Research Article
Print
Research Article
Computational design and validation of siRNA molecules to silence oncogenic CTNNB1 mRNA as a potential therapeutic strategy against hepatitis B/C virus-associated hepatocellular carcinoma
expand article infoLauren Emily Fajardo, Mark Andrian Macalalad, Nyzar Mabeth Odchimar, John Christian de Guzman, Fredmoore Orosco§|
‡ Industrial Technology Development Institute, Taguig, Philippines
§ S&T Fellows Program, Department of Science and Technology, Taguig, Philippines
| University of the Philippines Manila, Manila, Philippines
Open Access

Abstract

The majority of hepatocellular carcinoma cases are caused by infection with hepatitis B (HBV) or C (HCV) viruses. CTNNB1 is the most mutated oncogene in HBV- and HCV-associated tumors. CTNNB1 mutations can lead to β-catenin accumulation, resulting in tumor progression. Small interfering RNAs (siRNAs) can be used to silence CTNNB1 mRNA. After prediction and evaluation, four siRNAs were found to have the highest silencing potential. All four siRNAs had an acceptable GC content, no palindromic sequences, no off-targets, were thermostable, and had accessible target sites. Molecular docking of the siRNAs to Argonaute 2 demonstrated favorable docking scores within the binding pocket for three siRNAs. Molecular dynamics simulations and binding energy calculations demonstrated that the siRNAs steadily remained in the binding pocket. In this study, three siRNAs were successfully designed to silence oncogenic CTNNB1 mRNA as a therapeutic strategy against hepatocellular carcinoma and warrant further in vitro and in vivo validation.

Keywords

Argonaute 2 (Ago2), molecular docking, molecular dynamics, small interfering RNA (siRNA), RNA interference (RNAi)

Introduction

Hepatocellular carcinoma (HCC) is the most common type of primary liver cancer and accounts for approximately 90% of all cases (Llovet et al. 2021). HCC progression involves a series of genetic, epigenetic, and molecular alterations in hepatocytes that are mainly associated with chronic inflammation in the liver (Panneerselvam et al. 2023). Risk factors linked to the development of HCC include chronic hepatitis B virus (HBV) and hepatitis C virus (HCV) infections, which represent the most predominant etiology (Farazi and DePinho 2006; Akinyemiju et al. 2017). In particular, HCV accounts for 29.5% of HCC cases, compared to 8.2% of HCC cases caused by HBV (Shiels et al. 2019).

Somatic mutations in various oncogenes and tumor suppressor genes are considered drivers of HCC development. Some of these mutations may be caused by the integration of HBV DNA into the host genome (Yeh et al. 2023) or by dysregulation of important cellular processes by HCV-associated proteins (Irshad et al. 2017). CTNNB1 is the most frequently mutated oncogene in HCC, with mutations occurring in more than 10% of tumors. CTNNB1 encodes β-catenin, an important intracellular signal transducer in the Wnt signaling pathway. In the absence of Wnt, β-catenin is phosphorylated at the N-terminal domain and subsequently degraded. However, mutations in serine and threonine residues in exon 3 of CTNNB1 impair the phosphorylation site of β-catenin, preventing its degradation. This results in the aberrant activation of the Wnt pathway, which drives HCC progression (Tornesello et al. 2013) (Fig. 1).

Figure 1. 

Graphical representation of how HBV/HCV-induced CTNNB1 mutations drive HCC progression. HBV/HCV infection can introduce somatic mutations in exon 3 of CTNNB1, impairing the phosphorylation site and subsequently preventing the degradation of β-catenin, allowing aberrant activation of transcription factors, and driving HCC progression. hepatitis B virus; HCV, hepatitis C virus; TCF/LEF, T-cell factor/lymphoid enhancer factor; HCC, hepatocellular carcinoma.

RNA interference (RNAi) is a naturally occurring biological process in which short, double-stranded RNAs, such as microRNAs (miRNAs) and small interfering RNAs (siRNAs), recognize and degrade mRNA via post-transcriptional gene silencing. In the siRNA RNAi pathway, double-stranded RNA (dsRNA) is recognized by Dicer and cleaved into a siRNA duplex (Katoh and Suzuki 2007). The siRNA duplex is incorporated into the endonuclease Argonaute 2 (Ago2), a component of the RNA-induced silencing complex (RISC), and is unwound by an ATP-dependent helicase. One siRNA strand, known as the passenger strand, is discarded and degraded, while the other strand, known as the guide strand, recognizes and binds to a complementary mRNA sequence, allowing it to be cleaved by RISC and degraded (Leung and Whittaker 2005) (Fig. 2).

Figure 2. 

Graphical representation of the RNAi pathway for siRNAs. The siRNA guide strand is loaded onto Ago2, while the passenger strand is discarded. Ago2 associates with other proteins to form RISC and facilitates binding of the guide strand to a complementary mRNA sequence. The endonuclease activity of Ago2 cleaves the mRNA, thereby silencing its expression. Ago2, Argonaute 2.

The use of computational tools to predict and evaluate potential siRNA candidates allows for the efficient screening of hundreds to thousands of short RNA sequences at once. The selection of siRNAs with the highest silencing potential was successfully validated experimentally in previous studies. As such, siRNAs that were computationally designed against the MERS-CoV-orf1ab gene using siRNA prediction algorithms and evaluation of GC content, off-targets, target site accessibility, thermodynamic stability, and secondary structure formation were validated and were found to inhibit viral replication in Vero cells (Sohrab et al. 2021). Another study by Sohrab et al. (2022) designed siRNAs against SARS-CoV-2-S-RBD using similar prediction and evaluation methods and molecular docking of the predicted siRNAs to their target sequences. The predicted siRNAs demonstrated antiviral efficiency without cytotoxicity in Vero E6 cells. Similar methods were used to predict and evaluate siRNAs against a highly conserved region of the MERS-CoV membrane (M), with the top candidates demonstrating antiviral efficacy without cytotoxicity in Vero cells (El-Sayed et al. 2023).

The regulation of β-catenin expression through siRNA-mediated silencing of CTNNB1 may serve as a potential therapeutic strategy against HCC. To rapidly screen and evaluate all possible sequences, this study aimed to design siRNAs targeting the coding sequence (CDS) of oncogenic CTNNB1 using computational methods. In particular, this study aimed to predict and evaluate potential siRNAs targeting CTNNB1 using known algorithms and model the behavior of the top siRNA candidates through molecular docking with Ago2 and molecular dynamics simulation.

Materials and methods

Retrieval of CTNNB1 CDS and multiple sequence alignment

Thirty-nine (39) transcript variants of the CDS of CTNNB1 were retrieved from Ensembl (Martin et al. 2023) and aligned using ClustalW in MEGA11 (Tamura et al. 2021). Regions that were conserved in the majority of transcripts were used for siRNA prediction.

Prediction of siRNAs

The conserved regions of the CTNNB1 CDS were used to predict potential siRNAs in siDirect 2.1 (Naito et al. 2004, 2009) using the Ui-Tei, Reynolds, and Amarzguioui algorithms (Amarzguioui and Prydz 2004; Reynolds et al. 2004; Ui-Tei 2004) (Table 1). The Homo sapiens transcript Refseq release 223 from September 2023 was used to limit specificity to humans, and the maximum melting temperature of the seed duplex (Max Tm) was set at 21.5 °C to minimize seed-dependent off-target effects. The predicted siRNAs were cross-validated by using the siExplorer (Katoh and Suzuki 2007), OligoWalk (Lu and Mathews 2008), and siPRED (Pan et al. 2011) servers. siRNAs that scored ≥60 in siExplorer, ≥0.7 in OligoWalk, and ≥70% in siPRED were retained for further analysis.

Table 1.

Algorithms used in siDirect 2.1. To be predicted as a potential siRNA, an RNA sequence must satisfy all four Ui-Tei rules, at least six Reynolds rules, or at least three Amarzguioui rules.

Ui-Tei Reynolds Amarzguioui
1. A/U at 5’ end of antisense strand 1. 30–52% GC content 1. A/U content differential between the terminal 3 nucleotides of both ends
2. G/C at 5’ end of sense strand 2. ≥3 A/U at positions 15–19 of sense strand 2. No U at position 1 of sense strand
3. ≥5 A/U residues in 5’ terminal one-third of antisense strand 3. No internal repeats 3. A at position 6 of sense strand
4. No GC stretch ≥9 nt 4. A at position 19 of sense strand 4. Strong attachment of 5’ end of sense strand
5. A at position 3 of sense strand 5. Weak binding of 3’ end of sense strand
6. U at position 10 of sense strand 6. No G at position 19 of sense strand
7. No G/C at position 19 of sense strand 7. G/C at position 1 of sense strand
8. No G at position 13 of sense strand 8. A/U at position 1 of sense strand

GC content, presence of palindromic sequences, and off-target similarity search

The GC content of the siRNA duplexes and the presence of palindromic sequences in both the guide and passenger strands of each siRNA were determined using the OligoCalc server (Kibbe 2007). siRNAs with a GC content of 31.6–57.9% and no palindromic sequences of at least four (4) nucleotides in length on either strand were kept for off-target analysis in BLAST (Altschul et al. 1990). Both strands of each siRNA were checked for similarity against the RefSeq RNA database of Homo sapiens (taxid:9606). The program was optimized for somewhat similar sequences (blastn), the expected threshold was set to 100, and the word size was set to 7. siRNAs that had at least two (2) mismatches to all non-targeted transcripts in the 19-nt regions spanning positions 2–20 of both strands were cross-validated by the genome-wide enrichment of seed sequence (GESS) tool (Yilmazel et al. 2014) to further check for off-target effects against the built-in Human CDS database. siRNAs that were found to have no off-target effects in both BLAST and GESS were retained for further evaluation.

Prediction of target site accessibility

Target site accessibility within the target mRNA of each guide siRNA was predicted using the siRNA module of Sfold (Ding and Lawrence 2001, 2003; Ding et al. 2004). siRNAs with an antisense siRNA binding energy of ≤10, differential stability of siRNA duplex ends (DSSE) of >0, duplex feature score of ≥7, average internal stability (AIS) at the cleavage site of ≥-8.6, duplex thermodynamics score of 2, and total score of ≥12 were considered to have good target accessibility. The target accessibility score was also noted. The siRNAs that passed all evaluations were chosen as final candidates.

Secondary structure prediction of siRNA and mRNA-siRNA duplexes and thermodynamics analysis

The secondary structure and free energy of folding of the guide strand of each siRNA were predicted using MaxExpect (Lu et al. 2009), while the secondary structure and free energy of binding of the guide siRNA and the target mRNA were predicted using DuplexFold, which are available on the RNAstructure web server (Reuter and Mathews 2010). Heat capacity and concentration plots for each mRNA-siRNA duplex were generated using DINAMelt (Markham and Zuker 2005).

Structure retrieval and optimization of human Argonaute 2

The crystal structure of human Ago2 was retrieved from the Research Collaboratory for Structural Bioinformatics Protein Data Bank (RCSB PDB) (PDB ID: 4Z4D) (Schirle and MacRae 2015). To fill in the missing loops, homology modeling was performed using SWISS-MODEL (Waterhouse et al. 2018), using the crystal structure as a template against the complete Ago2 sequence (UniProt ID: Q9UKV8) (Bateman et al. 2023). The predicted structure was refined using GalaxyRefine (Heo et al. 2013). The refined model with the lowest MolProbity was evaluated using the ERRAT and PROCHECK modules in SAVES 6.0 (Colovos and Yeates 1993; Laskowski et al. 1993). The overall model quality and the local model quality were validated using the Protein Structural Analysis web server (ProSA-web) (Sippl 1993; Wiederstein and Sippl 2007).

siRNA tertiary structure modeling

The tertiary structures of the guide siRNAs were modeled using BIOVIA Discovery Studio Visualizer v21.1.0.20298 based on the template miRNA bound to the crystal structure of Ago2. The template RNA sequence was mutated based on the siRNA sequence. The geometry of the siRNA was then optimized using the CHARMM27 force field.

Molecular docking of siRNAs to human Argonaute 2

The optimized siRNA was then docked to Ago2 using HDOCK (Yan et al. 2020). The docked complex with the lowest (most negative) docking energy score, where the siRNA was bound to both the PIWI/Argonaute/Zwille (PAZ) and middle (MID) domains of Ago2 and had a similar conformation to the template RNA, was chosen for molecular dynamics simulations. The molecular interactions between the siRNA and Ago2 were evaluated using PDBSum (Laskowski et al. 2018).

Molecular dynamics of the siRNA-Ago2 complex

Molecular dynamics (MD) simulations of the docked siRNA-Ago2 complexes were performed using GROMACS 2023.3 to assess their stability. The topology and parameter files were created using the CHARMM36 force field (Vanommeslaeghe et al. 2010). The complex was initially enclosed in a box with periodic boundary conditions, maintaining a distance of 10 Å between the solute and the box edge. The box was then filled with TIP3P water molecules and neutralized by the addition of sodium and chloride ions. Following neutralization, an additional 0.15 M salt concentration was introduced to mimic human physiological conditions. Once the system was assembled, minimization using the steepest descent algorithm was conducted for 5000 steps to relax the system and prevent any steric clashes between atoms. The system equilibration was conducted in two phases to achieve the desired temperature and pressure: NVT and NPT equilibration. During NVT equilibration, the volume, temperature, and number of atoms were kept constant until the target temperature of 310 K. Similarly, in the subsequent NPT equilibration, the pressure, temperature, and number of atoms were held constant until the desired pressure of 1 bar was achieved. The modified Berendsen thermostat and the Parrinello-Rahman thermostat were used to regulate the temperature and pressure, respectively. Finally, the production run was carried out for 100 ns or 50,000,000 steps, saving snapshots every 100 ps or 50,000 steps, generating a total of 1000 frames.

Binding energy calculation

The siRNA molecules that exhibited stable behavior were further assessed for their binding affinities against Ago2. The GROMACS-compatible gmx_MMPBSA package (Valdés-Tresanco et al. 2021) was used to estimate the binding energy, which was determined from the sum of the van der Waals, electrostatic, polar, and nonpolar solvation energies (Equation 1). The molecular mechanics components (van der Waals and electrostatic forces) are from the intramolecular interactions of the siRNA-Ago2 complex, while the solvation energies are the interactions between the continuum solvent and the complex.

ΔGbinding = ΔGvdW + ΔGelectrostatic + ΔGpolar + ΔGnonpolar (1)

The van der Waals and electrostatic energies were calculated using the Lennard-Jones potential and Coulomb’s law, respectively. The polar solvation energy was estimated using the Poisson-Boltzmann and generalized Born equations, while the nonpolar solvation energy was assumed to be linearly dependent on the solvent-accessible surface area (SASA). Only the last 10 ns, representing the stable phase of the trajectory, were investigated.

Results

Retrieval of CTNNB1 CDS and multiple sequence alignment

A total of 39 CTNNB1 CDS transcripts were retrieved from Ensembl and aligned using ClustalW in MEGA11. A complete list of the transcripts is shown in Suppl. material 1: table S1. The transcripts ranged in length from 2051 to 2352 nt, with most transcript variants being either 2325 or 2346 nt long. The 2325 nt region making up the entirety of nine of the transcript variants was conserved in the majority of the transcript variants and was used as the region for siRNA prediction.

siRNA prediction

From the CTNNB1 CDS consensus sequence, a total of 45 siRNAs were predicted in siDirect 2.1 using the Ui-Tei, Reynolds, and Amarzguioui algorithms. Each siRNA fulfilled the requirements of at least one of the three algorithms, with nine (9) siRNAs fulfilling the requirements of all three. Of the 45 initially predicted siRNAs, 27 were cross-validated using siExplorer, OligoWalk, and siPRED.

GC content, presence of palindromic sequences, and off-target similarity search

Of the 27 predicted siRNAs, 19 had a GC content within the range of 31.6–57.9%, and 16 siRNAs did not have palindromic sequences in both their guide and passenger strands, as predicted by OligoCalc. Ten (10) siRNAs did not have any off-targets as they had at least two (2) mismatches in positions 2–20 of both the guide and passenger strands to any non-targeted transcripts in BLAST. This was further confirmed in GESS as off-targets were not significant by Fisher, Bonferroni, Bonferroni step-down (Holm) correction, and Benjamini and Hochberg methods.

Prediction of target site accessibility

The target site accessibility of the remaining ten (10) siRNAs was predicted using Sfold. The target accessibility scores of the siRNAs ranged between 5 and 8, with 0 being the lowest possible score and 8 being the highest possible score. The antisense binding energy of the siRNAs ranged between -27 and -11.8 kcal/mol. The DSSE of the siRNAs ranged from 0.8 to 7.5 kcal/mol. Five (5) out of ten (10) siRNAs had duplex feature scores ranging from 7–9 and were used for further analysis. The other five (5) siRNAs had scores ranging from 5 to 6 and were eliminated. The AIS of the siRNAs ranged from -8.2 to -5.1 kcal/mol. The duplex thermodynamics score can be 0, 1, or 2, and simply assigns one (1) point for each threshold met for the DSSE and AIS scores. Since all five (5) siRNAs met both thresholds, it follows that they all had duplex thermodynamics scores of 2. Finally, the total score is the sum of the accessibility, duplex feature, and duplex thermodynamics scores, with a maximum possible score of 20 and scores greater than or equal to 12 being acceptable (Ding and Lawrence 2001, 2003; Ding et al. 2004). The siRNAs had total scores ranging from 15–19. The detailed Sfold results are shown in Suppl. material 1: table S2.

Overall, five (5) siRNAs passed all evaluations: s3, s4, s19, s22, and s28. However, since s3 and s4 targeted overlapping regions in the CTNNB1 CDS (positions 205–227 and 208–230, respectively) and designing two such siRNAs would be redundant, s3 was chosen over s4 as it scored better overall in all evaluations. Therefore, four (4) siRNAs, s3, s19, s22, and s28, were chosen as the top four siRNA candidates (Table 2).

Table 2.

The top four siRNA candidates.

CTNNB1 CDS target position alias guide strand seed-duplex Tm (guide) (°C) seed-duplex Tm (passenger) (°C) GC content (%)
205–227 s3 AAUAUCAGCUACUUGUUCUUG 13.3 19.2 33
730–752 s19 UGUAAUGGCAUAAAACAACAC 20.0 5.6 33
770–792 s22 UUUUAGCUCCUUCUUGAUGUA 18.3 20.4 33
964–986 s28 UUUUUCGUAAGUAUAGGUCCU 15.3 19.8 35

Secondary structure prediction of siRNA and mRNA-siRNA duplexes and thermodynamics analysis

The secondary structures of the siRNA guide strands were predicted using MaxExpect (Fig. 3). None of the predicted secondary structures of the siRNAs had palindromic sequences ≥ 4 nt that formed hairpin loops, further supporting the findings of OligoCalc. Likewise, all siRNAs had a positive free energy of folding ranging between 1.6 and 1.8 kcal/mol. In contrast, all the predicted mRNA-siRNA duplex structures (Fig. 4) exhibited perfect complementarity and had a negative free energy of binding ranging from -32.4 to -31.5 kcal/mol. Concentration and heat capacity plots for each mRNA-siRNA duplex, as well as their respective melting temperatures (Tm), were generated using DINAMelt (Suppl. material 1: fig. S1). The Tm (Conc) of the siRNAs ranged from 71.5–74.6 °C, while the Tm (Conc) ranged from 72.8–75.8 °C.

Figure 3. 

Secondary structures of the siRNA guide strands and their free energies of folding. A positive free energy of folding indicates a higher binding efficiency, as the molecule is less prone to secondary structure formation. A. s3; B. s19; C. s22; D. s28.

Figure 4. 

Secondary structures of the mRNA-siRNA duplexes and their free energies of binding. A more negative binding energy indicates a stronger interaction between the mRNA and siRNA. A. s3; B. s19; C. s22; D. s28.

Structure retrieval and optimization of human Argonaute 2

The 3D structure of Ago2 was predicted using SWISS-MODEL and refined using GalaxyRefine for molecular docking and molecular dynamics with the siRNAs. From the five generated models listed in Table 3, Model 1 was chosen because it had the lowest MolProbity (1.549). The overall and local model qualities of the refined model were assessed using ProSA-web, as shown in Fig. 5. The refined model has a z-score of -13.92, and it can be seen that the overall model quality of the refined model falls within the range of experimentally derived structures of similar size. The local model quality graph shows that the knowledge-based energy of the refined model is almost entirely negative, indicating that nearly the entire refined model is of sound quality with minimal erroneous areas. The quality of the refined model was further validated using the ERRAT and PROCHECK modules of the SAVES 6.0 web server. The ERRAT score of the refined model was 95.865 (Suppl. material 1: fig. S2). The Ramachandran plot showed that, among the non-glycine and non-proline residues of the refined model, 94.6% (695) were in the most favored regions, 5.3% (39) were in additional allowed regions, 0% were in generously allowed regions, and 0.1% (1) were in disallowed regions, as predicted by PROCHECK (Suppl. material 1: fig. S3).

Figure 5. 

Quality plots of the refined Ago2 model. A. Overall model quality plot plotting the z-score of the refined Ago2 model (black dot) against z-scores of PDB structures experimentally determined by X-ray crystallography (light blue) or nuclear magnetic resonance (NMR) spectroscopy (dark blue); B. Local model quality plot showing the knowledge-based energy at each position of the refined Ago2 model.

Table 3.

Top five (5) refined structures for Ago2 as predicted by GalaxyRefine.

Model GDT-HA RMSD MolProbity Clash score Poor rotamers Rama favored
Initial 1.0000 0.000 1.182 1.6 0.5 95.9
Model 1 0.9902 0.264 1.549 10.7 0.4 98.6
Model 2 0.9890 0.276 1.582 11.7 0.1 98.7
Model 3 0.9878 0.286 1.582 11.7 0.4 98.7
Model 4 0.9884 0.275 1.589 11.9 0.1 99.0
Model 5 0.9869 0.283 1.569 11.3 0.4 98.9

Molecular docking of siRNAs to human Argonaute 2

After modeling Ago2, molecular docking was performed with each siRNA using HDOCK (Table 4, Fig. 6). From the top docking conformations of each siRNA, the conformation closest to that of the miRNA (designated as s0) bound to the PDB structure of Ago2 was selected for further analysis. Only one complex of each siRNA simultaneously interacted with the MID and PAZ domains. The interacting residues in each docked structure were analyzed using PDBSum. s3, s22, and s28 docked to Ago2 in a similar manner as s0, with the 5’ and 3’ ends of the siRNAs interacting with the MID and PAZ domains of Ago2, respectively. However, none of the docked structures of s19 had the expected conformation. Interestingly, the top-docked structure of s19 had a conformation completely opposite to what was expected, with the 5’ end of the siRNA interacting with the PAZ domain and the 3’ end of the siRNA interacting with the MID domain. The three correctly docked siRNAs and s0 interacted with certain residues in PAZ and MID, namely Y311 in PAZ and K525, Y529, K533, Q548, N562, and K566 in MID.

Figure 6. 

Top four siRNA candidates bound to Ago2 with their respective docking scores. All siRNA candidates had highly negative docking scores and interacted with residues in both the PAZ and MID domains. A. s3; B. s19; C. s22; D. s28; E. s0.

Table 4.

Molecular docking results for the top four siRNA candidates. All siRNA candidates had highly negative docking scores and interacted with residues in both the PAZ and MID domains.

Alias Interacting residues in PAZ and MID domains
s3 PAZ: Y279, F294, Q297, Y311 MID: G524, K525, T526, Y529, K533, C546, Q548, N562, K566
s19 PAZ: G264, R277, R280, Y311, F312, E333, K335, L339 MID: K550, N551, R554
s22 PAZ: R277, K278, R280, F294, Y311, K335, Y338 MID: K525, Y529, K533, Q545, Q548, N551, N562, K566
s28 PAZ: R277, R280, F294, P295, L296, Q297, Y311 MID: L522, G524, K525, T526, Y529, K533, C546, V547, Q548, T559, N562, K566
s0 PAZ: K266, R277, Y279, R280, L296, Y311, H316, Q332, K335, H336, Y338
MID: G524, K525, T526, Y529, K533, Q545, C546, V547, Q548, N551, N562, L563, K566

Molecular dynamics simulation

The stability of the siRNA complexes was investigated through a 100-ns simulation to determine how the structure changed over time relative to its starting conformation [see Suppl. materials 25]. From the RMSD trajectories shown in Fig. 7, the siRNA complexes achieved stability within the first 20 ns and remained stable for the rest of the simulation run. After equilibration, all complexes exhibited RMSD fluctuations within 2 Å, similar to s0. The complexes of s3 and s22 displayed trajectories with RMSD values most closely resembling those of s0. Additionally, s28 demonstrated stable binding with Ago2 throughout the run, albeit displaying more erratic RMSD fluctuations toward the end. Upon examining the RMSF plot in Fig. 8, the behavior of the s28 nucleotides closely resembled the more restrained fluctuations observed in s3 and s22. s19 recorded the highest RMSD values, indicating a more distinct conformational change from its initial structure due to the more pronounced movements at the unbound ends of the siRNA. These movements were also reflected in the RMSF plot (Fig. 8), which shows high RMSF values, especially for the first and last nucleotides of the siRNA molecule. The H-bonds diagram shown in Fig. 9 also shows that s19 produced the least number of hydrogen bonds. In contrast, the remaining three siRNA molecules demonstrated a relatively stable range of 40–50 hydrogen bond connections, similar to s0.

Figure 7. 

RMSD trajectories of the siRNA-Ago2 complexes and the reference miRNA (s0). All siRNA molecules exhibited structural stability to the same extent as the reference miRNA throughout the 100-ns simulation.

Figure 8. 

RMSF plot of the siRNA molecules and the reference miRNA (s0). s19 displayed high RMSF values, particularly for the 1st and 21st nucleotides, which contributed to the relatively higher values in its RMSD trajectory. The remaining siRNA molecules presented RMSF plots similar to those of the reference miRNA.

Figure 9. 

Number of hydrogen bonds formed between the siRNA molecules and Ago2. Shades closer to purple indicate fewer H-bonds, whereas shades closer to red indicate more frequent hydrogen bonding.

Binding energy calculations

The binding affinity of the four siRNA molecules to Ago2 was evaluated using both the MMPBSA and MMGBSA methods. In both methods, all four remaining siRNA molecules registered stronger binding energies (MMPBSA: -412.46 kcal/mol to -465.57 kcal/mol, MMGBSA: -313.0 kcal/mol to -379.57 kcal/mol) compared to the experimentally derived control miRNA (MMPBSA: -377.06 kcal/mol, MMGBSA: -293.69 kcal/mol), indicating a higher affinity for Ago2 (Fig. 10). s3 consistently scored the highest, while s19 obtained the lowest binding energy scores due to its less stable behavior compared to the other three siRNA molecules.

Figure 10. 

Estimated binding energies of the four siRNA molecules and the reference miRNA (s0) to Ago2. The binding energy was calculated as the sum of the van der Waals, electrostatic, polar solvation, and non-polar solvation energies.

Table 5 presents the amino acid residues of Ago2 with the most significant contribution during siRNA binding. Five residues (R68, R179, R710, R792, and R795) were identified to interact with all four siRNA molecules and s0, highlighting their importance in the complex formation between the siRNA molecules and Ago2.

Table 5.

Ago2 residues with the strongest binding energy contribution. Residues highlighted in bold and italicized are the common interacting residues among all four siRNAs and the control miRNA (s0). Values are in kcal/mol.

s0 s3 s19 s22 s28
R68 -7.081 R710 -6.366 R795 -5.737 R710 -7.028 R68 -7.501
R179 -4.425 R68 -5.959 R351 -5.720 R68 -5.869 R761 -5.181
R792 -4.189 R761 -5.332 R68 -5.422 R351 -4.988 P67 -5.010
R280 -4.126 P67 -5.034 R126 -5.033 R277 -4.742 R710 -4.767
R277 -3.792 R792 -4.407 I756 -4.751 R795 -4.682 R126 -4.189
Q757 -3.515 R795 -4.280 R710 -4.217 R179 -4.584 Y279 -3.812
R761 -3.420 R126 -4.255 K65 -4.169 R792 -4.572 R179 -3.663
K124 -3.065 Y279 -4.241 R315 -3.648 R761 -3.651 I756 -3.616
R795 -3.062 Y529 -4.025 I365 -3.426 P67 -3.637 R792 -3.593
Y529 -3.014 R280 -3.813 R179 -3.358 R280 -3.612 R635 -3.520
G123 -2.987 K124 -3.807 K124 -2.933 Y529 -3.364 I365 -3.348
G178 -2.870 R97 -3.677 R277 -2.903 I756 -3.080 R795 -3.126
R351 -2.811 I365 -3.621 R792 -2.774 Q757 -3.068 Q757 -3.074
R710 -2.775 R179 -3.597 G178 -2.552 F294 -3.027 Y529 -3.036

Discussion

The use of computational tools to efficiently predict and evaluate potential siRNAs has been previously employed in targeting various genes, whether within the host or the pathogen, for the treatment of several diseases. This study offers a novel, more stringent, and broadly targeting screening method for predicting and evaluating siRNAs that are highly likely to have high silencing efficacy in experimental setups. The consideration of several transcript variants of CTNNB1 instead of just the most common ones guarantees that the siRNAs will be able to treat a larger cohort of patients. Furthermore, extensive molecular docking and molecular dynamics analyses of the siRNA candidates with Ago2 and comparison to an experimentally derived control offer more extensive insights than prior siRNA design studies.

To facilitate knockdown of β-catenin, siRNAs were designed around the coding region of CTNNB1. The CDS of CTNNB1 was chosen as the region to design siRNAs as it is retained in the mature mRNA and is the region that is translated into β-catenin (Tornesello et al. 2013). Although the 3’UTR and 5’UTR are also retained after the maturation process and could be used for siRNA design, these regions were not conserved across the retrieved CTNNB1 transcript variants and could not have been used to design broadly active siRNAs.

Different siRNA prediction servers implement different algorithms that consider various parameters that affect the siRNA efficacy. A total of 45 siRNAs were initially predicted using siDirect 2.1. siDirect 2.1 uses the Ui-Tei, Reynolds, and Amarzguioui algorithms (Amarzguioui and Prydz 2004; Reynolds et al. 2004; Ui-Tei 2004). These algorithms are based on observations made regarding siRNA efficiency in the presence or absence of certain nucleotides at certain positions on both the guide and passenger strands (Table 1). While these algorithms are widely used in siRNA prediction studies, they do not consider the interaction of siRNAs with other elements in the RNAi pathway (Yasmin et al. 2022). Therefore, in this study, a siRNA candidate meeting the conditions of at least one algorithm was considered sufficient, as other siRNA prediction algorithms were employed. Out of the 45 initially predicted siRNAs, 27 were cross-validated by OligoWalk, siPRED, and siExplorer. OligoWalk and siPRED predict efficient siRNA candidates based on the mechanics of the interaction between the siRNA and the target mRNA (Lu and Mathews 2008; Pan et al. 2011), while the siExplorer algorithm predicts siRNAs based on the finding that specific nucleotides at every third position (3n + 1) greatly enhance their binding affinity to transactivation response element RNA-binding protein (TRBP), a partner protein of Dicer (Katoh and Suzuki 2007). Using this highly selective method of siRNA prediction, candidate siRNAs are guaranteed to generate favorable results in downstream evaluations.

The 27 predicted siRNAs underwent evaluation for GC content, presence of palindromic sequences, presence of off-targets, secondary structure formation, target site accessibility, and thermostability. The GC content and presence of palindromic sequences were checked using OligoCalc. A high GC content has been suggested to inhibit the unwinding of the siRNA duplex, preventing RISC loading (Liu et al. 2012). A GC range of 31.6–57.9%, is considered the optimal range for siRNAs, as suggested by Amarzguioui and Prydz (2004). Palindromic sequences can prohibit the proper formation of the mRNA-siRNA duplex and must be avoided (Fakhr et al. 2016). The presence of off-targets was checked using BLAST and cross-validated using GESS. siRNAs and other small RNAs can tolerate mismatches at the target site and may induce silencing even with imperfect complementarity (Yasmin et al. 2022).

Target site accessibility was assessed using Sfold. Sfold returns various scores that assess the functionality of the siRNA throughout the entire RNAi pathway, from loading onto RISC to binding and cleavage of the target mRNA. A siRNA that meets the thresholds for all scores is highly likely to be functional. The target accessibility score is based on the antisense (guide) siRNA binding energy, which is the weighted sum of the RNA/RNA stacking energies of the mRNA-siRNA duplex. The more negative the antisense siRNA binding energy, the higher the score. According to the target accessibility rule, the antisense siRNA binding energy must be less than or equal to -10 kcal/mol. The AIS values of the candidate siRNAs indicate that the mRNA-siRNA duplexes were predicted to have favorable stacking interactions.

To silence the intended target site, the correct siRNA strand must first be loaded onto RISC. The DSSE is the difference in stability between the 5’ end of the guide strand and the 5’ end of the passenger strand. The DSSE must be greater than 0 kcal/mol (i.e., the 5’ end of the guide strand is less stable than the 5’ end of the passenger strand), as RISC forms a complex with the strand with the less stable 5’ end (Elhefnawi and Mysar 2011). The DSSE values of the candidate siRNAs confirm that the desired guide strands will be preferably loaded into RISC. The duplex feature score is based on the Reynolds algorithm (Reynolds et al. 2004) and adds or deducts points based on whether certain criteria are met, with the highest possible score being 10 and the lowest possible score being -2. siRNAs with scores greater than 6 are considered acceptable.

After ensuring that the guide strand is loaded onto RISC, its ability to bind to the target mRNA so that Ago2 can be correctly positioned to perform cleavage activity is assessed. The AIS is the average of internal stability values for positions 9–14 of the guide strand and should be greater than or equal to -8.6 kcal/mol.

For the remaining four (4) siRNA candidates, the secondary structures of the siRNA guide strands, and the mRNA-siRNA duplexes were predicted. A positive free energy of folding for siRNA guide strands indicates higher binding efficiency as the molecule is less prone to secondary structure formation (Chowdhury et al. 2021). In contrast, a negative binding energy for mRNA-siRNA duplexes indicates a stronger interaction between the mRNA and siRNA (Yasmin et al. 2022). Tm (Conc) is the temperature at which the concentration of the double-stranded mRNA-siRNA duplex is half its maximum value, while Tm (Cp) is the temperature at which the collective heat capacity reaches its maximum value. Higher Tm values indicate the effectiveness of the siRNA, with a higher melting temperature denoting a more effective siRNA (Yasmin et al. 2022).

The tertiary structures of the siRNAs and Ago2 were modeled in preparation for molecular docking. The siRNAs were modeled by mutating the bases of the template miRNA so as to retain the conformation of the small RNA when it is within the binding pocket of Ago2. The tertiary structure of Ago2 was predicted and refined. GalaxyRefine returns various scores that describe the refined models with respect to the initial model. The global distance test-high accuracy (GDT-HA) and root mean square deviation (RMSD) are measures of similarity between the positions of the alpha carbons of two protein structures with the same amino acid sequence. GDT-HA calculates the proportion of alpha carbons located within a specified proximity to one another (Zemla 2003), while RMSD measures the displacement of alpha carbons from their initial positions (Sargsyan et al. 2017). The MolProbity score is a combination of the clash score, poor rotamers score, and Rama-favored scores. The clash score represents the number of unusually close atom pairs in a protein model that may result in steric clashes; the poor rotamers score is the number of amino acids with poor conformations; and the Rama-favored score is the percentage of residues with favorable backbone ϕ and ψ angles (Yasmin et al. 2022). A lower MolProbity score indicates a higher-quality refined structure; therefore, Model 1 was considered the best-refined structure of Ago2.

ProSA-web assesses the overall and local model qualities of the protein structure. The overall model quality plots the z-score of the query structure against that of the experimentally derived protein structures arranged by size. A good-quality structure is expected to have a z-score similar to that of protein structures of similar sizes. The local model quality plots the knowledge-based energy of the protein structure as a function of amino acid sequence position, where negative values indicate good quality regions (Wiederstein and Sippl 2007). ERRAT compares the distribution of non-bonded atomic interactions in the query structure to distributions in a database of known, high-resolution structures; a structure with a score of 95% or greater is considered to be high-resolution (Yasmin et al. 2022). PROCHECK plots each residue in the query structure in the Ramachandran plot based on their backbone ϕ and ψ angles and categorizes them based on their favorability. Protein structures with more than 90% of residues in the most favored regions are considered to be of good quality (Sumitha et al. 2020).

Out of the four (4) candidate siRNAs, three (3) docked to Ago2 in the expected orientation. These three siRNAs had common interacting residues within the MID and PAZ domains. Y529 in particular is highly conserved and serves as an insertion site for the phosphorylated 5’-end of the guide strand. This interaction is critical for the binding and cleavage activity of the guide strand. Meanwhile, the non-specific anchoring of the hydroxylated 3’ overhang of the guide strand to the PAZ domain is crucial for preventing its degradation (Wu et al. 2020).

The docked complexes of the four siRNAs underwent 100-ns molecular dynamics simulations. The results of the MD simulations showed that s3 and s22 behaved in a manner more similar to the reference miRNA, while s19 and s28 displayed more erratic fluctuations. The relatively higher RMSD values of the s28-Ago2 complex were most likely caused by the random movements of the nondescript loops of Ago2 interacting with the mobile water molecules. Meanwhile, the high RMSD and RMSF values of the s19-Ago2 complex are likely attributed to the incorrect binding pose of s19, where the 5’ end docked against the PAZ domain and the 3’ end docked against the MID domain. The interaction between the mismatched Ago2 domains and the ends of the siRNA might have generated repulsive forces that weakened the binding affinity and stability of the Ago2-s19 complex compared to the other siRNA complexes. Furthermore, the constant movement of s19 may have hindered its ability to establish a consistent number of hydrogen bond interactions with the protein. Despite its stronger affinity compared to the reference miRNA, as shown by the binding energy calculations, the incorrect orientation of s19 against Ago2 renders it unsuitable as a viable siRNA to silence CTNNB1.

It is interesting to note that the majority of the residues with the strongest binding energy contribution are arginine, a positively charged amino acid. Arginine residues in the PAZ, MID, and PIWI domains are known to significantly interact with siRNAs (Bhandare and Ramaswamy 2016a). These residues may have contributed to enhancing the electrostatic forces from the charge-dipole interaction against the polar siRNA molecules, thereby leading to stable and strong complex binding. An additional arginine R761 interacts with the three correct siRNA molecules and s0, as well as Y529, which, as mentioned earlier, is a highly conserved insertion site for the guide strand. Aside from the flipped conformation, the complete absence of an interaction between s19 and Y529 likely contributed to the stark difference in results.

These results show that in silico methods can be employed for the rational design of siRNAs targeting oncogenic CTNNB1. The computational design of therapeutic siRNAs has previously been employed to target wild-type or mutant human genes to combat various diseases, such as mutant tardbp against amyotrophic lateral sclerosis (Bhandare and Ramaswamy 2016b), CLEC5A against dengue and Japanese encephalitis virus (Yasmin et al. 2022), and oncogenic KRAS against cancer (Ramalingam and Arumugam 2023). Knockdown of CTNNB1 using siRNAs has previously been shown to decrease β-catenin activity in hepatoma cell lines (Zeng et al. 2007), although clinical trials have yet to be conducted. In silico screening allows us to rapidly characterize thousands of nucleotide sequences from a single gene and identify those with the highest silencing potential. Likewise, molecular docking and molecular dynamics studies with Ago2 provide a visualization of how siRNAs insert and interact with the binding pocket, as well as how this interaction changes throughout a minuscule timeframe under simulated conditions in the human body. The use of in silico methods in developing therapeutics is a rapidly growing field, and as technology evolves, computational results become more accurate, precise, and informative and greatly supplement experimental validation.

Conclusion

In this study, siRNAs were predicted from the CDS of CTNNB1 and evaluated for their silencing capability. The top four siRNA candidates underwent molecular docking and molecular dynamics simulation. s3, s22, and s28 demonstrated stability similar to the reference miRNA (s0), while s19 exhibited relatively weaker stability due to its incorrect binding orientation. Further analysis using MMPBSA and MMGBSA revealed that all four siRNA molecules exhibited stronger binding affinity towards the Ago2 protein compared to the miRNA, underscoring their potential for silencing CTNNB1. Although s19 showed promising binding energy scores, its mismatched binding pose indicates its unsuitability as a siRNA molecule, prompting the need for additional investigation. Additionally, we identified five arginine residues (R68, R179, R710, R792, and R795) with the highest binding energy contribution to siRNA binding. These specific residues hold potential as crucial targets for the optimization of siRNA design. Considering all evaluations, three (3) siRNAs were successfully designed to silence oncogenic CTNNB1, as demonstrated by their favorable properties and stable complex formation with Ago2. Although the results warrant further in vitro and in vivo validation, the current findings provide valuable insights and a strong foundation for further research.

Acknowledgements

The authors would like to thank the Philippine Council for Agriculture, Aquatic, and Natural Resources Research and Development (PCAARRD), which funded the project under the Virology and Vaccine Research and Development Program. The authors would also like to thank the Department of Science and Technology—S&T Fellows Program, the Department of Science and Technology—Science Education Institute (DOST-SEI) Career Incentive Program, the Department of Science and Technology—Industrial Technology Development Institute (DOST-ITDI), and the Department of Science and Technology—Advanced Science and Technology Institute (DOST-ASTI) Computing and Archiving Research Environment (COARE) for their support of this research.

References

  • Akinyemiju T, Abera S, Ahmed M, Alam N, Alemayohu MA, Allen C, Al-Raddadi R, Alvis-Guzman N, Amoako Y, Artaman A, Ayele TA, Barac A, Bensenor I, Berhane A, Bhutta Z, Castillo-Rivas J, Chitheer A, Choi J-Y, Cowie B, Dandona L, Dandona R, Dey S, Dicker D, Phuc H, Ekwueme DU, Zaki MES, Fischer F, Fürst T, Hancock J, Hay SI, Hotez P, Jee SH, Kasaeian A, Khader Y, Khang Y-H, Kumar GA, Kutz M, Larson H, Lopez A, Lunevicius R, Malekzadeh R, McAlinden C, Meier T, Mendoza W, Mokdad A, Moradi-Lakeh M, Nagel G, Nguyen Q, Nguyen G, Ogbo F, Patton G, Pereira DM, Pourmalek F, Qorbani M, Radfar A, Roshandel G, Salomon JA, Sanabria J, Sartorius B, Satpathy M, Sawhney M, Sepanlou S, Shackelford K, Shore H, Sun J, Mengistu DT, Topór-Madry R, Tran B, Ukwaja KN, Vlassov V, Vollset SE, Vos T, Wakayo T, Weiderpass E, Werdecker A, Yonemoto N, Younis M, Yu C, Zaidi Z, Zhu L, Murray CJL, Naghavi M, Fitzmaurice C (2017) The burden of primary liver cancer and underlying etiologies from 1990 to 2015 at the global, regional, and national level. JAMA Oncology 3: 1683. https://doi.org/10.1001/jamaoncol.2017.3055
  • Bateman A, Martin M-J, Orchard S, Magrane M, Ahmad S, Alpi E, Bowler-Barnett EH, Britto R, Bye-A-Jee H, Cukura A, Denny P, Dogan T, Ebenezer T, Fan J, Garmiri P, da Costa Gonzales LJ, Hatton-Ellis E, Hussein A, Ignatchenko A, Insana G, Ishtiaq R, Joshi V, Jyothi D, Kandasaamy S, Lock A, Luciani A, Lugaric M, Luo J, Lussi Y, MacDougall A, Madeira F, Mahmoudy M, Mishra A, Moulang K, Nightingale A, Pundir S, Qi G, Raj S, Raposo P, Rice DL, Saidi R, Santos R, Speretta E, Stephenson J, Totoo P, Turner E, Tyagi N, Vasudev P, Warner K, Watkins X, Zaru R, Zellner H, Bridge AJ, Aimo L, Argoud-Puy G, Auchincloss AH, Axelsen KB, Bansal P, Baratin D, Batista Neto TM, Blatter M-C, Bolleman JT, Boutet E, Breuza L, Gil BC, Casals-Casas C, Echioukh KC, Coudert E, Cuche B, de Castro E, Estreicher A, Famiglietti ML, Feuermann M, Gasteiger E, Gaudet P, Gehant S, Gerritsen V, Gos A, Gruaz N, Hulo C, Hyka-Nouspikel N, Jungo F, Kerhornou A, Le Mercier P, Lieberherr D, Masson P, Morgat A, Muthukrishnan V, Paesano S, Pedruzzi I, Pilbout S, Pourcel L, Poux S, Pozzato M, Pruess M, Redaschi N, Rivoire C, Sigrist CJA, Sonesson K, Sundaram S, Wu CH, Arighi CN, Arminski L, Chen C, Chen Y, Huang H, Laiho K, McGarvey P, Natale DA, Ross K, Vinayaka CR, Wang Q, Wang Y, Zhang J (2023) UniProt: the universal protein knowledgebase in 2023. Nucleic Acids Research 51: D523–D531. https://doi.org/10.1093/nar/gkac1052
  • Bhandare V, Ramaswamy A (2016a) Structural dynamics of human Argonaute2 and its interaction with siRNAs designed to target mutant tdp43. Advances in Bioinformatics 2016: 1–13. https://doi.org/10.1155/2016/8792814
  • Bhandare VV, Ramaswamy A (2016b) Identification of possible siRNA molecules for TDP43 mutants causing amyotrophic lateral sclerosis: in silico design and molecular dynamics study. Computational Biology and Chemistry 61: 97–108. https://doi.org/10.1016/j.compbiolchem.2016.01.001
  • Chowdhury UF, Sharif Shohan MU, Hoque KI, Beg MA, Sharif Siam MK, Moni MA (2021) A computational approach to design potential siRNA molecules as a prospective tool for silencing nucleocapsid phosphoprotein and surface glycoprotein gene of SARS-CoV-2. Genomics 113: 331–343. https://doi.org/10.1016/j.ygeno.2020.12.021
  • Ding Y, Lawrence C (2001) Statistical prediction of single-stranded regions in RNA secondary structure and application to predicting effective antisense target sites and beyond. Nucleic Acids Research 29: 1034–1046. https://doi.org/10.1093/nar/29.5.1034
  • Ding Y, Lawrence C (2003) A statistical sampling algorithm for RNA secondary structure prediction. Nucleic Acids Research 31: 7280–7301. https://doi.org/10.1093/nar/gkg938
  • Ding Y, Chan CY, Lawrence CE (2004) Sfold web server for statistical folding and rational design of nucleic acids. Nucleic Acids Research 32: W135–W141. https://doi.org/10.1093/nar/gkh449
  • Elhefnawi M, Mysar M (2011) In-silico approaches for RNAi post-transcriptional gene regulation: optimizing siRNA design and selection. In: Bioinformatics - Trends and Methodologies. InTech, 738 pp. https://doi.org/10.5772/18455
  • El-Sayed AY, Shehata M, Mahmoud SH, ElHefnawi M, Seoudi DM, Ali MA (2023) Evaluation of predicted siRNA as an antiviral against MERS-CoV targeting the membrane gene in the vero cell line. Microbiology Research 14: 1687–1701. https://doi.org/10.3390/microbiolres14040116
  • Fakhr E, Zare F, Teimoori-Toolabi L (2016) Precise and efficient siRNA design: a key point in competent gene silencing. Cancer Gene Therapy 23: 73–82. https://doi.org/10.1038/cgt.2016.4
  • Farazi PA, DePinho RA (2006) Hepatocellular carcinoma pathogenesis: from genes to environment. Nature Reviews Cancer 6: 674–687. https://doi.org/10.1038/nrc1934
  • Heo L, Park H, Seok C (2013) GalaxyRefine: protein structure refinement driven by side-chain repacking. Nucleic Acids Research 41: W384–W388. https://doi.org/10.1093/nar/gkt458
  • Irshad M, Gupta P, Irshad K (2017) Molecular basis of hepatocellular carcinoma induced by hepatitis C virus infection. World Journal of Hepatology 9: 1305–1314. https://doi.org/10.4254/wjh.v9.i36.1305
  • Katoh T, Suzuki T (2007) Specific residues at every third position of siRNA shape its efficient RNAi activity. Nucleic Acids Research 35: e27. https://doi.org/10.1093/nar/gkl1120
  • Laskowski RA, MacArthur MW, Moss DS, Thornton JM (1993) PROCHECK: a program to check the stereochemical quality of protein structures. Journal of Applied Crystallography 26: 283–291. https://doi.org/10.1107/S0021889892009944
  • Laskowski RA, Jabłońska J, Pravda L, Vařeková RS, Thornton JM (2018) PDBsum: structural summaries of PDB entries. Protein Science 27: 129–134. https://doi.org/10.1002/pro.3289
  • Liu Q, Zhou H, Cui J, Cao Z, Xu Y (2012) Reconsideration of in-silico siRNA design based on feature selection: a cross-platform data integration perspective. PLOS ONE 7: e37879. https://doi.org/10.1371/journal.pone.0037879
  • Llovet JM, Kelley RK, Villanueva A, Singal AG, Pikarsky E, Roayaie S, Lencioni R, Koike K, Zucman-Rossi J, Finn RS (2021) Hepatocellular carcinoma. Nature Reviews Disease Primers 7: 6. https://doi.org/10.1038/s41572-020-00240-3
  • Lu ZJ, Mathews DH (2008) OligoWalk: an online siRNA design tool utilizing hybridization thermodynamics. Nucleic Acids Research 36: W104–W108. https://doi.org/10.1093/nar/gkn250
  • Lu ZJ, Gloor JW, Mathews DH (2009) Improved RNA secondary structure prediction by maximizing expected pair accuracy. RNA 15: 1805–1813. https://doi.org/10.1261/rna.1643609
  • Martin FJ, Amode MR, Aneja A, Austine-Orimoloye O, Azov AG, Barnes I, Becker A, Bennett R, Berry A, Bhai J, Bhurji SK, Bignell A, Boddu S, Branco Lins PR, Brooks L, Ramaraju SB, Charkhchi M, Cockburn A, Da Rin Fiorretto L, Davidson C, Dodiya K, Donaldson S, El Houdaigui B, El Naboulsi T, Fatima R, Giron CG, Genez T, Ghattaoraya GS, Martinez JG, Guijarro C, Hardy M, Hollis Z, Hourlier T, Hunt T, Kay M, Kaykala V, Le T, Lemos D, Marques-Coelho D, Marugán JC, Merino GA, Mirabueno LP, Mushtaq A, Hossain SN, Ogeh DN, Sakthivel MP, Parker A, Perry M, Piližota I, Prosovetskaia I, Pérez-Silva JG, Salam AIA, Saraiva-Agostinho N, Schuilenburg H, Sheppard D, Sinha S, Sipos B, Stark W, Steed E, Sukumaran R, Sumathipala D, Suner M-M, Surapaneni L, Sutinen K, Szpak M, Tricomi FF, Urbina-Gómez D, Veidenberg A, Walsh TA, Walts B, Wass E, Willhoft N, Allen J, Alvarez-Jarreta J, Chakiachvili M, Flint B, Giorgetti S, Haggerty L, Ilsley GR, Loveland JE, Moore B, Mudge JM, Tate J, Thybert D, Trevanion SJ, Winterbottom A, Frankish A, Hunt SE, Ruffier M, Cunningham F, Dyer S, Finn RD, Howe KL, Harrison PW, Yates AD, Flicek P (2023) Ensembl 2023. Nucleic Acids Research 51: D933–D941. https://doi.org/10.1093/nar/gkac958
  • Naito Y, Yoshimura J, Morishita S, Ui-Tei K (2009) siDirect 2.0: updated software for designing functional siRNA with reduced seed-dependent off-target effect. BMC Bioinformatics 10: 392. https://doi.org/10.1186/1471-2105-10-392
  • Naito Y, Yamada T, Ui-Tei K, Morishita S, Saigo K (2004) siDirect: highly effective, target-specific siRNA design software for mammalian RNA interference. Nucleic Acids Research 32: W124–W129. https://doi.org/10.1093/nar/gkh442
  • Panneerselvam S, Wilson C, Kumar P, Abirami D, Pamarthi J, Reddy MS, Varghese J (2023) Overview of hepatocellular carcinoma: from molecular aspects to future therapeutic options. Cell Adhesion & Migration 17: 1–21. https://doi.org/10.1080/19336918.2023.2258539
  • Reynolds A, Leake D, Boese Q, Scaringe S, Marshall WS, Khvorova A (2004) Rational siRNA design for RNA interference. Nature Biotechnology 22: 326–330. https://doi.org/10.1038/nbt936
  • Sargsyan K, Grauffel C, Lim C (2017) How molecular size impacts RMSD applications in molecular dynamics simulations. Journal of Chemical Theory and Computation 13: 1518–1524. https://doi.org/10.1021/acs.jctc.7b00028
  • Shiels MS, Engels EA, Yanik EL, McGlynn KA, Pfeiffer RM, O’Brien TR (2019) Incidence of hepatocellular carcinoma among older Americans attributable to hepatitis C and hepatitis B: 2001 through 2013. Cancer 125: 2621–2630. https://doi.org/10.1002/cncr.32129
  • Sippl MJ (1993) Recognition of errors in three‐dimensional structures of proteins. Proteins: Structure, Function, and Bioinformatics 17: 355–362. https://doi.org/10.1002/prot.340170404
  • Sohrab SS, Aly El-Kafrawy S, Ibraheem Azhar E (2022) In silico prediction and experimental evaluation of potential siRNAs against SARS-CoV-2 inhibition in Vero E6 cells. Journal of King Saud University - Science 34: 102049. https://doi.org/10.1016/j.jksus.2022.102049
  • Sohrab SS, Aly El-Kafrawy S, Mirza Z, Hassan AM, Alsaqaf F, Azhar EI (2021) In silico prediction and experimental validation of siRNAs targeting ORF1ab of MERS-CoV in Vero cell line. Saudi Journal of Biological Sciences 28: 1348–1355. https://doi.org/10.1016/j.sjbs.2020.11.066
  • Sumitha A, Devi PB, Hari S, Dhanasekaran R (2020) Covid-19 - in silico structure prediction and molecular docking studies with doxycycline and quinine. Biomedical and Pharmacology Journal 13: 1185–1193. https://doi.org/10.13005/bpj/1986
  • Tornesello ML, Buonaguro L, Tatangelo F, Botti G, Izzo F, Buonaguro FM (2013) Mutations in TP53, CTNNB1 and PIK3CA genes in hepatocellular carcinoma associated with hepatitis B and hepatitis C virus infections. Genomics 102: 74–83. https://doi.org/10.1016/j.ygeno.2013.04.001
  • Ui-Tei K (2004) Guidelines for the selection of highly effective siRNA sequences for mammalian and chick RNA interference. Nucleic Acids Research 32: 936–948. https://doi.org/10.1093/nar/gkh247
  • Valdés-Tresanco MS, Valdés-Tresanco ME, Valiente PA, Moreno E (2021) gmx_MMPBSA: a new tool to perform end-state free energy calculations with GROMACS. Journal of Chemical Theory and Computation 17: 6281–6291. https://doi.org/10.1021/acs.jctc.1c00645
  • Vanommeslaeghe K, Hatcher E, Acharya C, Kundu S, Zhong S, Shim J, Darian E, Guvench O, Lopes P, Vorobyov I, Mackerell AD (2010) CHARMM general force field: a force field for drug‐like molecules compatible with the CHARMM all‐atom additive biological force fields. Journal of Computational Chemistry 31: 671–690. https://doi.org/10.1002/jcc.21367
  • Waterhouse A, Bertoni M, Bienert S, Studer G, Tauriello G, Gumienny R, Heer FT, de Beer TAP, Rempfer C, Bordoli L, Lepore R, Schwede T (2018) SWISS-MODEL: homology modelling of protein structures and complexes. Nucleic Acids Research 46: W296–W303. https://doi.org/10.1093/nar/gky427
  • Wiederstein M, Sippl MJ (2007) ProSA-web: interactive web service for the recognition of errors in three-dimensional structures of proteins. Nucleic Acids Research 35: W407–W410. https://doi.org/10.1093/nar/gkm290
  • Yasmin T, Adiba M, Saba A Al, Nabi AN (2022) In silico Design of siRNAs for silencing CLEC5A receptor as a potential therapeutic approach against dengue and Japanese encephalitis virus infection in human. Bioinformatics and Biology Insights 16: 117793222211421. https://doi.org/10.1177/11779322221142122
  • Yeh S-H, Li C-L, Lin Y-Y, Ho M-C, Wang Y-C, Tseng S-T, Chen P-J (2023) Hepatitis B virus DNA integration drives carcinogenesis and provides a new biomarker for HBV-related HCC. Cellular and Molecular Gastroenterology and Hepatology 15: 921–929. https://doi.org/10.1016/j.jcmgh.2023.01.001
  • Yilmazel B, Hu Y, Sigoillot F, Smith JA, Shamu CE, Perrimon N, Mohr SE (2014) Online GESS: prediction of miRNA-like off-target effects in large-scale RNAi screen data by seed region analysis. BMC Bioinformatics 15: 192. https://doi.org/10.1186/1471-2105-15-192
  • Zeng G, Apte U, Cieply B, Singh S, Monga SPS (2007) siRNA-mediated β-catenin knockdown in human hepatoma cells results in decreased growth and survival. Neoplasia 9: 951–959. https://doi.org/10.1593/neo.07469

Supplementary materials

Supplementary material 1 

Supplementary information

Lauren Emily Fajardo, Mark Andrian Macalalad, Nyzar Mabeth Odchimar, John Christian de Guzman, Fredmoore Orosco

Data type: pdf

Explanation note: table S1. CTNNB1 CDS transcript variants retrieved from Ensembl. table S2. Detailed Sfold results of top four siRNA candidates. fig. S1. Concentration plots (top) and heat capacity plots (bottom) of mRNA-siRNA duplexes of top four siRNA candidates. (A) s3, (B) s19, (C) s22, (D) s28. fig. S2. ERRAT plot of the predicted Ago2 tertiary structure showing error value for each residue. fig. S3. Ramachandran plot of the refined Ago2 model. fig. S3. Ramachandran plot of the refined Ago2 model.

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (399.71 kb)
Supplementary material 2 

s3-Ago2 MD simulation

Lauren Emily Fajardo, Mark Andrian Macalalad, Nyzar Mabeth Odchimar, John Christian de Guzman, Fredmoore Orosco

Data type: mov

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (14.77 MB)
Supplementary material 3 

s19-Ago2 MD simulation

Lauren Emily Fajardo, Mark Andrian Macalalad, Nyzar Mabeth Odchimar, John Christian de Guzman, Fredmoore Orosco

Data type: mov

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (14.21 MB)
Supplementary material 4 

s22-Ago2 MD simulation

Lauren Emily Fajardo, Mark Andrian Macalalad, Nyzar Mabeth Odchimar, John Christian de Guzman, Fredmoore Orosco

Data type: mov

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (15.37 MB)
Supplementary material 5 

s28-Ago2 MD simulation

Lauren Emily Fajardo, Mark Andrian Macalalad, Nyzar Mabeth Odchimar, John Christian de Guzman, Fredmoore Orosco

Data type: mov

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (10.26 MB)
login to comment