Next Article in Journal
Space Charge Characteristics at the Interface of Laminated Epoxy Resin
Next Article in Special Issue
Naive Prediction of Protein Backbone Phi and Psi Dihedral Angles Using Deep Learning
Previous Article in Journal
The Synthesis and Polymer-Reinforced Mechanical Properties of SiO2 Aerogels: A Review
Previous Article in Special Issue
Recent Advances in Deep Learning for Protein-Protein Interaction Analysis: A Comprehensive Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

RNA 3D Structure Prediction: Progress and Perspective

1
Department of Physics, Key Laboratory of Artificial Micro & Nano-Structures of Ministry of Education, School of Physics and Technology, Wuhan University, Wuhan 430072, China
2
School of Bioengineering and Health, Wuhan Textile University, Wuhan 430200, China
3
Research Center of Nonlinear Science, School of Mathematical and Physical Sciences, Wuhan Textile University, Wuhan 430200, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(14), 5532; https://doi.org/10.3390/molecules28145532
Submission received: 14 June 2023 / Revised: 5 July 2023 / Accepted: 13 July 2023 / Published: 20 July 2023

Abstract

:
Ribonucleic acid (RNA) molecules play vital roles in numerous important biological functions such as catalysis and gene regulation. The functions of RNAs are strongly coupled to their structures or proper structure changes, and RNA structure prediction has been paid much attention in the last two decades. Some computational models have been developed to predict RNA three-dimensional (3D) structures in silico, and these models are generally composed of predicting RNA 3D structure ensemble, evaluating near-native RNAs from the structure ensemble, and refining the identified RNAs. In this review, we will make a comprehensive overview of the recent advances in RNA 3D structure modeling, including structure ensemble prediction, evaluation, and refinement. Finally, we will emphasize some insights and perspectives in modeling RNA 3D structures.

1. Introduction

RNAs are a kind of fundamental biological macromolecule, in addition to proteins and DNAs. Numerous studies have demonstrated that RNA has many essential biological functions, such as regulating gene expressions [1], protein biosynthesis regulations [2], and catalytic biological reactions [3]. Furthermore, the multi-functional nature of RNAs significantly contributes to the design of related nanoscale biomedical and technological applications [4]. Generally, RNAs need to fold into their specific 3D structures to perform their specific functions [5,6,7]. For instance, ribozymes can only catalyze reaction functions when they fold into their native 3D structures [6], while riboswitches regulate gene expressions through dynamic changes in the 3D structures upon metabolite binding [5,7]. Therefore, the comprehensive knowledge of RNA 3D structures is of great significance for understanding and utilizing RNA biological functions.
Until now, experimental methods such as X-ray crystallography, nuclear magnetic resonance spectroscopy, and cryo-electron microscopy have been predominantly used to derive RNA 3D structures [8]. However, these methods are generally time-consuming and laborious, especially for RNAs with long sequences or complex structures [9]. Thus, the RNA 3D structures deposited in Protein Data Bank (PDB) [8] are still very limited relative to the vast number of RNA sequences stored in the central RNA [10] and the large number of 3D structures of proteins [8,9,10,11]. For example, there are only 1732 isolated RNA structures in the PDB database, while there are 34 million RNA sequences in the RNA central and 188,726 3D structures of proteins in the PDB database (until 23 May 2023). Hence, it is essentially necessary to develop computational models for modeling RNA 3D structures at a high resolution.
In recent decades, various computational models have been developed to predict RNA 3D structures [11,12,13,14,15], and the models are generally composed of three procedures: predicting 3D structure ensembles, identifying near-native structures that are close to the experimentally determined native structures through a reliable scoring function/statistical potential, and refining identified near-native structures [11,15]. Here, we will go through them.
First, a computational model generates an ensemble of RNA 3D structure candidates based on input information, such as RNA sequences solely or RNA sequences and secondary structures for some models [15]. For improving the prediction accuracy, some models can also involve additional constraint information, such as experimental information and the distance between specific atom pairs. Through physics-based force field and conformation sampling, or knowledge-based fragment assembly, or deep-learning-based information, a computational model can generate a structure candidate ensemble with near-native structures. Moreover, generally, these prediction models can generate a structure ensemble for different types of RNA structures based on sequence information or secondary structure information, such as hairpins, hairpins/duplexes with bulges/internal loops, pseudoknots, kissing complexes, and multi-way junctions. However, it is still difficult for existing models to predict the 3D structures of RNA G-quadruplexes from their sequences [11,15], while some models can be used to study RNA G-quadruplexes [16], i.e., HiRE-RNA can almost stabilize the native 3D structure of an RNA G-quadruplexes for 3 µs and a convolutional neural network-based model can identify potential RNA G-quadruplexes from transcriptomics data [17,18]. Moreover, some computational models display a competitive performance in RNA-Puzzles [19,20,21,22,23] and CASP-RNA [24]. Here, RNA-Puzzles is a CASP-like competition for RNA 3D structure prediction [19,20,21,22,23], and CASP-RNA is a prediction competition for RNA 3D structures newly present in CASP15 [24]. Depending on the methods for generating the 3D structure ensemble, the existing computational models can be roughly divided into physics-based [25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43], knowledge-based fragment assembly [44,45,46,47,48,49,50,51,52,53,54,55,56], and deep-learning-based ones [57,58,59,60]; see Figure 1.
Second, a structure evaluation is required to identify the top near-native structures from the precedingly predicted 3D structure candidate ensemble [61,62], and a high-quality scoring function/energy function for structure evaluation should be able to identify the candidate structures closest to the native ones in the predicted ensemble [61,62]. For a blind prediction of an RNA 3D structure, a reliable scoring function/statistical potential is definitely required as the native structure of the target RNA is not available. These scoring functions/energy functions are generally developed based on the native structures in the PDB database, which would contain the major structure features of the native structures and would provide a reliable structure evaluation to identify the top structures in the predicted structure ensemble, even if the native structures are not available, and the existing scoring functions/energy functions for RNA 3D structure evaluation typically include physics-based energy functions [26], knowledge-based statistical potentials [63,64,65,66,67,68,69], deep-learning-based scoring functions [70,71], and clustering-involved algorithms [36,37].
Finally, structure refinement may need to be performed for the identified top structures so as to obtain high-quality RNA 3D structures as the structures may contain some unreasonable defects, such as unnatural bonds and serious conflicts between atoms [72,73,74]. In addition to modifying the local defects in RNA structures, a good refinement approach can also improve the global 3D RNA structures [73,75], making the structures to be overall closer to their native ones.
In this review, we provide a comprehensive overview of the recent advances in modeling RNA 3D structures, including structure ensemble prediction, structure evaluation, and structure refinement. The main text is organized as follows. First, we give a detailed overview of the existing RNA 3D structure prediction models, including physics-based, knowledge-based, and deep-learning-based models. Second, we describe the recent progress in RNA 3D structure evaluation. Third, we introduce recently developed methods for RNA 3D structure refinement. Finally, we discuss the challenges in modeling RNA 3D structures and offer some insights in modeling RNA 3D structures.

2. RNA 3D Structure Prediction Models

2.1. Physics-Based Models

The physics-based models are based on the fundamental physics principle that the native structures are those states with the lowest energies for RNAs [15]. In principle, all-atom molecular dynamics simulations such as Amber [76,77,78] and Charmm [79,80] can be used to predict RNA 3D structures, while due to the huge computation from the all-atom representation of RNAs, solvents, and ions, such all-atom-based simulation methods are only limited to very small RNAs with very simple topological structures such as duplex and hairpin [81]. To reduce the atom-representation-concerned complexity, some physics-based coarse-grained (CG) models with varying CG levels have been developed to predict RNA 3D structures in a reasonable time [25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42]. The typical CG representations are illustrated in Figure 2, and the existing physics-based CG models are summarized in Table 1, as well as the corresponding references and the available websites [25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42]. Generally, a physics-based CG model guides RNA structure folding through a specific force field (energy function) and a conformation sampling algorithm such as Monte Carlo (MC) [82] or molecular dynamics (MD) sampling [83,84]. A CG force field in these models is generally composed of bonded and non-bonded energy terms [30,31,32,33,34,35].
E = E bond + E angle + E dihedral   angle + E exclusion   volume + E base   paring + E base   stacking ,
where the bonded energy term includes bond length, bond angle, and dihedral/torsion angle energies, and the non-bonded energy term usually includes exclusion volume, base pairing, and base stacking energies. In the following, we mainly introduce physics-based predictive models with different CG representations for RNA 3D structures rather than the all-atom molecular dynamic simulations.

2.1.1. One-Bead Nucleotide Model

YUP. In the YUP model, a bead located at the P atom is used to represent a nucleotide. Through the one-bead energy function and MC sampling algorithm, YUP can be used to simulate RNA structure folding, and the structure with the lowest energy in the conformational ensemble is chosen as the predicted final structure [25,26]. Although the model may require additional experimental information such as a secondary structure to predict RNA 3D structures, YUP has been developed into an adaptive package for the automatic CG modeling of RNA structures. Furthermore, the model can also predict protein and DNA structures with additional constraint information. In addition, the YUP model can be found at http://rumour.biology.gatech.edu/YammpWeb/.
NAST. Similar to YUP [26], NAST is also a one-bead CG model, and a bead located at the C3 atom is used to represent a whole nucleotide [27]. The model utilizes RNA-specific knowledge-based potential energy and MD algorithms to predict the RNA 3D structures, and the predicted structures can be obtained through the clustering method for 1500 low-energy conformations. By involving the secondary and tertiary contact information, NAST can predict the RNA 3D structures. Moreover, NAST predicts the 3D structures of the yeast phenylalanine tRNA (76 nt) and the P4-P6 domain of the Tetrahymena thermophile group I intron (158 nt) with ~8 Å and ~16 Å RMSDs from the native structures, respectively [26]. In addition, NAST can combine small-angle X-ray scattering data [85] and experimental solvent accessibility data as filters to rank the clusters of similar structures. The package of NAST can be downloaded at https://simtk.org/home/nast.

2.1.2. Three-Bead Nucleotide Model

iFoldRNA/iFoldRNA2. In the iFoldRNA model, three beads are used to represent each nucleotide: the three beads are located at the mass center of a phosphate group, at the center of the five-atom ring of a sugar, and at the center of the six-atom ring of a base [28,29]. Moreover, the bonded terms are composed of bond length, bond angle, and dihedral angle interactions, and non-bonded terms are composed of base pairing, base stacking, short-range phosphate–phosphate repulsion, and hydrophobic interactions. Through the discrete MD algorithm and clustering method for low-energy conformations, iFoldRNA can predict RNA 3D structures from their sequences. iFoldRNA allows for the prediction of the structure of short RNAs (<50 nt) within 4 Å RMSD from the corresponding experimental structures. Additionally, iFoldRNA can predict complex RNAs by involving experimental information such as base-pairing and hydroxyl-radical probing. iFoldRNA has been developed into the iFoldRNA2 [29] webserver at http://ifoldrna.dokhlab.org.
A CG model with salt effect. A three-bead CG model with a salt effect has been developed for predicting RNA 3D structures in the presence of monovalent/divalent ions [30,31,32,33,34,35] due to the polyanionic nature of RNAs [86,87,88]. In contrast, most existing 3D structure prediction models are generally focused on predicting the 3D structures of RNAs, and the effects of salt ions and temperature are seldom involved in the models. In the three-bead CG model with salt effect, the three beads are located at P, C4 , and N9 (N1) atoms of purine (pyrimidine), respectively. In the energy function of the model, the salt effect is implicitly involved through a combination of the counterion condensation theory [89] and the tightly bound ion model [90,91,92], and the effect of temperature is accounted for through involving the experimental parameters of base stacks from Turner et al. [93]. Using the MC simulated annealing or replica-exchange MC (REMC) algorithm [94], the model can predict 3D structures for RNA hairpins, kissing complexes, and minimal H-type pseudoknots and complex pseudoknots from sequences and the thermal stabilities of the RNAs in monovalent/divalent ion solutions, including SARS-CoV-2 programming-1 ribosomal frameshifting element and Zika virus xrRNA [30,31,32,33,34,35]. The mean RMSD of the predicted structures is less than 7 Å for RNAs (<~90 nt) [30,31,32,33,34,35]. In addition, the mean deviations of the melting temperatures from the experimental data are ~2 °C for RNA pseudoknots and kissing complexes [30,31,32,33,34,35] over extensive Na+/Mg2+ concentrations. However, the model still cannot make good predictions for RNAs with multi-way ( 3) junctions. Anyway, this model can capture the effects of salt and temperature on predicted RNA 3D structures, which are seldom captured by other existing RNA 3D prediction models.

2.1.3. Five-Bead Nucleotide Model

SimRNA. In the SimRNA model, the five beads are located at the P atom, C4 atom of the sugar, C2 atom, N1 atom, and C4 atoms of pyrimidine (N9 and C6 atoms of purine) [36,37]. The energy function of SimRNA comprises sequence-independent local-ranged terms and sequence-dependent long-ranged terms. The local terms consist of bond length, flat angle, and torsional angle interactions, and the long-ranged terms include base–base, base–backbone, and backbone–backbone interactions [36]. Using the REMC algorithm, SimRNA can predict RNA 3D structures solely from sequences in a reasonable time, and the predicted final structures are obtained by the clustering method for the low-energy conformations. Additionally, secondary structures or other constraint information can be added into SimRNA to improve the prediction accuracy and efficiency, especially for complex RNAs. SimRNA can effectively predict RNA with relatively short RNA sequences (< ~50 nt). However, for RNAs with longer sequences (>70 nt), the prediction accuracy of the model can be improved if the secondary structures are provided [36]. SimRNA has been developed into the versions of software package and webserver, which can be found at https://genesilico.pl/SimRNAweb.
IsRNA. In the IsRNA1 model, four/five beads are used to represent a nucleotide [38,39,40]. Specifically, two beads are located at the P atom of the phosphate group and the C4 atom of sugar, and three/two beads are positioned at the center of mass of the constituent heavy-atom groups to represent purine/pyrimidine bases, respectively. The energy functions of IsRNA1 [39] were derived through an iterative simulated reference state approach [38,39], and the MD or REMD algorithm was employed to accelerate conformational sampling. In the model, the predicted final structures are obtained by a clustering method for the 10% conformations with the lowest energy in the candidate ensemble. IsRNA1 can predict the structures of small RNAs based on their sequences and more complex RNAs based on their secondary structures. IsRNA1 has been examined against a large-scale benchmark dataset (containing 40–161 nts) with secondary structure constraints. The dataset includes 44 stem loops, 43 multi-way junctions, and 43 structures of long-range tertiary interactions, and the mean RMSD of the predicted top-1 structure is less than 10 Å [39]. Moreover, the model can be used to optimize the structures generated by two fragment assembly methods [54,55]. Very recently, IsRNA was developed into IsRNA2 to predict 3D structures for RNAs with noncanonical base pairs [40], and the web server of IsRNA is available at http://rna.physics.missouri.edu/IsRNA/index.html.
RNAJP. Very recently, Li et al. developed a five-bead CG model called RNAJP to predict RNA 3D structures, especially multi-way junction structures [41]. Specifically, two CG beads are placed at the P and C4 atoms, representing the phosphate and sugar groups, and three beads are located at the N1/N9, C2, and C4/C6 atoms for pyrimidine/purine, respectively. In addition to the conventional bonded and non-bonded energy terms, similar to other physics-based CG models discussed above, the RNAJP model explicitly considers the interactions between adjacent helices and between adjacent strands in the junction structures and the long-ranged interactions between loops. With the use of the toolkit OpenMM [95], RNAJP can reliably predict RNA 3D structures with secondary structures as the input, and the top-1 structure predicted by RNAJP is identified through a specific energy function [41]. Although the model cannot predict RNA 3D structures solely from sequences, it can predict high-quality RNA 3D structures with given sequences and their corresponding secondary structures, especially for three- and four-way junctions, and the predictions of RNAJP can reach a mean RMSD of ~8 Å for the 22 RNA three-way junction structures (53–160 nts) and 5 RNA four-way junction structures (68–155 nts). However, the model cannot create reliable predictions for five-way or higher-way junctions due to the extremely high structural complexity [41]. The source code of the RNAJP model can be downloaded at http://rna.physics.missouri.edu/RNAJP/index.html.

2.1.4. Six/Seven-Bead Nucleotide Model

HiRE-RNA. The HiRE-RNA model is a high-resolution CG model that uses six to seven beads to represent a nucleotide [42]. Specifically, five CG beads are assigned to the phosphate group and sugar rings (namely, P, O5 , C5 , C4 , and C1 atoms), and another one/two beads are located at the mass centers of nonhydrogen atoms of the pyrimidine/purine base. In the force field of HiRE-RNA, harmonic form potentials were used to describe the bond length and bond angle energies in the bonded terms, while the torsional angle energy was described with a cosine form potential. Moreover, non-bonded terms include exclusion volume, electrostatic, base pairing, and base stacking potentials. Using the REMD algorithm [96] for sampling conformations and the clustering method for low-energy conformations, HiRE-RNA can predict RNA 3D structures from sequences and complex RNAs with secondary structure information, and HiRE-RNA has been tested on 13 RNAs (12–71 nts), and the RMSDs are less than ~8 Å from the corresponding native structures [42]. Additionally, HiRE-RNA can be used to examine the stability of RNA hairpins and duplexes, while the energy function of the model may require further optimization to improve the predictions for the thermodynamic properties of RNAs.

2.1.5. Coarse-Grain Helix-Centered Model

Ernwin. Unlike the above models, Ernwin is a helix-centered CG model, which uses one line segment and two vectors to represent a helix and considers elements (loops) linking helices as the degrees of freedom [43]. The energy function of Ernwin comprises five separate terms: two terms attributed to clash and junction closure, which serve as constraint energies to exclude impossible structures based on physical forces; the other three terms accounting for radius of gyration, A-minor energy, and loop–loop interaction energy serve as non-constraint energies, which are knowledge-based potentials derived from the known structures in the PDB database [8]. Combining the energy function with Markov chain Monte Carlo simulation, the Ernwin model can efficiently predict RNA 3D structures based on their secondary structures, and the structure with the lowest energy in the conformational ensemble is chosen as the predicted final structure. The package of the model is freely available at http://github.com/pkerpedjiev/ernwin.
As described above, an important advantage of various physics-based CG models is that they can predict RNA 3D structures solely from sequences based on different specific force fields and various MC/REMC or MD/REMD conformation sampling algorithms. However, even with such CG simplification for nucleotides, to predict the 3D structures of large RNA at high accuracy solely from sequences remains challenging for a physics-based CG model. The reliable predictions from the existing physics-based CG models are still limited to RNAs of medium size (<~90 nt with inputting only sequences or <~150 nt with inputting secondary structures). Namely, the accuracy of the CG force field and the conformation sampling algorithm are still required to be refined and developed to improve the prediction accuracy and efficiency. In addition, due to the polyanionic nature of RNAs, the structures and stabilities of RNAs are usually sensitive to solution ion conditions, especially to multivalent ions [90], while the effect of ions such as Mg2+ are rarely involved in the existing 3D structure prediction models. Although the effects of monovalent/divalent ions on RNA thermodynamics can be captured by the three interaction sites (TIS) model developed by Thirumalai et al. [97,98,99], the TIS model is a Gö-like model and cannot make predictions for RNA 3D structures from sequences. The CG model with salt effect proposed by Tan et al. can create reliable predictions for the thermal stabilities of (complex) RNA pseudoknots in monovalent/divalent ions [33,35], while it is still a challenge for this model to predict more complex RNA structures (e.g., multi-way junctions). It is necessary to develop a physics-based CG model to efficiently and reliably predict 3D structures and stabilities of complex RNA structures in monovalent/divalent ion solutions.

2.2. Knowledge-Based Fragment Assembly Models

As macromolecule 3D structures evolve much slower than their sequences, evolutionarily related macromolecules typically retain similar 3D structures, despite their differences at the sequence level [11,15]. Moreover, the number of experimental structures deposited in the PDB database increases gradually with the development of various experimental technologies [11]. Therefore, RNA 3D structures can be predicted through so-called knowledge-based methods [11,15]. The early knowledge-based methods were graphics-based methods such as MANIP [100] and RNA2D3D [101], which could intuitively and quickly predict large-scale RNA 3D structures, but required users to have professional knowledge of a high level. Later, some fully automated fragment-assembly methods were developed and widely employed for RNA 3D structure predictions [44,45,46,47,48,49,50,51,52,53,54,55,56]. Typically, the fragment assembly models predict RNA 3D structures based on RNA sequences and secondary structures as inputs. Figure 2 shows the typical schematic diagram for existing fragment assembly models, which can be roughly classified into two categories based on the size of the fragments used: small motifs as fragments and medium motifs as fragments. In the following, we introduce the existing knowledge-based, fully automated fragment assembly models for RNA 3D structures, which are summarized in Table 2.

2.2.1. Small Motifs as Fragments

FARNA/FARFAR/FARFAR2. Das et al. proposed a fragment assembly model named FARNA/FARFAR for predicting RNA 3D structures based on three-nucleotide fragments [44,45]. FARNA/FARFAR assemble RNA 3D structures guided by the specific knowledge-based energy function and MC algorithm. As an early RNA 3D structure assembly model, FARNA/FARFAR has a good prediction accuracy for small-size RNAs. Later, Das et al. proposed FARFAR2 [46], in which four treatments were implemented to improve the prediction accuracy and efficiency, including (a) an updated fragment library, (b) a fractional filter for the fragment assembly process, (c) a special Monte Carlo movement, and (d) a new all-atom scoring function. Compared with FARFAR, FARFAR2 recovers near-native structures more accurately and predicts the 3D structures of adenovirus virus-associated RNA and five riboswitch complexes with RMSDs of ~3–14 Å [46]. FARFAR2 has been developed into a user-friendly webserver at https://rosie.rosettacommons.org/farfar2.

2.2.2. Medium Motifs as Fragments

MC-Fold/MC-Sym. Parisien et al. proposed an RNA structure prediction pipeline consisting of two computational models: MC-Fold and MC-Sym [47]. In the pipeline, MC-Fold predicts RNA secondary structures, and then MC-Sym assembles an ensemble of 3D structures based on the secondary structures. Different thermodynamic models such as Mfold [103], MC-Fold can predict secondary structures for RNAs with noncanonical base pairs with the use of knowledge-based scoring functions based on NCM (nucleotide cyclic motif) databases. The NCM database contains lone-pair loops up to six nucleotides, including the flanking lone base pair, and double-stranded NCMs up to eight nucleotides including both flanking base pairs. MC-Sym performs the fragment insertion simulation with the 3D NCMs and the Las Vegas algorithm. The MC-Fold/MC-Sym pipeline has been tested by building 3D structures of precursor microRNA and proposing a new structure of the human immunodeficiency virus (HIV) cis-1 frameshift segment, and the model has been tested against 13 different types of RNAs (29–47 nts) with a mean RMSD of ~2 Å [47]. The web server of the MC-Fold/MC-sym pipeline can be found at http://www.major.iric.ca.
RNAComposer. Popenda et al. developed a fast and fully automated fragment assembly model named RNAComposer for predicting RNA 3D structures based on secondary structures [48,49]. The model employs smallest secondary elements (SSE) as blocks, which helps the model to achieve RNA 3D structure prediction at a relatively high accuracy. The predicted top-1 structure in RNAComposer is assembled from the best fragments, which are evaluated based on four criteria: secondary structure topology, sequence similarity, source structure resolution, and energy based on the Charmm force field. RNAComposer was examined for 40 different types of RNAs (31–161 nts), which include hairpins without/with internal loops, pseudoknots, and multi-way junctions; the mean RMSD of the top-1 structure is ~5 Å [48]. RNAComposer has been developed into a user-friendly webserver version at http://rnacomposer.ibch.poznan.pl.
3dRNA. Similar to RNAComposer [48], 3dRNA also employs SSEs as building blocks to predict RNA 3D structures [50,51,52,53], while in 3dRNA, the 3D structures of the SSEs extracted from the experimental structures contain one more base pair at their 5 -end for more accurate superposition between different fragments in the global structure assembly. Notably, the assembled 3D structures can be further optimized by a specific CG energy function with the information from direct coupling analysis [75]. In 3dRNA, the final predicted structure can be identified by the clustering algorithm for the assembled structures and a specific scoring function named 3dRNAscore [68] to rank the candidates in the different clusters. 3dRNA exhibits relatively high prediction accuracy for different types of RNA structures, particularly pseudoknots and large RNAs with tertiary contacts. 3dRNA has been tested against extensive RNAs including five very large RNAs (500–3000 nts), and it is encouraging that the predicted structures of four of the five large RNAs have good performance with RMSDs within 15 Å [52]. In addition, 3dRNA can predict the 3D structures of circular RNAs [53]. 3dRNA has been developed into a friendly webserver version at http://biophy.hust.edu.cn/new/3dRNA.
Vfold3D/VfoldLA. Cao et al. proposed two different fragment assembly models, namely Vfold3D [54] and VfoldLA [55]. With a given secondary structure, Vfold3D automatically assembles RNA 3D structures with the 3D fragments from the PDB database based on resolved secondary motifs such as hairpin loops and multi-way junction loops [54]. However, Vfold3D is limited by the small number of known RNA structures, especially for the templates of multi-way junctions. Xu et al. developed a new model of VfoldLA with template search and the assembly algorithm to build RNA 3D structures [55], where templates with single-stranded loops/junctions are searched for instead of whole motifs, and a whole multi-junction loop can be assembled based on single-stranded loops in the absence of such a whole multi-junction loop [55]. Furthermore, a hybrid method has been proposed that combines Vfold3D and VfoldLA to predict the RNA 3D structures [104]. The hybrid method is focused on the definition of motifs and loops, the processing of template-free motifs, and the 3D structure assembly based on motifs and loops templates [104]. Moreover, the Vfold-pipeline was developed by integrating Vfold2D (secondary structure prediction model), Vfold3D, and VfoldLA [105].The pipeline has been examined on 92 RNAs, including hairpin/internal loops, 3-, 4-, and 5-way junctions, and pseudoknots, and the mean RMSDs of the predicted structure are ~6 Å, ~10 Å, ~7 Å, ~15 Å, and ~16 Å, respectively [106]. Vfold3D/VfoldLA are available for users on their webservers, and the Vfold-pipeline can be found at http://rna.physics.missouri.edu/vfoldPipeline/index.html.
FebRNA. Very recently, Zhou et al. proposed a fragment-ensemble-based model named FebRNA for building RNA 3D structures with secondary structures as the input [56]. This model selects almost all of the templates according to the types of secondary motifs and the lengths, regardless of the sequences and transfers the all-atom fragments into CG ones according to the CG model with salt effect [35], resulting in a global CG 3D candidate ensemble with a vast number of assembled structures (up to ~25,000). This naturally increases the likelihood of including the structures very close to the native structures [56]. Moreover, when building the global structures, FebRNA prefers loop templates with more end base pairs, which helps to better describe the orientation of the stems and improves the accuracy of the final structures. The predicted structures are identified using a specific scoring function from an efficient CG scoring function of cgRNASP [67]. Afterwards, the identified top CG structures are rebuilt into all-atom structures. It has been demonstrated that FebRNA can reliably and efficiently predict the 3D structures of different types of RNAs, including 14 RNA hairpins (17–31 nts), 8 pseudoknots (28–127 nts), 25 multi-way junctions (54–393 nts), and 16 RNA-Puzzles (37–189 nts), and the mean RMSD of the predicted top-1 structures is ~6 Å [56]. The package of FebRNA can be available at https://github.com/Tan-group/FebRNA.
As described above, the knowledge-based fragment assembly model can efficiently predict RNA 3D structures with a relatively high accuracy. However, the performance of the models is strongly dependent on the quality and completeness of the template library and the performance for the structure evaluation method, in addition to secondary structures, which are required as the input. Although the number of RNA 3D structures deposited in the PDB database is continuously increasing, it may still be very difficult to find suitable templates for some special target RNAs, especially for multi-way junction templates, which are generally very crucial for building global RNA structures.

2.3. The Deep-Learning-Based Approaches

The advent of artificial intelligence has significantly advanced science and technology worldwide in recent years. A typical example is AlphaFold2, a deep-learning-based method for accurate protein 3D structure prediction [107,108,109]. However, it is essential to note that accurate prediction of macromolecule 3D structures generally requires extensive experimental structure data [107]. Based on the successful experience of deep learning in protein 3D structure prediction, some deep-learning-based methods have been developed to predict RNA 3D structures, although the available RNA 3D structures in the PDB database are rather limited compared with proteins [8]. In the following, we will introduce four deep-learning-based approaches for RNA 3D structure prediction developed very recently, which have been summarized in Table 3.
RhoFold. Shen et al. proposed an end-to-end deep-learning-based de novo RNA 3D structure prediction approach named RhoFold, which consists of three modules: structure feature extraction, structure prediction, and structure refinement [57]. The feature extraction module combines infernal and rMSA protocol to extract the MSA and pairwise residue features [57]. The structure prediction module, which is the core of RhoFold, predicts the rotation and translation matrices of the main frames based on the sequence and pair presentation from the feature extraction module. Finally, the structure refinement module modifies possible remaining structural conflicts, and the predicted structure is further relaxed through a constraint energy minimization algorithm. The prediction performance of RhoFold has been demonstrated by a test against the non-redundant RNA-Puzzle test dataset, and the average RMSD is less than 4 Å [57]. Additionally, RhoFold can also achieve promising predictions for the 3D structure of RNA complexes. RhoFold is available at https://github.com/RFOLD/RhoFold.
DeepFoldRNA. DeepFoldRNA is a fully automated end-to-end deep-learning-based method for RNA 3D structure prediction, composed of two consecutive modules: a constraint generation module and a structure construction module [58]. In the constraint generation module, multiple sequence alignments (MSAs) of RNAs are collected by iteratively searching multiple nucleic acid sequence databases using rMSA [110]. Afterwards, a self-attention neural network is utilized to predict pairwise distances and inter-residue/backbone torsion angles, and the predicted geometric constraints are transformed into composite potentials using the negative logarithmic likelihood of loading probability prediction in the structure construction module. Through the limited-memory Broyden–Letcher–Goldfarb–Shanno (L-BFGS) minimization algorithm, DeepFoldRNA achieves the end-to-end prediction for RNA 3D structures [58]. DeepFoldRNA was tested against two independent benchmark datasets from Rfam families (105 RNAs) [111] and RNA-Puzzle experiments (17 RNAs), where DeepFoldRNA predicts the structures with a mean RMSD of ~3 Å [58]. DeepFoldRNA is available at https://zhanggroup.org/DeepFoldRNA.
trRosettaRNA. Feng et al. proposed trRosettaRNA [59], a deep-learning-based de novo approach for RNA 3D structure prediction through the transformer network. trRosettaRNA follows a two-step procedure of trRosetta [112,113]: the first step is to predict 1D and 2D geometric shapes with a transformer network [59], and the geometrics include 1D orientations and 2D contacts, distances, and orientations. The second step is to generate 3D structures through energy minimization. The prediction performance of trRosettaRNA was demonstrated on two independent datasets from RNA-Puzzle datasets (30 RNAs) and Rfam families (101 RNAs), and the mean RMSDs of the predicted structures by trRosettaRNA are ~6 Å and <~4 Å, respectively [59]. Notably, trRosettaRNA performs similarly to DeepFoldRNA [58] in the all-atom RMSD, but predicts more realistic side-chain atoms. trRosettaRNA is available at https://yanglab.nankai.edu.cn/trRosettaRNA/.
epRNA. Sha et al. developed a neural network Euclidean parametrization-based method (epRNA) to predict RNA 3D structures solely from sequences, using the state-of-the-art neural network architecture and symmetries [60]. epRNA utilizes the parameterization of Euclidean distance matrices [114] to enable the neural network to directly output the distances between all of the residues. Subsequently, the structure predicted by the neural network is converted into an all-atom structure using DMD with constraints [28,115]. It is noted that epRNA achieves high accuracy predictions on the 3D structures of RNAs of up to 100 nucleotides in length. epRNA is available at https://bitbucket.org/dokhlab/eprna-euclidean-parametrization-of-rna/src/master/.
The above deep-learning-based methods can achieve high-quality and efficient prediction for RNA 3D structures, as tested against the available structure datasets. However, the reliability and performance of a deep-learning-based method strongly relies on the number and the structure spectrum of known RNAs, and due to the limited RNA 3D structures in the PDB database, it would be a great challenge for a deep-learning-based method to make blind predictions for the target RNAs, whose structures do not reside in the known structure spectrum. Nevertheless, these deep-learning models provide an alternative way for modeling RNA 3D structures, in addition to the physics-based methods and traditional knowledge-based models.

3. RNA 3D Structure Evaluation

Generally, a predictive model for RNA 3D structures would generate a candidate ensemble with more than one candidate for a target RNA, and consequently, a high-quality 3D structure evaluation is critical for an RNA 3D structure prediction model [61,62]. Some evaluation methods have been involved in the existing RNA 3D structure prediction models, such as selecting the cluster centers as the representative structures using the clustering algorithm, or directly evaluating the structures by the energies obtained from a statistical potential/scoring function/force field, or combining the clustering algorithm and energy functions/force fields to select the representative near-native structures [26,35,36,41]. However, an efficient scoring function with a high performance is still lacking and consequently is definitely required for RNA 3D structure evaluation. In the last two decades, several statistical potentials/scoring functions have been proposed for RNA 3D structure evaluation, owing to the prior progress made for proteins [61,62], which can be classified into knowledge-based scoring functions/statistical potentials [63,64,65,66,67,68,69] and deep-learning-based scoring functions [70,71], and are summarized in Table 4.

3.1. Knowledge-Based Scoring Functions/Statistical Potentials

In principle, any kind of geometrical parameters, such as distances or angles between atoms/atom groups, which can be utilized to distinguish a native conformation from a decoy one can be adopted to derive a statistical potential/scoring function [61,120]. According to Boltzmann’s law, a general expression of a statistical potential/scoring function can be obtained as a function of the geometrical parameter s [116]:
Δ E ( s ) = k B T ln [ P obs ( s ) P ref ( s ) ] ,  
where k B and T are the Boltzmann constant and the temperature in Kelvin, respectively. P obs ( s ) and P ref ( s ) are the probability of the geometrical parameter s in the native and reference states, respectively. As shown in Equation (2), the reference state and the geometrical parameter s, which may involve two or more atoms, are crucial for building statistical potentials, and the core difference between various statistical potentials is attributed to the choice of them.
The native state is the non-redundant native structure ensemble, while an ideal reference state is a conformation ensemble composed of a non-redundant and complete spectrum of conformations in phase space and without interactions between atoms [120]. Nevertheless, an ideal reference state could not be obtained. Thus, several simulated reference states have been proposed by various approximations, including averaging [121], quasi-chemical approximation [117], atom-shuffled [122], finite-ideal-gas [118], spherical-non-interacting [123], and random-walk-chain [119] reference states. Moreover, geometrical parameters can be the inter-atom contact, inter-atom distance, inter-atom angle, inter-block orientation, and so on [61].
After deriving a statistical potential, the total energy Δ E ( S , C ) for a conformation C of a given sequence S can be given by the following [116]:
Δ E ( S , C ) = Δ E ( s ) ,  
where the summation is overall applicable items for the geometrical parameter s with the additive assumption for the statistical potential [116].

3.1.1. Two-Body Distance-Dependent Statistical Potentials

RASP. Based on the averaging reference state [121], in which the distribution of different atom pair types is approximately represented by the distribution averaged over all of the atom pair types in native structures, RASP was derived by Capriotti et al. at both CG and all-atom (23 clustered atom types) levels, and the distance between atom pairs was considered as the geometrical parameter [63]. Capriotti et al. showed that RASP had a better performance than NAST [27], which is a nucleotide-level CG statistical potential composed of bond, angle, dihedral, and non-bond terms. The package of RASP can be found at http://melolab.org/webrasp/home.php.
KB potential. Based on the quasi-chemical approximation reference state in which the number of certain atom−pair types should be proportional to the molar fraction of the corresponding ones from the native structures [117], Bernauer et al. proposed the distance-dependent statistical potentials of the KB potential at both CG and all-atom levels [64]. As a result of adopting Dirichlet process mixture models, the KB potentials are fully differentiable, which makes them applicable for molecular dynamics simulations [64].
DFIRE-RNA. Zhang et al. proposed an all-atom (85 atom types) distance-dependent statistical potential of DFIRE-RNA based on the finite-ideal-gas reference state [65], in which the pair distribution function in ideal gas is used to simulate the atomic pair distribution of RNAs in the reference state [65,118]. In DFIRE-RNA, a dimension parameter α is involved, which can help to better match the spatial scale of physical models to that of a realistic RNA or protein system [62,118]. For the RNA-Puzzles dataset, DFIRE-RNA was shown to have a consistently better performance than 3dRNAscore, RASP, and Rosetta energy function, which was combined through a series of knowledge-based and classical physics energy terms [45,63,65]. In addition, the package of DFIRE-RNA can be downloaded at https://github.com/tcgriffith/dfire_rna.
rsRNASP. Tan et al. recently developed an all-atom (85 atom types) distance-dependent statistical potential of rsRNASP by distinguishing local-ranged and non-local-ranged interactions at the residue separation level [66]. The averaging [121] and random-walk-chain [119] reference states were applied for extracting local-ranged and non-local-ranged potentials, respectively. For two test datasets from various structure prediction models, including the RNA-Puzzles dataset, rsRNASP showed an overall superior performance over the existing traditional statistical potentials (RASP, 3dRNAscore, and DFIRE-RNA) and deep-learning-based scoring functions (RNA3DCNN and ARES), which will be introduced below [66]. The package of rsRNASP is available at: https://github.com/Tan-group/rsRNASP.
cgRNASP. For high efficiency and direct applicability to CG-based RNA 3D structure prediction models, a series of residue-separation-based CG statistical potentials at different CG levels were recently proposed by Tan et al., and three-bead cgRNASP (12 CG atom types) is regarded as being representative [67]. Compared with the all-atom rsRNASP, the local-ranged interaction in cgRNASP was involved more subtly and completely through explicitly adding the interactions between the nearest neighbor residues and between the next-nearest ones [67]. Compared with rsRNASP, cgRNASP can have a similarly good or slightly better performance for extensive test datasets, while cgRNASP is strikingly more efficient than all-atom potentials such as rsRNASP, 3dRNAscore, and DFIRE-RNA. cgRNASP can be directly applicable to some existing CG-based RNA 3D structure prediction models [67]. The package of cgRNASP can be downloaded at https://github.com/Tan-group/cgRNASP.

3.1.2. Two-Body Distance-Dependent and Angle-Dependent Statistical Potentials

3dRNAscore. Different from the above-described distance-dependent potentials such as RASP and KB potential, 3dRNAscore is composed of the distance- and torsion angle-dependent potentials based on the averaging reference state [68]. In 3dRNAscore, 85 atom types and 7 torsion angle types were involved [68], and a weight factor was optimized by the decoys of four typical RNAs generated by 3dRNA to balance the contributions of the two kinds of energy terms [68]. The involvement of the dihedral-dependent potential in 3dRNAscore has been shown to improve the evaluation performance [68]. 3dRNAscore has been embedded in 3dRNA, and the standalone package of 3dRNAscore can be found at http://biophy.hust.edu.cn/new/resources/3dRNAscore.

3.1.3. Four-Body Contact Statistical Potential

RAMP. For capturing higher-order interactions beyond two-body potentials, Masso developed a four-body contact potential of RAMP, which is the first multi-body statistical potential for RNA 3D structure evaluation [69]. In RAMP, atomic four-body nearest-neighbors were divided by the Delaunary tessellation [124] for an RNA 3D structure, and each RNA was represented by four atom types of C, N, O, and P [69]. Thus, 35 distinct quadruplet types can be produced by the four-letter atomic alphabet. However, overall, RAMP has a worse performance than RASP-ALL when identifying native structures [69], which might be attributed to the fact that RAMP is a coarse-grained-level contact potential with only four clustered atom types, while RASP-ALL is a distance-dependent potential with 23 clustered atom types [61].
The further development of knowledge-based scoring functions/statistical potentials would benefit from the continuously increasing experimental structures deposited in the PDB database, which are currently inadequate. Moreover, the reference state and the geometrical parameter are crucial for building a knowledge-based scoring function. Thus, expanding the native structure database, modeling more realistic reference states for RNAs, and exploring unique RNA structure features as geometrical parameters could help improve the performance of a statistical potential/scoring function.

3.2. Deep-Learning-Based Scoring Functions

Beyond the above-described traditional statistical potentials/scoring functions, deep-learning techniques have recently been employed to develop scoring functions for RNA 3D structure evaluation. Generally, to develop a deep-learning-based scoring function does not require an artificial definition for RNA structure-related features or an explicit involvement of the reference state [70,71].
RNA3DCNN. Li et al. employed 3D convolutional neural network (3DCNN) to develop two 3DCNN-based scoring functions, named RNA3DCNN_MD and RNA3DCNN_MDMC, for assessing near-native RNA decoys and RNA decoys with large structure fluctuation, respectively [70]. 3DCNNs can directly use a 3D grid representation of RNA structures as the input without extracting RNA structure-related features manually, and the training sets of RNA3DCNN were generated by MD simulations and MC structure prediction for 414 RNAs [70]. RNA3DCNN_MD was trained by the first decoy set, while RNA3DCNN_MDMC was trained by the two decoy sets together [70]. For decoys with RMSDs less than 1.0 Å, RNA3DCNN was shown to have a similar or worse performance compared with the traditional statistical potentials (3dRNAscore, KB, RASP, and Rosetta), while for the RNA-Puzzles dataset, RNA3DCNN performed obviously better in identifying the native structures [70]. The package of RNA3DCNN can be downloaded at https://github.com/lijunRNA/RNA3DCNN.
ARES. Townshend et al. designed a neural network, the Atomic Rotationally Equivariant Scorer (ARES), to obtain the RMSDs of predicted structures from unknown native structures [71]. The initial layers of the ARES network with the 3D coordinates and chemical element type of each atom as the input were designed to recognize structural motifs, which were learned by training rather than being specified in advance. Furthermore, each layer of ARES was rotationally and translationally equivariant, which ensures that the corresponding transformation of its output could be achieved with the rotation or translation of its input. Thus, the orientation and position of an identified motif can be passed on to next layer of the network. Therefore, the initial layers of ARES can gather information locally, which can help to recognize finer-scale motifs (e.g., base pairs) and further recognize coarser-scale motifs (e.g., helices), and the remaining layers aggregate information across all atoms and capture the RNA global property, namely RMSD. Notably, the parameters of ARES were optimized with the training set, which contains decoy structures for 18 target RNAs generated by FARFAR2 [46]. ARES shows an excellent performance for evaluating structures from FARFAR2 [71], while becoming ordinary for the existing test datasets such as the RNA-Puzzles dataset [66,67]. ARES has been developed into a friendly webserver version: http://drorlab.stanford.edu/ares.html.
The deep-learning-based scoring functions could have outstanding performance in some aspects. For example, compared with previous statistical potentials/scoring functions, RNA3DCNN has an apparently better performance for identifying native/near-native structures [70], and ARES performs uniquely better for candidate structures from FARFAR2 [71]. Deep-learning-based scoring functions are generally free of artificially defined features and free of the reference states, which have natural advantages over the traditional statistical potentials/scoring functions [70,71]. However, the performance of a deep-learning-based scoring function is severely limited by its training dataset, such as the structure spectrum of the native structures and that of decoy structures. The incomplete spectrums of the native structures may lead to an unreliable performance of a scoring function for those structures not contributing to the native structure spectrum, and the spectrum of decoy structures may also cause a strong bias for the trained scoring function to the model for generating the decoy structures.

4. RNA 3D Structure Refinement

As described in Section 2 and Section 3, a predictive model for RNA 3D structures generally generates a structure ensemble through either force-field-guided conformation sampling of chains at different CG levels or structure assembly based on different-sized fragments [11,15]. Thus, a structure prediction model may produce many structure candidates with apparently abnormal structural domains containing imperfect stereochemistry information, such as unnatural bond lengths/angles and steric atom conflicts or incorrect tertiary interactions [72,75]. Afterwards, it is generally necessary to make structure refinement (optimization) for the identified (sampled/assembled) structures through a scoring function/evaluation algorithm. By combining the sampling algorithm (such as Gradient Descent, MC) with certain energy functions, the local and global structures of RNAs can be improved to their near-native states through adjusting the abnormal structure domains. Until now, most existing models for refining RNA 3D structures are physics-based ones [72,73,74,75], which are summarized in Table 5 and will be introduced below.
QRNAS. The model employs a modified version of Amber force field [76,77], incorporating four additional energy terms that are not explicitly present in the standard Amber force field, namely hydrogen bonds, base pair co-planarity, backbone regularization, and custom distance constraints [72]. By combining the fastest descent with golden section search and Polak−Ribiere conjugate gradient algorithm [125], QRNAS can significantly improve the local quality of RNA 3D structures while maintaining the global quality of RNA structures [72]. Additionally, QRNAS can refine not only RNA structures, but also DNA structures, chimeras, hybrids, and nucleic acids containing modified residues. QRNAS has been developed into a user-friendly software and is available at http://genesilico.pl/software/stand-alone/qrnas.
BRiQ refinement. BRiQ is a knowledge-based energy function at an atom level, which is corrected with quantum mechanics calculations on base−base interactions, and includes bonded and non-bonded terms [73]. The bonded term includes bond length, bond angle, torsion angle, and backbone rotameric interactions, and the non-bonded term comprises base−base, base−oxygen, oxygen−oxygen, and atomic clash interactions [73]. With utilizing the nucleobase-centric tree (NuTree) algorithm, BRiQ refinement achieves an atom-level refinement for RNA 3D structures, and the BRiQ refinement improves 81% Rosetta-SWM structures with RMSD < 2 Å, 100% RNA puzzle structures with RMSD < 4 Å, and 83% FARFAR2 structures with RMSD < 6 Å [73]. The package of BRiQ refinement is available at https://github.com/Jian-Zhan/RNA-BRiQ.
RNAfitme. Different from QRNAS [72] and BRiQ refinement [73], the RNAfitme model [74] refines RNA 3D structures by reconstructing them with fixed main chains and simulating them guided by the Charmm force field [79,80]. To refine an RNA structure with RNAfitme, five backbone atoms (O5 , C5 , C4 , C3 , and O3 ) or nine backbone and sugar ring atoms (O5 , C5 , C4 , C3 , O3 , O4 , C1 , C2 , and O2 ) need to be involved [74]. In the all-atom reconstruction process, the optimal fragments are selected using an adaptive matching algorithm, and a final all-atom RNA structure can be generated. Finally, NAMD [126] is used to relax the structure and reduce the spatial conflicts to improve the geometric shape of the preliminary RNA structure [74]. Although RNAfitme can hardly improve the global quality of an RNA structure, it can effectively solve the local spatial conflicts for RNA 3D structures. The web server of RNAfitme can be found at http://rnafitme.cs.put.poznan.pl/.
3dRNA optimization. Different from the above models [72,73,74], Wang et al. proposed a method to optimize the global quality of RNA 3D structures [75]. In this method, each nucleotide is represented by six CG beads at the atoms of P, C4 , C1 , C2, C4, and C6. The direct coupled analysis (DCA) [127,128] is used to extract evolutionary constraint information of RNA through multiple sequence alignments, and these constraints are complemented to a physics-based CG force field that includes bonded energies (bond length, bond angle, and dihedral angle) and non-bonded energies (base pairing, base stacking). Combining with the MC annealing algorithm, the method can be used to optimize an RNA 3D structure [75]. It is worth noting that the method can improve the global backbone structures of RNAs, especially for large RNA structures with tertiary contacts. The method has been embedded in 3dRNA and is available online at http://biophy.hust.edu.cn/new/3dRNA.
As described above, structure refinement would improve the local or global structures for RNAs. RNAfitme can reduce spatial conflicts in RNA structures [74], while compared with RNAfitme, QRNAS can also enforce backbone regularization and improve the base pair planarity [72]. However, RNAfitme and QRNAS can hardly improve the global quality of an RNA structure. BRiQ refinement can improve the base pairing structure and repair the RNA backbone structure, but it cannot make improvements for poor-quality RNA structures generated by structure prediction models. The 3dRNA optimization method [75] can significantly improve the backbone structure of large RNAs with tertiary contacts, while it may be difficult to achieve a high-precision RNA structure for those structures without tertiary contacts or no available DCA information. Thus, a high-precision refinement/optimization for RNA structures is still highly required to obtain high-precision RNA 3D structures.

5. Conclusions and Perspectives

Understanding the 3D structures of RNAs is crucial for unraveling the mysteries of the RNA world. As introduced above, a great progress has been made in modeling RNA 3D structures, including structure ensemble prediction, structure evaluation, and structure refinement. However, the existing models for RNA 3D structure prediction are still far away from ab initio predictions in terms of high accuracy solely from sequences, especially compared with protein 3D structure prediction. Here, we will discuss the major challenges and perspectives in modeling RNA 3D structures in different aspects.

5.1. On Physics-Based Structure Modeling

As introduced above, the existing physics-based models can make ab initio predictions for RNA 3D structures only relying on sequences. However, there are still challenges remaining for developing a reliable and applicable physics-based model.
First, the existing physics-based models are severely limited to the prediction accuracy and efficiency, and can generally be inapplicable for the ab initio structure predictions of large RNAs (e.g., >tens of nucleotides) and RNAs with complex topology such as multi-way junctions. The different-level CG approximations in the physics-based models would significantly reduce the structure representation complexity to improve simulation efficiency, while would also losing certain structure accuracy. Thus, achieving both high prediction accuracy and efficiency in CG representation modeling is an important challenge for physics-based RNA 3D structure modeling. Second, deriving an accurate force field/energy function is another critical challenge for a physics-based model to predict complex RNA 3D structures, especially for RNA structures with multi-way junctions and tertiary contacts. Third, developing an efficient and effective conformation sampling algorithm is important for a model to predict 3D structures of large RNAs as the employed simulated annealing, REMC, and REMD algorithms generally have huge computations despite their excellent ability to simulate structure folding on rugged energy landscapes. Fourth, a physics-based model essentially requires a high-quality rebuilding of all-atom structures due to the conventional CG simplification.
Therefore, multi-scale modeling may be very applicable for the further development of physics-based models, where a highly simplified CG-level representation would ensure the basic global structures and computation efficiency, and a final all-atom representation with an exact all-atom force field would ensure the predicted structure accuracy. Moreover, the direct involvement of a fragment assembly treatment after secondary structures fold may also be helpful for improving prediction efficiency for a physics-based model.

5.2. On Knowledge-Based Fragment-Assembly Structure Modeling

As described above, the fragment-assembly models are generally significantly more efficient than the physics-based models, while they are also limited to some severe challenges. First, the prediction accuracy of the existing fragment-assembly models is definitely limited to the quality of the input secondary structures. Thus, a blind prediction of such models would certainly require high-performance secondary structure prediction models. Second, the prediction performance severely depends on the completeness and quality of different types of RNA fragments, while, due to the limited RNA 3D structures in the PDB database, establishing a high-quality library of different types of fragments with a complete spectrum is impossible when only depending on the PDB database. Thus, it is very important to manually build the structures of core fragments, such as multi-way junctions and tertiary contacts for a fragment-assembly model. Third, this kind of model would generally generate structures with severe stereochemistry defects, such as abnormal backbone bond lengths/angles and severe atom conflicts, which require excellent structure refinement at the atom level.
Therefore, for this kind of model, (i) a benchmark survey may be first required on the existing secondary structure prediction methods in order for a blind prediction solely from sequences through a fragment-assembly model based on a more reliable secondary structures as an input, and (ii) combination with reliable physics-based models may be required to generate the 3D structures of core fragments that cannot be obtained through the RNA structures in the PDB database, and such combination may also help to diminish the severe stereochemistry defects in assembled 3D structures.

5.3. On Deep-Learning-Based Structure Modeling

It has been shown that the existing deep-learning-based models are rather effective and efficient at predicting the known RNA 3D structures, while the models also have severe limitations.
First, the performance of deep-learning-based models is severely limited by the structure spectrum of the RNA structures in the PDB database. Due to possible “over-fitting” via the neural network, a deep-learning-based model may make accurate predictions for the RNA structures within the structure spectrum of training (known) structures, while it may completely fail to predict those outside of the training structure spectrum [129]. Second, the “black-box” training process makes a deep-learning-based model difficult to understand. Consequently, it becomes very hard to extend a deep-learning-based model in methodology rather than only extending training datasets.
Therefore, a combination with physics-based models may be a possible way to reduce the effect of the ‘over-fitting’ training process and to improve the performance for deep-learning-based prediction, as a similar combination has shown its performance in RNA secondary structure modeling [130].

5.4. On Overall Modeling for RNA 3D Structures

In addition to the above-described specific limitations for different types of RNA 3D structure prediction models, there are still general challenges for modeling RNA 3D structures.
First, because of the highly polyanionic nature of RNAs, RNA 3D structures are strongly sensitive to the solution environment, such as ions, ligands, and other small molecules [131,132,133,134,135,136,137,138], while the existing structure prediction models rarely consider the effects of solution environments. Due to the critical role of ions in RNA structures and functions, it is still a challenge to properly involve the effects of ions in a structure prediction model, especially for the effect of Mg2+ in complex RNA structures [86,87]. Second, most existing structure prediction models are focused on static RNA 3D structures, while RNA functions can depend on proper structure changes upon the binding of other ions or ligands [139,140,141]. Thus, the knowledge of static 3D structures can sometimes become inadequate for understanding the functions of some RNAs, and it is still a challenge for a structure prediction model to predict RNA 3D structure change due to the interactions of ions/ligands/molecules or due to the change in temperature [15]. Third, non-canonical base pairs and nucleotide modifications such as m6A methylation and pseudouridylation are very important for RNA structures and functions [142,143], while it is still a challenge to predict RNA 3D structures with non-canonical base pairs and modified nucleotides. Fourth, the existing models essentially make in vitro structure predictions for RNAs, while cells can contain up to ~40% volume of various macromolecules such as proteins, DNA, and RNA [133,144,145]. The presence of crowding macromolecules can strongly affect the structures and stabilities of RNAs [135]. Thus, it is very necessary to make the predictions for RNA 3D structures in vivo, and it is still a challenge for structure prediction models to be able to involve the effect of in vivo.
The effects of ions and temperature for complex RNA structures can be possibly captured in a multi-scale physics-based model with the involvement of applicable polyelectrolyte theories and experimental thermodynamic parameters, which has been shown by a CG physics-based model for predicting the thermal stability and ion effect for RNAs with simple topology such as hairpins and pseudoknots [33,35]. Moreover, the non-canonical base pairs and those with nucleotide modifications can be possibly involved in a physics-based structure prediction model through deriving the bond energy parameters for the base pairs based on the corresponding structures in the PDB database and using the corresponding Turner nearest-neighbor thermodynamic parameters [93]. Finally, the effect of an in vivo environment on RNA structures may be modeled based on an effective/efficient physics-based model by involving the effects of ions and temperature through explicitly including the crowders and the associated interactions [146].

5.5. On RNA 3D Structure Evaluation

Although important progress has been made in RNA 3D structure evaluation, there are several apparent limitations to the existing scoring functions. First, the performance of scoring functions is still apparently lower than satisfactory for realistic datasets such as the RNA-Puzzles dataset from various structure prediction models. For example, the Pearson correlation coefficients (PCC) between RMSDs and energies by the existing top scoring functions (e.g., rsRNASP [66], cgRNASP [67], DFIRE-RNA [65], and ARES [71]) are still less than 0.6 for the RNA-Puzzles dataset, a value far less than the ideal one of 1 [66,67]. Second, a structure prediction model can generate a huge number of structure candidates for structure evaluation. Thus, it is still a great challenge to develop a high-performance and efficient scoring function that is applicable to various structure prediction models. Third, the performance and applicability of a scoring function are also subject to the limited RNA 3D structures deposited in the PDB database. Thus, it is challenging to develop a high-performance scoring function based on an RNA structure database with an incomplete structure spectrum. Fourth, as RNA 3D structures can be strongly dependent on ion conditions and temperature, it is also necessary to develop a scoring function involving the effects of ion conditions and temperature to evaluate RNA structures at varying ion conditions and temperatures.
The further developments on a universal and efficient scoring function with a high performance may come from: (i) developing more realistic reference state or circumventing reference state, (ii) finding more proper geometrical parameters to better capture the relations between two atoms or among multiple atoms rather than the inter-atom distance, (iii) involving multi-body interactions beyond two-body potential, (iv) only keeping key (CG) atoms when developing scoring functions to improve the evaluation efficiency, and (v) specifically developing an effective and specific scoring function for a specific structure prediction model based on the training data from the model.

5.6. On RNA 3D Structure Refinement/Optimization

Although the existing structure refinement/optimization models can make the modeled RNA structures closer to their native states, they are also limited to some challenges. First, the existing models either mainly improve the local structures rather than the global structures for RNAs, or they improve the global structures rather than local structures for RNAs, and thus cannot effectively improve both the local and global structures. Second, a model for improving the both global and local structures at the atom level is computationally challenging [75].
Reliable structure refinement and optimization would significantly benefit structure prediction models as more near-native structures can be generated through the procedure. In this sense, a procedure of combining structure optimization with structure refinement may be applicable to obtain more accurate RNA 3D structures as the preceding one can generate more near-native global structures and the follow-up one can generate more local structures accurately. Moreover, a combination of a CG-level structure optimization and an atom-level refinement would also improve the computation efficiency of the procedure.
In summary, great progress has been made in the recent two decades in modeling RNA 3D structures, while many efforts are still required in order to archive accurate predictions for RNA 3D structures and to understand the associated RNA functions in vivo. Aided by the increase in RNA structure data and the advance in physics-based modeling techniques and computational technology, we anticipate exciting developments in modeling RNA 3D structures in the near future.

Author Contributions

Z.-J.T., Y.-L.T. and X.W. designed the research; X.W., S.Y. and E.L. collected the information of the computational models; X.W., Y.-L.T. and Z.-J.T. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants from the National Science Foundation of China (12075171, 11774272).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this review come from the corresponding references.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Roundtree, I.A.; Evans, M.E.; Pan, T.; He, C. Dynamic RNA Modifications in Gene Expression Regulation. Cell 2017, 169, 1187–1200. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Bergkessel, M. Regulation of protein biosynthetic activity during growth arrest. Curr. Opin. Microbiol. 2020, 57, 62–69. [Google Scholar] [CrossRef]
  3. Mortimer, S.A.; Kidwell, M.A.; Doudna, J.A. Insights into RNA structure and function from genome-wide studies. Nat. Rev. Genet. 2014, 15, 469–479. [Google Scholar] [CrossRef]
  4. Chen, F.; Wassie, A.T.; Cote, A.J.; Sinha, A.; Alon, S.; Asano, S.; Daugharthy, E.R.; Chang, J.-B.; Marblestone, A.; Church, G.; et al. Nanoscale imaging of RNA with expansion microscopy. Nat. Methods 2016, 13, 679–684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Serganov, A.; Yuan, Y.-R.; Pikovskaya, O.; Polonskaia, A.; Malinina, L.; Phan, A.T.; Hobartner, C.; Micura, R.; Breaker, R.R.; Patel, D.J. Structural Basis for Discriminative Regulation of Gene Expression by Adenine- and Guanine-Sensing mRNAs. Chem. Biol. 2004, 11, 1729–1741. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Scheitl, C.P.M.; Mieczkowski, M.; Schindelin, H.; Höbartner, C. Structure and mechanism of the methyltransferase ribozyme MTR1. Nat. Chem. Biol. 2022, 18, 547–555. [Google Scholar] [CrossRef]
  7. Kavita, K.; Breaker, R.R. Discovering riboswitches: The past and the future. Trends Biochem. Sci. 2023, 48, 119–141. [Google Scholar] [CrossRef]
  8. Rose, P.W.; Prlić, A.; Altunkaya, A.; Bi, C.; Bradley, A.R.; Christie, C.H.; Di Costanzo, L.; Duarte, J.M.; Dutta, S.; Feng, Z.; et al. The RCSB protein data bank: Integrative view of protein, gene and 3D structural information. Nucleic Acids Res. 2017, 45, D271–D281. [Google Scholar] [CrossRef]
  9. Schlick, T.; Pyle, A.M. Opportunities and Challenges in RNA Structural Modeling and Design. Biophys. J. 2017, 113, 225–234. [Google Scholar] [CrossRef] [Green Version]
  10. RNAcentral Consortium. RNAcentral 2021: Secondary structure integration, improved sequence search and new member databases. Nucleic Acids Res. 2021, 49, D212–D220. [Google Scholar] [CrossRef]
  11. Ou, X.; Zhang, Y.; Xiong, Y.; Xiao, Y. Advances in RNA 3D Structure Prediction. J. Chem. Inf. Model. 2022, 62, 5862–5874. [Google Scholar] [CrossRef] [PubMed]
  12. Capriotti, E.; Marti-Renom, M.A. Computational RNA structure prediction. Curr. Bioinform. 2008, 3, 32–45. [Google Scholar]
  13. Laing, C.; Schlick, T. Computational approaches to 3D modeling of RNA. J. Physics Condens. Matter 2010, 22, 283101. [Google Scholar] [CrossRef] [PubMed]
  14. Li, J.; Chen, S.-J. RNA 3D Structure Prediction Using Coarse-Grained Models. Front. Mol. Biosci. 2021, 8, 720937. [Google Scholar] [CrossRef] [PubMed]
  15. Shi, Y.-Z.; Wu, Y.-Y.; Wang, F.-H.; Tan, Z.-J. RNA structure prediction: Progress and perspective. Chin. Phys. B 2014, 23, 078701. [Google Scholar] [CrossRef] [Green Version]
  16. Puig Lombardi, E.; Londoño-Vallejo, A. A guide to computational methods for G-quadruplex prediction. Nucleic Acids Res. 2020, 48, 1603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Turner, M.; Danino, Y.M.; Barshai, M.; Yacovzada, N.S.; Cohen, Y.; Olender, T.; Rotkopf, R.; Monchaud, D.; Hornstein, E.; Orenstein, Y. rG4detector, a novel RNA G-quadruplex predictor, uncovers their impact on stress granule formation. Nucleic Acids Res. 2022, 50, 11426–11441. [Google Scholar] [CrossRef]
  18. Cragnolini, T.; Derreumaux, P.; Pasquali, S. Coarse-Grained Simulations of RNA and DNA Duplexes. J. Phys. Chem. B 2013, 117, 8047–8060. [Google Scholar] [CrossRef] [Green Version]
  19. Cruz, J.A.; Blanchet, M.-F.; Boniecki, M.; Bujnicki, J.M.; Chen, S.-J.; Cao, S.; Das, R.; Ding, F.; Dokholyan, N.V.; Flores, S.C.; et al. RNA-Puzzles: A CASP-like evaluation of RNA three-dimensional structure prediction. RNA 2012, 18, 610–625. [Google Scholar] [CrossRef] [Green Version]
  20. Miao, Z.; Adamiak, R.W.; Blanchet, M.-F.; Boniecki, M.; Bujnicki, J.M.; Chen, S.-J.; Cheng, C.; Chojnowski, G.; Chou, F.-C.; Cordero, P.; et al. RNA-Puzzles Round II: Assessment of RNA structure prediction programs applied to three large RNA structures. RNA 2015, 21, 1066–1084. [Google Scholar] [CrossRef] [Green Version]
  21. Miao, Z.; Adamiak, R.W.; Antczak, M.; Batey, R.T.; Becka, A.J.; Biesiada, M.; Boniecki, M.J.; Bujnicki, J.M.; Chen, S.-J.; Cheng, C.Y.; et al. RNA-Puzzles Round III: 3D RNA structure prediction of five riboswitches and one ribozyme. RNA 2017, 23, 655–672. [Google Scholar] [CrossRef] [Green Version]
  22. Magnus, M.; Antczak, M.; Zok, T.; Wiedemann, J.; Lukasiak, P.; Cao, Y.; Bujnicki, J.M.; Westhof, E.; Szachniuk, M.; Miao, Z. RNA-Puzzles toolkit: A computational resource of RNA 3D structure benchmark datasets, structure manipulation, and evaluation tools. Nucleic Acids Res. 2020, 48, 576–588. [Google Scholar] [CrossRef]
  23. Miao, Z.; Adamiak, R.W.; Antczak, M.; Boniecki, M.J.; Bujnicki, J.M.; Chen, S.-J.; Cheng, C.Y.; Cheng, Y.; Chou, F.-C.; Das, R.; et al. RNA-Puzzles Round IV: 3D structure predictions of four ribozymes and two aptamers. RNA 2020, 26, 982–995. [Google Scholar] [CrossRef]
  24. Kryshtafovych, A.; Antczak, M.; Szachniuk, M.; Zok, T.; Kretsch, R.C.; Rangan, R.; Pham, P.; Das, R.; Robin, X.; Studer, G.; et al. New prediction categories in CASP15. Proteins Struct. Funct. Bioinform. 2023. [Google Scholar] [CrossRef]
  25. Malhotra, A.; Tan, R.; Harvey, S. Modeling large RNAs and ribonucleoprotein particles using molecular mechanics techniques. Biophys. J. 1994, 66, 1777–1795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Tan, R.K.Z.; Petrov, A.S.; Harvey, S.C. YUP: A Molecular Simulation Program for Coarse-Grained and Multiscaled Models. J. Chem. Theory Comput. 2006, 2, 529–540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Jonikas, M.A.; Radmer, R.J.; Laederach, A.; Das, R.; Pearlman, S.; Herschlag, D.; Altman, R.B. Coarse-grained modeling of large RNA molecules with knowledge-based potentials and structural filters. RNA 2009, 15, 189–199. [Google Scholar] [CrossRef] [Green Version]
  28. Ding, F.; Sharma, S.; Chalasani, P.; Demidov, V.V.; Broude, N.E.; Dokholyan, N.V. Ab initio RNA folding by discrete molecular dynamics: From structure prediction to folding mechanisms. RNA 2008, 14, 1164–1173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Krokhotin, A.; Houlihan, K.; Dokholyan, N.V. iFoldRNA v2: Folding RNA with constraints. Bioinformatics 2015, 31, 2891–2893. [Google Scholar] [CrossRef] [Green Version]
  30. Shi, Y.-Z.; Wang, F.-H.; Wu, Y.-Y.; Tan, Z.-J. A coarse-grained model with implicit salt for RNAs: Predicting 3D structure, stability and salt effect. J. Chem. Phys. 2014, 141, 105102. [Google Scholar] [CrossRef] [Green Version]
  31. Shi, Y.-Z.; Jin, L.; Wang, F.-H.; Zhu, X.-L.; Tan, Z.-J. Predicting 3D Structure, Flexibility, and Stability of RNA Hairpins in Monovalent and Divalent Ion Solutions. Biophys. J. 2015, 109, 2654–2665. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Jin, L.; Shi, Y.-Z.; Feng, C.-J.; Tan, Y.-L.; Tan, Z.-J. Modeling Structure, Stability, and Flexibility of Double-Stranded RNAs in Salt Solutions. Biophys. J. 2018, 115, 1403–1416. [Google Scholar] [CrossRef] [Green Version]
  33. Shi, Y.-Z.; Jin, L.; Feng, C.-J.; Tan, Y.-L.; Tan, Z.-J. Predicting 3D structure and stability of RNA pseudoknots in mon-ovalent and divalent ion solutions. PLoS Comput. Biol. 2018, 14, e1006222. [Google Scholar] [CrossRef]
  34. Jin, L.; Tan, Y.-L.; Wu, Y.; Wang, X.; Shi, Y.-Z.; Tan, Z.-J. Structure folding of RNA kissing complexes in salt solutions: Predicting 3D structure, stability, and folding pathway. RNA 2019, 25, 1532–1548. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, X.; Tan, Y.-L.; Yu, S.; Shi, Y.-Z.; Tan, Z.-J. Predicting 3D structures and stabilities for complex RNA pseudoknots in ion solutions. Biophys. J. 2023, 122, 1503–1516. [Google Scholar] [CrossRef] [PubMed]
  36. Boniecki, M.J.; Lach, G.; Dawson, W.; Tomala, K.; Lukasz, P.; Soltysinski, T.; Rother, K.M.; Bujnicki, J.M. SimRNA: A coarse-grained method for RNA folding simulations and 3D structure prediction. Nucleic Acids Res. 2016, 44, e63. [Google Scholar] [CrossRef]
  37. Magnus, M.; Boniecki, M.J.; Dawson, W.; Bujnicki, J.M. SimRNAweb: A web server for RNA 3D structure modeling with optional restraints. Nucleic Acids Res. 2016, 44, W315–W319. [Google Scholar] [CrossRef]
  38. Zhang, D.; Chen, S.-J. IsRNA: An Iterative Simulated Reference State Approach to Modeling Correlated Interactions in RNA Folding. J. Chem. Theory Comput. 2018, 14, 2230–2239. [Google Scholar] [CrossRef]
  39. Zhang, D.; Li, J.; Chen, S.-J. IsRNA1: De novo prediction and blind screening of RNA 3D structures. J. Chem. Theory Comput. 2021, 17, 1842–1857. [Google Scholar] [CrossRef]
  40. Zhang, D.; Chen, S.-J.; Zhou, R. Modeling Noncanonical RNA Base Pairs by a Coarse-Grained IsRNA2 Model. J. Phys. Chem. B 2021, 125, 11907–11915. [Google Scholar] [CrossRef]
  41. Li, J.; Chen, S.J. RNAJP: Enhanced RNA 3D structure predictions with non-canonical interactions and global to-pology sampling. Nucleic Acids Res. 2023, 51, 3341–3356. [Google Scholar] [CrossRef]
  42. Cragnolini, T.; Laurin, Y.; Derreumaux, P.; Pasquali, S. Coarse-grained HiRE-RNA model for ab initio RNA folding beyond simple molecules, including noncanonical and multiple base pairings. J. Chem. Theory Comput. 2015, 11, 3510–3522. [Google Scholar] [CrossRef] [PubMed]
  43. Kerpedjiev, P.; zu Siederdissen, C.H.; Hofacker, I.L. Predicting RNA 3D structure using a coarse-grain helix-centered model. RNA 2015, 21, 1110–1121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Das, R.; Baker, D. Automated de novo prediction of native-like RNA tertiary structures. Proc. Natl. Acad. Sci. USA 2007, 104, 14664–14669. [Google Scholar] [CrossRef] [PubMed]
  45. Das, R.; Karanicolas, J.; Baker, D. Atomic accuracy in predicting and designing noncanonical RNA structure. Nat. Methods 2010, 7, 291–294. [Google Scholar] [CrossRef]
  46. Watkins, A.M.; Rangan, R.; Das, R. FARFAR2: Improved De Novo Rosetta Prediction of Complex Global RNA Folds. Structure 2020, 28, 963–976. [Google Scholar] [CrossRef] [PubMed]
  47. Parisien, M.; Major, F. The MC-Fold and MC-Sym pipeline infers RNA structure from sequence data. Nature 2008, 452, 51–55. [Google Scholar] [CrossRef]
  48. Popenda, M.; Szachniuk, M.; Antczak, M.; Purzycka, K.J.; Lukasiak, P.; Bartol, N.; Blazewicz, J.; Adamiak, R.W. Automated 3D structure composition for large RNAs. Nucleic Acids Res. 2012, 40, e112. [Google Scholar] [CrossRef]
  49. Biesiada, M.; Pachulska-Wieczorek, K.; Adamiak, R.W.; Purzycka, K.J. RNAComposer and RNA 3D structure prediction for nanotechnology. Methods 2016, 103, 120–127. [Google Scholar] [CrossRef]
  50. Zhao, Y.; Huang, Y.; Gong, Z.; Wang, Y.; Man, J.; Xiao, Y. Automated and fast building of three-dimensional RNA structures. Sci. Rep. 2012, 2, 734. [Google Scholar] [CrossRef] [Green Version]
  51. Wang, J.; Wang, J.; Huang, Y.; Xiao, Y. 3dRNA v2.0: An updated web server for RNA 3D structure prediction. Int. J. Mol. Sci. 2019, 20, 4116. [Google Scholar] [CrossRef] [Green Version]
  52. Zhang, Y.; Wang, J.; Xiao, Y. 3dRNA: Building RNA 3D structure with improved template library. Comput. Struct. Biotechnol. J. 2020, 18, 2416–2423. [Google Scholar] [CrossRef] [PubMed]
  53. Zhang, Y.; Wang, J.; Xiao, Y. 3dRNA: 3D Structure Prediction from Linear to Circular RNAs. J. Mol. Biol. 2022, 434, 167452. [Google Scholar] [CrossRef] [PubMed]
  54. Cao, S.; Chen, S.-J. Physics-Based De Novo Prediction of RNA 3D Structures. J. Phys. Chem. B 2011, 115, 4216–4226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Xu, X.; Chen, S.-J. Hierarchical Assembly of RNA Three-Dimensional Structures Based on Loop Templates. J. Phys. Chem. B 2017, 122, 5327–5335. [Google Scholar] [CrossRef]
  56. Zhou, L.; Wang, X.; Yu, S.; Tan, Y.-L.; Tan, Z.-J. FebRNA: An automated fragment-ensemble-based model for building RNA 3D structures. Biophys. J. 2022, 121, 3381–3392. [Google Scholar] [CrossRef]
  57. Shen, T.; Hu, Z.; Peng, Z.; Chen, J.; Xiong, P.; Hong, L.; Zheng, L.; Wang, Y.; King, I.; Wang, S. E2Efold-3D: End-to-End Deep Learning Method for accurate de novo RNA 3D Structure Prediction. arXiv 2022, arXiv:2207.01586. [Google Scholar]
  58. Pearce, R.; Omenn, G.S.; Zhang, Y. De novo RNA tertiary structure prediction at atomic resolution using geometric potentials from deep learning. BioRxiv 2022. [Google Scholar] [CrossRef]
  59. Feng, C.; Wang, W.; Han, R.; Wang, Z.; Ye, L.; Du, Z.; Wei, H.; Zhang, F.; Peng, Z.; Yang, J. Accurate de novo prediction of RNA 3D structure with transformer network. BioRxiv 2022. [Google Scholar] [CrossRef]
  60. Sha, C.M.; Wang, J.; Dokholyan, N.V. Predicting 3D RNA structure from solely the nucleotide sequence using Euclidean distance neural networks. Biophys. J. 2023, 122, 444A. [Google Scholar] [CrossRef]
  61. Tan, Y.-L.; Feng, C.-J.; Wang, X.; Zhang, W.; Tan, Z.-J. Statistical potentials for 3D structure evaluation: From proteins to RNAs. Chin. Phys. B 2021, 30, 028705. [Google Scholar] [CrossRef]
  62. Tan, Y.-L.; Feng, C.-J.; Jin, L.; Shi, Y.-Z.; Zhang, W.; Tan, Z.-J. What is the best reference state for building statistical potentials in RNA 3D structure evaluation? RNA 2019, 25, 793–812. [Google Scholar] [CrossRef]
  63. Capriotti, E.; Norambuena, T.; Marti-Renom, M.A.; Melo, F. All-atom knowledge-based potential for RNA structure prediction and assessment. Bioinformatics 2011, 27, 1086–1093. [Google Scholar] [CrossRef] [Green Version]
  64. Bernauer, J.; Huang, X.; Sim, A.Y.; Levitt, M. Fully differentiable coarse-grained and all-atom knowledge-based potentials for RNA structure evaluation. RNA 2011, 17, 1066–1075. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Zhang, T.; Hu, G.; Yang, Y.; Wang, J.; Zhou, Y. All-Atom Knowledge-Based Potential for RNA Structure Discrimination Based on the Distance-Scaled Finite Ideal-Gas Reference State. J. Comput. Biol. 2020, 27, 856–867. [Google Scholar] [CrossRef] [PubMed]
  66. Tan, Y.-L.; Wang, X.; Shi, Y.-Z.; Zhang, W.; Tan, Z.-J. rsRNASP: A residue-separation-based statistical potential for RNA 3D structure evaluation. Biophys. J. 2022, 121, 142–156. [Google Scholar] [CrossRef] [PubMed]
  67. Tan, Y.-L.; Wang, X.; Yu, S.; Zhang, B.; Tan, Z.-J. cgRNASP: Coarse-grained statistical potentials with residue separation for RNA structure evaluation. NAR Genom. Bioinform. 2023, 5, lqad016. [Google Scholar] [CrossRef]
  68. Wang, J.; Zhao, Y.; Zhu, C.; Xiao, Y. 3dRNAscore: A distance and torsion angle dependent evaluation function of 3D RNA structures. Nucleic Acids Res. 2015, 43, e63. [Google Scholar] [CrossRef]
  69. Masso, M. All-atom four-body knowledge-based statistical potential to distinguish native tertiary RNA structures from nonnative folds. J. Theor. Biol. 2018, 453, 58–67. [Google Scholar] [CrossRef]
  70. Li, J.; Zhu, W.; Wang, J.; Li, W.; Gong, S.; Zhang, J.; Wang, W. RNA3DCNN: Local and global quality assessments of RNA 3D structures using 3D deep convolutional neural networks. PLOS Comput. Biol. 2018, 14, e1006514. [Google Scholar] [CrossRef] [Green Version]
  71. Townshend, R.J.L.; Eismann, S.; Watkins, A.M.; Rangan, R.; Karelina, M.; Das, R.; Dror, R.O. Geometric deep learning of RNA structure. Science 2021, 373, 1047–1051. [Google Scholar] [CrossRef] [PubMed]
  72. Stasiewicz, J.; Mukherjee, S.; Nithin, C.; Bujnicki, J.M. QRNAS: Software tool for refinement of nucleic acid structures. BMC Struct. Biol. 2019, 19, 5. [Google Scholar] [CrossRef] [PubMed]
  73. Xiong, P.; Wu, R.; Zhan, J.; Zhou, Y. Pairing a high-resolution statistical potential with a nucleobase-centric sampling algorithm for improving RNA model refinement. Nat. Commun. 2021, 12, 2777. [Google Scholar] [CrossRef]
  74. Antczak, M.; Zok, T.; Osowiecki, M.; Popenda, M.; Adamiak, R.W.; Szachniuk, M. RNAfitme: A webserver for modeling nucleobase and nucleoside residue conformation in fixed-backbone RNA structures. BMC Bioinform. 2018, 19, 304. [Google Scholar] [CrossRef]
  75. Wang, J.; Mao, K.; Zhao, Y.; Zeng, C.; Xiang, J.; Zhang, Y.; Xiao, Y. Optimization of RNA 3D structure prediction using evolutionary restraints of nucleotide–nucleotide interactions from direct coupling analysis. Nucleic Acids Res. 2017, 45, 6299–6309. [Google Scholar] [CrossRef] [Green Version]
  76. Weiner, S.J.; Kollman, P.A.; Singh, U.C.; Case, D.A.; Ghio, C.; Alagona, G.; Profeta, S.; Weiner, P. A New Force Field for Molecular Mechanical Simulation of Nucleic Acids and Proteins. J. Am. Chem. Soc. 1984, 106, 765–784. [Google Scholar] [CrossRef]
  77. Perez, A.; Marchán, I.; Svozil, D.; Sponer, J.; Cheatham, T.E., III; Laughton, C.A.; Orozco, M. Refinement of the AMBER Force Field for Nucleic Acids: Improving the Description of α/γ Conformers. Biophys. J. 2007, 92, 3817–3829. [Google Scholar] [CrossRef] [Green Version]
  78. He, J.H.; Tao, H.Y.; Yan, Y.M.; Huang, S.Y.; Xiao, Y. Molecular Mechanism of Evolution and Human Infection with SARS-CoV-2. Viruses 2020, 12, 428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. MacKerell, A.D., Jr.; Banavali, N.; Foloppe, N. Development and current status of the CHARMM force field for nucleic acids. Biopolymers 2000, 56, 257–265. [Google Scholar] [CrossRef]
  80. Vanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I. CHARMM general force field: A force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fields. J. Comput. Chem. 2010, 31, 671–690. [Google Scholar] [CrossRef] [Green Version]
  81. Wu, Y.-Y.; Zhang, Z.-L.; Zhang, J.-S.; Zhu, X.-L.; Tan, Z.-J. Multivalent ion-mediated nucleic acid helix-helix interac-tions: RNA versus DNA. Nucleic Acids Res. 2015, 43, 6156–6165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Wang, F.-H.; Wu, Y.-Y.; Tan, Z.-J. Salt contribution to the flexibility of single-stranded nucleic acid of finite length. Biopolymers 2013, 99, 370–381. [Google Scholar] [CrossRef] [Green Version]
  83. Fu, H.; Zhang, C.; Qiang, X.-W.; Yang, Y.-J.; Dai, L.; Tan, Z.-J.; Zhang, X.-H. Opposite Effects of High-Valent Cations on the Elasticities of DNA and RNA Duplexes Revealed by Magnetic Tweezers. Phys. Rev. Lett. 2020, 124, 058101. [Google Scholar] [CrossRef] [PubMed]
  84. Qiang, X.-W.; Zhang, C.; Dong, H.-L.; Tian, F.-J.; Fu, H.; Yang, Y.-J.; Dai, L.; Zhang, X.-H.; Tan, Z.-J. Multivalent Cations Reverse the Twist-Stretch Coupling of RNA. Phys. Rev. Lett. 2022, 128, 108103. [Google Scholar] [CrossRef] [PubMed]
  85. Lipfert, J.; Doniach, S. Small-Angle X-Ray Scattering from RNA, Proteins, and Protein Complexes. Annu. Rev. Biophys. Biomol. Struct. 2007, 36, 307–327. [Google Scholar] [CrossRef] [Green Version]
  86. Woodson, S.A. Metal ions and RNA folding: A highly charged topic with a dynamic future. Curr. Opin. Chem. Biol. 2005, 9, 104–109. [Google Scholar] [CrossRef] [PubMed]
  87. Tan, Z.-J.; Chen, S.-J. Salt Contribution to RNA Tertiary Structure Folding Stability. Biophys. J. 2011, 101, 176–187. [Google Scholar] [CrossRef] [Green Version]
  88. Xi, K.; Wang, F.-H.; Xiong, G.; Zhang, Z.-L.; Tan, Z.-J. Competitive Binding of Mg2+ and Na+ Ions to Nucleic Acids: From Helices to Tertiary Structures. Biophys. J. 2018, 114, 1776–1790. [Google Scholar] [CrossRef] [Green Version]
  89. Manning, G.S. The molecular theory of polyelectrolyte solutions with applications to the electrostatic properties of polynucleotides. Q. Rev. Biophys. 1978, 11, 179–246. [Google Scholar] [CrossRef]
  90. Tan, Z.-J.; Chen, S.-J. Electrostatic free energy landscapes for nucleic acid helix assembly. Nucleic Acids Res. 2006, 34, 6629–6639. [Google Scholar] [CrossRef] [Green Version]
  91. Tan, Z.-J.; Chen, S.-J. Nucleic Acid Helix Stability: Effects of Salt Concentration, Cation Valence and Size, and Chain Length. Biophys. J. 2006, 90, 1175–1190. [Google Scholar] [CrossRef] [Green Version]
  92. Tan, Z.-J.; Chen, S.-J. RNA helix stability in mixed Na+/Mg2+ solution. Biophys J. 2007, 92, 3615–3632. [Google Scholar] [CrossRef] [Green Version]
  93. Xia, T.; SantaLucia, J., Jr.; Burkard, M.E.; Kierzek, R.; Schroeder, S.J.; Jiao, X.; Cox, C.; Turner, D.H. Thermodynamic Parameters for an Expanded Nearest-Neighbor Model for Formation of RNA Duplexes with Watson−Crick Base Pairs. Biochemistry 1998, 37, 14719–14735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Hukushima, K.; Nemoto, K. Exchange Monte Carlo Method and Application to Spin Glass Simulations. J. Phys. Soc. Jpn. 1996, 65, 1604–1608. [Google Scholar] [CrossRef] [Green Version]
  95. Eastman, P.; Swails, J.; Chodera, J.D.; McGibbon, R.T.; Zhao, Y.; Beauchamp, K.A.; Wang, L.-P.; Simmonett, A.C.; Harrigan, M.P.; Stern, C.D. OpenMM 7: Rapid development of high performance algorithms for molecular dynamics. PLoS Comput. Biol. 2017, 13, e1005659. [Google Scholar] [CrossRef] [Green Version]
  96. Sugita, Y.; Okamoto, Y. Replica-exchange molecular dynamics method for protein folding. Chem. Phys. Lett. 1999, 314, 141–151. [Google Scholar] [CrossRef]
  97. Hori, N.; Denesyuk, N.A.; Thirumalai, D. Salt Effects on the Thermodynamics of a Frameshifting RNA Pseudoknot under Tension. J. Mol. Biol. 2016, 428, 2847–2859. [Google Scholar] [CrossRef] [Green Version]
  98. Denesyuk, N.A.; Hori, N.; Thirumalai, D. Molecular Simulations of Ion Effects on the Thermodynamics of RNA Folding. J. Phys. Chem. B 2018, 122, 11860–11867. [Google Scholar] [CrossRef] [PubMed]
  99. Denesyuk, N.A.; Thirumalai, D. Coarse-Grained Model for Predicting RNA Folding Thermodynamics. J. Phys. Chem. B 2013, 117, 4901–4911. [Google Scholar] [CrossRef] [Green Version]
  100. Massire, C.; Westhof, E. MANIP: An interactive tool for modelling RNA. J. Mol. Graph. Model. 1998, 16, 197–205. [Google Scholar] [CrossRef]
  101. Martinez, H.M.; Maizel, J.V.; Shapiro, B.A. RNA2D3D: A program for Generating, Viewing, and Comparing 3-Dimensional Models of RNA. J. Biomol. Struct. Dyn. 2008, 25, 669–683. [Google Scholar] [CrossRef] [Green Version]
  102. Wang, J.; Cieplak, P.; Kollman, P.A. How well does a restrained electrostatic potential (RESP) model perform in calculating conformational energies of organic and biological molecules? J. Comput. Chem. 2000, 21, 1049–1074. [Google Scholar] [CrossRef]
  103. Mathews, D.H.; Sabina, J.; Zuker, M.; Turner, D.H. Expanded sequence dependence of thermodynamic parameters improves prediction of RNA secondary structure. J. Mol. Biol. 1999, 288, 911–940. [Google Scholar] [CrossRef] [Green Version]
  104. Xu, X.; Chen, S.-J. Predicting RNA Scaffolds with a Hybrid Method of Vfold3D and VfoldLA. In RNA Scaffolds; Springer: New York, NY, USA, 2021; Volume 2323, pp. 1–11. [Google Scholar] [CrossRef]
  105. Xu, X.; Zhao, P.; Chen, S.-J. Vfold: A Web Server for RNA Structure and Folding Thermodynamics Prediction. PLoS ONE 2014, 9, e107504. [Google Scholar] [CrossRef] [PubMed]
  106. Li, J.; Zhang, S.; Zhang, D.; Chen, S.-J. Vfold-Pipeline: A web server for RNA 3D structure prediction from sequences. Bioinformatics 2022, 38, 4042–4043. [Google Scholar] [CrossRef] [PubMed]
  107. Cramer, P. AlphaFold2 and the future of structural biology. Nat. Struct. Mol. Biol. 2021, 28, 704–705. [Google Scholar] [CrossRef]
  108. Bryant, P.; Pozzati, G.; Elofsson, A. Improved prediction of protein-protein interactions using AlphaFold2. Nat. Commun. 2022, 13, 1265. [Google Scholar] [CrossRef]
  109. Jones, D.T.; Thornton, J.M. The impact of AlphaFold2 one year on. Nat. Methods 2022, 19, 15–20. [Google Scholar] [CrossRef] [PubMed]
  110. Zhang, C.; Zhang, Y.; Pyle, A.M. rMSA: A Sequence Search and Alignment Algorithm to Improve RNA Structure Modeling. J. Mol. Biol. 2023, 435, 167904. [Google Scholar] [CrossRef]
  111. Kalvari, I.; Nawrocki, E.P.; Ontiveros-Palacios, N.; Argasinska, J.; Lamkiewicz, K.; Marz, M.; Griffiths-Jones, S.; Toffano-Nioche, C.; Gautheret, D.; Weinberg, Z.; et al. Rfam 14: Expanded coverage of metagenomic, viral and microRNA families. Nucleic Acids Res. 2021, 49, D192–D200. [Google Scholar] [CrossRef] [PubMed]
  112. Yang, J.; Anishchenko, I.; Park, H.; Peng, Z.; Ovchinnikov, S.; Baker, D. Improved protein structure prediction using predicted interresidue orientations. Proc. Natl. Acad. Sci. USA 2020, 117, 1496–1503. [Google Scholar] [CrossRef]
  113. Du, Z.; Su, H.; Wang, W.; Ye, L.; Wei, H.; Peng, Z.; Anishchenko, I.; Baker, D.; Yang, J. The trRosetta server for fast and accurate protein structure prediction. Nat. Protoc. 2021, 16, 5634–5651. [Google Scholar] [CrossRef] [PubMed]
  114. Hoffmann, M.; Noé, F. Generating valid euclidean distance matrices. arXiv 2019, arXiv:1910.03131. [Google Scholar]
  115. Shirvanyants, D.; Ding, F.; Tsao, D.; Ramachandran, S.; Dokholyan, N.V. Discrete Molecular Dynamics: An Efficient and Versatile Simulation Method for Fine Protein Characterization. J. Phys. Chem. B 2012, 116, 8375–8382. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Sippl, M.J. Calculation of conformational ensembles from potentials of mena force: An approach to the knowledge-based prediction of local structures in globular proteins. J. Mol. Biol. 1990, 213, 859–883. [Google Scholar] [CrossRef] [PubMed]
  117. Lu, H.; Skolnick, J. A distance-dependent atomic knowledge-based potential for improved protein structure selection. Proteins: Struct. Funct. Bioinform. 2001, 44, 223–232. [Google Scholar] [CrossRef]
  118. Zhou, H.; Zhou, Y. Distance-scaled, finite ideal-gas reference state improves structure-derived potentials of mean force for structure selection and stability prediction. Protein Sci. 2002, 11, 2714–2726. [Google Scholar] [CrossRef] [Green Version]
  119. Zhang, J.; Zhang, Y. A Novel Side-Chain Orientation Dependent Potential Derived from Random-Walk Reference State for Protein Fold Selection and Structure Prediction. PLoS ONE 2010, 5, e15386. [Google Scholar] [CrossRef] [Green Version]
  120. Deng, H.; Jia, Y.; Wei, Y.; Zhang, Y. What is the best reference state for designing statistical atomic potentials in protein structure prediction? Proteins Struct. Funct. Bioinform. 2012, 80, 2311–2322. [Google Scholar] [CrossRef] [Green Version]
  121. Samudrala, R.; Moult, J. An all-atom distance-dependent conditional probability discriminatory function for protein structure prediction. J. Mol. Biol. 1998, 275, 895–916. [Google Scholar] [CrossRef] [Green Version]
  122. Rykunov, D.; Fiser, A. New statistical potential for quality assessment of protein models and a survey of energy functions. BMC Bioinform. 2010, 11, 128. [Google Scholar] [CrossRef] [Green Version]
  123. Shen, M.-Y.; Sali, A. Statistical potential for assessment and prediction of protein structures. Protein Sci. 2006, 15, 2507–2524. [Google Scholar] [CrossRef] [Green Version]
  124. Barber, C.B.; Dobkin, D.P.; Huhdanpaa, H. The quickhull algorithm for convex hulls. ACM Trans. Math. Softw. 1996, 22, 469–483. [Google Scholar] [CrossRef] [Green Version]
  125. Press, W.H.; Teukolsky, S.A.; Vetterling, W.T.; Flannery, B.P. Numerical Recipes, 3rd ed.; The Art of Scientific Computing; Cambridge University Press: Cambridge, UK, 2007. [Google Scholar]
  126. Phillips, J.C.; Braun, R.; Wang, W.; Gumbart, J.; Tajkhorshid, E.; Villa, E.; Chipot, C.; Skeel, R.D.; Kalé, L.; Schulten, K. Scalable molecular dynamics with NAMD. J. Comput. Chem. 2005, 26, 1781–1802. [Google Scholar] [CrossRef] [Green Version]
  127. Morcos, F.; Pagnani, A.; Lunt, B.; Bertolino, A.; Marks, D.S.; Sander, C.; Zecchina, R.; Onuchic, J.N.; Hwa, T.; Weigt, M. Direct-coupling analysis of residue coevolution captures native contacts across many protein families. Proc. Natl. Acad. Sci. USA 2011, 108, E1293–E1301. [Google Scholar] [CrossRef] [PubMed]
  128. Morcos, F.; Hwa, T.; Onuchic, J.N.; Weigt, M. Direct Coupling Analysis for Protein Contact Prediction. Methods Mol. Biol. 2014, 1137, 55–70. [Google Scholar] [CrossRef] [PubMed]
  129. Das, R.; Kretsch, R.C.; Simpkin, A.J.; Mulvaney, T.; Pham, P.; Rangan, R.; Bu, F.; Keegan, R.; Topf, M.; Rigden, D. Assessment of three-dimensional RNA structure prediction in CASP15. bioRxiv 2023. [Google Scholar] [CrossRef]
  130. Sato, K.; Akiyama, M.; Sakakibara, Y. RNA secondary structure prediction using deep learning with thermodynamic integration. Nat. Commun. 2021, 12, 941. [Google Scholar] [CrossRef] [PubMed]
  131. Tan, Z.-J.; Chen, S.-J. Predicting Ion Binding Properties for RNA Tertiary Structures. Biophys. J. 2010, 99, 1565–1576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Gong, Z.; Zhao, Y.; Chen, C.; Xiao, Y. Role of Ligand Binding in Structural Organization of AddA-riboswitch Aptamer: A Molecular Dynamics Simulation. J. Biomol. Struct. Dyn. 2011, 29, 403–416. [Google Scholar] [CrossRef]
  133. Ellis, R. Macromolecular crowding: An important but neglected aspect of the intracellular environment. Curr. Opin. Struct. Biol. 2001, 11, 114–119. [Google Scholar] [CrossRef] [PubMed]
  134. Tan, Z.-J.; Chen, S.-J. Ion-Mediated RNA Structural Collapse: Effect of Spatial Confinement. Biophys. J. 2012, 103, 827–836. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Zhou, H.-X. Protein folding in confined and crowded environments. Arch. Biochem. Biophys. 2008, 469, 76–82. [Google Scholar] [CrossRef] [Green Version]
  136. Draper, D.E. RNA Folding: Thermodynamic and Molecular Descriptions of the Roles of Ions. Biophys. J. 2008, 95, 5489–5495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Feng, Y.; Zhang, K.; Wu, Q.; Huang, S.-Y. NLDock: A fast nucleic acid–ligand docking algorithm for modeling RNA/DNA–ligand complexes. J. Chem. Inf. Model. 2021, 61, 4771–4782. [Google Scholar] [CrossRef]
  138. Yan, Y.; Wen, Z.; Zhang, D.; Huang, S.-Y. Determination of an effective scoring function for RNA–RNA interactions with a physics-based double-iterative method. Nucleic Acids Res. 2018, 46, e56. [Google Scholar] [CrossRef]
  139. Sun, L.-Z.; Zhang, D.; Chen, S.-J. Theory and Modeling of RNA Structure and Interactions with Metal Ions and Small Molecules. Annu. Rev. Biophys. 2017, 46, 227–246. [Google Scholar] [CrossRef] [Green Version]
  140. Zhang, Y.; Zhang, J.; Wang, W. Atomistic Analysis of Pseudoknotted RNA Unfolding. J. Am. Chem. Soc. 2011, 133, 6882–6885. [Google Scholar] [CrossRef]
  141. He, J.; Wang, J.; Tao, H.; Xiao, Y.; Huang, S.-Y. HNADOCK: A nucleic acid docking server for modeling RNA/DNA–RNA/DNA 3D complex structures. Nucleic Acids Res. 2019, 47, W35–W42. [Google Scholar] [CrossRef]
  142. Tanzer, A.; Hofacker, I.L.; Lorenz, R. RNA modifications in structure prediction—Status quo and future challenges. Methods 2019, 156, 32–39. [Google Scholar] [CrossRef]
  143. Zhao, X.-C.; Dong, H.-L.; Li, X.-L.; Yang, H.-Y.; Chen, X.-F.; Dai, L.; Wu, W.-Q.; Tan, Z.-J.; Zhang, X.-H. 5-Methyl-cytosine stabilizes DNA but hinders DNA hybridization revealed by magnetic tweezers and simulations. Nucleic Acids Res. 2022, 50, 12344–12354. [Google Scholar] [CrossRef] [PubMed]
  144. Yan, Y.; Huang, S.-Y. Modeling protein–protein or protein–DNA/RNA complexes using the HDOCK webserver. Methods Mol. Biol. 2020, 2165, 217–229. [Google Scholar] [PubMed]
  145. Yan, Y.; Huang, S.-Y. RRDB: A comprehensive and non-redundant benchmark for RNA–RNA docking and scoring. Bioinformatics 2018, 34, 453–458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Feng, C.; Tan, Y.-L.; Cheng, Y.-X.; Shi, Y.-Z.; Tan, Z.-J. Salt-Dependent RNA Pseudoknot Stability: Effect of Spatial Confinement. Front. Mol. Biosci. 2021, 8, 666369. [Google Scholar] [CrossRef]
Figure 1. The general workflow for modeling RNA 3D structures: structure generation, structure evaluation, and structure refinement. There are three types of predictive models for RNA 3D structures, namely physics-based, knowledge-based, and deep-learning-based models.
Figure 1. The general workflow for modeling RNA 3D structures: structure generation, structure evaluation, and structure refinement. There are three types of predictive models for RNA 3D structures, namely physics-based, knowledge-based, and deep-learning-based models.
Molecules 28 05532 g001
Figure 2. (A) Major coarse-grained (CG) representations for existing physics-based models. (B) A typical schematic diagram for building an RNA 3D structure through knowledge-based fragment assembly models.
Figure 2. (A) Major coarse-grained (CG) representations for existing physics-based models. (B) A typical schematic diagram for building an RNA 3D structure through knowledge-based fragment assembly models.
Molecules 28 05532 g002
Table 1. The existing physics-based CG models for RNA 3D structure prediction.
Table 1. The existing physics-based CG models for RNA 3D structure prediction.
ModelsRefs.CG BeadsSampling aFinal StructuresFrom
Sequence e
Availability
YUP[25,26]1-beadMCLowest-energy structure bNohttp://rumour.biology.gatech.edu/YammpWeb/
NAST[27]1-beadMDCentroid structures of clusters cNohttps://simtk.org/home/nast
iFoldRNA[28,29]3-beadDMDCentroid structures of clusters cYeshttps://dokhlab.med.psu.edu/ifoldrna
CG model with salt effect[30,31,32,33,34,35]3-beadREMCLowest-energy structure dYesNo
SimRNA[36,37]5-beadREMCCentroid structures of clusters cYeshttps://genesilico.pl/SimRNAweb
IsRNA/IsRNA1[38,39]4/5-beadREMDCentroid structures of clusters cYeshttp://rna.physics.missouri.edu/IsRNA/index.html
IsRNA2[40]5-beadREMDCentroid structures of clusters cYeshttp://rna.physics.missouri.edu/IsRNA/index.html
RNAJP[41]5-beadMDLowest-energy structure bNohttp://rna.physics.missouri.edu/RNAJP/index.html
HiRE-RNA[42]6/7-beadREMDCentroid structures of clusters cYesNo
Ernwin[43]helix-centeredMCMCLowest-energy structure bNohttp://github.com/pkerpedjiev/ernwin
a MC, MD, DMD, REMC, REMD, and MCMC represent Monte Carlo, molecular dynamics, discrete molecular dynamics, replica exchange Monte Carlo, replica exchange molecular dynamics, and Markov chain Monte Carlo simulation, respectively. b The lowest-energy structures are determined by the energy functions from the corresponding prediction models. c The centroid structures of clusters are obtained through clustering the low-energy structures by the energy functions from the corresponding prediction models. d The lowest-energy structures are determined by a knowledge-based statistical potential of cgRNASP [67]. e Can the model make predictions solely from sequence?
Table 2. The existing knowledge-based fragment assembly models for RNA 3D structure prediction.
Table 2. The existing knowledge-based fragment assembly models for RNA 3D structure prediction.
ModelsRefs.Fragment FeatureFinal StructuresAvailability
FARNA/FARFAR2[44,45,46]3-nucleotide fragmentsCentroid structures of clusters chttps://rosie.rosettacommons.org/farfar2
MC-Fold/MC-Sym[47]SSELowest-energy structures dhttp://www.major.iric.ca
RNAComposer[48,49]SSEThe representative structure is assembled from the best templateshttp://rnacomposer.ibch.poznan.pl
3dRNA[50,51,52,53]SSE aLowest-energy structures ehttp://biophy.hust.edu.cn/new/3dRNA
Vfold3D[54]CG SSE bThe representative structure is assembled from the best templateshttp://rna.physics.missouri.edu/vfold3D/
VfoldLA[55]SSE aCentroid structures of clustershttp://rna.physics.missouri.edu/vfoldLA/
FebRNA[56]CG SSE bLowest-energy structure fhttps://github.com/Tan-group/FebRNA
The smallest secondary elements (SSE) are defined as base pair, hairpin loop, internal loop, bulge loop, pseudoknot loop, and junction. a SSEs contain additional base pairs at their ends. b Coarse-grained (CG) SSEs contain base pairs at their ends. c Centroid structures of clusters are determined through clustering the low-energy structures by the energy functions from the corresponding prediction models. d The lowest-energy structures are determined by Amber’99 force field [102]. e The lowest-energy structures are determined through clustering the assembled structures and ranking the cluster centers using 3dRNAscore [68]. f The lowest-energy structures are determined by a knowledge-based statistical potential [67].
Table 3. The existing deep-learning-based approaches for RNA 3D structure prediction.
Table 3. The existing deep-learning-based approaches for RNA 3D structure prediction.
ApproachesRefs.Neural Network Learning InformationFinal StructuresAvailability
RhoFold[57]Sequence representations and interactions between different nucleotidesLowest-energy structurehttps://github.com/RFOLD/RhoFold
DeepFoldRNA[58]Structural information from evolutionary profilesLowest-energy structurehttps://zhanggroup.org/DeepFoldRNA
trRosettaRNA[59]MSA and secondary structure representationsLowest-energy structurehttps://yanglab.nankai.edu.cn/trRosettaRNA/
epRNA[60]RNA sequencesCentroid structures of clustershttps://bitbucket.org/dokhlab/eprna-euclidean-parametrization-of-rna/src/master/
Table 4. The existing scoring functions/statistical potentials for RNA 3D structure evaluation.
Table 4. The existing scoring functions/statistical potentials for RNA 3D structure evaluation.
Knowledge-Based Scoring Functions
Scoring FunctionsRefs.Reference StatesGeometrical ParametersAtom TypesAvailability
RASP-ALL[63]Averaging [116]Distance between atom pairs23http://melolab.org/webrasp/home.php
All-atom KB potential[64]Quasi-chemical approximation [117]Distance between atom pairs85No
DFIRE-RNA[65]Finite-ideal-gas [118]Distance between atom pairs85https://github.com/tcgriffith/dfire_rna
rsRNASP[66]Averaging [116] + Random-walk-chain [119] Distance between atom pairs85https://github.com/Tan-group/rsRNASP
cgRNASP[67]Averaging [116] + Finite-ideal-gas [118]Distance between atom pairs12https://github.com/Tan-group/cgRNASP
3dRNAscore[68]Averaging [116]Distance between atom pairs and torsional angles of backbone85http://biophy.hust.edu.cn/new/resources/3dRNAscore
RAMP[69]Multinomial reference distribution Atomic quadruplet interaction4No
Deep-Learning-Based Scoring Functions
Scoring FunctionsRefs.Reference StatesGeometrical ParametersAtom TypesAvailability
RNA3DCNN[70]Free Free, and the 3D grid representation of RNA structure as the input85https://github.com/lijunRNA/RNA3DCNN
ARES[71]FreeFree and the 3D coordinates and chemical element type of each atom as the input.85http://drorlab.stanford.edu/ares.html
Table 5. The existing physics-based approaches for RNA 3D structure refinement.
Table 5. The existing physics-based approaches for RNA 3D structure refinement.
ApproachesRefs.Force FieldRefinement CharacteristicsAvailability
QRNAS[72]Amber with four optional energy termsReducing clash, enforcing backbone regularization, explicit hydrogen bonds, base pair co-planarity.
co-planarity
http://genesilico.pl/software/stand-alone/qrnas
BRiQ refinement[73]A fully knowledge-based atom-level force filedReducing clash, improving base pairing and backbone structureshttps://github.com/Jian-Zhan/RNA-BRiQ
RNAfitme[74]Charmm force fieldReducing clash and smoothing the structurehttp://rnafitme.cs.put.poznan.pl/
3dRNA optimization[75]CG force field with evolutionary restraints from DCAImproving global backbone structurehttp://biophy.hust.edu.cn/new/3dRNA
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, X.; Yu, S.; Lou, E.; Tan, Y.-L.; Tan, Z.-J. RNA 3D Structure Prediction: Progress and Perspective. Molecules 2023, 28, 5532. https://doi.org/10.3390/molecules28145532

AMA Style

Wang X, Yu S, Lou E, Tan Y-L, Tan Z-J. RNA 3D Structure Prediction: Progress and Perspective. Molecules. 2023; 28(14):5532. https://doi.org/10.3390/molecules28145532

Chicago/Turabian Style

Wang, Xunxun, Shixiong Yu, En Lou, Ya-Lan Tan, and Zhi-Jie Tan. 2023. "RNA 3D Structure Prediction: Progress and Perspective" Molecules 28, no. 14: 5532. https://doi.org/10.3390/molecules28145532

Article Metrics

Back to TopTop