Next Article in Journal
Applications of Infrared Thermography in Ophthalmology
Previous Article in Journal
Liver Transcriptome Analysis of the Black Porgy (Acanthopagrus schlegelii) under Acute Low-Temperature Stress
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanism of Chiral-Selective Aminoacylation of an RNA Minihelix Explored by QM/MM Free-Energy Simulations

1
Department of Applied Electronics, Tokyo University of Science, 6-3-1 Niijuku, Katsushika-ku, Tokyo 125-8585, Japan
2
Research Institute for Science and Technology, Tokyo University of Science, 2641 Yamazaki, Noda, Chiba 278-8510, Japan
3
Department of Biological Science and Technology, Tokyo University of Science, 6-3-1 Niijuku, Katsushika-ku, Tokyo 125-8585, Japan
*
Authors to whom correspondence should be addressed.
Life 2023, 13(3), 722; https://doi.org/10.3390/life13030722
Submission received: 7 February 2023 / Revised: 3 March 2023 / Accepted: 3 March 2023 / Published: 7 March 2023
(This article belongs to the Section Origin of Life)

Abstract

:
Aminoacylation of a primordial RNA minihelix composed of D-ribose shows L-amino acid preference over D-amino acid without any ribozymes or enzymes. This preference in the amino acylation reaction likely plays an important role in the establishment of homochirality in L-amino acid in modern proteins. However, molecular mechanisms of the chiral selective reaction remain unsolved mainly because of difficulty in direct observation of the reaction at the molecular scale by experiments. For seeking a possible mechanism of the chiral selectivity, quantum mechanics/molecular mechanics (QM/MM) umbrella sampling molecular dynamics (MD) simulations of the aminoacylation reactions in a modeled RNA were performed to investigate differences in their free-energy profiles along the reactions for L- and D-alanine and its physicochemical origin. The reaction is initiated by approaching a 3′-oxygen of the RNA minihelix to the carbonyl carbon of an aminoacyl phosphate oligonucleotide. The QM/MM umbrella sampling MD calculations showed that the height of the free-energy barrier for L-alanine aminoacylation reaction was 17 kcal/mol, which was 9 kcal/mol lower than that for the D-alanine system. At the transition state, the distance between the negatively charged 3′-oxygen and the positively charged amino group of L-alanine was shorter than that of D-alanine, which was caused by the chirality difference of the amino acid. These results indicate that the transition state for L-alanine is more electrostatically stabilized than that for D-alanine, which would be a plausible mechanism previously unexplained for chiral selectivity in the RNA minihelix aminoacylation.

Graphical Abstract

1. Introduction

The origin of amino acid homochirality in biological systems has long been a mystery and continues to intrigue many people. Parity violation observed in the β-decay of the nucleus suggests a slight enrichment of the L-enantiomer over the D-enantiomer (10−11) [1]. Studies have also suggested the enantiomeric enrichment of the L-enantiomer in an interstellar environment [2,3,4], including that caused by polarized synchrotron radiation from neutron stars [5] and enantioselective autocatalysis by chiral materials [6,7,8]. However, L-amino acids were slightly predominant in α-methyl amino acids in meteorites but not in α-H amino acids (natural components of proteins) that are more easily racemized [9]. Among these controversies, tRNA aminoacylation by aminoacyl-tRNA synthetase [10] is the most critical phenomenon because natural proteins produced in the ribosomes are composed exclusively of L-amino acid chains, which are derived from aminoacyl-tRNAs. Thus, the origin of the chiral-selective process in tRNA aminoacylation would be more important in the continuity of the evolution of biological systems on Earth. An RNA minihelix recapitulates the domain within tRNA that harbors the amino acid attachment site and may have been the progenitor of modern tRNA [11,12]. Multiple kissing-loop interaction-mediated conformational changes starting from short-hairpin RNAs possibly produced minihelix-like and then tRNA-like structures [13]. On the contrary, the duplication of minihelix-like RNA molecules could have evolved to the peptidyl transferase center on the ribosome [14,15].
Tamura and Schimmel designed a model system to achieve aminoacylation of an RNA minihelix without any help from enzymes and ribosomes, but with a bridging oligonucleotide and an aminoacyl phosphate oligonucleotide (5′-Ala-p-dT6dA2) (Figure 1A) [16], in consideration of the contemporary systems that use aminoacyl phosphate (mononucleotide) adenylates as intermediates for aminoacyl-tRNA synthesis (here, deoxynucleotides were used just for technical convenience). In this system, the L-aminoacyl-minihelix formation was preferred over that of a D-aminoacyl-minihelix formation in a ratio of approximately 4:1 [16]. RNA components with the opposite chirality (L-ribose) resulted in the preference of the formation of D-aminoacyl-minihelix over L-aminoacyl-minihelix [16]. The experimental result is important because it fits well with the “RNA world” hypothesis [17]. RNA replication could have arisen before protein synthesis and RNA could have already become chiral before the origin of protein synthesis. Thus, early tRNAs would have already used D-ribose, and this experiment shows that a D-ribose bias leads to an L-amino acid bias in the peptides that are synthesized. The chiral selectivity in the experiment was suggested to be caused by the steric clash of the amino acid side chain in the constraint of a double helical conformation [18,19]. However, in contrast to the experimental fact, the chiral-selectivity mechanism has not been fully explained.
To elucidate the possible mechanisms, so far, we have performed molecular dynamics (MD) simulations of the modeled RNA minihelix with L/D-alanine (L/D-Ala) [20]. The aminoacylation reaction was initiated by the nucleophile attack of O of 3′-OH (O3′) in the terminal CCA onto the trigonal unsaturated sp2 carbon of the carbonyl center C=O (Ccarb=Ocarb) in an acyl phosphate linkage (Figure 1B). The approaching geometry of the reaction is described by two important quantities: Bürgi–Dunitz (BD) [21,22] and Flippin–Lodge (FL) [23] angles. The BD angle was defined as the O3′…Ccarb=Ocarb angle that should be approximately 105° (Figure 1C), which is measured in small-molecule crystal structures that contained amine and ketone carbonyl in the original paper of Bürgi and Dunitz [22]. The value was also validated by a theoretical study [21]. The FL angle was defined as the angle between the plane formed by O3′, Ccarb and Ocarb atoms and that perpendicular to the carbonyl plane formed by Ccarb, Ocarb, α-carbon of alanine (Cα) and bridging phosphate oxygen (Ob) (Figure 1D). If the size difference between Cα and Ob is small, the FL angle is simply expected to be approximately 0°. Our simulation study showed that the frequency for adopting the geometry required for the reaction with L-Ala was much higher than that with D-Ala, which was likely caused by the difference in the combination of stable dihedral angles along L- and D-alanyl phosphate backbones [20]. In general, the energy of a transition state plays a key role in determining its reaction mechanism. Therefore, for deep understanding of the chiral-selective reaction mechanism in the RNA minihelix, it is essential to describe structures and energetics along the whole reaction pathway, including the transition state of the nucleophilic substitution reaction at the trigonal carbon center. However, we could not investigate a reaction pathway in our previous study due to the use of classical MD simulation based on Newtonian mechanics [20].
Here, to elucidate the mechanism of the chiral-selective aminoacylation found in the RNA minihelix, by employing quantum mechanics/molecular mechanics (QM/MM) umbrella sampling MD simulations, we described structure and energetics along the whole pathway of the reaction. The QM/MM method provides atomically detailed structural information of reaction intermediate (s) and transition state (s) for a chemical reaction involving bond formation and breaking, which are not obtained by classical MD simulations [24]. Umbrella sampling allows us to estimate a free-energy profile along a given reaction coordinate, called a potential mean force (PMF) [25]. Hybrid QM/MM umbrella sampling MD simulations have been performed for various enzymes to investigate their reaction mechanisms [26,27,28,29,30,31]. This QM/MM umbrella sampling MD simulation study for aminoacylation reactions of the RNA minihelix provides key knowledge for explaining a chiral-selectivity mechanism observed in experiments [16].

2. Materials and Methods

2.1. Modeling of the Initial Structure for QM/MM

The model of the RNA minihelix, simulation system and force field were the same as those used in our previous study [20]. Atomic structures with six base pairs centered at the reaction site composed of the RNA minihelix, a bridging oligonucleotide and 5′-Ala-p-dT6dA2 were modeled, where L/D-Ala was covalently attached at 5′-phosphate via an acyl phosphate linkage (Figure 1A). The nucleotides in the RNA minihelix consisted of D-(deoxy) ribose. The modeled RNA minihelix, which had nine excess negative charges in electron units, was solvated with approximately 2900 water molecules and 10 Mg2+, 63 Cl and 52 Na+ ions to neutralize the simulation system at approximately 1 M NaCl solution to mimics and experimental conditions (Figure 2). For DNA and RNA parts, Amber OL15 [32,33,34] and OL3 [35,36] force fields were used, respectively. For D-Ala linked to a phosphate atom, the N-terminal Ala in the Amber ff14SB force field [37] was applied, and the charge of the ester oxygen atom was assigned to −0.3079e to give a total charge of zero for D-Ala-dT, where e is the elementary charge. Amber ff14SB contained the parameters for bond, angle and dihedral involving the ester oxygen atom. The TIP3P model was used for water [38]. For monovalent and divalent ions, the parameters developed by Joung and Cheatham [39] and Li and Merz [40], respectively, were used. The system was minimized, followed by equilibration at 300 K and 1 bar conditions with the same procedures as reported in Ref. [20], except for applying distance restraints between non-hydrogen atoms involving hydrogen bonds in all base pairs to their ideal distances by a harmonic potential of the form 1 2 k r r 0 2 with a force constant k of 1 kcal/mol/Å2. Subsequently, 20-ns NPT simulation was performed with an additional harmonic distance restraint between O3′ and Ccarb atoms to 5 Å using a force constant k of 50 kcal/mol/Å2. The coordinate obtained at the end of this NPT simulation of the D-Ala system was used as the initial point for QM/MM MD simulations with umbrella sampling, as described below. An initial structure of the L-Ala system was generated by exchanging the coordinates between the amino and methyl groups of the D-Ala system. Therefore, the initial configurations between the L-Ala and D-Ala systems for QM/MM MD simulations only differed in their chirality. All classical MD simulations were performed using the PMEMD module in the AMBER 18 software package [41] on NVIDIA GPU RTX-2080Ti.

2.2. QM/MM Umbrella Sampling MD Simulations

The QM calculations were performed using Self-Consistent Charge Density Functional Tight Binding method (SCC-DFTB) [43,44,45] including the third order term of the Taylor series expansion of the DFT total energy, called DFTB3 [46], which is implemented in AMBER. The QM region comprises Ala-dT and 3′-A, whose O3′ attacks the carbonyl carbon Ccarb in the acyl phosphate linkage, which is shown in Figure 2. The QM region consists of 79 atoms. All remaining atoms of the nucleic acid and solvent including ions were treated using MM with the AMBER force field.
For umbrella sampling, a reaction coordinate (RC) was chosen as the difference in the distances between H3′ and O3′, d (H3′…O3′) (Figure 1B, green-dotted line) and between Ccarb and O3′, d (Ccarb…O3′) (Figure 1B, red-dotted line): RC = d (H3′…O3′) − d (Ccarb…O3′). The equilibrium distance between H3′ and O3′ atoms in the AMBER force field was 0.96 Å. Therefore, the initial coordinate for QM/MM umbrella sampling MD corresponded to the state of reaction coordinate = −4.0 Å. In total, 101 windows were computed along the reaction coordinate (an increment of reaction coordinate, ∆RC = 0.1 Å for −4.0 Å ≤ RC < −1.5 Å and 1.5 Å < RC ≤ 3.0 Å and ∆RC = 0.05 Å for −1.5 Å ≤ RC ≤ 1.5 Å) with a harmonic restraint force constant k of 500 kcal/mol/Å2. For each window, 10 ps of equilibration and 20 ps of the production run with harmonic restraints were performed, where the last coordinate of the equilibration run in the previous window was used as the starting point. A time step of 0.5 fs was used. Long-range electrostatic interactions were calculated using the particle mesh Ewald method [47,48], and short-range electrostatic and Lennard–Jones interactions were truncated at a cut-off radius of 9 Å. The temperature was controlled using the Langevin thermostat [49] at 300 K with a 1.0 ps−1 collision frequency. Pressure was regulated with the Berendsen algorithm [50] at 1 bar with a pressure relaxation time of 1.0 ps. Coordinates, energies and values of reaction coordinate were collected every 5 fs during the production run. Next, the weighted histogram analysis method [51,52,53] was used to compute the PMF from the simulation data for all windows with a convergence tolerance of 10−7. All QM/MM umbrella sampling MD simulations were performed using the SANDER module in the AMBER 18 [41].

3. Results

The model system by Tamura and Schimmel was composed of the extended double helix with the CCA of an RNA minihelix, an aminoacyl phosphate donor nucleotide (mimic of aminoacyl-AMP) and a bridging nucleotide [16,18]. The reaction occurs under the restriction that the nucleophile attack of O3′ in the terminal CCA onto Ccarb in acyl phosphate linkage was performed with BD and FL angles to be approximately 105° and 0°, respectively. In umbrella sampling, a reaction coordinate was defined as RC = d (H3′…O3′) − d (Ccarb…O3′), where d (H3′…O3′) is the distance between H3′ and O3′ (Figure 1B, green-dotted line) and d (Ccarb…O3′) is that between Ccarb and O3′ (Figure 1B, red-dotted line).
Figure 3 shows a conceptual image of the QM/MM umbrella sampling MD method used in this study. Table S1 briefly explains the simulation methods related to our method. Figure 4 shows the PMFs for aminoacylation reactions with the L/D-Ala obtained from the QM/MM umbrella sampling MD simulations. The heights of the reaction-free-energy barriers at this level of theory were 17.0 kcal/mol for L-Ala at reaction coordinate of −0.5 Å and 26.2 kcal/mol for D-Ala at reaction coordinate of −0.55 Å. The reaction barrier for L-Ala was 9 kcal/mol lower than that for D-Ala. The overall free energies were −7.0 and −1.0 kcal/mol for the L-Ala and D-Ala systems, respectively. The intermediate state was not observed for both systems.
Representative geometries of the aminoacylation sites for the L/D-Ala systems at various reaction coordinates are shown in Figure 5. Movies of the simulations throughout the whole reaction coordinates are also provided in Supplementary Materials (Movie S1 and Movie S2 for the L-Ala and D-Ala systems, respectively). The L/D-Ala system started with similar coordinates (reaction coordinate of −4.0 Å), and no distinct difference in coordinates around the reaction site was observed between them, except for the chirality of the terminal Ala. The overall structures of the L-/D-Ala system after the complete transfer of alanine moieties from 5′-dT to 3′-A (reaction coordinate of 3.0 Å) were also comparable. For both systems, the H3′ atom transferred from 3′-oxygen to one of the oxygens in the 5′-end phosphate group simultaneously with the nucleophile attack of O3′.
Figure 6A,B shows the BD and FL angles, respectively, averaged over each umbrella window for reaction coordinate of −4 to 0 Å. In this analysis, the FL angle has positive values when the plane formed O3′, Ccarb and Ocarb tilts to the bridging phosphate oxygen Ob. The BD angles gradually increased to the experimentally measured value of 105° just before the transition states, with a reaction coordinate of −1.0 Å for the L- and D-Ala systems. For the FL angles in both systems, the values fluctuated approximately 0° with small deviations for reaction coordinate of −2 to −1 Å. The simulation models used in this study well reproduced the restrictions of BD and FL angles for the aminoacylation reaction.
To quantify structural differences between L-/D-Ala systems, three geometrical features around the active site were measured. The first measure is the distance between the nitrogen atom N of the positively charged amino group and the O3′ atom of the relatively high electronegative hydroxyl group, d (N…O3′) (Figure 1B, blue-dotted line). The second measure is the distance between the N atom and the P5′ atom of the negatively charged 5′-phospate group, d (N…P5′) (Figure 1B, black-dotted line). The third measure is a dihedral angle defined by Ob-Ccarb-Cα-N, τ (Ob-Ccarb-Cα-N). Figure 7 shows changes in d (N…O3′), d (N…P5′) and τ (Ob-Ccarb-Cα-N) along the reaction coordinate. These values are averaged over each umbrella window for a reaction coordinate of −4 to 0 Å. Until reaction coordinate of −2.5 Å, the trajectories and absolute values of the measures for the L-/D-Ala systems were comparable. At reaction coordinate of −2.2 Å (d (Ccarb…O3′) ≈ 3.2 Å), the averaged value of d (N…O3′) for the D-Ala system suddenly dropped from 5 Å to 3.5 Å (Figure 7A), which was coupled with a rotation of the dihedral angle τ (Ob-Ccarb-Cα-N) from 0° (cis conformation) to ±180° (trans conformation) (Figure 7C). Contrarily, for the L-Ala system, the distance d (N…O3′) gradually decreased until reaction coordinate of −0.75 Å (Figure 7A), and the dihedral angle τ (Ob-Ccarb-Cα-N) also gradually rotated from 0 to 60° during the reaction coordinate from −2.2 to −0.5 Å (Figure 7C). As a result of the difference in the mode of conformational changes, the d (N…O3′) of the D-Ala system was ~1 Å shorter than that of the L-Ala system during reaction coordinate of −2.3 to −1.6 Å; however, after reaction coordinate of −1.6 Å, the d (N…O3′) for the L-Ala system was slightly but clearly shorter than that for the D-Ala system, which continued until the transition states at reaction coordinate ≈ −0.5 Å. For the d (N…P5′), its difference between L-Ala and D-Ala systems is more evident (Figure 7B). Until reaction coordinate of −2.5 Å, the d (N…P5′) of the L-/D-Ala systems were 4.2–4.3 Å. For the D-Ala system, the averaged value of d (N…P5′) increased to 5.1 Å at reaction coordinate of −2.2 Å, which was also coupled with the cis/trans isomerization of the dihedral angle τ (Ob-Ccarb-Cα-N) (Figure 7C). The distance for D-Ala system remained around 5.1 Å until the transition state at reaction coordinate of −0.55 Å (See also Figure 5 showing the d (N…P5′) at the transition state). After passing the transition state, the d (N…P5′) in D-Ala system jumped up to 6.3 Å. On the other hand, the d (N…P5′) in L-Ala system stayed around 4.25 Å until reaction coordinate of −1.1 Å (Figure 7B). Around the transition state, the d (N…P5′) in L-Ala system, which is shown in Figure 5, fluctuated around 4.0 Å. Therefore, the distance between positively charged amino group and the negatively charged 5′-phosphate group of L-Ala system is approximately 1.1 Å shorter than that of D-Ala system at their transition states.

4. Discussion

In the original experiments by Tamura and Schimmel, the chiral preference for L-amino acid in the RNA minihelix was observed not only for Ala but also for Leu and Phe with approximately fourfold selectivity, which was measured by the ratio of the product formed at 30 min of reaction time [16]. As discussed in Ref. [18], an energetic difference of <1 kcal/mol in the rate-determining step is sufficient to give the fourfold preference for L- vs. D-specific aminoacylation. The calculated energy difference of 9 kcal/mol at the transition state (reaction coordinate ≈ −0.5 Å) was too large compared with the experimental estimation (Figure 4); however, the chiral preference estimated by the QM/MM simulation study is qualitatively consistent with the experimental results. The primary goal of our simulations at the SCC-DFTB3 level of theory was to provide qualitative results to investigate chiral-selectivity mechanisms, and we cannot expect that the absolute quantitative results are comparable to experimental results. Therefore, our simulation results would reflect the experimental facts.
The BD and FL angles define the approaching geometry of the nucleophilic oxygen O3′ atom toward the electrophile carbonyl carbon Ccarb atom in the aminoacylation reaction, which are described by the relative positions of O3′, Ccarb, Ocarb, Cα and Ob atoms. In QM/MM MD simulations, both angles were gradually adopting their adequate values for the reaction until the transition states L- and D-Ala systems (Figure 6). Therefore, the observed free-energy difference between the L- and D-Ala systems at the transition state is not attributed to the local geometry difference described by only these atom positions but to the effects of surrounding atoms and functional groups around the reactive carbonyl group.
We measured the distances between the N atom of the positively charged amino group in Ala and the O3′ atom of the relatively high electronegative hydroxyl group, d (N…O3′), and between the N atom and P5′ atom of negatively charged 5′-phosphate group, d (N…P5′), to find the structural origin that generates the energy difference at the transition state between L- and D-Ala systems. Electrostatic interactions between the amino group and the negatively charged groups could largely contribute to the transition state energy. At the transition states for L-Ala (reaction coordinate of −0.5 Å) and D-Ala (reaction coordinate of −0.55 Å), the d (N…O3′) and d (N…P5′) for the L-Ala system was roughly 0.5 and 1.1 Å, respectively, shorter than those for the D-Ala system (Figure 5 and Figure 7A,B). Therefore, the transition state would be more stabilized in the L-Ala system than in the D-Ala system. This is a possible mechanism for the chiral selectivity in the RNA minihelix found in the QM/MM umbrella sampling MD simulations.
The homochirality of proteins and nucleic acids is intrinsically required in their structures: the secondary structures of proteins, α-helix and β-sheet, are likely to be formed only if the constituent amino acids are homochiral (all-L or all-D) [54], and a template-directed elongation of nucleotide chains occurs properly only when the ribose sugars of the template are homochiral [55,56]. However, why natural proteins and nucleic acids are composed of L-amino acids and D-sugars, respectively, remains unclear. It has been shown that oligomerization of all possible combinations of short homochiral and heterochiral RNA diastereomers proceeded chiral-selectively [57], forming two libraries consisting of equal amounts of homochiral all-L and all-D RNAs. However, symmetry breaking occurs inevitably because the number of possible sequences with growing oligomer length is beyond the number of actually formed sequences. D-ribose RNA would have been selected by chance due to an important chemical ability acquired through the symmetry breaking [19]. Furthermore, although the hypothesis that amino acids were chirally biased first and that the chiral bias in RNA arose secondarily cannot also be denied, ribose has four asymmetric centers (C1′, C2′, C3′ and C4′), and Watson-Crick helices can be formed properly if any two of the three asymmetric carbons (C1′, C3′ and C4′) are correctly positioned relative to each other [58]. In considering the complex set of asymmetric centers in ribose, the probability that L-amino acids could have selected proper formation of the chain of D-ribose with proper configurations linked through 5′-3′ phosphodiester bond would be quite low [59]. The discovery of ribozymes raised the possibility of the “RNA world” as a plausible stage in Earth’s history [17]. Thus, the transition from the putative “RNA world” to the “protein world” would have been a key step [14].
In the Tamura and Schimmel model, the RNA chirality has been shown to correlate with recognition of chiral amino acids through RNA aminoacylation: D-ribose RNA aminoacylates L-amino acids, whereas L-ribose RNA aminoacylates D-amino acids [16]. All tRNA molecules possess a single-stranded CCA sequence at their 3′-ends [14]. Although it has been proposed that tRNA-like structural motifs with CCA first evolved as 3′-terminal tags in RNA genomes for replication in the “RNA world” [60], the CCA would have also been a target for tRNA charging by the interactions with other oligomers. The primordial tRNAs could have been somewhat charged by transferring the amino acid from another oligomer as shown in this model. In this stage, a D-ribose based “RNA world” would have already been formed. In fact, an amino acid activation ribozyme was selected in vitro [61] and this reaction could have happened evolutionarily. However, in terms of the continuity of the biological evolution toward peptide synthesis on the ribosome, the transfer of amino acid from phosphate (in the form of acyl phosphate) to hydroxyl group (in the form of ester) would have been necessary. Although the chiral-selective experiments using mutants [18] and MD simulations [20] suggested the steric clash of the side chain of amino acids in the original chiral-selective model reaction, the quantum mechanical calculations must be included to provide a clear energetics of the proposed mechanism.
Thus, the present attempt was conducted, and finally, the energetic and structural differences of the active sites of the L-/D-Ala systems at the transition states were observed in our simulations. An origin that causes differences would be the chirality of amino acids. The differences in d (N…O3′) and d (N…P5′) at the transition states were initially generated by the cis to trans transition of Ob-Ccarb-Cα-N dihedral angle at a reaction coordinate of −2.2 Å only in the D-Ala system (Figure 7). Chirality may perturb the rotational probability of the dihedral angle.

5. Conclusions

Chiral-selective aminoacylation in the primordial RNA minihelix is a key reaction for homochirality of L-amino acid. In this study, we used QM/MM MD simulation techniques and explored atomistic mechanisms of the chiral-selective aminoacylation observed in experiments. Our computational study presented the model mechanism that the L-Ala moiety stabilizes the transition state more than D-Ala, resulting in L-Ala preference in the aminoacylation reaction in the RNA. For a more detailed mechanism, longer and larger QM/MM simulation studies with high-level QM theory are necessary. Specifically, the inclusion of solvent atoms around the reaction site as the QM region would be important for more quantitative evaluations of the height of the barrier at transition states, which is our further study. The computational studies presented here highlight that the QM/MM MD simulations have the potential to solve evolutionary problems in terms of molecular quantum mechanics.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/life13030722/s1. Table S1: Brief explanations for the methods related with the hybrid QM/MM umbrella sampling MD. Movie S1: QM/MM umbrella sampling MD simulation for the L-Ala system. Movie S2: QM/MM umbrella sampling MD simulation for the D-Ala system.

Author Contributions

Conceptualization, T.A. and K.T.; Data curation; T.A., Funding acquisition, T.A. and K.T.; Investigation, T.A.; Project administration, T.A.; Supervision, K.T.; Writing–original draft, T.A. and K.T.; Writing–review and editing, T.A. and K.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by JSPS KAKENHI Grant Numbers JP20K06592 (to T.A.) and JP21K06293 (to K.T.).

Data Availability Statement

The data presented in this study are available from the corresponding authors upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hegstrom, R.A. Parity violation and chiral symmetry breaking of a racemic mixture. Biosystems 1987, 20, 49–56. [Google Scholar] [CrossRef] [PubMed]
  2. Oró, J. Comets and the formation of biochemical compounds on the primitive earth. Nature 1961, 190, 389–390. [Google Scholar] [CrossRef]
  3. Chyba, C.F.; Thomas, P.J.; Brookshaw, L.; Sagan, C. Cometary delivery of organic molecules to the early Earth. Science 1990, 249, 366–373. [Google Scholar] [CrossRef]
  4. Chyba, C.; Sagan, C. Endogenous production, exogenous delivery and impact-shock synthesis of organic molecules: An inventory for the origins of life. Nature 1992, 355, 125–132. [Google Scholar] [CrossRef] [PubMed]
  5. Bonner, W.A. The quest for chirality. In Proceedings of the AIP Conference Proceedings; AIP Press: New York, NY, USA, 1996; Volume 379, pp. 17–49. [Google Scholar]
  6. Soai, K.; Shibata, T.; Morioka, H.; Choji, K. Asymmetric autocatalysis and amplification of enantiomeric excess of a chiral molecule. Nature 1995, 378, 767–768. [Google Scholar] [CrossRef]
  7. Blackmond, D.G. Asymmetric autocatalysis and its implications for the origin of homochirality. Proc. Natl. Acad. Sci. USA 2004, 101, 5732–5736. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Kawasaki, T.; Matsumura, Y.; Tsutsumi, T.; Suzuki, K.; Ito, M.; Soai, K. Asymmetric autocatalysis triggered by carbon isotope (13C/12C) chirality. Science 2009, 324, 492–495. [Google Scholar] [CrossRef] [PubMed]
  9. Cronin, J.R.; Pizzarello, S. Enantiomeric excesses in meteoritic amino acids. Science 1997, 275, 951–955. [Google Scholar] [CrossRef]
  10. Schimmel, P. Aminoacyl tRNA synthetases: General scheme of structure-function relationships in the polypeptides and recognition of transfer RNAs. Annu. Rev. Biochem. 1987, 56, 125–158. [Google Scholar] [CrossRef]
  11. Schimmel, P.; Giegé, R.; Moras, D.; Yokoyama, S. An operational RNA code for amino-acids and possible relationship to genetic-code. Proc. Natl. Acad. Sci. USA 1993, 90, 8763–8768. [Google Scholar] [CrossRef] [Green Version]
  12. Schimmel, P.; Ribas de Pouplana, L. Transfer RNA: From minihelix to genetic code. Cell 1995, 81, 983–986. [Google Scholar] [CrossRef] [Green Version]
  13. Mutsuro-Aoki, H.; Hamachi, K.; Kurihara, R.; Tamura, K. Aminoacylation of short hairpin RNAs through kissing-loop interactions indicates evolutionary trend of RNA molecules. Biosystems 2020, 197, 104206. [Google Scholar] [CrossRef]
  14. Tamura, K. Origins and early evolution of the tRNA molecule. Life 2015, 5, 1687–1699. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Kawabata, M.; Kawashima, K.; Mutsuro-Aoki, H.; Ando, T.; Umehara, T.; Tamura, K. Peptide bond formation between aminoacyl-minihelices by a scaffold derived from the peptidyl transferase center. Life 2022, 12, 573. [Google Scholar] [CrossRef] [PubMed]
  16. Tamura, K.; Schimmel, P. Chiral-selective aminoacylation of an RNA minihelix. Science 2004, 305, 1253. [Google Scholar] [CrossRef] [PubMed]
  17. Gilbert, W. Origin of life: The RNA world. Nature 1986, 319, 618. [Google Scholar] [CrossRef]
  18. Tamura, K.; Schimmel, P.R. Chiral-selective aminoacylation of an RNA minihelix: Mechanistic features and chiral suppression. Proc. Natl. Acad. Sci. USA 2006, 103, 13750–13752. [Google Scholar] [CrossRef] [Green Version]
  19. Tamura, K. Origin of amino acid homochirality: Relationship with the RNA world and origin of tRNA aminoacylation. Biosystems 2008, 92, 91–98. [Google Scholar] [CrossRef]
  20. Ando, T.; Takahashi, S.; Tamura, K. Principles of chemical geometry underlying chiral selectivity in RNA minihelix aminoacylation. Nucleic Acids Res. 2018, 46, 11144–11152. [Google Scholar] [CrossRef] [Green Version]
  21. Bürgi, H.B.; Dunitz, J.D.; Shefter, E. Geometrical reaction coordinates. II. Nucleophilic addition to a carbonyl group. J. Am. Chem. Soc. 1973, 95, 5065–5067. [Google Scholar] [CrossRef]
  22. Bürgi, H.B.; Dunitz, J.D.; Lehn, J.M.; Wipff, G. Stereochemistry of reaction paths at carbonyl centers. Tetrahedron 1974, 30, 1563–1572. [Google Scholar] [CrossRef]
  23. Heath cock, C.H. Understanding and controlling diastereofacial selectivity in carbon-carbon bond-forming reactions. Aldrichimica Acta 1990, 23, 90–111. [Google Scholar]
  24. Warshel, A.; Levitt, M. Theoretical studies of enzymic reactions: Dielectric, electrostatic and steric stabilization of the carbonium ion in the reaction of lysozyme. J. Mol. Biol. 1976, 103, 227–249. [Google Scholar] [CrossRef] [PubMed]
  25. Torrie, G.M.; Valleau, J.P. Nonphysical sampling distributions in Monte Carlo free-energy estimation: Umbrella sampling. J. Comput. Phys. 1977, 23, 187–199. [Google Scholar] [CrossRef]
  26. Senn, H.M.; Thiel, W. QM/MM methods for biomolecular systems. Angew. Chem. Int. Ed. Engl. 2009, 48, 1198–1229. [Google Scholar] [CrossRef]
  27. Saharay, M.; Guo, H.; Smith, J.C. Catalytic mechanism of cellulose degradation by a cellobiohydrolase, CelS. PLoS ONE 2010, 5, e12947. [Google Scholar] [CrossRef]
  28. Lence, E.; van der Kamp, M.W.; González-Bello, C.; Mulholland, A.J. QM/MM simulations identify the determinants of catalytic activity differences between type II dehydroquinase enzymes. Org. Biomol. Chem. 2018, 16, 4443–4455. [Google Scholar] [CrossRef] [Green Version]
  29. Wang, X.; Li, R.R.; Cui, W.C.; Li, Q.; Yao, J.Z. QM/MM free energy simulations of an efficient gluten hydrolase (Kuma030) implicate for a reactant-state based protein-design strategy for general acid/base catalysis. Sci. Rep. 2018, 8, 7042. [Google Scholar] [CrossRef] [Green Version]
  30. Pereira, C.S.; Silveira, R.L.; Skaf, M.S. QM/MM simulations of enzymatic hydrolysis of cellulose: Probing the viability of an endocyclic mechanism for an inverting cellulase. J. Chem. Inf. Model. 2021, 61, 1902–1912. [Google Scholar] [CrossRef]
  31. Zalloum, W.A.; Zalloum, N. Comparative QM/MM molecular dynamics and umbrella sampling simulations: Interaction of the zinc-bound intermediate gem-diolate trapoxin A inhibitor and acetyl-l-lysine substrate with histone deacetylase 8. J. Phys. Chem. B 2021, 125, 5321–5337. [Google Scholar] [CrossRef]
  32. Krepl, M.; Zgarbová, M.; Stadlbauer, P.; Otyepka, M.; Banáš, P.; Koča, J.; Cheatham, T.E., 3rd; Jurečka, P.; Šponer, J. Reference simulations of noncanonical nucleic acids with different chi variants of the AMBER force field: Quadruplex DNA, quadruplex RNA and Z-DNA. J. Chem. Theor. Comput. 2012, 8, 2506–2520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Zgarbová, M.; Luque, F.J.; Šponer, J.; Cheatham, T.E., 3rd; Otyepka, M.; Jurečka, P. Toward improved description of DNA backbone: Revisiting epsilon and zeta torsion force field parameters. J. Chem. Theor. Comput. 2013, 9, 2339–2354. [Google Scholar] [CrossRef] [PubMed]
  34. Zgarbová, M.; Šponer, J.; Otyepka, M.; Cheatham, T.E., 3rd; Galindo-Murillo, R.; Jurečka, P. Refinement of the sugar-phosphate backbone torsion beta for AMBER force fields improves the description of Z- and B-DNA. J. Chem. Theor. Comput. 2015, 11, 5723–5736. [Google Scholar] [CrossRef]
  35. Pérez, A.; Marchán, I.; Svozil, D.; Sponer, J.; Cheatham, T.E., 3rd; Laughton, C.A.; Orozco, M. Refinement of the AMBER force field for nucleic acids: Improving the description of alpha/gamma conformers. Biophys. J. 2007, 92, 3817–3829. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Zgarbová, M.; Otyepka, M.; Šponer, J.; Mládek, A.; Banáš, P.; Cheatham, T.E., 3rd; Jurečka, P. Refinement of the Cornell et al. nucleic acids force field based on reference quantum chemical calculations of glycosidic torsion profiles. J. Chem. Theor. Comput. 2011, 7, 2886–2902. [Google Scholar] [CrossRef]
  37. Maier, J.A.; Martinez, C.; Kasavajhala, K.; Wickstrom, L.; Hauser, K.E.; Simmerling, C. ff14SB: Improving the accuracy of protein side chain and backbone parameters from ff99SB. J. Chem. Theor. Comput. 2015, 11, 3696–3713. [Google Scholar] [CrossRef] [Green Version]
  38. Price, D.J.; Brooks, C.L. A modified TIP3P water potential for simulation with Ewald summation. J. Chem. Phys. 2004, 121, 10096–10103. [Google Scholar] [CrossRef]
  39. Joung, I.S.; Cheatham, T.E. Determination of alkali and halide monovalent ion parameters for use in explicitly solvated biomolecular simulations. J. Phys. Chem. B 2008, 112, 9020–9041. [Google Scholar] [CrossRef] [Green Version]
  40. Li, P.F.; Merz, K.M. Taking into account the ion-induced dipole interaction in the nonbonded model of ions. J. Chem. Theor. Comput. 2014, 10, 289–297. [Google Scholar] [CrossRef] [Green Version]
  41. Case, D.A.; Ben-Shalom, I.Y.; Brozell, S.R.; Cerutti, D.S.; Cheatham, I.T.E.; Cruzeiro, V.W.D.; Darden, T.A.; Duke, R.E.; Ghoreishi, D.; Giambasu, G.; et al. AMBER 2019. University of California, San Francisco. Available online: ambermd.org (accessed on 6 March 2023).
  42. Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graph. Modell. 1996, 14, 33–38. [Google Scholar] [CrossRef]
  43. Elstner, M.; Porezag, D.; Jungnickel, G.; Elsner, J.; Haugk, M.; Frauenheim, T.; Suhai, S.; Seifert, G. Self-consistent-charge density-functional tight-binding method for simulations of complex materials properties. Phys. Rev. B 1998, 58, 7260–7268. [Google Scholar] [CrossRef]
  44. Krüger, T.; Elstner, M.; Schiffels, P.; Frauenheim, T. Validation of the density-functional based tight-binding approximation method for the calculation of reaction energies and other data. J. Chem. Phys. 2005, 122, 114110. [Google Scholar] [CrossRef]
  45. Seabra, G.D.d.M.; Walker, R.C.; Elstner, M.; Case, D.A.; Roitberg, A.E. Implementation of the SCC-DFTB method for hybrid QM/MM simulations within the amber molecular dynamics package. J. Phys. Chem. A 2007, 111, 5655–5664. [Google Scholar] [CrossRef] [Green Version]
  46. Gaus, M.; Cui, Q.; Elstner, M. DFTB3: Extension of the self-consistent-charge density-functional tight-binding method (SCC-DFTB). J. Chem. Theor. Comput. 2012, 7, 931–948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Darden, T.; York, D.; Pedersen, L. Particle mesh Ewald–an N⸱log(N) method for Ewald sums in large systems. J. Chem. Phys. 1993, 98, 10089–10092. [Google Scholar] [CrossRef] [Green Version]
  48. Essmann, U.; Perera, L.; Berkowitz, M.L.; Darden, T.; Lee, H.; Pedersen, L.G. A smooth particle mesh Ewald method. J. Chem. Phys. 1995, 103, 8577–8593. [Google Scholar] [CrossRef] [Green Version]
  49. Loncharich, R.J.; Brooks, B.R.; Pastor, R.W. Langevin dynamics of peptides: The frictional dependence of isomerization rates of N-acetylalanyl-N′-methylamide. Biopolymers 1992, 32, 523–535. [Google Scholar] [CrossRef]
  50. Berendsen, H.J.C.; Postma, J.P.M.; van Gunsteren, W.F.; Dinola, A.; Haak, J.R. Molecular dynamics with coupling to an external bath. J. Chem. Phys. 1984, 81, 3684–3690. [Google Scholar] [CrossRef] [Green Version]
  51. Grossfield, A. Version 2.0.10.1. Available online: http://membrane.urmc.rochester.edu/content/wham/ (accessed on 6 May 2021).
  52. Kumar, S.; Rosenberg, J.M.; Bouzida, D.; Swendsen, R.H.; Kollman, P.A. Multidimensional free-energy calculations using the weighted histogram analysis method. J. Comput. Chem. 1995, 16, 1339–1350. [Google Scholar] [CrossRef]
  53. Roux, B. The calculation of the potential of mean force using computer-simulations. Comput. Phys. Commun. 1995, 91, 275–282. [Google Scholar] [CrossRef]
  54. Bonner, W.A. Parity violation and the evolution of biomolecular homochirality. Chirality 2000, 12, 114–126. [Google Scholar] [CrossRef]
  55. Joyce, G.F.; Visser, G.M.; van Boeckel, C.A.; van Boom, J.H.; Orgel, L.E.; van Westrenen, J. Chiral selection in poly(C)-directed synthesis of oligo(G). Nature 1984, 310, 602–604. [Google Scholar] [CrossRef] [PubMed]
  56. Tupper, A.S.; Shi, K.; Higgs, P.G. The role of templating in the emergence of RNA from the prebiotic chemical mixture. Life 2017, 7, 41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Bolli, M.; Micura, R.; Eschenmoser, A. Pyranosyl-RNA: Chiroselective self-assembly of base sequences by ligative oligomerization of tetranucleotide-2′,3′-cyclophosphates (with a commentary concerning the origin of biomolecular homochirality). Chem. Biol. 1997, 4, 309–320. [Google Scholar] [CrossRef] [Green Version]
  58. Bada, J.L.; Miller, S.L. Racemization and the origin of optically active organic compounds in living organisms. Biosystems 1987, 20, 21–26. [Google Scholar] [CrossRef]
  59. Tamura, K. Perspectives on the origin of biological homochirality on Earth. J. Mol. Evol. 2019, 87, 143–146. [Google Scholar] [CrossRef]
  60. Weiner, A.M.; Maizels, N. tRNA-like structures tag the 3′ ends of genomic RNA molecules for replication: Implications for the origin of protein synthesis. Proc. Natl. Acad. Sci. USA 1987, 84, 7383–7387. [Google Scholar] [CrossRef] [Green Version]
  61. Kumar, R.K.; Yarus, M. RNA-catalyzed amino acid activation. Biochemistry 2001, 40, 6998–7004. [Google Scholar] [CrossRef]
Figure 1. (A) Schematic view of aminoacylation reaction in an RNA minihelix. An aminoacyl moiety is transferred from 5′-phosphate of deoxythymine (dT) of the oligonucleotide to the 3′-hydroxyl group of adenosine (A) of the minihelix. Here aa represents amino acid. The 6-base pair nucleotide sequences surrounded by the blue-dotted rectangle with aa of alanine were modeled in this simulation study. (B) A reaction scheme for the aminoacylation found in the RNA minihelix. Distances between carbonyl carbon (Ccarb) and 3′-hydroxyl oxygen (O3′) atoms, d (Ccarb…O3′), and between 3′-hydrogen atom (H3′) and O3′, d (H3′ …O3′) are drawn with red- and green-dotted lines, respectively, which are used for defining the reaction coordinate for umbrella sampling. The distance between the nitrogen atom of alanine (N) and O3′, d (N…O3′), and between N and the phosphorus atom of 5′-phosphate group (P5′) are drawn with blue- and black-dotted lines, respectively. The abbreviations of atoms used in this study are also described. (C) The Bürgi–Dunitz (BD) angle is defined as the O3′…Ccarb=Ocarb angle in the model. (D) The Flippin–Lodge (FL) angle, defined as the angle between the plane formed by O3′, Ccarb and Ocarb atoms and that perpendicular to the carbonyl plane formed by Ccarb, Ocarb, Cα and Ob atoms (shown as the orange parallelogram in (C)), where Ob is the bridging phosphate oxygen atom. In the study, the FL angle has positive values when the plane formed O3′, Ccarb and Ocarb tilts to the Ob atom.
Figure 1. (A) Schematic view of aminoacylation reaction in an RNA minihelix. An aminoacyl moiety is transferred from 5′-phosphate of deoxythymine (dT) of the oligonucleotide to the 3′-hydroxyl group of adenosine (A) of the minihelix. Here aa represents amino acid. The 6-base pair nucleotide sequences surrounded by the blue-dotted rectangle with aa of alanine were modeled in this simulation study. (B) A reaction scheme for the aminoacylation found in the RNA minihelix. Distances between carbonyl carbon (Ccarb) and 3′-hydroxyl oxygen (O3′) atoms, d (Ccarb…O3′), and between 3′-hydrogen atom (H3′) and O3′, d (H3′ …O3′) are drawn with red- and green-dotted lines, respectively, which are used for defining the reaction coordinate for umbrella sampling. The distance between the nitrogen atom of alanine (N) and O3′, d (N…O3′), and between N and the phosphorus atom of 5′-phosphate group (P5′) are drawn with blue- and black-dotted lines, respectively. The abbreviations of atoms used in this study are also described. (C) The Bürgi–Dunitz (BD) angle is defined as the O3′…Ccarb=Ocarb angle in the model. (D) The Flippin–Lodge (FL) angle, defined as the angle between the plane formed by O3′, Ccarb and Ocarb atoms and that perpendicular to the carbonyl plane formed by Ccarb, Ocarb, Cα and Ob atoms (shown as the orange parallelogram in (C)), where Ob is the bridging phosphate oxygen atom. In the study, the FL angle has positive values when the plane formed O3′, Ccarb and Ocarb tilts to the Ob atom.
Life 13 00722 g001
Figure 2. Simulation system for QM/MM umbrella sampling MD calculations for the modeled 6-base pair RNA minihelix. The image is for the D-Ala system. The residues treated with QM calculation are represented by a thick stick model. Atom colors are: oxygen (red), nitrogen (blue), carbon (cyan), phosphorus (brown), and hydrogen (white). The ions of sodium (blue), chloride (cyan) and magnesium (pink) are shown by spheres. Solvent water molecules are rendered as a cyan surface. The image was created using VMD [42].
Figure 2. Simulation system for QM/MM umbrella sampling MD calculations for the modeled 6-base pair RNA minihelix. The image is for the D-Ala system. The residues treated with QM calculation are represented by a thick stick model. Atom colors are: oxygen (red), nitrogen (blue), carbon (cyan), phosphorus (brown), and hydrogen (white). The ions of sodium (blue), chloride (cyan) and magnesium (pink) are shown by spheres. Solvent water molecules are rendered as a cyan surface. The image was created using VMD [42].
Life 13 00722 g002
Figure 3. Conceptual image of QM/MM umbrella sampling MD simulations. The atom coloring scheme is the same as in Figure 2.
Figure 3. Conceptual image of QM/MM umbrella sampling MD simulations. The atom coloring scheme is the same as in Figure 2.
Life 13 00722 g003
Figure 4. Potential of mean forces for aminoacylation reactions with L/D-Ala in the modeled RNA minihelix. The reaction coordinate (RC) is defined as follows: RC = d (H3′…O3′) − d (Ccarb…O3′), where d (H3′…O3′) is the distance between H3′ and O3′ (Figure 1B, green-dotted line) and d (Ccarb…O3′) is that between Ccarb and O3′ (Figure 1B, red-dotted line). Statistical uncertainties of data point estimated by Monte Carlo bootstrap error analysis are shown with error bars, but their magnitudes are too small to see in the plot. The values in the plot are the barrier heights for L-Ala (black) and D-Ala (red) systems.
Figure 4. Potential of mean forces for aminoacylation reactions with L/D-Ala in the modeled RNA minihelix. The reaction coordinate (RC) is defined as follows: RC = d (H3′…O3′) − d (Ccarb…O3′), where d (H3′…O3′) is the distance between H3′ and O3′ (Figure 1B, green-dotted line) and d (Ccarb…O3′) is that between Ccarb and O3′ (Figure 1B, red-dotted line). Statistical uncertainties of data point estimated by Monte Carlo bootstrap error analysis are shown with error bars, but their magnitudes are too small to see in the plot. The values in the plot are the barrier heights for L-Ala (black) and D-Ala (red) systems.
Life 13 00722 g004
Figure 5. Representative structures of the aminoacylation sites for the L/D-Ala systems at various reaction coordinates (RCs). The upper panel shows chemical structures corresponding to the progress of the reaction. The 3′-oxygen (O3′) and carbonyl carbon (Ccarb) atoms are indicated by asterisk and hash, respectively, in snapshots at RC of −4.0 Å and the corresponding chemical structure at the upper panel. Structures at RC = −0.5 Å correspond to the transition states for L and D-Ala systems. Distances between nitrogen atom of amino group in alanine and phosphorous atom of 5′-phosphate group are shown with white arrows in the snapshots at the transition states. The coloring scheme in the snapshots is the same as in Figure 3. In the chemical structures at the upper panel, oxygen and nitrogen atoms are drawn in red and blue, respectively, and carbon, phosphorus, and hydrogen atoms are drawn in black.
Figure 5. Representative structures of the aminoacylation sites for the L/D-Ala systems at various reaction coordinates (RCs). The upper panel shows chemical structures corresponding to the progress of the reaction. The 3′-oxygen (O3′) and carbonyl carbon (Ccarb) atoms are indicated by asterisk and hash, respectively, in snapshots at RC of −4.0 Å and the corresponding chemical structure at the upper panel. Structures at RC = −0.5 Å correspond to the transition states for L and D-Ala systems. Distances between nitrogen atom of amino group in alanine and phosphorous atom of 5′-phosphate group are shown with white arrows in the snapshots at the transition states. The coloring scheme in the snapshots is the same as in Figure 3. In the chemical structures at the upper panel, oxygen and nitrogen atoms are drawn in red and blue, respectively, and carbon, phosphorus, and hydrogen atoms are drawn in black.
Life 13 00722 g005
Figure 6. (A) Bürgi–Dunitz (BD) angles averaged in each window as a function of the reaction coordinate from −4 to 0 Å. The black broken line corresponds to the angle of 105°, which is measured in small-molecule crystal structures that contained amine and ketone carbonyl in the original paper of Bürgi and Dunitz [21]. (B) Flippin–Lodge (FL) angles averaged in each window as a function of the reaction coordinate from −4 to 0 Å. An angle of 0° is indicated by the black broken line. Error bars in (A) and (B) indicate the standard deviations.
Figure 6. (A) Bürgi–Dunitz (BD) angles averaged in each window as a function of the reaction coordinate from −4 to 0 Å. The black broken line corresponds to the angle of 105°, which is measured in small-molecule crystal structures that contained amine and ketone carbonyl in the original paper of Bürgi and Dunitz [21]. (B) Flippin–Lodge (FL) angles averaged in each window as a function of the reaction coordinate from −4 to 0 Å. An angle of 0° is indicated by the black broken line. Error bars in (A) and (B) indicate the standard deviations.
Life 13 00722 g006
Figure 7. (A) Distances between N and O3′ atoms (d (N…O3′) shown in Figure 1B with blue-dotted line), (B) distance between N and P5′ atoms (d (N…P5′) shown in Figure 1B with black-dotted line) (C) dihedral angle for Ob-Ccarb-Cα-N (τ (Ob-Ccarb-Cα-N)) averaged in each window as a function of the reaction coordinate (RC) from −4 to 0 Å. (D) Schematic representation of the dihedral angle τ (Ob-Ccarb-Cα-N), defining the rotation of bond Ob-Ccarb around Ccarb-Cα with respect to Cα-N. A positive sign corresponds to a clockwise rotation. In this figure, the τ (Ob-Ccarb-Cα-N) is −60°. The coloring scheme is the same as in Figure 3. Error bars in (A), (B) and (C) indicate the standard deviations.
Figure 7. (A) Distances between N and O3′ atoms (d (N…O3′) shown in Figure 1B with blue-dotted line), (B) distance between N and P5′ atoms (d (N…P5′) shown in Figure 1B with black-dotted line) (C) dihedral angle for Ob-Ccarb-Cα-N (τ (Ob-Ccarb-Cα-N)) averaged in each window as a function of the reaction coordinate (RC) from −4 to 0 Å. (D) Schematic representation of the dihedral angle τ (Ob-Ccarb-Cα-N), defining the rotation of bond Ob-Ccarb around Ccarb-Cα with respect to Cα-N. A positive sign corresponds to a clockwise rotation. In this figure, the τ (Ob-Ccarb-Cα-N) is −60°. The coloring scheme is the same as in Figure 3. Error bars in (A), (B) and (C) indicate the standard deviations.
Life 13 00722 g007
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

Ando, T.; Tamura, K. Mechanism of Chiral-Selective Aminoacylation of an RNA Minihelix Explored by QM/MM Free-Energy Simulations. Life 2023, 13, 722. https://doi.org/10.3390/life13030722

AMA Style

Ando T, Tamura K. Mechanism of Chiral-Selective Aminoacylation of an RNA Minihelix Explored by QM/MM Free-Energy Simulations. Life. 2023; 13(3):722. https://doi.org/10.3390/life13030722

Chicago/Turabian Style

Ando, Tadashi, and Koji Tamura. 2023. "Mechanism of Chiral-Selective Aminoacylation of an RNA Minihelix Explored by QM/MM Free-Energy Simulations" Life 13, no. 3: 722. https://doi.org/10.3390/life13030722

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

Article Metrics

Back to TopTop