Replicative refers to the need for sustained expression via constant replication. CAR_Ma’s replicative modality contains elements that must be continuously maintained to allow for safe and sustained expression and efficacy of our immunotherapy: a 5th generation nanobody-powered Chimeric Antigen Receptor (CAR), the self-replicating RNA (srRNA) components, and the TMP small-molecule mediated control system.

The self-replicative ability of srRNA components enables persistent and maintained expression of CAR on the cell surface, replenishing receptors lost during receptor-mediated phagocytosis. Continued and tightly controlled expression of the 5th generation nanobody-powered CAR facilitates improved recognition and targeting of hepatocellular carcinoma (HCC) tumor cells, by preventing depletion and stress during receptor-mediated phagocytosis. Simultaneously, the small-molecule mediated control system was implemented to ensure the self-replicating system could be regulated when necessary to avoid depletion of cellular resources. In combination, these elements of CAR_Ma act together to maintain and control the expression of CAR on macrophages.


Release Elements in the release modality consist of RNA molecules that need to be released to confer a function for CAR_Ma, typically when used in conjunction with proteins encoded in other modalities. This includes the guide RNA (gRNA) of CRISPR/Cas systems, as well as short-hairpin RNA (shRNA) acting as an alternative method for SIRPα silencing, based on the principle of RNA interference (RNAi). The release of gRNA enables the targeting of a Cas protein for permanent genomic knockout, while the release of shRNA enables entirely transient transcriptomic silencing of immune evasion (Karim et al., 2018) during the active period of the multimodal RNA. Rather than relying on a permanent method targeting the genome like CRISPR, the use of RNAi and shRNA will allow the system to be completely transient. This means that after its controlled shutdown, the cell will be able to return to a completely native state, enhancing long-term safety and stability for clinical uses. The strategic release of RNA molecules ensures CAR_Ma can effectively modulate gene expression, enabling both transient and permanent solutions to the problems of immune evasion.


Chimeric Antigen Receptor (CAR)

CAR-M vs CAR-T For the next generation CAR-M (CAR_MultiModal Advanced or CAR_Ma) developed by our HKU iGEM 2024 team, the chimeric antigen receptor is expressed on macrophages instead of the most common T-cells, effectively addressing significant challenges in cancer therapy.

One major advantage of this approach is that macrophages can penetrate the tumor microenvironment (TME) more efficiently, overcoming the physical barriers that generate chemokines suppressing T-cell functionality (Bule et al., 2021; Nagarsheth et al., 2017). Additionally, this strategy mitigates the adverse effects of M2-like tumor-associated macrophages (TAMs) and the chemokines that often promote tumor growth and lead to CAR-T immunosuppression (Li et al., 2019; Rodriguez-Garcia et al., 2021). Macrophages can adopt various functional states, enabling them to adapt to the TME and exert anti-tumor effects while reducing immune evasion (Mantovani et al., 2022). This flexibility enhances the persistence of CAR-M cells and improves coordination with other immune cells, resulting in a more robust immune response.

By harnessing the innate capabilities of macrophages for tumor migration and phagocytosis in synergy with the power of vector multi-modality, CAR_Ma not only enhances the potential for a robust anti-tumor response by addressing the limitations of traditional CAR-T systems, but also offers a more adaptable and resilient therapeutic strategy in the fight against cancer.


The Heart of CAR_Ma: Crafting the 5th Generation CAR The CAR receptor introduced this year is an advanced iteration of the one developed by the HKU iGEM team last year, marking the launch of the fifth-generation CAR.

5th Generation CAR: A Nanobody-Based Approach The driving force behind our novel fifth-generation CAR is our humanized nanobody, derived from Camelidae antibodies. This project presents an innovative single-body nanobody-based design that moves beyond the conventional use of single-chain fragment variable (scFv) antibodies. By harnessing the unique properties of nanobodies, this strategy offers key advantages, including increased stability, reduced immunogenicity, enhanced infiltration into the TME of solid tumors, and a more efficient humanization process. Moreover, it offers great benefits over scFvs, such as lower immunogenicity, decreased non-specific binding, and lower immunogenicity (Maali et al., 2023). These characteristics enable precise targeting of glypican-3 (GPC3), a cell-surface glycoprotein that is highly expressed in diverse cancers, including hepatocellular carcinoma – the second leading cause of cancer deaths globally in 2020 (Guo et al. 2020). GPC3, along with other nanobodies, serve as a reliable immunohistochemical marker given its high expression, where nanobodies – which are light chain-free modular domains that can easily be switched or multiplexed – enable the simultaneous detection of multiple antigens, thereby allowing for the targeted treatment of a wide array of cancers (Naidoo & Chuturgoon, 2021; Yang & Shah, 2020).


4th Generation CAR: TAM Repolarization A continuation from our 2023 project, low-dose of interferon-gamma (IFN-γ) is included to address the tumor associated macrophages (TAMs), the dominant immune cell in the TME and a major defense mechanism for solid tumors that provides nutrition to tumor ells and maintaining an environment that suppresses immune responses (Bied et al., 2023). IFN-γ, a pleiotropic cytokine known for its cytostatic, pro-apoptotic, antiproliferative, and stimulatory effects on M1 proinflammatory macrophages – acts as a potent polarization factor, steering transfected macrophages and TAMs toward an anti-cancer, pro-inflammatory state (Jorgovanovic et al., 2020).

To further improve phagocytosis, trogocytosis (the engulfment of large targets) and coordination within the broader immune system, the alternate costimulatory domains CD19, FcγR and MegF10 are used. Together, they promote antigen presentation to T-cells, promote phagocytosis and trogocytosis, improve macrophage migration, inflammatory response and survival (Morrissey et al., 2018).


The Specifics – Chimeric Antigen Receptor (CAR) Construct Design: Typical CARs are composed of four main components: a) an extracellular target antigen-binding domain, b) a hinge region that provides flexibility, c) a transmembrane domain that anchors the CAR in the cell membrane, and d) one or more intracellular signaling (co-stimulatory) domains that activate host cell responses upon antigen recognition (Sterner & Sterner, 2021).

The newly designed fifth-generation CAR features a sophisticated assembly of components: 2x K-Turns, a signal peptide, an anti-GPC3 nanobody, a CD8 hinge, a CD28 transmembrane domain, CD19, FcγR, MegF10, P2A sequences, IFN-gamma, additional P2A sequences, DD-L7Ae, and mCherry2 – a design that draws inspiration from previous research on CARs for phagocytosis (Morrissey et al., 2018).

The kink-turn (k-turn) is a prevalent structural motif in RNA that introduces a tight bend into the helical axis, serving as an element of RNA architectural stability and a crucial binding site for various proteins (Huang & Lilley, 2016; Schroeder et al., 2010). In cancer research and our CAR development, k-turns regulate gene expression to modulate CAR component levels, thereby enhancing the transcription and translation efficiency of therapeutic proteins and improving CAR functionality (Morrissey et al., 2018). In our case, we use our k-turn motifs as the foundation for achieving variable control within an in vivo expression system. By co-expressing DD-L7Ae, a fusion of a degradation domain to a k-turn binding protein (Mc Cafferty et al., 2021), we can use Trimethoprim (TMP) to stabilize a desired proportion of the DD-L7Ae. This will in turn trigger the strong conformational change of the k-turns in a desired proportion of the mRNA. The effect of this allows us to control the expression of the CAR receptor itself, allowing for variable immune response and control over the usage of intracellular resources, with the aim of extending the longevity of the CAR cells.

The CD8 hinge region, an extracellular structural component, enhances the functionality of CAR_Ma by promoting the longevity of CARs in the TME (Qin et al., 2017). It provides the necessary flexibility to navigate steric hindrance and optimizes the length and composition of the hinge region for greater access of the antigen-binding domain to the targeted epitope. This ultimately influences overall CAR signaling efficacy, expression, strength of activation outputs, and epitope recognition (Sterner & Sterner, 2021).

The CD28 and CD19 transmembrane domains, recognized for their high stability, anchor the CAR to the macrophage membrane and influence CAR expression levels, stability, signaling activities, and synapse formation (Sterner & Sterner, 2021). Furthermore, they dimerize with endogenous signaling molecules, enhancing overall functionality, and may also serve as biomarkers for specific types of cancer (Zhang et al., 2017).

In CAR models, Fc gamma receptors (FcγRs) facilitate anti-cancer and anti-tumor activities by mediating antibody-dependent cellular cytotoxicity, polarizing macrophages and other immune cells in TME toward a more proinflammatory state, and potentially serving as biomarkers for treatment efficacy (Caratelli et al., 2017).

Self Replicating RNA (srRNA)

Synthetic mRNA has become a major area of interest when designing gene editing systems with clinical applications, especially considering the development of methods for in vitro synthesis and in vivo delivery of RNA (Wagner & Mutschler, 2023). The non-integrative nature of RNA-based gene editing systems allows for transient, more regulatable changes, rather than permanent genomic ones associated with risks of tumorigenicity (Warren & Lin, 2019). However, RNA is transient due to constant enzymatic degradation - with foreign RNAs suffering further degradation through the innate immune response - causing RNA-based applications to require frequent re-transfections to maintain effective expression of the system (Wagner & Mutschler, 2023). With the complications brought about by frequent transfection, would it be possible to optimize RNA-based systems to increase their stability and longevity after transfection?
(Correction: the orange box at 1:02 should say “nsP1234 Replicase Complex” instead of “nsP123 Replicase Complex”)

Mechanisms of srRNA Self-replicating RNA (srRNA) offers a solution to the short life span of traditional mRNA-based therapies with the ability of in vivo replication, effectively mitigating the degradation experienced by RNA in the body (Wagner & Mutschler, 2023). As a result, introducing a srRNA module to an engineered RNA-based therapy - such as our CAR_Ma - allows the system's expression to increase to the medium or long-term (Wagner & Mutschler, 2023). For use in genetic engineering, srRNA is commonly sourced from the Venezuelan Equine Encephalitis (VEE) alphavirus due to being effective and safe (Lundstrom, 2016).

During a viral infection, the initial invasion of a cell is facilitated by viral structural proteins. The virus disassembles in the cytoplasm, releasing viral RNA encoding structural and nonstructural proteins (nsPs) (Kim & Diamond, 2022). While the structural proteins are responsible for viral pathogenicity and virulence to other cells, nsPs are responsible for effectively increasing the lifespan of the virus in the cell by replicating viral genetic material continuously (Kim & Diamond, 2022). Alphaviruses encode for four nsPs: nsP1, nsP2, nsP3, and nsP4; which form a replicase complex that allows for continuous replication of viral RNA, thus increasing its expression in the cell (Kim & Diamond, 2022). Incorporating these nsPs into an RNA construct effectively grants it this same ability.


Engineering srRNA CAR_Ma is designed to utilize an evolved srRNA vector, with a VEE replicase complex modified to enable significantly longer expression of our RNA-based immunotherapy, as well as a small-molecule modulated kill-switch control system to enhance safety by providing greater control to a self-replicative cancer immunotherapy. The basis of CAR_Ma’s srRNA production is the T7-VEE-GFP plasmid, containing the necessary nsPs for in vivo replication (Yoshioka et al., 2013). Using NEBuilder, the vector has been modified and designed to act as the foundation for CAR_Ma’s immunotherapy.

To enhance the inherent ability of the nsP123 replicase complex, an 18bp insertion mutation was introduced to allow for more efficient and long-term replication (Li et al., 2019). Thus, its addition to a sRNA vector allows for increased persistence in the cell, which provides stable and robust expression for a significantly longer timeframe.

To enhance the inherent ability of the nsP123 replicase complex, an 18bp insertion mutation was introduced to allow for more efficient and long-term replication (Li et al., 2019). Thus, its addition to a sRNA vector allows for increased persistence in the cell, which provides stable and robust expression for a significantly longer timeframe.

Furthermore, two additional nsPs: nsP14 and nsP10 were introduced to the srRNA construct to facilitate the maintenance of the replicating RNA. While srRNA is capable of self-replication, the duration of its efficacy is limited by the in vivo replicative fidelity of VEE’s nsP1, nsP2, nsP3, and nsP4, and the lack of a proofreading domain to counteract the errors of the replicase complex (Kautz et al., 2018). To overcome this limitation and increase the potential persistence of the srRNA construct in the cell, nsP14 and nsP10 were added to the design.

Originating from Coronaviruses, nsP14 is a viral proofreading exonuclease domain and nsP10 is its associated cofactor, which together decreases the mutational error rate of RNA-dependent RNA polymerases, allowing for enhanced maintenance of the RNA genome (Rona et al., 2021). Adding these nsPs to the VEE vector lends the replicase the same decreased mutation rate, increasing fidelity and effectively improving persistence, stability, and strength of expression.

CAR_Ma was designed to be an RNA-based immunotherapy to leverage the unique advantages of RNA. RNA molecules are genomically non-integrative, which removes concerns of genetic permanence and insertional mutagenesis attributed to plasmid and viral platforms, and mitigates the problem of potential malignancies observed in recent BMCA-mediated CAR-T therapies (U.S. Food and Drug Administration, 2023). With modifications such as the introduction of nsPs to produce a higher fidelity, persistent, self-amplifying RNA molecule, our design improves upon the inherent benefits of RNA, by addressing the problems of persistence and fidelity observed with other RNA therapies.

TMP Control System

While (sr)RNA has emerged at the forefront of gene therapy, it has generally lacked the small-molecule expression control systems that DNA-based systems have (Wagner et al., 2018). Given CAR_Ma’s in vivo human therapeutic goal, a reliable and robust expression control system incorporated in the RNA is absolutely essential. To address this, a variable expression control system based on trimethoprim (TMP), a common human antibiotic, is added in the mmRNA to allow for rapid expression shutdown of self-replicating functions and various genes of interest like the CAR construct and shRNA functions (Wagner et al., 2018). The system can be expanded into a variable switch, allowing adjustments in expression and therapeutic intensity even after administration.

The system is based on two main components, the first being DD-L7Ae, a protein complex encoded in the mmRNA. DDd is a destabilization domain protein taken from Escherichia coli, which will be automatically targeted for degradation by the proteasome if left untreated. Yet when stabilized by binding to TMP, DDd can disrupt the function of the protein it is bound to. This is combined with L7Ae, a protein from Archaeoglobus fulgidus that binds to RNA kink-turns (k-turns), a common RNA motif and the second component of the control system. When L7Ae binds to a K-turn, it initiates a sharp conformational change that in the mmRNA, blocks the expression of downstream open reading frames (Turner et al., 2005; Wagner et al., 2018). By adding 2x k-turns into our mmRNA, an off-switch for CAR_Ma expression is achieved.

In our mmRNA, 2x k-turns are placed in 3 positions: before nsP1 (before all replicons), between nsP3 and nsP4 (since srRNA replicons can either work as nsP1234 or nsP123 and nsP4), and before our genes of interest. Due to the complexity and unknown mechanisms of the replicons, the multiple 2x k-turns inserted ensure that we can shut down both the self-replicating functions and genes of interest, providing reliable and uncompromising control over CAR_Ma expression. The control system is assembled along with other parts of the plasmid with NEBuilder HiFi DNA Assembly.


Furthermore, akin to an antibiotic kill curve experiment, a TMP expression curve can be conducted to map how TMP concentration affects CAR_Ma expression by tracking fluorescence signals present after THP-1 transfection. This turns the off-switch into a variable switch, capable of easy adjustment and personalisation of the therapeutic at the point-of-care. This provides the flexibility that healthcare professionals require to better respond to different patient needs. Additionally, the widespread oral administration of TMP and its low cost makes it an ideal small molecule for the control system, ensuring the accessibility of this therapeutic (Kemnic, 2022).

SIRPα Knockout

CAR-T cell therapy has revolutionized the cancer immunotherapy field, especially in treating some hematological malignancies (Chen, 2024). However, in the case of solid tumors, CAR-T cells face crucial challenges. These CAR-T cells grapple to penetrate the dense and immunosuppressive tumor microenvironment (TME) and their survivability in this hostile environment is very limited (Chen, 2024). This brings up an important question: Could there be a more effective immune cell for targeting solid tumors?

The Promise of CAR-Macrophages Unlike T-cells, macrophages are natural inhabitants of the TME. They can navigate and survive the harsh environment of TME more efficiently than other immune cells. According to Brempelis et al. (2020), macrophages have been proven to survive longer in the TME than any other cells. Their engineered GFP-expressing macrophages were found to persist in the tumor cells for 22 days post-injection, which is 39-fold greater than the initial levels (Brempelis et al., 2020). Our 5th Gen CAR construct not only targets specific solid tumor cells but also has the ability to repolarize tumor-associated macrophages (TAMs) back to the M1 stage, which has been shown to exert significant anti-tumor and inflammatory effects. Additionally, TAMs are found in nearly all solid tumors providing numerous therapeutic targets and enhancing the potential of Car_Ma.

The Importance of SIRPα-CD47 Pathway When CAR-Macrophages enter the TME, they must bind to tumor cells through antigen recognition before exerting its effects. However, this process is often overtaken by tumor cells, leading to reduced phagocytosis and evasion from the immune system. One prominent pathway the tumor cells use is the SIRPα-CD47 signaling pathway. A key mechanism it uses is the overexpression of CD47, a protein on tumor cells that produces a “Don’t Eat Me” signal when interacting with the SIRPα protein on macrophages (Jiang et al., 2024). This triggers the phosphorylation of Immunoreceptor Tyrosine-Based Inhibitory Motifs (ITIMs) within SIRPα’s cytoplasmic domain (Qu et al., 2022). This in turn sends inhibitory signals that reduce phagocytosis and tumor-antigen presentation (Qu et al., 2022). Therefore, targeting the SIRPα-CD47 pathway is a promising strategy to induce the elimination of tumor cells.


CRISPR-mediated SIRPα Knockout In order to address the inhibition by the SIRPα-CD47 pathway, CRISPR genome editing can be used to knock out the SIRPα domain within the macrophages, namely with the Cas9 and Cas12a mechanisms. Through preventing the SIRPα-CD47 interaction, the phagocytic abilities of the macrophages are able to increase by 4-fold (Ray et al., 2018).

The Cas9 mechanism involves the formation of a Cas9 complex, consisting of the Cas9 protein and a guide RNA (gRNA). The gRNA is composed of the target-specific crRNA, whose sequence is complementary to the target domain, and the scaffolding tracrRNA, which allows for binding with the Cas9 endonuclease. When formed, the gRNA is used to direct the Cas9 enzyme to a specific gene of interest, where the Cas9 enzyme is then able to induce a double-strand break at the specific site, being the SIRPα domain (Kim et al., 2014). With the use of cell’s natural DNA repair pathways, such as non-homologous end joining, the excised DNA is repaired, however, small insertions and deletions (indels) are introduced at the junction. As such, the error-prone DNA sequence results in disrupted protein expression (Ishibashi et al., 2020). The final product is a SIRPα-deficient macrophage, able to interrupt the SIRPα-CD47 pathway and thus preventing the immune evasion of tumor cells, while allowing for phagocytosis and an overall increase in the efficacy of CAR macrophages.

The Cas12a mechanism works similarly to Cas9, with the difference of four separate guide RNAs composed of only the crRNA, thus being able to knock out up to 4 targets, rather than just 1 target with Cas9 (Esmaeili Anvar et al., 2024). In our design, aside from the SIRPα domain, three other targets have been chosen to increase the efficacy of our CAR macrophage. The PD1-PDL1 pathway, which acts as a T-cell activation checkpoint, STAB1, which involves inhibition of phagocytosis, as well as the VTCN1 which plays a role in T-cell inhibition. Targeting these domains allows our macrophage to increase phagocytic ability and boost T-cell activity, overall targeting immune evasion with an increased efficacy.


Limitations of CRISPR While CRISPR/Cas systems like Cas9 and Cas12a offer a powerful means to achieve permanent knockout of genes at the genomic level, they come with numerous limitations including potential off-target effects, and the prolonged expression of Cas9 or Cas12a enzymes. These off-targets occur when the CRISPR system unintentionally edits DNA sequences that are similar but not identical to the target sequence, which might lead to unforeseen genetic modifications causing undesirable phenotypic changes or other mutations (Yang et al., 2021). Prolonged expression of the Cas9 enzyme can increase the chances of having off-target effects and our body may also recognize the Cas9 protein as a foreign property, leading to an immune response that can have adverse effects. Our immune system recognizes the Cas9 protein as a non-self entity, triggering the production of antibodies against it (Wienert et al., 2018). Moreover, the gRNAs used in conjugation with Cas9 can activate innate immune pathways such as the RIG-I pathway producing type-1 interferons and inflammatory cytokines (Xu, 2024). The type-1 IFNs can interfere with and affect the proper functioning of the gRNA-Cas9 complex, leading to a much lower efficiency in terms of its ability to knock out genes (Wienert et al., 2018). Initially, CAR_Ma was built with CRISPR in mind, but these design concerns made us rethink and redesign CAR_Ma.

Transient Gene Modulation

Short Hairpin RNA (shRNA) To prevent the immune evasion strategy mediated by the SIRPα/CD47 pathway, we propose using shRNA to knock down the expression of SIRPα, by which the CD47 on tumor cells cannot bind to SIRPα and therefore exposing itself to our immune system. Unlike CRISPR, which induces permanent genetic changes, shRNA-mediated knockdown reduces gene expression without altering the DNA sequence. It knocks down genes on a transcriptomic level by following the working principle of RNA interference (RNAi) rather than CRISPR’s genomic-level gene knockout.

Short Hairpin RNA (shRNA) is indeed a single-stranded RNA that folds back on itself forming a stem-loop structure. This loop is what gives the shRNA the appearance of a double-stranded RNA (dsRNA). The non-complementary region of the RNA sequence forms the loop that connects the two ends of the stem, allowing it to fold back on itself.

Once the shRNA is introduced into the host cell, the shRNA is transcribed by its promoter. The transcribed shRNA forms a stem-loop structure which is then transported from the nucleus to the cytoplasm by the action of Exportin 5. Then it is processed by an endoribonuclease known as “DICER” which cleaves the loop and converts the shRNA into short dsRNAs or small interfering RNA (siRNA).

Upon getting processed by DICER, the dsRNA is then loaded into the RNA-induced silencing complex (RISC), where the sense (passenger) strand gets degraded, leaving the antisense (guide) strand to guide the RISC complex towards the complementary target mRNA sequence (O’Keefe, 2013). A crucial component of RISC that plays a central role in the gene silencing mechanism is the presence of Argonaute 2 (Ago2) protein in RISC. Ago2 possesses endonuclease activity which allows it to cleave the target mRNA at the target site when the guide strand perfectly complements the target strand (Ruda et al., 2014). This cleavage results in the degradation of the target mRNA, leading to effective silencing of the target gene. According to GeneCards, the cleaved mRNA fragments are further degraded by exonucleases present in our cells (GeneCards, 2024).

In our case, the guide strand of the designed shRNA binds to the mRNA of SIRPα leading to its cleavage and degradation by RISC. This cleavage of SIRPα results in a significant reduction in the expression levels of SIRPα which enhances the phagocytic activity

Designing shRNA for CAR_Ma The design of our shRNA sequence is critical to ensure specific and efficient gene knockdown. The process begins with selecting a target sequence within the mRNA of the gene of interest. This target sequence is typically 19-29 nucleotides long and is extremely specific to the target gene to minimize off-target effects (Taxman et al., 2010). The shRNA sequence is then designed to form a stem-loop structure, which includes the sense and anti-sense strands of the target sequence (O’Keefe, 2022). In our construct, the 26s subgenomic promoter has been incorporated as the promoter which will drive the expression of the shRNA. This allows us to continually express our shRNA like the existing viral and plasmid based approaches while maintaining controllability and safety with a controllable and transient RNA based vector.


Engineering shRNA into multi-modal RNA using Ribozymes Self-cleaving ribozymes are RNA molecules that can catalyze their own cleavage at specific sites. By flanking the self-cleaving ribozymes around the shRNA sequence, we can facilitate a precise and efficient expression of our shRNA construct to mediate effective gene silencing. As we know, the presence of post-transcriptional modifications like the 5’ cap and poly(A) tail are critical features of an mRNA to enhance its stability and promote efficient translation. However, according to McIntyre and Fanning (2006), in a shRNA, the presence of these PTM features can hinder the effectiveness of shRNA. The 5’ cap can interfere with the recognition and processing of the shRNA by the RNAi machinery, as it may prevent the loading of the dsRNA into the RISC complex (McIntyre and Fanning, 2006). The presence of these PTMs increases the overall length of the stem regions on the shRNA. These can result in misfolding, or processing failure by DICER, leading to lowered silencing efficiencies.

To address these issues, self-cleaving ribozymes can be flanked around the ends of our shRNA construct. In our case, the ribozymes used were hammerhead (HH) ribozymes and hepatitis delta virus (HDV) ribozymes which catalyze the precise cleavage of the RNA, removing the 5’ cap and poly(A) tail respectively, and preserving the optimal shRNA sequence.

Nanoparticle Delivery

Nanostructured Lipid Carrier In the rapidly evolving field of RNA therapeutics, srRNA has emerged as a powerful tool for the development of various therapeutic cancer vaccines. Clinical trials have demonstrated that srRNA is able to encode tumor-associated antigens such as human epidermal growth factor receptor 2 (HER2), and carcinoembryonic antigen (CEA) (Davis et al, 2021).

In terms of non-viral carriers, lipid nanoparticles were selected as the most effective delivery method due to their superior biocompatibility, reduced cytotoxicity, and enhanced cellular uptake and internalization. Lipid nanoparticles are the most effective due to their ability to prevent endosomal escape– the unique composition of LNPs, specifically the incorporation of pH-sensitive lipids, allows a disruption of endosomal membranes, leading to successful cytoplasmic delivery of our srRNA (Zhao, 2014).

While srRNA has been delivered through a variety of different methods, such as viral vectors, the innovative use of lipid nanoparticles (LNPs) remains a large piece of uncharted territory. Following COVID-19, the remarkable success of mRNA vaccines such as Pfizer and Moderna, exemplifies the excellent biocompatibility of lipid nanoparticle delivery.

However, packaging srRNA is inherently more complex than its mRNA counterparts, due to its unique structural characteristics. To ensure that the nanoparticle we are using is the most biocompatible, while maintaining environmental biodegradability and human safety, we examined a myriad of different lipid-based carriers. Amongst them, we decided that the Nanostructured Lipid Carrier (NLC) fits our project the best. The imperfect crystal model of the NLC, and its amorphous quality allows a higher drug loading capacity, controlled release, and its own edge of higher stability during storage. NLC can be stored at room temperature for at least 8 months and in 4°C for up to 21 months (Gerhardt et al., 2022). Allowing us to simplify logistics and storage making this one of our ideal candidates for CAR_Ma enabling our simplified administration via intravenous injection simplifying and improving accessibility of immunotherapies.

At the current moment, NLCs are currently being used in a multitude of medical trials, such as UV protection cosmetics, dermal drug delivery, and topical nanomedicines (Ferreira et al., 2021). NLC and rvRNAs complexes are successfully designed for Zika vaccines (Erasmus et al., 2018). Similar ingredients were used, of which Tween 80 is replaced with DMG-PEG. They perform similar functions in our case, for the sake of aggregation prevention.

Though NLCs are far more capable of delivery, storage and logistics enabling CAR_Ma simple administration and accessibility. NLCs like other nanoparticle delivery systems are relatively indiscriminate in their transformation targets. Though recipe optimization can better target proportions of the NLCs administered to certain tissue groups, it will be inevitable that other tissues will be transformed. In our case, that could risk complications with incorrect tissues expressing the CAR_Ma system. To optimize this further we functionalized our nanoparticles with anti-CD14 antibodies (Marques et al., 2023) or anti-CD14 nanobodies, to ensure with maximal efficiency that we are able to specifically transfect monocytes and macrophages. Allowing us to maximize efficiency while minimizing unwanted off-target effects, and completing the CAR_Ma package.


Zein Nanoparticle In addition to our work with Nanostructured Lipid Carriers (NLC), our team has innovatively developed zein nanoparticles for the targeted delivery of srRNA into macrophages. Zein, being edible, biocompatible, and biodegradable, stands out as an ideal material for delivering drug and genetic material (Zhang, 2021).

The production process for zein nanoparticles involves transitioning zein from an organic phase (alcohol) to aqueous phase or via altering the pH. This allows aggregation, and rearranges the peptide chains in zein (Zhang, 2021; Dong, 2021). Zein has shown promise in various applications, such as the sustained release of plasmid DNA (pDNA) (Regier et al, 2012). In addition, zein nanofibers have been developed to explore their potential to deliver siRNA for gene silencing in skin fibroblasts (Karthikeyan et al, 2015).

Despite these advancements, aggregation of zein nanoparticles poses a significant challenge, which will limit the viability and shelf-life of zein nanoparticles. To tackle this issue, we added Tween 80, a surfactant that promotes electrostatic repulsion amongst zein nanoparticles. This addition effectively prevents aggregation, and enhances the stability for our formulations.

By leveraging the unique properties of zein, and addressing any potential challenges, we are paving the way for effective, biodegradable, and biocompatible delivery systems for srRNA delivery for CAR_Ma and beyond.

CAR_Ma’s Engineering Cycles

In many ways, the development of CAR_Ma has been very iterative, cycling through several rounds of design, building, testing and learning. Since CAR_Ma is a ground-up rearchitecting of CAR_Ma from HKU iGEM 2023, there are major engineering cycles and smaller iterative ones that have occurred.


Rearchitecting 2023’s CAR_Ma Proof-of-Concept While the work conducted by the HKU iGEM team 2023 was rudimentary, tt allowed us to gain incredible insight into the meticulous design and short-commings of our previous work.


CD3ζ-based CAR From the beginning an iterative immunotherapy was the original goal, in doing so the designs that we used borrowed heavily from one of our advisor’s work with CAR-T, a receptor with a CD3ζ costimulatory domain typically seen with CAR-T therapies. The benefit at the time is that we were quickly able to work with this due to the part already existing with us, unfortunately it was found that with macrophages CD3ζ was a costimulatory domain that did not greatly improve phagocytosis. Our testing, review of literature by Morrissey et al. (2018) and feedback from our PI and others from 2023 iGEM Giant Jamboree. Learning from this undersight, we completely rearchitected our CAR based on the work by Morrissey et al. (2018), integrating CD19-FcγR Tandem and MegF10 costimulatory domains into our CAR enhancing phagocytosis, trogocytosis and antigen presentation resulting in a far more effective system.


Overwhelming cellular resources, IFN-γ overexpression and TMP shutdown Besides the core of our CAR being based on costimulatory domains we integrated IFN-γ as an armouring factor to reinforce our CAR_Ma cells and repolarize tumor associated macrophages (TAMs) disrupting the stability of the tumor microenvironment. Another system we integrated was a small-molecule based shutdown system based on trimethoprim (TMP) to terminate the expression upon the completion of the treatment regime. In our initial experiments from 2023, there were difficulties with cloning our control system into our plasmids, delaying their completion. We also noted that CAR_Ma cells demonstrated an impressive effect in initial tests but exhibited shorter than expected lifespans. After further investigation, and discussion with our PIs and their colleagues we were able to determine that the CAR expression with srRNA if left uncontrolled would quickly runaway and induce excessive stress on cellular machinery and resources. One PI also noted that IFN-γ in small doses resulted in the armoring of CAR-M and repolarization TAMs, however the sustained expression via our self-replicating RNA (srRNA) vector would lead to severe downstream immune responses which could risk patient health.

As it turns out, the solution to these problems all could be solved by our small-molecule shutdown system. After these learnings, the shutdown system was broken-down and rearchitected completely into a system that allows for variable control. In a recent paper on the in vivo validation of the TMP-based shutdown system on mice, Cafferty et al. (2021) found that there was a distinct curve relating the dose of TMP given to the strength of expression. After amending our cloning scheme to streamline the integration of the shutdown system’s functional units we designed our shutdown system to regulate both the expression of CAR but also the replication of the srRNA itself. Our shutdown system initially designed for termination in case of emergencies and completion of the treatment regime was now evolved to allow for the control of the CAR expression and IFN-γ release, by simply pairing each dose of CAR_Ma with a corresponding dosage of TMP. With this, our new CAR_Ma cells are able to retain overall effectiveness, but gain increased longevity and reduce immunological side-effects with tighter control.


In vivo vs Ex vivo: Delivery, CRISPR and Gene Modulation When 2023’s proof-of-concept moved away from Sendai viral delivery to self-replicating RNA based vectors, the auxiliary engineering aside from CAR expression was not fully adapted to the new system. The importance of mitigating immune evasion via pathways such as SIRPα/CD47 have been well established in enhancing phagocytosis of tumor cells (Ray et al., 2018), but since common methods currently of delivering CRISPR operations for therapeutics were delivered ex vivo such as those via ribonucleoprotein (RNP) complexes, CAR_Ma cells were actually engineered with a two-step process that required the knockout of SIRPα first by the electroporation of an RNP, then the transformation with srRNA for CAR expression also via electroporation. Though effective for use in the laboratory, this negated a significant advantage of using RNA as a vector as the cell must still be engineered ex vivo. This was acceptable for a proof-of-concept, but to not only be a viable but superior therapeutic, this would be insufficient.

Ultimately, the Holy Grail of CAR_Ma would be in vivo expression of CAR but also the in vivo gene editing of SIRPα via CRISPR. When we rearchitected CAR_Ma, we went even a step further by integrating both these functions on to the same RNA molecule dramatically simplifying production by only requiring one template but also vastly simplifying packaging and screening by allowing for a homogenous RNA payload. The advancement we required was born out of engineering our self-replicating RNA and using the functional blocks to allow for transient expression of Cas proteins, sustained expression of CAR and the release of small non-coding RNA such as gRNA on a unified multi-modal RNA vector. Once this breakthrough was achieved, we needed a vector to preserve the goal of in vivo delivery. As the world recovered from the COVID-19 pandemic, inspiration struck. The mRNA vaccines used for COVID-19 used lipid nanoparticles for in vivo delivery. When discussing with our PIs and collaborators in microfluidics, we engineered two cutting edge nanoparticles to package our system: A nanostructured lipid carrier and a zein-based protein nanoparticle. Allowing us to achieve the goal of an in vivo immunotherapy, while setting the stage for a large array of gene therapies and immunotherapies by combining CRISPR, our multi-modal RNA and nanoparticle delivery systems.



Minor Iterations enhancing CAR_Ma Aside from breakthroughs from the rearchitecting of 2023’s CAR_Ma, there were several crucial iterations done through minor engineering cycles during the experimentation done this year.


CRISPR vs RNAi When originally conceived, CAR_Ma and the 2023 proof-of-concept were built to use CRISPR. Since knockout studies that facilitated the knowledge of tumor evasion, knockout improving phagocytosis were all using CRISPR and even modern gene therapeutics based upon CRISPR it seemed as the obvious choice. Especially since the tool sets to use CRISPR were more comprehensive and robust. However, a tenet of CAR_Ma was transience. With the system itself having the ability to be modulated, shutoff and even removed. The persistence of the CRISPR knockout long-term broke this tenet and further posed significant risks downstream. Mainly, since CRISPR knockouts are permanent the activity of the knocked out gene in our case would remain off even after the CAR_Ma system expires or is shut down via TMP, this could lead to severe risks and complications where these cells no longer ignore friendly cells leading to an autoimmune risk. Further, by acting at a genomic level CRISPR knockout would only have a complete shutdown of SIRPα, lacking any control or variability.

Countering this, we moved away from genomic gene modulation towards transcriptomic regulation instead. By leveraging RNA interference, the SIRPα gene would remain untouched producing SIRPα RNA transcripts. By transcribing shRNA from our multi-modal RNA vector, we are able to trigger RNA interference causing the specific degradation of targeted SIRPα RNA transcripts, effectively silencing the gene expression transiently. Furthermore, by leveraging our TMP-modulated variable switch, we are able to adjust the expression of the shRNA effectively controlling the strength of the SIRPα silencing granting the system even more flexibility.

Unfortunately, the shRNA sequences are unable to directly be added onto our RNA release system. When initially designed, our RNA release system had a severe deficiency in that the RNA-dependent RNA polymerase from the Venezuelan equine encephalitis (VEE) virus replicase that powers the CAR_Ma inherently has capping and tailing capabilities. With our Cas9-mediated knockout, the gRNA was expressed enough to allow for the knockout to occur. But when moving to our Cas12a multiplexed gRNA and especially with shRNA, we noticed that the knockouts were unsuccessful. Upon further investigation we discovered the capping and tailing capability of the VEE polymerase. We quickly adapted, designing self-cleaving ribozymes to the 5’ and 3’ ends of the shRNA. By adding a hammerhead ribozyme to the 5’ and a HDV ribozyme to the 3’ end, we are able to perfectly preserve the sequence of the gRNA or shRNA during the subgenomic release of the non-coding RNA, furthermore the ribozyme sequences will not cleave as effectively on the parent multi-modal RNA due to the length and structural characteristics of the longer RNA molecule inhibiting effective cleavage. The resulting system allows gRNA or shRNA to be replicated, and its subgenomic release sustained while preserving the exact sequence needed for maximal efficiency for CRISPR or RNAi operations.


PCR, Cloning and Transformation Through designing the complexities and intricacies of CAR_Ma, there were a lot of areas for errors to occur. The complexities required a cloning system with high-specificity and the ability to assemble plasmids from over 8 fragments. Eventually we landed on cloning with NEBuilder as it was a system that was both accessible to us and matched our needs perfectly. However using NEBuilder would require designing over 30 different gene fragments to be assembled in various ways along with over 100 different primers.

Cloning and assembly of fragments was laborious and required frequent use of PCR, oftentimes requiring optimization as our PCRs would be run in gels, sometimes resulting in mispriming, smearing, or in the worst case lacking amplification. By optimizing PCR recipes and run operations: Adjusting amount of template, primers, including additives such as DMSO. Running touchdown, using hot-start polymerases, running longer extensions. We were able to optimize a majority of our PCR processes to create an optimal amount of fragment for cloning. NEBuilder itself also often required optimization, adjusting incubation times or fragments to be in equivalent molarities.

The final step in our cloning would be transformation into cloning strains of E. coli. We found that this process has been the most troublesome throughout all of our experiments. Initially we worked with E. coli DH10b for its capability to work with larger plasmids, however we found that they would never be any transformants. After optimizing upstream processes and verifying that our plasmids would successfully assemble, we narrowed down the process to the transformation. Eventually, we were able to diagnose that the large plasmid would be too heavy on cellular resources, thus requiring an extended recovery time of 2 hours to successfully express the relevant antibiotic resistance gene for selection, thus we were able to finally get transformants. However, after miniprepping our transformants and sending our plasmids for sequencing, we noticed significant portions of our plasmid were missing. Including CAR or shRNA, essential parts of CAR_Ma. We were able to soon diagnose that the repeated sequences, especially with shRNA were toxic to the E. coli resulting in the cell excising the DNA. To remedy this, we moved away from E. coli DH10b and used E. coli Stable, which provided us with the ability to clone toxic and repeated sequences. Eventually, we were able to optimize our overall cloning and transformation process to enable us to get the required plasmids we needed to develop and characterize CAR_Ma.

CAR_Ma Parts List

Basic Parts Kozak sequence - BBa_K4344028
Signal peptide BBa_K2549044
CD8 hinge: BBa_K5017003
CD28 transmembrane domain BBa_K4803006
CD19 BBa_K5062003
FcyR BBa_K5062004
MegF10 Costim - BBa_K5062005
Anti-gpc3 nanbody BBa_K5062000
Anti-GPC3 scFv BBa_K5062001
DDd BBa_K4696007
L7Ae - BBa_K3113009
2x kTurn - BBa_K4696005
P2A - BBa_K1442039
Ifn-gamma BBa_K4696020
mCherry XL BBa_K4241792
5' Hammerhead Ribozyme BBa_K5062009
3' Hepatitis D Virus Ribozyme BBa_K5062010
SIRPa shRNA BBa_K5062011
PD1 shRNA BBa_K5062012
SIRPa flank sequence BBa_K5062013
PD1 flank sequence BBa_K5062014
26s BBa_K4696004
-26s BBa_K5062033
nsP10 BBa_K5062031
nsP14 BBa_K5062029
Cas9 gRNA BBa_K5062020
Cas9 gRNA flank sequence BBa_K5062032
IRES BBa_K5062038
Enhanced IRES BBa_K5062037
nsP1 - BBa_K4696000
nsP2 - BBa_K4696001
nsP3 - BBa_K4696002
nsP4 - BBa_K4696003
Cas12a gRNA flank sequence - BBa_K5062034
DR1 - BBa_K5062023
DR2 - BBa_K5062025
DR3 - BBa_K5062027
SIRPa gRNA - BBa_K5062022
PD-1 gRNA - Part:BBa_K5062024
VTCN1 gRNA - Part:BBa_K5062026
STAB1 gRNA - Part:BBa_K5062028
AsCpf1 (Cas12a) - BBa_K5062021
Puromycin resistance - BBa_K4696009


Composite Parts DD-L7Ae - BBa_K5062002
CD19/ FcyR Tandem - BBa_K5062006
5th Generation NanoCAR - BBa_K5062007
5th Generation CAR - BBa_K5062008
SIRPa Ribo-shRNA - BBa_K5062016
PD-1 Ribo-shRNA - BBa_K5062017
SIRPa/PD1 Multiplex shRNA - BBa_K5062018
Cas9-GFP - BBa_K5062030
Cas9 gRNA Ribo - BBa_K5062039
Multiplex Cas12a gRNA - BBa_K5062045
nsP123 - BBa_K5062046


Multi-modal RNA CAR_Ma - BBa_K5062047
CAR_Ma shRNA Multiplex - BBa_K5062048
CAR_Ma Cas9 - BBa_K5062049
CAR_Ma Cas12a Multiplex - BBa_K5062050

References

Baghban, R., Roshangar, L., Jahanban-Esfahlan, R., Seidi, K., Ebrahimi-Kalan, A., Jaymand, M., Kolahian, S., Javaheri, T., & Zare, P. (2020). Tumor microenvironment complexity and therapeutic implications at a glance. Cell Communication and Signaling, 18(1). https://doi.org/10.1186/s12964-020-0530-4


Bied, M., Ho, W. W., Ginhoux, F., & Blériot, C. (2023). Roles of macrophages in tumor development: A Spatiotemporal perspective. Cellular & Molecular Immunology, 20(9), 983–992. https://doi.org/10.1038/s41423-023-01061-6


Brempelis, K. J., Cowan, C. M., Kreuser, S. A., Labadie, K. P., Prieskorn, B. M., Lieberman, N. A., Ene, C. I., Moyes, K. W., Chinn, H., DeGolier, K. R., Matsumoto, L. R., Daniel, S. K., Yokoyama, J. K., Davis, A. D., Hoglund, V. J., Smythe, K. S., Balcaitis, S. D., Jensen, M. C., Ellenbogen, R. G., … Crane, C. A. (2020). Genetically engineered macrophages persist in solid tumors and locally deliver therapeutic proteins to activate immune responses. Journal for ImmunoTherapy of Cancer, 8(2). https://doi.org/10.1136/jitc-2020-001356


Bule, P., Aguiar, S. I., Aires-Da-Silva, F., & Dias, J. N. (2021). Chemokine-directed tumor microenvironment modulation in cancer immunotherapy. International Journal of Molecular Sciences, 22(18), 9804. https://doi.org/10.3390/ijms22189804


Cafferty, S. M., De Temmerman, J., Kitada, T., Becraft, J. R., Weiss, R., Irvine, D. J., Devreese, M., De Baere, S., Combes, F., & Sanders, N. N. (2021). In vivo validation of a reversible small Molecule-Based switch for synthetic Self-Amplifying mRNA regulation. Molecular Therapy, 29(3), 1164–1173. https://doi.org/10.1016/j.ymthe.2020.11.010


Caratelli, S., Sconocchia, T., Arriga, R., Coppola, A., Lanzilli, G., Lauro, D., Venditti, A., Del Principe, M. I., Buccisano, F., Maurillo, L., Ferrone, S., & Sconocchia, G. (2017). FCγ chimeric receptor-engineered T cells: Methodology, advantages, limitations, and clinical relevance. Frontiers in Immunology, 8. https://doi.org/10.3389/fimmu.2017.00457


Chen, K., Liu, M., Wang, J., & Fang, S. (2024). Car-macrophage versus CAR-t for solid tumors: The race between a rising star and a superstar. Biomolecules and Biomedicine, 24(3), 465–476. https://doi.org/10.17305/bb.2023.9675


Davies, N., Hovdal, D., Edmunds, N., Nordberg, P., Dahlén, A., Dabkowska, A., Arteta, M. Y., Radulescu, A., Kjellman, T., Höijer, A., Seeliger, F., Holmedal, E., Andihn, E., Bergenhem, N., Sandinge, A.-S., Johansson, C., Hultin, L., Johansson, M., Lindqvist, J., … Andersson, S. (2021). Functionalized lipid nanoparticles for subcutaneous administration of mrna to achieve systemic exposures of a therapeutic protein. Molecular Therapy - Nucleic Acids, 24, 369–384. https://doi.org/10.1016/j.omtn.2021.03.008


de Visser, K. E., & Joyce, J. A. (2023). The evolving tumor microenvironment: From cancer initiation to metastatic outgrowth. Cancer Cell, 41(3), 374–403. https://doi.org/10.1016/j.ccell.2023.02.016


Dong, Fangyuan, et al. “Doxorubicin-Loaded Biodegradable Self-Assembly Zein Nanoparticle and Its Anti-Cancer Effect: Preparation, in Vitro Evaluation, and Cellular Uptake.” Colloids and Surfaces B: Biointerfaces, vol. 140, Apr. 2016, pp. 324–331, https://doi.org/10.1016/j.colsurfb.2015.12.048


Light, D. W., & Lexchin, J. (2021). The costs of coronavirus vaccines and their pricing. Journal of the Royal Society of Medicine, 114(11), 502–504. https://doi.org/10.1177/01410768211053006


Maeda, H., & Khatami, M. (2018). Analyses of repeated failures in cancer therapy for solid tumors: poor tumor‐selective drug delivery, low therapeutic efficacy and unsustainable costs. Clinical and Translational Medicine, 7(1). https://doi.org/10.1186/s40169-018-0185-6


Erasmus, J. H., Khandhar, A. P., Guderian, J., Granger, B., Archer, J., Archer, M., Gage, E., Fuerte-Stone, J., Larson, E., Lin, S., Kramer, R., Coler, R. N., Fox, C. B., Stinchcomb, D. T., Reed, S. G., & Van Hoeven, N. (2018). A nanostructured lipid carrier for delivery of a replicating viral RNA provides single, low-dose protection against zika. Molecular Therapy, 26(10), 2507–2522. https://doi.org/10.1016/j.ymthe.2018.07.010


Esmaeili Anvar, N., Lin, C., Ma, X., Wilson, L. L., Steger, R., Sangree, A. K., Colic, M., Wang, S. H., Doench, J. G., & Hart, T. (2024). Efficient gene knockout and genetic interaction screening using the in4mer CRISPR/CAS12A multiplex knockout platform. Nature Communications, 15(1). https://doi.org/10.1038/s41467-024-47795-3


Ferreira, K. C., Valle, A. B., Paes, C. Q., Tavares, G. D., & Pittella, F. (2021). Nanostructured lipid carriers for the formulation of topical anti-inflammatory nanomedicines based on natural substances. Pharmaceutics, 13(9), 1454. https://doi.org/10.3390/pharmaceutics13091454


Fridovich, I., Fridovich-Keil, J. L., & Robinson, A. (1998, August 23). Genetics of cancer. Encyclopædia Britannica. https://www.britannica.com/science/human-genetic-disease/Genetics-of-cancer


GeneCards. (2024, August 5). AGO2 Gene - Argonaute RISC Catalytic Component 2. Genecards. https://www.genecards.org/cgi-bin/carddisp.pl?gene=AGO2


Gerhardt, A., Voigt, E., Archer, M., Reed, S., Larson, E., Van Hoeven, N., Kramer, R., Fox, C., & Casper, C. (2022). A flexible, thermostable nanostructured lipid carrier platform for RNA vaccine delivery. Molecular Therapy - Methods & Clinical Development, 25, 205–214. https://doi.org/10.1016/j.omtm.2022.03.009


Guo, M., Zhang, H., Zheng, J., & Liu, Y. (2020). Glypican-3: A new target for diagnosis and treatment of hepatocellular carcinoma. Journal of Cancer, 11(8), 2008–2021. https://doi.org/10.7150/jca.39972


Ishibashi, A., Saga, K., Hisatomi, Y., Li, Y., Kaneda, Y., & Nimura, K. (2020). A simple method using CRISPR-Cas9 to knock-out genes in murine cancerous cell lines. Scientific Reports, 10(1). https://doi.org/10.1038/s41598-020-79303-0


Jiang, C., Sun, H., Jiang, Z., Tian, W., Cang, S., & Yu, J. (2024). Targeting the CD47/SIRPΑ pathway in malignancies: Recent progress, difficulties and future perspectives. Frontiers in Oncology, 14. https://doi.org/10.3389/fonc.2024.1378647


Jorgovanovic, D., Song, M., Wang, L., & Zhang, Y. (2020). Roles of IFN-γ in tumor progression and regression: A Review. Biomarker Research, 8(1). https://doi.org/10.1186/s40364-020-00228-x


Kamrud, K. I., Custer, M., Dudek, J. M., Owens, G., Alterson, K. D., Lee, J. S., Groebner, J. L., & Smith, J. F. (2007). Alphavirus replicon approach to promoterless analysis of IRES elements. Virology, 360(2), 376–387. https://doi.org/10.1016/j.virol.2006.10.049


Karginov, T. A., Pastor, D. P., Semler, B. L., & Gomez, C. M. (2017). Mammalian polycistronic mrnas and disease. Trends in Genetics, 33(2), 129–142. https://doi.org/10.1016/j.tig.2016.11.007


Karim, E., Tha, K. K., Othman, L., Uddin, M. B., & Chowdhury, E. H. (2018). Therapeutic potency of nanoformulations of siRNAs and shRNAs in animal models of cancers. Pharmaceutics, 10(2), 65. https://doi.org/10.3390/pharmaceutics10020065


Karthikeyan, K., Krishnaswamy, V. R., Lakra, R., Kiran, M. S., & Korrapati, P. S. (2015). Fabrication of Electrospun Zein nanofibers for the sustained delivery of Sirna. Journal of Materials Science: Materials in Medicine, 26(2). https://doi.org/10.1007/s10856-015-5439-x


Kautz, T. F., Guerbois, M., Khanipov, K., Patterson, E. I., Langsjoen, R. M., Yun, R., Warmbrod, K. L., Fofanov, Y., Weaver, S. C., & Forrester, N. L. (2018). Low-fidelity Venezuelan equine encephalitis virus polymerase mutants to improve live-attenuated vaccine safety and efficacy. Virus Evolution, 4(1). https://doi.org/10.1093/ve/vey004


Kemnic, T. R. (2022, November 28). Trimethoprim sulfamethoxazole. StatPearls. https://www.ncbi.nlm.nih.gov/books/NBK513232/


Kim, A. S., & Diamond, M. S. (2022). A molecular understanding of alphavirus entry and antibody protection. Nature Reviews Microbiology, 21(6), 396–407. https://doi.org/10.1038/s41579-022-00825-7


Kim, S. K., & Cho, S. W. (2022). The evasion mechanisms of cancer immunity and drug intervention in the tumor microenvironment. Frontiers in Pharmacology, 13. https://doi.org/10.3389/fphar.2022.868695


Kim, S., Kim, D., Cho, S. W., Kim, J., & Kim, J.-S. (2014). Highly efficient RNA-guided genome editing in human cells via delivery of purified Cas9 ribonucleoproteins. Genome Research, 24(6), 1012–1019. https://doi.org/10.1101/gr.171322.113


Li, Y., Teague, B., Zhang, Y., Su, Z., Porter, E., Dobosh, B., Wagner, T., Irvine, D. J., & Weiss, R. (2019). In vitro evolution of enhanced RNA replicons for immunotherapy. Scientific Reports, 9(1). https://doi.org/10.1038/s41598-019-43422-0


Lundstrom, K. (2016). Replicon RNA viral vectors as vaccines. Vaccines, 4(4), 39. https://doi.org/10.3390/vaccines4040039


Marques, A. C., Costa, P. C., Velho, S., & Amaral, M. H. (2023). Lipid Nanoparticles Functionalized with Antibodies for Anticancer Drug Therapy. Pharmaceutics, 15(1), 216. https://doi.org/10.3390/pharmaceutics15010216


Mantovani, A., Allavena, P., Marchesi, F., & Garlanda, C. (2022). Macrophages as tools and targets in cancer therapy. Nature Reviews Drug Discovery, 21(11), 799–820. https://doi.org/10.1038/s41573-022-00520-5


Mc Cafferty, S., De Temmerman, J., Kitada, T., Becraft, J. R., Weiss, R., Irvine, D. J., Devreese, M., De Baere, S., Combes, F., & Sanders, N. N. (2021). In vivo validation of a reversible small molecule-based switch for synthetic self-amplifying mrna regulation. Molecular Therapy, 29(3), 1164–1173. https://doi.org/10.1016/j.ymthe.2020.11.010


McIntyre, G. J., & Fanning, G. C. (2006). Design and cloning strategies for constructing Shrna Expression Vectors. BMC Biotechnology, 6(1). https://doi.org/10.1186/1472-6750-6-1


Morrissey, M. A., Williamson, A. P., Steinbach, A. M., Roberts, E. W., Kern, N., Headley, M. B., & Vale, R. D. (2018). Chimeric antigen receptors that trigger phagocytosis. eLife, 7. https://doi.org/10.7554/elife.36688


Nagarsheth, N., Wicha, M. S., & Zou, W. (2017). Chemokines in the cancer microenvironment and their relevance in cancer immunotherapy. Nature Reviews Immunology, 17(9), 559–572. https://doi.org/10.1038/nri.2017.49


Naidoo, D. B., & Chuturgoon, A. A. (2021). Nanobodies enhancing cancer visualization, diagnosis and therapeutics. International Journal of Molecular Sciences, 22(18), 9778. https://doi.org/10.3390/ijms22189778


O’Keefe, E. P. (2013). SiRNAs and shRNAs: Tools for protein knockdown by gene silencing. Materials and Methods, 3. https://doi.org/10.13070/mm.en.3.197


Qin, L., Lai, Y., Zhao, R., Wei, X., Weng, J., Lai, P., Li, B., Lin, S., Wang, S., Wu, Q., Liang, Q., Li, Y., Zhang, X., Wu, Y., Liu, P., Yao, Y., Pei, D., Du, X., & Li, P. (2017). Incorporation of a hinge domain improves the expansion of chimeric antigen receptor T cells. Journal of Hematology & Oncology, 10(1). https://doi.org/10.1186/s13045-017-0437-8


Qu, T., Li, B., & Wang, Y. (2022). Targeting CD47/SIRPΑ as a therapeutic strategy, where we are and where we are headed. Biomarker Research, 10(1). https://doi.org/10.1186/s40364-022-00373-5


Ray, M., Lee, Y.-W., Hardie, J., Mout, R., Yeşilbag Tonga, G., Farkas, M. E., & Rotello, V. M. (2018). CRISPRED macrophages for cell-based cancer immunotherapy. Bioconjugate Chemistry, 29(2), 445–450. https://doi.org/10.1021/acs.bioconjchem.7b00768


Regier, M. C., Taylor, J. D., Borcyk, T., Yang, Y., & Pannier, A. K. (2012). Fabrication and characterization of DNA-loaded zein nanospheres. Journal of Nanobiotechnology, 10(1). https://doi.org/10.1186/1477-3155-10-44


Rodriguez-Garcia, A., Lynn, R. C., Poussin, M., Eiva, M. A., Shaw, L. C., O’Connor, R. S., Minutolo, N. G., Casado-Medrano, V., Lopez, G., Matsuyama, T., & Powell, D. J. (2021). Car-T cell-mediated depletion of immunosuppressive tumor-associated macrophages promotes endogenous antitumor immunity and augments adoptive immunotherapy. Nature Communications, 12(1). https://doi.org/10.1038/s41467-021-20893-2


Rona, G., Zeke, A., Miwatani-Minter, B., de Vries, M., Kaur, R., Schinlever, A., Garcia, S. F., Goldberg, H. V., Wang, H., Hinds, T. R., Bailly, F., Zheng, N., Cotelle, P., Desmaële, D., Landau, N. R., Dittmann, M., & Pagano, M. (2021). The NSP14/NSP10 RNA repair complex as a pan-coronavirus therapeutic target. Cell Death & Differentiation, 29(2), 285–292. https://doi.org/10.1038/s41418-021-00900-1


Ruda, V. M., Chandwani, R., Sehgal, A., Bogorad, R. L., Akinc, A., Charisse, K., Tarakhovsky, A., Novobrantseva, T. I., & Koteliansky, V. (2014). The roles of individual mammalian Argonautes in RNA interference in vivo. PLoS ONE, 9(7). https://doi.org/10.1371/journal.pone.0101749


Schroeder, K. T., McPhee, S. A., Ouellet, J., & Lilley, D. M. J. (2010). A structural database for k-turn motifs in RNA. RNA, 16(8), 1463–1468. https://doi.org/10.1261/rna.2207910


Sterner, R. C., & Sterner, R. M. (2021). Car-T cell therapy: Current limitations and potential strategies. Blood Cancer Journal, 11(4). https://doi.org/10.1038/s41408-021-00459-7


Taxman, D. J., Moore, C. B., Guthrie, E. H., & Huang, M. T.-H. (2010). Short hairpin RNA (shrna): Design, delivery, and assessment of Gene Knockdown. Methods in Molecular Biology, 139–156. https://doi.org/10.1007/978-1-60761-657-3_10


U.S. Food and Drug Administration. (2023, November 28). FDA investigating serious risk of T-cell malignancy.


U.S. Food and Drug Administration. (2023, November 28). FDA investigating serious risk of T-cell malignancy. https://www.fda.gov/vaccines-blood-biologics/safety-availability-biologics/fda-investigating-serious-risk-t-cell-malignancy-following-bcma-directed-or-cd19-directed-autologous


U.S. Food and Drug Administration. (2024, April 18). FDA requires boxed warning for T cell malignancies. FDA Requires Boxed Warning for T cell Malignancies Following Treatment with BCMA-Directed or CD19-Directed Autologous Chimeric Antigen Receptor (CAR) T cell Immunotherapies. https://www.fda.gov/vaccines-blood-biologics/safety-availability-biologics/fda-requires-boxed-warning-t-cell-malignancies-following-treatment-bcma-directed-or-cd19-directed


Wagner, A., & Mutschler, H. (2023). Design principles and applications of synthetic self‐replicating RNAs. WIREs RNA, 14(6). https://doi.org/10.1002/wrna.1803


Wagner, T. E., Becraft, J. R., Bodner, K., Teague, B., Zhang, X., Woo, A., Porter, E., Alburquerque, B., Dobosh, B., Andries, O., Sanders, N. N., Beal, J., Densmore, D., Kitada, T., & Weiss, R. (2018). Small-molecule-based regulation of RNA-delivered circuits in mammalian cells. Nature Chemical Biology, 14(11), 1043–1050. https://doi.org/10.1038/s41589-018-0146-9


Warren, L., & Lin, C. (2019). MRNA-based genetic reprogramming. Molecular Therapy, 27(4), 729–734. https://doi.org/10.1016/j.ymthe.2018.12.009


Wienert, B., Shin, J., Zelin, E., Pestal, K., & Corn, J. E. (2018). In vitro–transcribed guide RNAS trigger an innate immune response via the rig-I pathway. PLOS Biology, 16(7). https://doi.org/10.1371/journal.pbio.2005840


Xu, C. (2024). CRISPR/Cas9-mediated knockout strategies for enhancing immunotherapy in breast cancer. Naunyn-Schmiedeberg’s Archives of Pharmacology. https://doi.org/10.1007/s00210-024-03208-2


Yang, E. Y., & Shah, K. (2020). Nanobodies: Next generation of cancer diagnostics and therapeutics. Frontiers in Oncology, 10. https://doi.org/10.3389/fonc.2020.01182


Yang, Y., Xu, J., Ge, S., & Lai, L. (2021). CRISPR/Cas: Advances, limitations, and applications for Precision Cancer Research. Frontiers in Medicine, 8. https://doi.org/10.3389/fmed.2021.649896


Yoshioka, N., Gros, E., Li, H.-R., Kumar, S., Deacon, D. C., Maron, C., Muotri, A. R., Chi, N. C., Fu, X.-D., Yu, B. D., & Dowdy, S. F. (2013). Efficient generation of human ipscs by a synthetic self-replicative RNA. Cell Stem Cell, 13(2), 246–254. https://doi.org/10.1016/j.stem.2013.06.001


Zhang, C., Liu, J., Zhong, J. F., & Zhang, X. (2017). Engineering car-T cells. Biomarker Research, 5(1). https://doi.org/10.1186/s40364-017-0102-y


Zhang, X., Dong, C., Hu, Y., Gao, M., & Luan, G. (2021). Zein as a structural protein in gluten-free systems: An Overview. Food Science and Human Wellness, 10(3), 270–277. https://doi.org/10.1016/j.fshw.2021.02.018


Zhao, Y., & Huang, L. (2014). Lipid nanoparticles for gene delivery. Advances in Genetics, 13–36. https://doi.org/10.1016/b978-0-12-800148-6.00002-x