Refinement of the Cornell et al. Nucleic Acids Force Field Based on Reference Quantum Chemical Calculations of Glycosidic Torsion ProfilesClick to copy article linkArticle link copied!
Abstract
We report a reparameterization of the glycosidic torsion χ of the Cornell et al. AMBER force field for RNA, χOL. The parameters remove destabilization of the anti region found in the ff99 force field and thus prevent formation of spurious ladder-like structural distortions in RNA simulations. They also improve the description of the syn region and the syn–anti balance as well as enhance MD simulations of various RNA structures. Although χOL can be combined with both ff99 and ff99bsc0, we recommend the latter. We do not recommend using χOL for B-DNA because it does not improve upon ff99bsc0 for canonical structures. However, it might be useful in simulations of DNA molecules containing syn nucleotides. Our parametrization is based on high-level QM calculations and differs from conventional parametrization approaches in that it incorporates some previously neglected solvation-related effects (which appear to be essential for obtaining correct anti/high-anti balance). Our χOL force field is compared with several previous glycosidic torsion parametrizations.
Introduction
Methods
Selection of Model Molecules
Levels of Theory
Geometry Optimizations and Constraints
Solvent Models
Obtaining the Torsion Profiles
Derivation of χ Parameters
χOL parameter | ||||
---|---|---|---|---|
nucleoside | torsion (atom types) | n | Vn/2 | ϕ |
A | O4′–C1′–N9–C8 | 1 | 0.9656 | 68.79 |
(OS-CT-N*-C2) | 2 | 1.0740 | 15.64 | |
3 | 0.4575 | 171.58 | ||
4 | 0.3092 | 19.09 | ||
G | O4′–C1′–N9–C8 | 1 | 0.7051 | 74.76 |
(OS-CT-N*-CK) | 2 | 1.0655 | 6.23 | |
3 | 0.4427 | 168.65 | ||
4 | 0.2560 | 3.97 | ||
C | O4′–C1′–N1–C6 | 1 | 1.2251 | 146.99 |
(OS-CT-N*-C1) | 2 | 1.6346 | 16.48 | |
3 | 0.9375 | 185.88 | ||
4 | 0.3103 | 32.16 | ||
U(T) | O4′–C1′–N1–C6 | 1 | 1.0251 | 149.88 |
(OS-CT-N*-CM) | 2 | 1.7488 | 16.76 | |
3 | 0.5815 | 179.35 | ||
4 | 0.3515 | 16.00 |
C1 and C2 are new atom types for C introduced to distinguish A from G and C from U (T). The parameters can be downloaded from http://fch.upol.cz/en/rna_chi_ol/.
MD Simulations of RNA and DNA Duplexes
Results and Discussion
Choice of the Method for Geometry Optimization
Choice of Method for Single-Point Calculations
Dependence of the χ Profile and Dihedral Term on Sugar Conformation and Type
Effects of Geometry Relaxation
Comparing χ Parameters
Anti Minimum and Relative Anti/High-Anti Stability
χ Contribution to Relative Anti/High-Anti Stability and Ladder-Like Structures
ΔEanti/high-anti,dih [kcal/mol] | |||||
---|---|---|---|---|---|
parameterization | A | G | C | U(T) | average |
ff94 | 1.9 | 1.9 | 1.9 | 1.9 | 1.9 |
ff98/99 | 1.3 | 1.3 | 1.3 | 1.3 | 1.3 |
χODE | 2.0 | 2.0 | 1.8 | 1.8 | 1.9 |
χYIL | 0.8 | 0.5 | 0.0 | 0.5 | 0.5 |
χvacb | 1.8 | 1.7 | 1.1 | 1.9 | 1.6 |
χOL-DFT | 0.5 | 0.5 | 0.8 | 1.2 | 0.8 |
χOL | 0.9 | 0.8 | 0.4 | 0.9 | 0.8 |
The more positive the anti/high-anti value, the stronger the stabilization of the high-anti conformation. Results with and without bsc0 correction are identical.
The χ dihedral term was derived in the same way as in χOL-DFT, but based on gas phase QM data; see text.
Syn Region
Torsion Barriers
MD Simulations of A-RNA Duplexes
parameter | X-ray | no χ correction | χYIL | χOL-DFT | χOL |
---|---|---|---|---|---|
χ/deg | 197.1 ± 4.4 | 203.1 ± 9.2 | 196.0 | 197.4 | 199.1 |
209.4 ± 12.7 | 194.0 | 196.3 | 196.7 | ||
P/deg | 17.7 ± 6.0 | 19.3 ± 13.5 | 15.4 | 19.2 | 17.4 |
27.5 ± 16.9 | 13.4 | 17.5 | 17.1 | ||
minor groove width/Å | 15.4 ± 0.1 | 15.3 ± 0.6 | 15.2 | 15.3 | 15.3 |
15.0 ± 0.6 | 14.9 | 15.1 | 14.8 | ||
major groove width /Å | 14.7 ± 1.5 | 15.9 ± 2.9 | 19.0 | 17.5 | 17.9 |
18.9 ± 3.2 | 22.1 | 19.8 | 22.3 | ||
slide/Å | –1.70 ± 0.25 | –1.69 ± 0.50 | –2.07 | –1.94 | –1.90 |
-1.89 ± 0.57 | -2.35 | -2.11 | -2.30 | ||
roll/deg | 8.1 ± 4.1 | 9.7 ± 6.1 | 4.6 | 7.1 | 6.7 |
8.5 ± 6.2 | 3.0 | 6.4 | 4.0 | ||
propeller/deg | –12.5 ± 4.5 | –13.7 ± 8.5 | –6.3 | –10.7 | –9.7 |
-12.5 ± 8.7 | -4.3 | -9.8 | -7.4 | ||
X-displ./Å | –4.45 ± 1.18 | –4.85 ± 1.60 | –5.01 | –5.07 | –4.95 |
-5.35 ± 2.18 | -5.50 | -5.49 | -5.91 | ||
inclination/deg | 15.2 ± 8.3 | 18.0 ± 11.0 | 8.8 | 13.4 | 12.7 |
16.3 ± 11.7 | 5.5 | 12.2 | 8.0 | ||
helical twist/deg | 32.3 ± 3.6 | 31.7 ± 4.1 | 29.7 | 30.5 | 30.4 |
31.1 ± 4.9 | 28.6 | 29.8 | 28.6 | ||
rmsd/Å | 1.04 | 1.21 | 1.06 | 1.07 | |
1.36 | 1.85 | 1.43 | 1.90 |
Standard deviations are shown for the unmodified force fields for orientation, and they are very similar for the other force fields. RMSD is mass weighted for all atoms.
Sensitivity of the A-RNA Structure to χ Potential
Basic Sampling of the A-RNA Conformational Space
MD Simulation of B-DNA
parameter | X-ray | ff99bsc0 | ff99bsc0 χYIL | ff99bsc0 χODE | ff99bsc0 χOL-DFT | ff99bsc0 χOL |
---|---|---|---|---|---|---|
χ/deg | 243.6 ± 14.7 | 243.3 ± 18.2 | 223.1 | 244.4 | 229.1 | 231.4 |
P/deg | 129.2 ± 26.7 | 130.4 ± 31.6 | 105.1 | 133.5 | 118.4 | 115.6 |
minor groove width /Å | 10.3 ± 1.0 | 11.5 ± 1.1 | 12.6 | 11.4 | 11.7 | 12.3 |
major groove width /Å | 17.3 ± 0.7 | 19.1 ± 1.9 | 21.5 | 18.7 | 20.5 | 20.2 |
slide/Å | 0.07 ± 0.53 | –0.41 ± 0.58 | –1.20 | –0.36 | –0.90 | –0.83 |
roll/deg | 1.98 ± 3.41 | 3.64 ± 5.22 | 2.76 | 3.53 | 3.03 | 4.24 |
propeller/deg | –13.3 ± 5.94 | –12.5 ± 7.9 | –8.5 | –11.5 | –11.1 | –11.0 |
X-displ./Å | –0.23 ± 0.53 | –1.65 ± 1.73 | –2.82 | –1.44 | –2.19 | –2.33 |
inclination/deg | 4.0 ± 7.2 | 7.8 ± 10.3 | 5.4 | 6.9 | 5.7 | 8.0 |
helical twist/deg | 35.6 ± 5.2 | 33.5 ± 5.7 | 31.5 | 34.2 | 33.0 | 32.6 |
rmsd/Å | 1.58 | 2.52 | 1.46 | 1.95 | 2.15 |
Standard deviations are only shown for the ff99bsc0 force fields because they are very similar for the other force fields. RMSD is mass weighted for all atoms.
Conclusions
Using the same (usually QM-optimized) geometry for deriving the torsion parameters as differences between the QM and MM χ energies may introduce significant errors in the resulting profiles. Instead, geometry for the MM single-point calculations should be optimized at the MM level.
Solvation-related effects considerably influence the resulting χ torsion profile. For instance, their inclusion results in stabilization of the anti region typical for A-RNA with respect to the high-anti region typical for B-DNA. It appears that appropriate balance of the anti and high-anti structures in RNA systems can only be obtained when the solvation effects are considered.
The χ torsion profile is quite sensitive to the level of theory. On the basis of comparisons with estimated reference CCSD(T)/CBS data, we suggest that the MP2/CBS method provides results of sufficient accuracy in this case, while using small basis sets such as 6-31G* with the MP2 method introduces significant errors. The PBE DFT functional does not provide sufficiently accurate results, even when a large (6-311++G(3df,3pd)) basis set is used and a dispersion correction (D-1.06-23) is applied. Results obtained with M06 and M06-2X functionals of Zhao and Truhlar are of similar quality to the PBE-D-1.06-23/LP results and also insufficiently accurate for force-field derivation. Thus, it appears that despite the impressive recent progress in DFT methodology, DFT-based calculations cannot currently match the accuracy of high-quality wave function theory calculations for modeling DNA and RNA backbone segments.
Supporting Information
Torsion profiles of the studied ribo- and deoxyribonucleosides in vacuo and in COSMO and PB solvent models, χOL-DFT torsion parameters, and tables of structural parameters for 1RNA and 2R20′ structures. This material is available free of charge via the Internet at http://pubs.acs.org.
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgment
The authors thank F. Javier Luque for valuable discussions and suggestions regarding the continuum solvation models. This work was supported by the Academy of Sciences of the Czech Republic (grants nos. AV0Z50040507 (J.S.), AV0Z50040702 (J.S.), and GACR 203/09/1476 and P208/11/1822 (J.S.)), the Grant Agency of the Academy of Sciences of the Czech Republic (grants no. P208/10/1742 (P.J.), P301/11/P558 (P.B.), and IAA400040802 (J.S., M.O.)), the Ministry of Education of the Czech Republic (grant 203/09/H046 (P.J., M.O., J.S., and M.Z.)), the NIH R01-GM59306890 (TEC3), and NSF MCA01S027 (TEC3), Student Project PrF_2011_020 of Palacky University, Operational Program Research and Development for Innovations-European Regional Development Fund (projects CZ.1.05/2.1.00/03.0058 and CZ.1.07/2.3.00/20.0017 of the Ministry of Education, Youth and Sports of the Czech Republic), and the HPC-EUROPA2 project (project no. 228398) with the support of the European Community-Research Infrastructure Action of the FP7.
References
This article references 93 other publications.
- 1Weiner, S. J.; Kollman, P. A.; Case, D. A.; Singh, U. C.; Ghio, C.; Alagona, G.; Profeta, S.; Weiner, P. J. Am. Chem. Soc. 1984, 106 (3) 765– 784Google ScholarThere is no corresponding record for this reference.
- 2Weiner, S. J.; Kollman, P. A.; Nguyen, D. T.; Case, D. A. J. Comput. Chem. 1986, 7 (2) 230– 252Google ScholarThere is no corresponding record for this reference.
- 3Mackerell, A. D. J. Comput. Chem. 2004, 25 (13) 1584– 1604Google ScholarThere is no corresponding record for this reference.
- 4Orozco, M.; Noy, A.; Perez, A. Curr. Opin. Struct. Biol. 2008, 18 (2) 185– 193Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXkvFCls74%253D&md5=81ed332ed9f1903d8e13b42d1d1106bfRecent advances in the study of nucleic acid flexibility by molecular dynamicsOrozco, Modesto; Noy, Agnes; Perez, AlbertoCurrent Opinion in Structural Biology (2008), 18 (2), 185-193CODEN: COSBEF; ISSN:0959-440X. (Elsevier B.V.)A review. The recent use of mol. dynamics (MD) simulations to study flexibility of nucleic acids has been reviewed from an anal. of the publications appearing in the past two years (from 2005 till date). Despite the existence of some unsolved problems in the methodologies, these years have been witness to major advances in the field. Based on a crit. review of the most recent contributions, excitement exists on the expected evolution of the field in the next years.
- 5Fulle, S.; Gohlke, H. J. Mol. Recognit. 2010, 23 (2) 220– 231Google Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXitVShsb0%253D&md5=1f0f7c926ad5672ed3765328a2ea0888Molecular recognition of RNA: challenges for modeling interactions and plasticityFulle, Simone; Gohlke, HolgerJournal of Molecular Recognition (2010), 23 (2), 220-231CODEN: JMORE4; ISSN:0952-3499. (John Wiley & Sons Ltd.)A review. There is growing interest in mol. recognition processes of RNA because of RNA's widespread involvement in biol. processes. Computational approaches are increasingly used for analyzing and predicting binding to RNA, fuelled by encouraging progress in developing simulation, free energy, and docking methods for nucleic acids. These developments take into account challenges regarding the energetics of RNA-ligand binding, RNA plasticity, and the presence of water mols. and ions in the binding interface. Accordingly, the authors detail advances in force field and scoring function development for mol. dynamics (MD) simulations, free energy computations, and docking calcns. of nucleic acid complexes. Furthermore, the authors present methods that can detect moving parts within RNA structures based on graph-theor. approaches or normal mode anal. (NMA). As an example of the successful use of these developments, the authors discuss recent structure-based drug design approaches that focus on the bacterial ribosomal A-site RNA as a drug target.
- 6MacKerell, A. D. J. Phys. Chem. B 2009, 113 (10) 3235– 3244Google ScholarThere is no corresponding record for this reference.
- 7Foloppe, N.; Hartmann, B.; Nilsson, L.; MacKerell, A. D. Biophys. J. 2002, 82 (3) 1554– 1569Google ScholarThere is no corresponding record for this reference.
- 8Cornell, W. D.; Cieplak, P.; Bayly, C. I.; Gould, I. R.; Merz, K. M.; Ferguson, D. M.; Spellmeyer, D. C.; Fox, T.; Caldwell, J. W.; Kollman, P. A. J. Am. Chem. Soc. 1995, 117 (19) 5179– 5197Google ScholarThere is no corresponding record for this reference.
- 9Mackerell, A. D.; Wiorkiewiczkuczera, J.; Karplus, M. J. Am. Chem. Soc. 1995, 117 (48) 11946– 11975Google ScholarThere is no corresponding record for this reference.
- 10Cheatham, T. E.; Kollman, P. A. J. Mol. Biol. 1996, 259 (3) 434– 444Google ScholarThere is no corresponding record for this reference.
- 11Yang, L. Q.; Pettitt, B. M. J. Phys. Chem. 1996, 100 (7) 2564– 2566Google ScholarThere is no corresponding record for this reference.
- 12Cheatham, T. E.; Kollman, P. A. Structure 1997, 5 (10) 1297– 1311Google ScholarThere is no corresponding record for this reference.
- 13Cheatham, T. E.; Crowley, M. F.; Fox, T.; Kollman, P. A. Proc. Natl. Acad. Sci. U.S.A. 1997, 94 (18) 9626– 9630Google ScholarThere is no corresponding record for this reference.
- 14Langley, D. R. J. Biomol. Struct. Dyn. 1998, 16 (3) 487– 509Google Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXhtlanu7g%253D&md5=73908d8477291848c077e630d87273bfMolecular dynamic simulations of environment and sequence dependent DNA conformations: the development of the BMS nucleic acid force field and comparison with experimental resultsLangley, David R.Journal of Biomolecular Structure & Dynamics (1998), 16 (3), 487-509CODEN: JBSDD6; ISSN:0739-1102. (Adenine Press)Mol. dynamic (MD) simulations using the BMS nucleic acid force field produce environment and sequence dependent DNA conformations that closely mimic exptl. derived structures. The parameters were initially developed to reproduce the potential energy surface, as defined by quantum mechanics, for a set of small mols. that can be used as the building blocks for nucleic acid macromols. (di-Me phosphate, cyclopentane, THF, etc.). Then the dihedral parameters were fine tuned using a series of condensed phase MD simulations of DNA and RNA (in zero added salt, 4M NaCl, and 75% ethanol solns.). In the tuning process the free energy surface for each dihedral was derived from the MD ensemble and fitted to the conformational distributions and populations obsd. in 87 A- and B-DNA x-ray and 17 B-DNA NMR structures. Over 41 ns of MD simulations are presented which demonstrate that the force field is capable of producing stable trajectories, in the correct environments, of A-DNA, double stranded A- form RNA, B-DNA, Z-DNA, and a netropsin-DNA complex that closely reproduce the exptl. detd. and/or canonical DNA conformations. Frequently the MD averaged structure is closer to the exptl. detd. structure than to the canonical DNA conformation. MD simulations of A- to B- and B- to A-DNA transitions are also shown. A-DNA simulations in a low salt environment cleanly convert into the B-DNA conformation and converge into the RMS space sampled by a low salt simulation of the same sequence starting from B-DNA. In MD simulations using the BMS force field the B-form of d(GGGCCC)2 in a 75% ethanol soln. converts into the A-form. Using the same methodol., parameters, and conditions the A-form of d(AAATTT)2 correctly converts into the B-DNA conformation. These studies demonstrate that the force field is capable of reproducing both environment and sequence dependent DNA structures. The 41 ns of MD simulations presented in this paper paint a global picture which suggests that the DNA structures obsd. in low salt solns. are largely due to the favorable internal energy brought about by the nearly uniform screening of the DNA electrostatics. While the conformations sampled in high salt or mixed solvent environments occur from selective and asym. screening of the phosphate groups and DNA grooves, resp., brought about by sequence induced ion and solvent packing.
- 15MacKerell, A. D. In Molecular Modeling of Nucleic Acids; Leontis, N. B.; SantaLucia, J., Eds.; American Chemical Society: Washington, DC, 1998; Vol. 682, p 304– 311.Google ScholarThere is no corresponding record for this reference.
- 16Foloppe, N.; MacKerell, A. D. J. Phys. Chem. B 1998, 102 (34) 6669– 6678Google ScholarThere is no corresponding record for this reference.
- 17Cheatham, T. E.; Cieplak, P.; Kollman, P. A. J. Biomol. Struct. Dyn. 1999, 16 (4) 845– 862Google Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXitF2jsb8%253D&md5=d3789449f1dd0d2ed769a22316469d4bA modified version of the Cornell et al. force field with improved sugar pucker phases and helical repeatCheatham, Thomas E., III; Cieplak, Piotr; Kollman, Peter A.Journal of Biomolecular Structure & Dynamics (1999), 16 (4), 845-862CODEN: JBSDD6; ISSN:0739-1102. (Adenine Press)We have examd. some subtle parameter modifications to the Cornell et al. force field, which has proven quite successful in reproducing nucleic acid properties, but whose C2'-endo sugar pucker phase and helical repeat for B DNA appear to be somewhat underestimated. Encouragingly, the addn. of a single V2 term involving the atoms C(sp3)-O-(sp3)-C(sp3)-N(sp2), which can be nicely rationalized because of the anomeric effect (lone pairs on oxygen are preferentially oriented relative to the electron withdrawing N), brings the sugar pucker phase of C2'-endo sugars to near perfect agreement with ab initio calcns. (W near 162°). Secondly, the use of high level ab initio calcns. on entire nucleosides (in contrast to smaller model systems necessitated in 1994-95 by computer limitations) lets one improve the χ torsional potential for nucleic acids. Finally, the O(sp3)-C(sp3)-C(sp3)-O(sp3) V2 torsional potential has been empirically adjusted to reproduce the ab initio calcd. relative energy of C2'-endo and C3'-endo nucleosides. These modifications are tested in mol. dynamics simulations of mononucleosides (to assess sugar pucker percentages) and double helixes of DNA and RNA (to assess helical and sequence specific structural properties). In both areas, the modified force field leads to improved agreement with exptl. data.
- 18MacKerell, A. D.; Banavali, N. K. J. Comput. Chem. 2000, 21 (2) 105– 120Google ScholarThere is no corresponding record for this reference.
- 19Foloppe, N.; MacKerell, A. D. J. Comput. Chem. 2000, 21 (2) 86– 104Google ScholarThere is no corresponding record for this reference.
- 20Wang, J. M.; Cieplak, P.; Kollman, P. A. J. Comput. Chem. 2000, 21 (12) 1049– 1074Google ScholarThere is no corresponding record for this reference.
- 21Cheatham, T. E.; Young, M. A. Biopolymers 2000, 56 (4) 232– 256Google ScholarThere is no corresponding record for this reference.
- 22Varnai, P.; Djuranovic, D.; Lavery, R.; Hartmann, B. Nucleic Acids Res. 2002, 30 (24) 5398– 5406Google ScholarThere is no corresponding record for this reference.
- 23Dixit, S. B.; Beveridge, D. L.; Case, D. A.; Cheatham, T. E.; Giudice, E.; Lankas, F.; Lavery, R.; Maddocks, J. H.; Osman, R.; Sklenar, H.; Thayer, K. M.; Varnai, P. Biophys. J. 2005, 89 (6) 3721– 3740Google ScholarThere is no corresponding record for this reference.
- 24Perez, A.; Marchan, I.; Svozil, D.; Sponer, J.; Cheatham, T. E.; Laughton, C. A.; Orozco, M. Biophys. J. 2007, 92 (11) 3817– 3829Google ScholarThere is no corresponding record for this reference.
- 25Svozil, D.; Sponer, J. E.; Marchan, I.; Perez, A.; Cheatham, T. E.; Forti, F.; Luque, F. J.; Orozco, M.; Sponer, J. J. Phys. Chem. B 2008, 112 (27) 8188– 8197Google ScholarThere is no corresponding record for this reference.
- 26Perez, A.; Luque, F. J.; Orozco, M. J. Am. Chem. Soc. 2007, 129 (47) 14739– 14745Google ScholarThere is no corresponding record for this reference.
- 27Lavery, R.; Zakrzewska, K.; Beveridge, D.; Bishop, T. C.; Case, D. A.; Cheatham, T.; Dixit, S.; Jayaram, B.; Lankas, F.; Laughton, C.; Maddocks, J. H.; Michon, A.; Osman, R.; Orozco, M.; Perez, A.; Singh, T.; Spackova, N.; Sponer, J. Nucleic Acids Res. 2010, 38 (1) 299– 313Google ScholarThere is no corresponding record for this reference.
- 28Lankas, F.; Spackova, N.; Moakher, M.; Enkhbayar, P.; Sponer, J. Nucleic Acids Res. 2010, 38 (10) 3414– 3422Google ScholarThere is no corresponding record for this reference.
- 29Fadrna, E.; Spackova, N.; Sarzynska, J.; Koca, J.; Orozco, M.; Cheatham, T. E.; Kulinski, T.; Sponer, J. J. Chem. Theory Comput. 2009, 5 (9) 2514– 2530Google Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhtVWku7vN&md5=7ae168c141e60affa2fcda871a3b3057Single Stranded Loops of Quadruplex DNA As Key Benchmark for Testing Nucleic Acids Force FieldsFadrna, Eva; Spackova, Nada; Sarzynska, Joanna; Koca, Jaroslav; Orozco, Modesto; Cheatham, Thomas E., III; Kulinski, Tadeusz; Sponer, JiriJournal of Chemical Theory and Computation (2009), 5 (9), 2514-2530CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)We have carried out a set of explicit solvent mol. dynamics (MD) simulations on two DNA quadruplex (G-DNA) mols., namely the antiparallel d(G4T4G4)2 dimeric quadruplex with diagonal loops and the parallel-stranded human telomeric monomol. quadruplex d[AGGG(TTAGGG)3] with three propeller loops. The main purpose of the paper was testing of the capability of the MD simulation technique to describe single-stranded topologies of G-DNA loops, which represent a very challenging task for computational methods. The total amt. of conventional and locally enhanced sampling (LES) simulations analyzed in this study exceeds 1.5 μs, while we tested several versions of the AMBER force field (parm99, parmbsc0, and a version with modified glycosidic χ torsion profile) and the CHARMM27 force field. Further, we compared minimal salt and excess salt simulations. Post-processing MM-PBSA (Mol. Mechanics, Poisson-Boltzmann, Surface Area) free energy calcns. are also reported. None of the presently available force fields is accurate enough in describing the G-DNA loops. The imbalance is best seen for the propeller loops, as their exptl. structure is lost within a few ns of std. simulations with all force fields. Among them, parmbsc0 provides results that are clearly closest to the exptl. target values but still not in full agreement. This confirms that the improvement of the γ torsional profile penalizing the γ trans substates in the parmbsc0 parametrization was a step in the right direction, albeit not sufficient to treat all imbalances. The modified χ parameterization appears to rigidify the studied systems but does not change the ultimate outcome of the present simulations. The structures obtained in simulations with the modified χ profile are predetd. by its combination with either parm99 or parmbsc0. Exptl. geometries of diagonal loops of d(G4T4G4)2 are stable in std. simulations on the ∼10 ns time scale but are becoming progressively lost in longer and LES simulations. In addn., the d(G4T4G4)2 quadruplex contains, besides the three genuine binding sites for cations in the channel of its stem, also an ion binding site at each stem-loop junction. This arrangement of five cations in the quadruplex core region is entirely unstable in all 24 simulations that we attempted. Overall, our results confirm that G-DNA loops represent one of the most difficult targets for mol. modeling approaches and should be considered as ref. structures in any future studies aiming to develop or tune nucleic acids force fields.
- 30Reblova, K.; Lankas, F.; Razga, F.; Krasovska, M. V.; Koca, J.; Sponer, J. Biopolymers 2006, 82 (5) 504– 520Google ScholarThere is no corresponding record for this reference.
- 31Blount, K. F.; Breaker, R. R. Nat. Biotechnol. 2006, 24 (12) 1558– 1564Google ScholarThere is no corresponding record for this reference.
- 32Strobel, S. A.; Cochrane, J. C. Curr. Opin. Chem. Biol. 2007, 11 (6) 636– 643Google Scholar32https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXhtlyrtbrP&md5=4fcdfa52636d776969f6c5d939991fb5RNA catalysis: ribozymes, ribosomes, and riboswitchesStrobel, Scott A.; Cochrane, Jesse C.Current Opinion in Chemical Biology (2007), 11 (6), 636-643CODEN: COCBF4; ISSN:1367-5931. (Elsevier B.V.)A review. The catalytic mechanisms employed by RNA are chem. more diverse than initially suspected. Divalent metal ions, nucleobases, ribosyl hydroxyl groups, and even functional groups on metabolic cofactors all contribute to the various strategies employed by RNA enzymes. This catalytic breadth raises intriguing evolutionary questions about how RNA lost its biol. role in some cases, but not in others, and what catalytic roles RNA might still be playing in biol.
- 33Montange, R. K.; Batey, R. T. Annu. Rev. Biophys. 2008, 37, 117– 133Google ScholarThere is no corresponding record for this reference.
- 34Steitz, T. A. Nat. Rev. Mol. Cell Biol. 2008, 9 (3) 242– 253Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXitlGnu7w%253D&md5=e82078dbacb84a56b48507e6b703df8dA structural understanding of the dynamic ribosome machineSteitz, Thomas A.Nature Reviews Molecular Cell Biology (2008), 9 (3), 242-253CODEN: NRMCBP; ISSN:1471-0072. (Nature Publishing Group)A review. Ribosomes, which are central to protein synthesis and convert transcribed mRNAs into polypeptide chains, have been the focus of structural and biochem. studies for >50 yr. The structure of its larger subunit has revealed that the ribosome is a ribozyme with RNA at the heart of its enzymic activity that catalyzes peptide bond formation. Numerous initiation, elongation, and release factors ensure that protein synthesis occurs progressively and with high specificity. In the past few years, high-resoln. structures have provided mol. snapshots of different intermediates in ribosome-mediated translation in at. detail. Together, these studies have revolutionized the understanding of the mechanism of protein synthesis.
- 35Paulsen, R. B.; Seth, P. P.; Swayze, E. E.; Griffey, R. H.; Skalicky, J. J.; Cheatham, T. E., 3rd; Davis, D. R. Proc. Natl. Acad. Sci. U.S.A. 2010, 107 (16) 7263– 7268Google ScholarThere is no corresponding record for this reference.
- 36Reddy, S. Y.; Leclerc, F.; Karplus, M. Biophys. J. 2003, 84 (3) 1421– 1449Google ScholarThere is no corresponding record for this reference.
- 37Besseova, I.; Otyepka, M.; Reblova, K.; Sponer, J. Phys. Chem. Chem. Phys. 2009, 11 (45) 10701– 10711Google ScholarThere is no corresponding record for this reference.
- 38Deng, N. J.; Cieplak, P. Biophys. J. 2010, 98 (4) 627– 636Google ScholarThere is no corresponding record for this reference.
- 39Ricci, C. G.; de Andrade, A. S. C.; Mottin, M.; Netz, P. A. J. Phys. Chem. B 2010, 114 (30) 9882– 9893Google ScholarThere is no corresponding record for this reference.
- 40Auffinger, P.; Westhof, E. Curr. Opin. Struct. Biol. 1998, 8 (2) 227– 236Google Scholar40https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXjt1Gktr0%253D&md5=b13df97d5a7f78d19b8dad6f28b68eceSimulations of the molecular dynamics of nucleic acidsAuffinger, Pascal; Westhof, EricCurrent Opinion in Structural Biology (1998), 8 (2), 227-236CODEN: COSBEF; ISSN:0959-440X. (Current Biology Ltd.)A review with 91 refs. The growing amt. of high quality mol. dynamics simulations generated using the latest methodol. developments and force fields has led to a sharper understanding of the forces underlying the dynamics of biomol. systems, as well as to stimulating insights into the structure and catalysis of nucleic acids. It is now clear that inclusion of long-range electrostatic interactions and of the aq. and ionic environment is necessary for producing realistic and accurate simulations. Yet, many papers hint at a force field and protocol dependence of the results and thus contain the seeds for the future improvements that will e necessary for deepening our understanding of recognition phenomena and folding of nucleic acids.
- 41Bosch, D.; Foloppe, N.; Pastor, N.; Pardo, L.; Campillo, M. J. Mol. Struc.: THEOCHEM 2001, 537, 283– 305Google ScholarThere is no corresponding record for this reference.
- 42Foloppe, N.; MacKerell, A. D. J. Phys. Chem. B 1999, 103 (49) 10955– 10964Google ScholarThere is no corresponding record for this reference.
- 43Mlynsky, V.; Banas, P.; Hollas, D.; Reblova, K.; Walter, N. G.; Sponer, J.; Otyepka, M. J. Phys. Chem. B 2010, 114 (19) 6642– 6652Google ScholarThere is no corresponding record for this reference.
- 44Ode, H.; Matsuo, Y.; Neya, S.; Hoshino, T. J. Comput. Chem. 2008, 29 (15) 2531– 2542Google ScholarThere is no corresponding record for this reference.
- 45Yildirim, I.; Stern, H. A.; Kennedy, S. D.; Tubbs, J. D.; Turner, D. H. J. Chem. Theory Comput. 2010, 6 (5) 1520– 1531Google Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXkvV2lurg%253D&md5=43f9653ecb996aecc3274434b4a24c3eReparameterization of RNA χ Torsion Parameters for the AMBER Force Field and Comparison to NMR Spectra for Cytidine and UridineYildirim, Ilyas; Stern, Harry A.; Kennedy, Scott D.; Tubbs, Jason D.; Turner, Douglas H.Journal of Chemical Theory and Computation (2010), 6 (5), 1520-1531CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)A reparameterization of the torsional parameters for the glycosidic dihedral angle, χ, for the AMBER99 force field in RNA nucleosides is used to provide a modified force field, AMBER99χ. Mol. dynamics simulations of cytidine, uridine, adenosine, and guanosine in aq. soln. using the AMBER99 and AMBER99χ force fields are compared with NMR results. For each nucleoside and force field, 10 individual mol. dynamics simulations of 30 ns each were run. For cytidine with AMBER99χ force field, each mol. dynamics simulation time was extended to 120 ns for convergence purposes. NMR spectroscopy, including one-dimensional (1D) 1H, steady-state 1D 1H nuclear Overhauser effect (NOE), and transient 1D 1H NOE, was used to det. the sugar puckering and preferred base orientation with respect to the ribose of cytidine and uridine. The AMBER99 force field overestimates the population of syn conformations of the base orientation and of C2'-endo sugar puckering of the pyrimidines, while the AMBER99χ force field's predictions are more consistent with NMR results. Moreover, the AMBER99 force field prefers high anti conformations with glycosidic dihedral angles around 310° for the base orientation of purines. The AMBER99χ force field prefers anti conformations around 185°, which is more consistent with the quantum mech. calcns. and known 3D structures of folded ribonucleic acids (RNAs). Evidently, the AMBER99χ force field predicts the structural characteristics of ribonucleosides better than the AMBER99 force field and should improve structural and thermodn. predictions of RNA structures.
- 46Banáš, P.; Hollas, D.; Zgarbova, M.; Jurecka, P.; Orozco, M.; Cheatham, T.; Sponer, J.; Otyepka, M. J. Chem. Theory Comput. 2010, 6 (12) 3836– 3849Google Scholar46https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhsVSisLnK&md5=b02fa19c2fb52e1d36476274c86aaeeaPerformance of Molecular Mechanics Force Fields for RNA Simulations: Stability of UUCG and GNRA HairpinsBanas, Pavel; Hollas, Daniel; Zgarbova, Marie; Jurecka, Petr; Orozco, Modesto; Cheatham, Thomas E., III; Sponer, Jiri; Otyepka, MichalJournal of Chemical Theory and Computation (2010), 6 (12), 3836-3849CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The RNA hairpin loops represent important RNA topologies with indispensable biol. functions in RNA folding and tertiary interactions. 5'-UNCG-3' and 5'-GNRA-3' RNA tetraloops are the most important classes of RNA hairpin loops. Both tetraloops are highly structured with characteristic signature three-dimensional features and are recurrently seen in functional RNAs and ribonucleoprotein particles. Explicit solvent mol. dynamics (MD) simulation is a computational technique which can efficiently complement the exptl. data and provide unique structural dynamics information on the at. scale. Nevertheless, the outcome of simulations is often compromised by imperfections in the parametrization of simplified pairwise additive empirical potentials referred to also as force fields. We have pointed out in several recent studies that a force field description of single-stranded hairpin segments of nucleic acids may be particularly challenging for the force fields. In this paper, we report a crit. assessment of a broad set of MD simulations of UUCG, GAGA, and GAAA tetraloops using various force fields. First, we utilized the three widely used variants of Cornell et al. (AMBER) force fields known as ff94, ff99, and ff99bsc0. Some simulations were also carried out with CHARMM27. The simulations reveal several problems which show that these force fields are not able to retain all characteristic structural features (structural signature) of the studied tetraloops. Then we tested four recent reparameterizations of glycosidic torsion of the Cornell et al. force field (two of them being currently parametrized in our labs.). We show that at least some of the new versions show an improved description of the tetraloops, mainly in the syn glycosidic torsion region of the UNCG tetraloop. The best performance is achieved in combination with the bsc0 parametrization of the α/γ angles. Another critically important region to properly describe RNA mols. is the anti/high-anti region of the glycosidic torsion, where there are significant differences among the tested force fields. The tetraloop simulations are complemented by simulations of short A-RNA stems, which are esp. sensitive to an appropriate description of the anti/high-anti region. While excessive accessibility of the high-anti region converts the A-RNA into a senseless "ladder-like" geometry, excessive penalization of the high-anti region shifts the simulated structures away from typical A-RNA geometry to structures with a visibly underestimated inclination of base pairs with respect to the helical axis.
- 47Fadrna, E.; Spackova, N.; Stefl, R.; Koca, J.; Cheatham, T. E.; Sponer, J. Biophys. J. 2004, 87 (1) 227– 242Google ScholarThere is no corresponding record for this reference.
- 48Reblova, K.; Fadrna, E.; Sarzynska, J.; Kulinski, T.; Kulhanek, P.; Ennifar, E.; Koca, J.; Sponer, J. Biophys. J. 2007, 93 (11) 3932– 3949Google ScholarThere is no corresponding record for this reference.
- 49Faustino, I.; Perez, A.; Orozco, M. Biophys. J. 2010, 99 (6) 1876– 1885Google ScholarThere is no corresponding record for this reference.
- 50Foloppe, N.; MacKerell, A. D. Biophys. J. 1999, 76 (6) 3206– 3218Google ScholarThere is no corresponding record for this reference.
- 51Jurecka, P.; Hobza, P. J. Am. Chem. Soc. 2003, 125 (50) 15608– 15613Google ScholarThere is no corresponding record for this reference.
- 52Mladek, A.; Sponer, J. E.; Jurecka, P.; Banas, P.; Otyepka, M.; Svozil, D.; Sponer, J. J. Chem. Theory Comput. 2010, 6 (12) 3817– 3835Google Scholar52https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhsVegurfK&md5=25b44243e0c15ce1315e85b277cd5aa0Conformational Energies of DNA Sugar-Phosphate Backbone: Reference QM Calculations and a Comparison with Density Functional Theory and Molecular MechanicsMladek, Arnost; Sponer, Judit E.; Jurecka, Petr; Banas, Pavel; Otyepka, Michal; Svozil, Daniel; Sponer, JiriJournal of Chemical Theory and Computation (2010), 6 (12), 3817-3835CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The study investigates electronic structure and gas-phase energetics of the DNA sugar-phosphate backbone via advanced quantum chem. (QM) methods. The anal. has been carried out on biol. relevant backbone conformations composed of 11 canonical BI-DNA structures, 8 pathol. structures with α/γ torsion angles in the g+/t region, and 3 real noncanonical γ-trans structures occurring in the loop region of guanine quadruplex DNA. The influence of backbone conformation on the intrinsic energetics was primarily studied using a model system consisting of two sugar moieties linked together via a phosphodiester bond (SPSOM model). To get the conformation of the studied system fully under control, for each calcn. we have frozen majority of the dihedral angles to their target values. CCSD(T) energies extrapolated to the complete basis set were utilized as ref. values. However, the calcns. show that inclusion of higher-order electron correlation effects for this system is not crucial and complete basis set second-order perturbation calcns. are sufficiently accurate. The ref. QM data are used to assess performance of 10 contemporary d. functionals with the best performance delivered by the PBE-D/TZVPP combination along with the Grimme's dispersion correction, and by the TPSS-D/6-311++G(3df,3pd) augmented by Jurecka's dispersion term. In addn., the QM calcns. are compared to mol. mechanics (MM) model based on the Cornell et al. force field. The destabilization of the pathol. g+/t conformers with respect to the ref. canonical structure and the network of intramol. CH···O interactions were investigated by means of natural bond orbital anal. (NBO) and atoms-in-mols. (AIM) Bader anal. Finally, four addnl. model systems of different sizes were assessed by comparing their energetics to that of the SPSOM system. Energetics of smaller MOSPM model consisting of a sugar moiety linked to a phosphate group and capped with Me and methoxy group on the 5'- and 3'-ends, resp., is fairly similar to that of SPSOM, while the role of undesired intramol. interactions is diminished.
- 53Halkier, A.; Helgaker, T.; Jorgensen, P.; Klopper, W.; Koch, H.; Olsen, J.; Wilson, A. K. Chem. Phys. Lett. 1998, 286 (3–4) 243– 252Google ScholarThere is no corresponding record for this reference.
- 54Halkier, A.; Helgaker, T.; Jorgensen, P.; Klopper, W.; Olsen, J. Chem. Phys. Lett. 1999, 302 (5–6) 437– 446Google ScholarThere is no corresponding record for this reference.
- 55Ahlrichs, R.; Bar, M.; Haser, M.; Horn, H.; Kolmel, C. Chem. Phys. Lett. 1989, 162 (3) 165– 169Google Scholar55https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3cXkt1yrtg%253D%253D&md5=b6aa32e6226a8e11b511e4d09cc60dc8Electronic structure calculations on workstation computers: the program system TURBOMOLEAhlrichs, Reinhart; Baer, Michael; Haeser, Marco; Horn, Hans; Koelmel, ChristophChemical Physics Letters (1989), 162 (3), 165-9CODEN: CHPLBC; ISSN:0009-2614.The basic structure of the program system TURBOMOLE for SCF - including first and second anal. derivs. with respect to nuclear coordinates - and MP2 calcns. is briefly described. The program takes full advantage of all discrete point group symmetries and has only modest - and (partially) adjustable - I/O and background storage requirements. The performance of TURBOMOLE is documented for demonstrative applications.
- 56Weigend, F.; Häser, M. Theor. Chem. Acc. 1997, 97 (1–4) 331– 340Google Scholar56https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXmvFCitL8%253D&md5=14ae7a8718188931367e3a192be50351RI-MP2. First derivatives and global consistencyWeigend, Florian; Haser, MarcoTheoretical Chemistry Accounts (1997), 97 (1-4), 331-340CODEN: TCACFW; ISSN:1432-881X. (Springer)The evaluation of RI-MP2 first derivs. with respect to nuclear coordinates or with respect to an external elec. field is described. The prefix RI indicates the use of an approx. resoln. of identity in the Hilbert space of interacting charge distributions (Coulomb metric), i.e., the use of an auxiliary basis set to approx. charge distributions. The RI technique is applied to first derivs. of the MP2 correlation energy expression while the (restricted) Hartree-Fock ref. is treated in the usual way. Computational savings by a factor of 10 over conventional approaches are demonstrated in an application to porphyrin. The RI approxn. to MP2 derivs. does not entail any significant loss in accuracy. Finally, the relative energetic stabilities of a representative sample of closed-shell mols. built from first and second row elements were investigated by the RI-MP2 approach, and thus it is tested whether such properties that refer to potential energy hypersurfaces in a more global way can be described with similar consistency to the more locally defined derivs.
- 57Werner, H. J.; Knowles, P. J.; Lindh, R.; Manby, F. R.; Schütz, M.; Celani, P.; Korona, T.; Rauhut, G.; Amos, R. D.; Bernhardsson, A.; Berning, A.; Cooper, D. L.; Deegan, M. J. O.; Dobbyn, A. J.; Eckert, F.; Hampel, C.; Hetzer, G.; Lloyd, A. W.; McNicholas, S. J.; Meyer, W.; Mura, M. E.; Nicklass, A.; Palmieri, P.; Pitzer, R.; Schumann, U.; Stoll, H.; Stone, A. J.; Tarroni, R.; Thorsteinsson, T. Molpro Version 2006.1, a package of ab initio programs; 2006; http://www.molpro.net (accessed July 2011).Google ScholarThere is no corresponding record for this reference.
- 58Krishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A. J. Chem. Phys. 1980, 72 (1) 650– 654Google ScholarThere is no corresponding record for this reference.
- 59Clark, T.; Chandrasekhar, J.; Spitznagel, G. W.; Schleyer, P. V. J. Comput. Chem. 1983, 4 (3) 294– 301Google ScholarThere is no corresponding record for this reference.
- 60Gill, P. M. W.; Johnson, B. G.; Pople, J. A. J. Chem. Phys. 1992, 96 (9) 7178– 7179Google ScholarThere is no corresponding record for this reference.
- 61Frisch, M. J.; Pople, J. A.; Binkley, J. S. J. Chem. Phys. 1984, 80 (7) 3265– 3269Google ScholarThere is no corresponding record for this reference.
- 62Jurecka, P.; Cerny, J.; Hobza, P.; Salahub, D. R. J. Comput. Chem. 2007, 28 (2) 555– 569Google ScholarThere is no corresponding record for this reference.
- 63Schafer, A.; Huber, C.; Ahlrichs, R. J. Chem. Phys. 1994, 100 (8) 5829– 5835Google ScholarThere is no corresponding record for this reference.
- 64Klamt, A.; Schuurmann, G. J. Chem. Soc., Perkin Trans. 2 1993, 5) 799– 805Google Scholar64https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXkvVSku7o%253D&md5=a8e9e41514d60ca7f11212cf5c281424COSMO: a new approach to dielectric screening in solvents with explicit expressions for the screening energy and its gradientKlamt, A.; Schueuermann, G.Journal of the Chemical Society, Perkin Transactions 2: Physical Organic Chemistry (1972-1999) (1993), (5), 799-805CODEN: JCPKBH; ISSN:0300-9580.Starting from the screening in conductors, an algorithm for the accurate calcn. of dielec. screening effects in solvents is presented, which leads to rather simple explicit expressions for the screening energy and its analytic gradient with respect to the solute coordinates. Thus geometry optimization of a solute within a realistic dielec. continuum model becomes practicable for the first time. The algorithm is suited for mol. mechanics as well as for any MO algorithm. The implementation into MOPAC and some example applications are reported.
- 65Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, J.A.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A. Gaussian 03, revision D.02; Gaussian, Inc.: Wallingford, CT, 2004.Google ScholarThere is no corresponding record for this reference.
- 66Case, D. A.; Cheatham, T. E.; Darden, T.; Gohlke, H.; Luo, R.; Merz, K. M.; Onufriev, A.; Simmerling, C.; Wang, B.; Woods, R. J. J. Comput. Chem. 2005, 26 (16) 1668– 1688Google ScholarThere is no corresponding record for this reference.
- 67Lu, Q.; Luo, R. J. Chem. Phys. 2003, 119 (21) 11035– 11047Google ScholarThere is no corresponding record for this reference.
- 68Luo, R.; David, L.; Gilson, M. K. J. Comput. Chem. 2002, 23 (13) 1244– 1253Google ScholarThere is no corresponding record for this reference.
- 69Guvench, O.; MacKerell, A. D., Jr. J. Mol. Model. 2008, 14 (8) 667– 679Google Scholar69https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhtVKgtLo%253D&md5=b5bd1c35ed00665a83be39345f88a6feAutomated conformational energy fitting for force-field developmentGuvench, Olgun; MacKerell, Alexander D., Jr.Journal of Molecular Modeling (2008), 14 (8), 667-679CODEN: JMMOFK; ISSN:0948-5023. (Springer GmbH)We present a general conformational-energy fitting procedure based on Monte Carlo simulated annealing (MCSA) for application in the development of mol. mechanics force fields. Starting with a target potential energy surface and an unparametrized mol. mechanics potential energy surface, an optimized set of either dihedral or grid-based correction map (CMAP) parameters is produced that minimizes the root mean squared error RMSE between the parametrized and targeted energies. The fitting is done using an MCSA search in parameter space and consistently converges to the same RMSE irresp. of the randomized parameters used to seed the search. Any no. of dihedral parameters can be simultaneously parametrized, allowing for fitting to multi-dimensional potential energy scans. Fitting options for dihedral parameters include non-uniform weighting of the target data, constraining multiple optimized parameters to the same value, constraining parameters to be no greater than a user-specified max. value, including all or only a subset of multiplicities defining the dihedral Fourier series, and optimization of phase angles in addn. to force consts. The dihedral parameter fitting algorithm's performance is characterized through multi-dimensional fitting of cyclohexane, tetrahydropyran, and hexopyranose monosaccharide energetics, with the latter case having a 30-dimensional parameter space. The CMAP fitting is applied in the context of polypeptides, and is used to develop a parametrization that simultaneously captures the .vphi.,ψ energetics of the alanine dipeptide and the alanine tetrapeptide. Because the dihedral energy term is common to many force fields, we have implemented the dihedral-fitting algorithm in the portable Python scripting language and have made it freely available as "fit_dihedral.py" for download at http://mackerell.umaryland.edu.
- 70Duan, Y.; Wu, C.; Chowdhury, S.; Lee, M. C.; Xiong, G. M.; Zhang, W.; Yang, R.; Cieplak, P.; Luo, R.; Lee, T.; Caldwell, J.; Wang, J. M.; Kollman, P. J. Comput. Chem. 2003, 24 (16) 1999– 2012Google Scholar70https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXovVygsbc%253D&md5=b4531cbf2d90deb9a173ec314dfd7b5dA point-charge force field for molecular mechanics simulations of proteins based on condensed-phase quantum mechanical calculationsDuan, Yong; Wu, Chun; Chowdhury, Shibasish; Lee, Mathew C.; Xiong, Guoming; Zhang, Wei; Yang, Rong; Cieplak, Piotr; Luo, Ray; Lee, Taisung; Caldwell, James; Wang, Junmei; Kollman, PeterJournal of Computational Chemistry (2003), 24 (16), 1999-2012CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)Mol. mechanics models have been applied extensively to study the dynamics of proteins and nucleic acids. Here we report the development of a third-generation point-charge all-atom force field for proteins. Following the earlier approach of Cornell et al., the charge set was obtained by fitting to the electrostatic potentials of dipeptides calcd. using B3LYP/cc-pVTZ/HF/6-31G** quantum mech. methods. The main-chain torsion parameters were obtained by fitting to the energy profiles of Ace-Ala-Nme and Ace-Gly-Nme di-peptides calcd. using MP2/cc-pVTZ//HF/6-31G** quantum mech. methods. All other parameters were taken from the existing AMBER data base. The major departure from previous force fields is that all quantum mech. calcns. were done in the condensed phase with continuum solvent models and an effective dielec. const. of ε = 4. We anticipate that this force field parameter set will address certain crit. short comings of previous force fields in condensed-phase simulations of proteins. Initial tests on peptides demonstrated a high-degree of similarity between the calcd. and the statistically measured Ramanchandran maps for both Ace-Gly-Nme and Ace-Ala-Nme di-peptides. Some highlights of our results include (1) well-preserved balance between the extended and helical region distributions, and (2) favorable type-II poly-proline helical region in agreement with recent expts. Backward compatibility between the new and Cornell et al. charge sets, as judged by overall agreement between dipole moments, allows a smooth transition to the new force field in the area of ligand-binding calcns. Test simulations on a large set of proteins are also discussed.
- 71Dockbregeon, A. C.; Chevrier, B.; Podjarny, A.; Johnson, J.; Debear, J. S.; Gough, G. R.; Gilham, P. T.; Moras, D. J. Mol. Biol. 1989, 209 (3) 459– 474Google ScholarThere is no corresponding record for this reference.
- 72Drew, H. R.; Wing, R. M.; Takano, T.; Broka, C.; Tanaka, S.; Itakura, K.; Dickerson, R. E. Proc. Natl. Acad. Sci.: Biol. 1981, 78 (4) 2179– 2183Google Scholar72https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3MXktVajt74%253D&md5=1278977a90de8b049e2577294d569c45Structure of a B-DNA dodecamer. I. Conformation and dynamicsDrew, Horace R.; Wing, Richard M.; Takano, Tsunehiro; Broka, Christopher; Tanaka, Shoji; Itakura, Keiichi; Dickerson, Richard E.Proceedings of the National Academy of Sciences of the United States of America (1981), 78 (4), 2179-83CODEN: PNASA6; ISSN:0027-8424.The crystal structure of the synthetic DNA dodecamer d(CpGpCpGpApApTpTpCpGpCpG) was refined to a residual error of R = 17.8% at 1.9-Å resoln. (2-σ data). The mol. forms slightly >1 complete turn of right-handed double-stranded B helix. The 2 ends of the helix overlap and interlock minor grooves with neighboring mols. up and down a 21 screw axis, producing a 19° bend in helix axis over the 11-base-pair steps of the dodecamer. In the center of the mol., where perturbation is least, the helix has a mean rotation of 36.9° per step, or 9.8 base pairs per turn. The mean propeller twist (total dihedral angle between base planes) between A·T base pairs in the center of the mol. is 17.3°, and that between C·G pairs on the 2 ends avs. 11.5°. Individual deoxyribose ring conformations, measured by the C5'-C4'-C3'-O3' torsion angle δ, exhibit an approx. Gaussian distribution centered around the C1'-exo position with δav. = 123° and a range of 79-157°. Purine sugars cluster at high δ values, and pyrimidine sugars cluster at lower δ. A tendency toward 2-fold symmetry in sugar conformation about the center of the mol. is detectable in spite of the destruction of ideal 2-fold symmetry by the mol. bending. More strikingly, sugar conformations of paired bases appear to follow a principle of anticorrelation, with δ values lying approx. the same distance to either side of the center value, δ = 123°. This same anticorrelation is also obsd. in other DNA and DNA·RNA structures.
- 73Timsit, Y.; Bombard, S. RNA 2007, 13 (12) 2098– 2107Google ScholarThere is no corresponding record for this reference.
- 74Klosterman, P. S.; Shah, S. A.; Steitz, T. A. Biochemistry 1999, 38 (45) 14784– 14792Google ScholarThere is no corresponding record for this reference.
- 75Aqvist, J. J. Phys. Chem. 1990, 94 (21) 8021– 8024Google ScholarThere is no corresponding record for this reference.
- 76Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R. W.; Klein, M. L. J. Chem. Phys. 1983, 79 (2) 926– 935Google Scholar76https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3sXksF2htL4%253D&md5=a1161334e381746be8c9b15a5e56f704Comparison of simple potential functions for simulating liquid waterJorgensen, William L.; Chandrasekhar, Jayaraman; Madura, Jeffry D.; Impey, Roger W.; Klein, Michael L.Journal of Chemical Physics (1983), 79 (2), 926-35CODEN: JCPSA6; ISSN:0021-9606.Classical Monte Carlo simulations were carried out for liq. H2O in the NPT ensemble at 25° and 1 atm using 6 of the simpler intermol. potential functions for the dimer. Comparisons were made with exptl. thermodn. and structural data including the neutron diffraction results of Thiessen and Narten (1982). The computed densities and potential energies agree with expt. except for the original Bernal-Fowler model, which yields an 18% overest. of the d. and poor structural results. The discrepancy may be due to the correction terms needed in processing the neutron data or to an effect uniformly neglected in the computations. Comparisons were made for the self-diffusion coeffs. obtained from mol. dynamics simulations.
- 77Lu, X. J.; Olson, W. K. Nucleic Acids Res. 2003, 31 (17) 5108– 5121Google ScholarThere is no corresponding record for this reference.
- 78Holroyd, L. F.; van Mourik, T. Chem. Phys. Lett. 2007, 442 (1–3) 42– 46Google ScholarThere is no corresponding record for this reference.
- 79Valdes, H.; Klusak, V.; Pitonak, M.; Exner, O.; Stary, I.; Hobza, P.; Rulisek, L. J. Comput. Chem. 2008, 29 (6) 861– 870Google ScholarThere is no corresponding record for this reference.
- 80Jensen, F. J. Chem. Theory Comput. 2010, 6 (1) 100– 106Google Scholar80https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhtlOksrzM&md5=dc6efcef8f646babf9de2136568222c5An Atomic Counterpoise Method for Estimating Inter- and Intramolecular Basis Set Superposition ErrorsJensen, FrankJournal of Chemical Theory and Computation (2010), 6 (1), 100-106CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)An at. counterpoise method is proposed to calc. ests. of inter- and intramol. basis set superposition errors. The method ests. the basis set superposition error as a sum of at. contributions and can be applied for both independent particle and electron correlation models. The at. counterpoise method provides results very similar to the mol. counterpoise method for intermol. basis set superposition errors at both the HF and MP2 levels of theory with a sequence of increasingly larger basis sets. The advantage of the at. counterpoise method is that it can be applied with equal ease to est. intramol. basis set superposition errors, for which few other methods exist. The at. counterpoise method is computationally quite efficient, requiring typically double the amt. of computer time as required for calcg. the uncorrected energy.
- 81Zhao, Y.; Truhlar, D. G. Theor. Chem. Acc. 2008, 120 (1–3) 215– 241Google Scholar81https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXltFyltbY%253D&md5=c31d6f319d7c7a45aa9b716220e4a422The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionalsZhao, Yan; Truhlar, Donald G.Theoretical Chemistry Accounts (2008), 120 (1-3), 215-241CODEN: TCACFW; ISSN:1432-881X. (Springer GmbH)We present two new hybrid meta exchange-correlation functionals, called M06 and M06-2X. The M06 functional is parametrized including both transition metals and nonmetals, whereas the M06-2X functional is a high-nonlocality functional with double the amt. of nonlocal exchange (2X), and it is parametrized only for nonmetals. The functionals, along with the previously published M06-L local functional and the M06-HF full-Hartree-Fock functionals, constitute the M06 suite of complementary functionals. We assess these four functionals by comparing their performance to that of 12 other functionals and Hartree-Fock theory for 403 energetic data in 29 diverse databases, including ten databases for thermochem., four databases for kinetics, eight databases for noncovalent interactions, three databases for transition metal bonding, one database for metal atom excitation energies, and three databases for mol. excitation energies. We also illustrate the performance of these 17 methods for three databases contg. 40 bond lengths and for databases contg. 38 vibrational frequencies and 15 vibrational zero point energies. We recommend the M06-2X functional for applications involving main-group thermochem., kinetics, noncovalent interactions, and electronic excitation energies to valence and Rydberg states. We recommend the M06 functional for application in organometallic and inorganometallic chem. and for noncovalent interactions.
- 82Zgarbova, M.; Otyepka, M.; Sponer, J.; Hobza, P.; Jurecka, P. Phys. Chem. Chem. Phys. 2010, 12, 10476− 10493.Google ScholarThere is no corresponding record for this reference.
- 83Neidle, S. In Nucleic Acid Structure and Recognition;Oxford University Press Inc.: Oxford, 2002.Google ScholarThere is no corresponding record for this reference.
- 84Sklenovsky, P.; Florova, P.; Banas, P.; Reblova, K.; Lankas, F.; Otyepka, M.; Sponer, J. J. Chem. Theory Comput. accepted for publication.Google ScholarThere is no corresponding record for this reference.
- 85Merz, K. M. J. Chem. Theory Comput. 2010, 6 (4) 1018– 1027Google ScholarThere is no corresponding record for this reference.
- 86Simmerling, C.; Strockbine, B.; Roitberg, A. E. J. Am. Chem. Soc. 2002, 124 (38) 11258– 11259Google ScholarThere is no corresponding record for this reference.
- 87Feig, M.; Pettitt, B. M. Biophys. J. 1998, 75 (1) 134– 149Google ScholarThere is no corresponding record for this reference.
- 88Altona, C.; Sundaralingam, M. J. Am. Chem. Soc. 1972, 94 (23) 8205– 8212Google ScholarThere is no corresponding record for this reference.
- 89Sponer, J.; Kypr, J. J. Mol. Biol. 1991, 221 (3) 761– 764Google ScholarThere is no corresponding record for this reference.
- 90Bhattacharyya, D.; Bansal, M. J. Biomol. Struct. Dyn. 1989, 6 (4) 635– 653Google Scholar90https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXhsFSqsro%253D&md5=9df52ea7d56288b5e508bad5858e4a08A self-consistent formulation for analysis and generation of non-uniform DNA structuresBhattacharyya, Dhananjay; Bansal, ManjuJournal of Biomolecular Structure & Dynamics (1989), 6 (4), 635-53CODEN: JBSDD6; ISSN:0739-1102.A fully self-consistent formulation is described here for the anal. and generation of base-pairs in nonuniform DNA structures, in terms of various local parameters. The internal wedge parameters are math. related to the parameters describing the base-pair orientation with respect to an external helix axis. Hence any one set of three translation and three rotation parameters are necessary and sufficient to completely describe the relative orientation of the base-pairs comprising a step (or doublet). A general procedure is outlined for obtaining an av. or global helix axis from the local helix axes for each step. A graphical representation of the local helix axes in the form of a polar plot is also shown and its application for estg. the curvature of oligonucleotide structures is illustrated, with examples of both A and B type structures.
- 91Besseova, I.; Reblova, K.; Leontis, N. B.; Sponer, J. Nucleic Acids Res. 2010, 38 (18) 6247– 6264Google ScholarThere is no corresponding record for this reference.
- 92Tolbert, B. S.; Miyazaki, Y.; Barton, S.; Kinde, B.; Starck, P.; Singh, R.; Bax, A.; Case, D. A.; Summers, M. F. J. Biomol. NMR 2010, 47 (3) 205– 219Google ScholarThere is no corresponding record for this reference.
- 93Tjandra, N.; Bax, A. Science 1997, 278 (5344) 1697– 1697Google ScholarThere is no corresponding record for this reference.
Cited By
This article is cited by 795 publications.
- Manas Mondal, Yi Qin Gao. Atomistic Insights into Sequence-Mediated Spontaneous Association of Short RNA Chains. Biochemistry 2024, Article ASAP.
- Rajwinder Kaur, Dylan J. Nikkel, Stacey D. Wetmore. Mechanism of Nucleic Acid Phosphodiester Bond Cleavage by Human Endonuclease V: MD and QM/MM Calculations Reveal a Versatile Metal Dependence. The Journal of Physical Chemistry B 2024, 128
(39)
, 9455-9469. https://doi.org/10.1021/acs.jpcb.4c05846
- Timothy J. Wilson, Erika McCarthy, Şölen Ekesan, Timothy J. Giese, Nan-Sheng Li, Lin Huang, Joseph A. Piccirilli, Darrin M. York, David M. J. Lilley. The Role of General Acid Catalysis in the Mechanism of an Alkyl Transferase Ribozyme. ACS Catalysis 2024, Article ASAP.
- Guodong Hu, Xue Yu, Zhaojun Li. Unveiling Putative Excited State and Transmission of Binding Information in the Fluoride Riboswitch. Journal of Chemical Information and Modeling 2024, Article ASAP.
- Avijit Mainan, Rimi Kundu, Rishabh K. Singh, Susmita Roy. Magnesium Regulates RNA Ring Dynamics and Folding in Subgenomic Flaviviral RNA. The Journal of Physical Chemistry B 2024, Article ASAP.
- Valerio Piomponi, Miroslav Krepl, Jiri Sponer, Giovanni Bussi. Molecular Simulations to Investigate the Impact of N6-Methylation in RNA Recognition: Improving Accuracy and Precision of Binding Free Energy Prediction. The Journal of Physical Chemistry B 2024, 128
(37)
, 8896-8907. https://doi.org/10.1021/acs.jpcb.4c03397
- Yuanzhe Zhou, Yangwei Jiang, Shi-Jie Chen. SPRank─A Knowledge-Based Scoring Function for RNA-Ligand Pose Prediction and Virtual Screening. Journal of Chemical Theory and Computation 2024, 20
(16)
, 7358-7369. https://doi.org/10.1021/acs.jctc.4c00681
- Stefano Muscat, Gianfranco Martino, Jacopo Manigrasso, Marco Marcia, Marco De Vivo. On the Power and Challenges of Atomistic Molecular Dynamics to Investigate RNA Molecules. Journal of Chemical Theory and Computation 2024, 20
(16)
, 6992-7008. https://doi.org/10.1021/acs.jctc.4c00773
- Lauren E. Raguette, Sarah S. Gunasekera, Rebeca I. Diaz Ventura, Ethan Aminov, Jason T. Linzer, Diksha Parwana, Qin Wu, Carlos Simmerling, Maria C. Nagan. Adjusting the Energy Profile for CH–O Interactions Leads to Improved Stability of RNA Stem-Loop Structures in MD Simulations. The Journal of Physical Chemistry B 2024, 128
(33)
, 7921-7933. https://doi.org/10.1021/acs.jpcb.4c01910
- Barbora Knappeová, Vojtěch Mlýnský, Martin Pykal, Jiří Šponer, Pavel Banáš, Michal Otyepka, Miroslav Krepl. Comprehensive Assessment of Force-Field Performance in Molecular Dynamics Simulations of DNA/RNA Hybrid Duplexes. Journal of Chemical Theory and Computation 2024, 20
(15)
, 6917-6929. https://doi.org/10.1021/acs.jctc.4c00601
- Jason T. Linzer, Ethan Aminov, Aalim S. Abdullah, Colleen E. Kirkup, Rebeca I. Diaz Ventura, Vinay R. Bijoor, Jiyun Jung, Sophie Huang, Chi Gee Tse, Emily Álvarez Toucet, Hugo P. Onghai, Arghya P. Ghosh, Alex C. Grodzki, Emilee R. Haines, Aditya S. Iyer, Mark K. Khalil, Alexander P. Leong, Michael A. Neuhaus, Joseph Park, Asir Shahid, Matthew Xie, Jan M. Ziembicki, Carlos Simmerling, Maria C. Nagan. Accurately Modeling RNA Stem-Loops in an Implicit Solvent Environment. Journal of Chemical Information and Modeling 2024, 64
(15)
, 6092-6104. https://doi.org/10.1021/acs.jcim.4c00756
- Gregory Mathez, Andrea Brancale, Valeria Cagno. Novel Inhibitors of SARS-CoV-2 RNA Identified through Virtual Screening. Journal of Chemical Information and Modeling 2024, 64
(15)
, 6190-6196. https://doi.org/10.1021/acs.jcim.4c00758
- Sélène Forget, Marie Juillé, Elise Duboué-Dijon, Guillaume Stirnemann. Simulation-Guided Conformational Space Exploration to Assess Reactive Conformations of a Ribozyme. Journal of Chemical Theory and Computation 2024, 20
(14)
, 6263-6277. https://doi.org/10.1021/acs.jctc.4c00294
- Zhengxin Li, Ge Song, Junjie Zhu, Junxi Mu, Yutong Sun, Xiaokun Hong, Taeyoung Choi, Xiaochen Cui, Hai-Feng Chen. Excited–Ground-State Transition of the RNA Strand Slippage Mechanism Captured by the Base-Specific Force Field. Journal of Chemical Theory and Computation 2024, 20
(14)
, 6082-6097. https://doi.org/10.1021/acs.jctc.4c00497
- Timothy J. Giese, Jinzhe Zeng, Lauren Lerew, Erika McCarthy, Yujun Tao, Şölen Ekesan, Darrin M. York. Software Infrastructure for Next-Generation QM/MM−ΔMLP Force Fields. The Journal of Physical Chemistry B 2024, 128
(26)
, 6257-6271. https://doi.org/10.1021/acs.jpcb.4c01466
- Stefano Bosio, Mattia Bernetti, Walter Rocchia, Matteo Masetti. Similarities and Differences in Ligand Binding to Protein and RNA Targets: The Case of Riboflavin. Journal of Chemical Information and Modeling 2024, 64
(11)
, 4570-4586. https://doi.org/10.1021/acs.jcim.4c00420
- Christian Kersten, Philippe Archambault, Luca P. Köhler. Assessment of Nucleobase Protomeric and Tautomeric States in Nucleic Acid Structures for Interaction Analysis and Structure-Based Ligand Design. Journal of Chemical Information and Modeling 2024, 64
(11)
, 4485-4499. https://doi.org/10.1021/acs.jcim.4c00520
- Afra Sabei, Cécilia Hognon, Juliette Martin, Elisa Frezza. Dynamics of Protein–RNA Interfaces Using All-Atom Molecular Dynamics Simulations. The Journal of Physical Chemistry B 2024, 128
(20)
, 4865-4886. https://doi.org/10.1021/acs.jpcb.3c07698
- Hong Huang, Dong-Xia Zhao, Jian Zhao, Xin Chen, Cui Liu, Zhong-Zhi Yang. Origin of Enantioselectivity in Engineered Cytochrome c-Catalyzed Carbon-Radical FePP Hydrolysis Revealed Using QM/MM (ABEEM Polarizable Force Field) and MD Simulations. The Journal of Physical Chemistry B 2024, 128
(16)
, 3807-3823. https://doi.org/10.1021/acs.jpcb.3c07158
- Elisa Frezza, Damien Laage, Elise Duboué-Dijon. Molecular Origin of Distinct Hydration Dynamics in Double Helical DNA and RNA Sequences. The Journal of Physical Chemistry Letters 2024, 15
(16)
, 4351-4358. https://doi.org/10.1021/acs.jpclett.4c00629
- Francesco Guidarelli Mattioli, Andrea Saltalamacchia, Alessandra Magistrato. Tracing Allostery in the Spliceosome Ski2-like RNA Helicase Brr2. The Journal of Physical Chemistry Letters 2024, 15
(13)
, 3502-3508. https://doi.org/10.1021/acs.jpclett.3c03538
- Taeyoung Choi, Zhengxin Li, Ge Song, Hai-Feng Chen. Comprehensive Comparison and Critical Assessment of RNA-Specific Force Fields. Journal of Chemical Theory and Computation 2024, 20
(6)
, 2676-2688. https://doi.org/10.1021/acs.jctc.4c00066
- Ilona C. Unarta, Siqin Cao, Eshani C. Goonetilleke, Jiani Niu, Samuel H. Gellman, Xuhui Huang. Submillisecond Atomistic Molecular Dynamics Simulations Reveal Hydrogen Bond-Driven Diffusion of a Guest Peptide in Protein–RNA Condensate. The Journal of Physical Chemistry B 2024, 128
(10)
, 2347-2359. https://doi.org/10.1021/acs.jpcb.3c08126
- Shiji Zhao, Piotr Cieplak, Yong Duan, Ray Luo. Assessment of Amino Acid Electrostatic Parametrizations of the Polarizable Gaussian Multipole Model. Journal of Chemical Theory and Computation 2024, 20
(5)
, 2098-2110. https://doi.org/10.1021/acs.jctc.3c01347
- Timothy J. Giese, Şölen Ekesan, Erika McCarthy, Yujun Tao, Darrin M. York. Surface-Accelerated String Method for Locating Minimum Free Energy Paths. Journal of Chemical Theory and Computation 2024, 20
(5)
, 2058-2073. https://doi.org/10.1021/acs.jctc.3c01401
- Giada Ciardullo, Angela Parise, Mario Prejanò, Tiziana Marino. Viral RNA Replication Suppression of SARS-CoV-2: Atomistic Insights into Inhibition Mechanisms of RdRp Machinery by ddhCTP. Journal of Chemical Information and Modeling 2024, 64
(5)
, 1593-1604. https://doi.org/10.1021/acs.jcim.3c01919
- Silvia Rinaldi, Elisabetta Moroni, Riccardo Rozza, Alessandra Magistrato. Frontiers and Challenges of Computing ncRNAs Biogenesis, Function and Modulation. Journal of Chemical Theory and Computation 2024, 20
(3)
, 993-1018. https://doi.org/10.1021/acs.jctc.3c01239
- Sana Akhter, Zhichao Tang, Jinan Wang, Mercy Haboro, Erik D Holmstrom, Jingxin Wang, Yinglong Miao. Mechanism of Ligand Binding to Theophylline RNA Aptamer. Journal of Chemical Information and Modeling 2024, 64
(3)
, 1017-1029. https://doi.org/10.1021/acs.jcim.3c01454
- Hana Dohnalová, Mona Seifert, Eva Matoušková, Misha Klein, Flávia S. Papini, Jan Lipfert, David Dulin, Filip Lankaš. Temperature-Dependent Twist of Double-Stranded RNA Probed by Magnetic Tweezer Experiments and Molecular Dynamics Simulations. The Journal of Physical Chemistry B 2024, 128
(3)
, 664-675. https://doi.org/10.1021/acs.jpcb.3c06280
- Olivia Love, Lauren Winkler, Thomas E. Cheatham, III. van der Waals Parameter Scanning with Amber Nucleic Acid Force Fields: Revisiting Means to Better Capture the RNA/DNA Structure through MD. Journal of Chemical Theory and Computation 2024, 20
(2)
, 625-643. https://doi.org/10.1021/acs.jctc.3c01164
- Gert-Jan Bekker, Yoshifumi Fukunishi, Junichi Higo, Narutoshi Kamiya. Binding Mechanism of Riboswitch to Natural Ligand Elucidated by McMD-Based Dynamic Docking Simulations. ACS Omega 2024, 9
(3)
, 3412-3422. https://doi.org/10.1021/acsomega.3c06826
- Santiago Antolínez, Peter Eugene Jones, James C. Phillips, Jodi A. Hadden-Perilla. AMBERff at Scale: Multimillion-Atom Simulations with AMBER Force Fields in NAMD. Journal of Chemical Information and Modeling 2024, 64
(2)
, 543-554. https://doi.org/10.1021/acs.jcim.3c01648
- Monsurat M. Lawal, Priti Roy, Martin McCullagh. Role of ATP Hydrolysis and Product Release in the Translocation Mechanism of SARS-CoV-2 NSP13. The Journal of Physical Chemistry B 2024, 128
(2)
, 492-503. https://doi.org/10.1021/acs.jpcb.3c06714
- Riccardo Rozza, Pavel Janoš, Alessandra Magistrato. Assessing the Binding Mode of a Splicing Modulator Stimulating Pre-mRNA Binding to the Plastic U2AF2 Splicing Factor. Journal of Chemical Information and Modeling 2023, 63
(23)
, 7508-7517. https://doi.org/10.1021/acs.jcim.3c01204
- Vojtěch Mlýnský, Petra Kührová, Petr Stadlbauer, Miroslav Krepl, Michal Otyepka, Pavel Banáš, Jiří Šponer. Simple Adjustment of Intranucleotide Base-Phosphate Interaction in the OL3 AMBER Force Field Improves RNA Simulations. Journal of Chemical Theory and Computation 2023, 19
(22)
, 8423-8433. https://doi.org/10.1021/acs.jctc.3c00990
- Xiaowei Wang, Tiantian Xu, Yuan Yao, Peter Pak-Hang Cheung, Xin Gao, Lu Zhang. SARS-CoV-2 RNA-Dependent RNA Polymerase Follows Asynchronous Translocation Pathway for Viral Transcription and Replication. The Journal of Physical Chemistry Letters 2023, 14
(45)
, 10119-10128. https://doi.org/10.1021/acs.jpclett.3c01249
- Vladimir S. Farafonov, Michael Stich, Dmitry A. Nerukh. Complete Virion Simulated: All-Atom Model of an MS2 Bacteriophage with Native Genome. Journal of Chemical Theory and Computation 2023, 19
(21)
, 7924-7933. https://doi.org/10.1021/acs.jctc.3c00846
- Yazdan Maghsoud, Vindi M. Jayasinghe-Arachchige, Pratibha Kumari, G. Andrés Cisneros, Jin Liu. Leveraging QM/MM and Molecular Dynamics Simulations to Decipher the Reaction Mechanism of the Cas9 HNH Domain to Investigate Off-Target Effects. Journal of Chemical Information and Modeling 2023, 63
(21)
, 6834-6850. https://doi.org/10.1021/acs.jcim.3c01284
- Mark A. Hix, Alice R. Walker. AutoParams: An Automated Web-Based Tool To Generate Force Field Parameters for Molecular Dynamics Simulations. Journal of Chemical Information and Modeling 2023, 63
(20)
, 6293-6301. https://doi.org/10.1021/acs.jcim.3c01049
- Pradeep Pant, Florian Leese. Probing the Nucleic Acid Flexibility to Disarm the Viral Counter-Defense Machinery: Design and Characterization of Potent p19 Inhibitors. The Journal of Physical Chemistry B 2023, 127
(41)
, 8842-8851. https://doi.org/10.1021/acs.jpcb.3c04788
- Joseph A. Makowski, Adam H. Kensinger, Caylee L. Cunningham, Caleb J. Frye, Morgan Shine, Patrick E. Lackey, Mihaela Rita Mihailescu, Jeffrey D. Evanseck. Delta SARS-CoV-2 s2m Structure, Dynamics, and Entropy: Consequences of the G15U Mutation. ACS Physical Chemistry Au 2023, 3
(5)
, 434-443. https://doi.org/10.1021/acsphyschemau.3c00008
- Xianwei Wang, Yiying Wang, Man Guo, Xuechao Wang, Yang Li, John Z.H. Zhang. Assessment of an Electrostatic Energy-Based Charge Model for Modeling the Electrostatic Interactions in Water Solvent. Journal of Chemical Theory and Computation 2023, 19
(18)
, 6294-6312. https://doi.org/10.1021/acs.jctc.3c00467
- Qiang Zhu, Yongxian Wu, Shiji Zhao, Piotr Cieplak, Yong Duan, Ray Luo. Streamlining and Optimizing Strategies of Electrostatic Parameterization. Journal of Chemical Theory and Computation 2023, 19
(18)
, 6353-6365. https://doi.org/10.1021/acs.jctc.3c00659
- Federica Battistini, Alba Sala, Adam Hospital, Modesto Orozco. Sequence-Dependent Properties of the RNA Duplex. Journal of Chemical Information and Modeling 2023, 63
(16)
, 5259-5271. https://doi.org/10.1021/acs.jcim.3c00741
- Jiaqi Liu, Jian Wan, Yanliang Ren, Xubo Shao, Xin Xu, Li Rao. DOX_BDW: Incorporating Solvation and Desolvation Effects of Cavity Water into Nonfitting Protein–Ligand Binding Affinity Prediction. Journal of Chemical Information and Modeling 2023, 63
(15)
, 4850-4863. https://doi.org/10.1021/acs.jcim.3c00776
- Raju Sarkar, Rishabh K. Singh, Susmita Roy. Hierarchical Hydration Dynamics of RNA with Nano-Water-Pool at Its Core. The Journal of Physical Chemistry B 2023, 127
(31)
, 6903-6919. https://doi.org/10.1021/acs.jpcb.3c03553
- Aimeric Dabin, Guillaume Stirnemann. Toward a Molecular Mechanism of Complementary RNA Duplexes Denaturation. The Journal of Physical Chemistry B 2023, 127
(27)
, 6015-6028. https://doi.org/10.1021/acs.jpcb.3c00908
- Sefora Naomi Agrò, Riccardo Rozza, Santiago Movilla, Jana Aupič, Alessandra Magistrato. Molecular Dynamics Simulations Elucidate the Molecular Basis of Pre-mRNA Translocation by the Prp2 Spliceosomal Helicase. Journal of Chemical Information and Modeling 2023, 63
(13)
, 4180-4189. https://doi.org/10.1021/acs.jcim.3c00585
- Nicola Calonaci, Mattia Bernetti, Alisha Jones, Michael Sattler, Giovanni Bussi. Molecular Dynamics Simulations with Grand-Canonical Reweighting Suggest Cooperativity Effects in RNA Structure Probing Experiments. Journal of Chemical Theory and Computation 2023, 19
(12)
, 3672-3685. https://doi.org/10.1021/acs.jctc.3c00084
- Kumar Sarthak, David Winogradoff, Yingda Ge, Sua Myong, Aleksei Aksimentiev. Benchmarking Molecular Dynamics Force Fields for All-Atom Simulations of Biological Condensates. Journal of Chemical Theory and Computation 2023, 19
(12)
, 3721-3740. https://doi.org/10.1021/acs.jctc.3c00148
- Ting Yu, Taigang Liu, Yujie Wang, Xuanlin Zhao, Wenbing Zhang. Effect of Cas9 Protein on the Seed-Target Base Pair of the sgRNA/DNA Hybrid Duplex. The Journal of Physical Chemistry B 2023, 127
(22)
, 4989-4997. https://doi.org/10.1021/acs.jpcb.3c00997
- Riccardo Rozza, Pavel Janoš, Alessandra Magistrato. Monovalent Ionic Atmosphere Modulates the Selection of Suboptimal RNA Sequences by Splicing Factors’ RNA Recognition Motifs. Journal of Chemical Information and Modeling 2023, 63
(10)
, 3086-3093. https://doi.org/10.1021/acs.jcim.3c00110
- Jon Uranga, Ricardo A. Mata. The Catalytic Mechanism of Acetoacetate Decarboxylase: A Detailed Study of Schiff Base Formation, Protonation States, and Their Impact on Catalysis. Journal of Chemical Information and Modeling 2023, 63
(10)
, 3118-3127. https://doi.org/10.1021/acs.jcim.3c00241
- Souvik Sinha, Chinmai Pindi, Mohd Ahsan, Pablo R. Arantes, Giulia Palermo. Machines on Genes through the Computational Microscope. Journal of Chemical Theory and Computation 2023, 19
(7)
, 1945-1964. https://doi.org/10.1021/acs.jctc.2c01313
- Petra Kührová, Vojtěch Mlýnský, Michal Otyepka, Jiří Šponer, Pavel Banáš. Sensitivity of the RNA Structure to Ion Conditions as Probed by Molecular Dynamics Simulations of Common Canonical RNA Duplexes. Journal of Chemical Information and Modeling 2023, 63
(7)
, 2133-2146. https://doi.org/10.1021/acs.jcim.2c01438
- Jinzhe Zeng, Yujun Tao, Timothy J. Giese, Darrin M. York. QDπ: A Quantum Deep Potential Interaction Model for Drug Discovery. Journal of Chemical Theory and Computation 2023, 19
(4)
, 1261-1275. https://doi.org/10.1021/acs.jctc.2c01172
- Shiji Zhao, Piotr Cieplak, Yong Duan, Ray Luo. Transferability of the Electrostatic Parameters of the Polarizable Gaussian Multipole Model. Journal of Chemical Theory and Computation 2023, 19
(3)
, 924-941. https://doi.org/10.1021/acs.jctc.2c01048
- Saikat Pal, Amit Kumar, Harish Vashisth. Role of Dynamics and Mutations in Interactions of a Zinc Finger Antiviral Protein with CG-rich Viral RNA. Journal of Chemical Information and Modeling 2023, 63
(3)
, 1002-1011. https://doi.org/10.1021/acs.jcim.2c01487
- Xiaoqing Wu, Remya Ramesh, Jinan Wang, Youguang Zheng, Ahlam M. Armaly, Lanjing Wei, Minli Xing, Sudeshna Roy, Lan Lan, Fei Philip Gao, Yinglong Miao, Liang Xu, Jeffrey Aubé. Small Molecules Targeting the RNA-Binding Protein HuR Inhibit Tumor Growth in Xenografts. Journal of Medicinal Chemistry 2023, 66
(3)
, 2032-2053. https://doi.org/10.1021/acs.jmedchem.2c01723
- Markéta Paloncýová, Martin Šrejber, Petra Čechová, Petra Kührová, Filip Zaoral, Michal Otyepka. Atomistic Insights into Organization of RNA-Loaded Lipid Nanoparticles. The Journal of Physical Chemistry B 2023, 127
(5)
, 1158-1166. https://doi.org/10.1021/acs.jpcb.2c07671
- Adam H. Kensinger, Joseph A. Makowski, Kendy A. Pellegrene, Joshua A. Imperatore, Caylee L. Cunningham, Caleb J. Frye, Patrick E. Lackey, Mihaela Rita Mihailescu, Jeffrey D. Evanseck. Structural, Dynamical, and Entropic Differences between SARS-CoV and SARS-CoV-2 s2m Elements Using Molecular Dynamics Simulations. ACS Physical Chemistry Au 2023, 3
(1)
, 30-43. https://doi.org/10.1021/acsphyschemau.2c00032
- Soma Roy, Santanu Bhattacharya. Chemical Information and Computational Modeling of Targeting Hybrid Nucleic Acid Structures of PIM1 Sequences by Synthetic Pyrrole-Imidazole Carboxamide Drugs. Journal of Chemical Information and Modeling 2022, 62
(24)
, 6411-6422. https://doi.org/10.1021/acs.jcim.1c01500
- Riccardo Rozza, Andrea Saltalamacchia, Clarissa Orrico, Pavel Janoš, Alessandra Magistrato. All-Atom Simulations Elucidate the Impact of U2AF2 Cancer-Associated Mutations on Pre-mRNA Recognition. Journal of Chemical Information and Modeling 2022, 62
(24)
, 6691-6703. https://doi.org/10.1021/acs.jcim.2c00511
- Jianzhong Chen, Qingkai Zeng, Wei Wang, Haibo Sun, Guodong Hu. Decoding the Identification Mechanism of an SAM-III Riboswitch on Ligands through Multiple Independent Gaussian-Accelerated Molecular Dynamics Simulations. Journal of Chemical Information and Modeling 2022, 62
(23)
, 6118-6132. https://doi.org/10.1021/acs.jcim.2c00961
- Klaudia Mráziková, Holger Kruse, Vojtěch Mlýnský, Pascal Auffinger, Jiří Šponer. Multiscale Modeling of Phosphate···π Contacts in RNA U-Turns Exposes Differences between Quantum-Chemical and AMBER Force Field Descriptions. Journal of Chemical Information and Modeling 2022, 62
(23)
, 6182-6200. https://doi.org/10.1021/acs.jcim.2c01064
- Jason M. Imamoto, Randy J. Zauhar, Michael F. Bruist. Sarcin/Ricin Domain RNA Retains Its Structure Better Than A-RNA in Adaptively Biased Molecular Dynamics Simulations. The Journal of Physical Chemistry B 2022, 126
(48)
, 10018-10033. https://doi.org/10.1021/acs.jpcb.2c05859
- Vangelis Daskalakis. Deciphering the QR Code of the CRISPR-Cas9 System: Synergy between Gln768 (Q) and Arg976 (R). ACS Physical Chemistry Au 2022, 2
(6)
, 496-505. https://doi.org/10.1021/acsphyschemau.2c00041
- Pavlína Pokorná, Miroslav Krepl, Sébastien Campagne, Jiří Šponer. Conformational Heterogeneity of RNA Stem-Loop Hairpins Bound to FUS-RNA Recognition Motif with Disordered RGG Tail Revealed by Unbiased Molecular Dynamics Simulations. The Journal of Physical Chemistry B 2022, 126
(45)
, 9207-9221. https://doi.org/10.1021/acs.jpcb.2c06168
- Sergio Cruz-León, Nadine Schwierz. RNA Captures More Cations than DNA: Insights from Molecular Dynamics Simulations. The Journal of Physical Chemistry B 2022, 126
(43)
, 8646-8654. https://doi.org/10.1021/acs.jpcb.2c04488
- Jiahui Zhang, Ashkan Fakharzadeh, Christopher Roland, Celeste Sagui. RNA as a Major-Groove Ligand: RNA–RNA and RNA–DNA Triplexes Formed by GAA and UUC or TTC Sequences. ACS Omega 2022, 7
(43)
, 38728-38743. https://doi.org/10.1021/acsomega.2c04358
- Angela Parise, Giada Ciardullo, Mario Prejanò, Aurélien de la Lande, Tiziana Marino. On the Recognition of Natural Substrate CTP and Endogenous Inhibitor ddhCTP of SARS-CoV-2 RNA-Dependent RNA Polymerase: A Molecular Dynamics Study. Journal of Chemical Information and Modeling 2022, 62
(20)
, 4916-4927. https://doi.org/10.1021/acs.jcim.2c01002
- Anastasia Croitoru, Alexey Aleksandrov. Parametrization of Force Field Bonded Terms under Structural Inconsistency. Journal of Chemical Information and Modeling 2022, 62
(19)
, 4771-4782. https://doi.org/10.1021/acs.jcim.2c00950
- Zhifeng Jing, Pengyu Ren. Molecular Dynamics Simulations of Protein RNA Complexes by Using an Advanced Electrostatic Model. The Journal of Physical Chemistry B 2022, 126
(38)
, 7343-7353. https://doi.org/10.1021/acs.jpcb.2c05278
- Kye Won Wang, Jumin Lee, Han Zhang, Donghyuk Suh, Wonpil Im. CHARMM-GUI Implicit Solvent Modeler for Various Generalized Born Models in Different Simulation Programs. The Journal of Physical Chemistry B 2022, 126
(38)
, 7354-7364. https://doi.org/10.1021/acs.jpcb.2c05294
- Sayyed Jalil Mahdizadeh, Emil Pålsson, Antonio Carlesso, Eric Chevet, Leif A. Eriksson. QM/MM Well-Tempered Metadynamics Study of the Mechanism of XBP1 mRNA Cleavage by Inositol Requiring Enzyme 1α RNase. Journal of Chemical Information and Modeling 2022, 62
(17)
, 4247-4260. https://doi.org/10.1021/acs.jcim.2c00735
- Emmanuelle Bignon, Antonio Monari. Modeling the Enzymatic Mechanism of the SARS-CoV-2 RNA-Dependent RNA Polymerase by DFT/MM-MD: An Unusual Active Site Leading to High Replication Rates. Journal of Chemical Information and Modeling 2022, 62
(17)
, 4261-4269. https://doi.org/10.1021/acs.jcim.2c00802
- Valerio Piomponi, Thorben Fröhlking, Mattia Bernetti, Giovanni Bussi. Molecular Simulations Matching Denaturation Experiments for N6-Methyladenosine. ACS Central Science 2022, 8
(8)
, 1218-1228. https://doi.org/10.1021/acscentsci.2c00565
- Şölen Ekesan, Erika McCarthy, David A. Case, Darrin M. York. RNA Electrostatics: How Ribozymes Engineer Active Sites to Enable Catalysis. The Journal of Physical Chemistry B 2022, 126
(32)
, 5982-5990. https://doi.org/10.1021/acs.jpcb.2c03727
- Matthew M. Copeland, Hung N. Do, Lane Votapka, Keya Joshi, Jinan Wang, Rommie E. Amaro, Yinglong Miao. Gaussian Accelerated Molecular Dynamics in OpenMM. The Journal of Physical Chemistry B 2022, 126
(31)
, 5810-5820. https://doi.org/10.1021/acs.jpcb.2c03765
- Shivangi Sharma, Vishal Singh, Parbati Biswas. Analysis of the Passage Times for Unfolding/Folding of the Adenine Riboswitch Aptamer. ACS Physical Chemistry Au 2022, 2
(4)
, 353-363. https://doi.org/10.1021/acsphyschemau.1c00056
- Thorben Fröhlking, Vojtěch Mlýnský, Michal Janeček, Petra Kührová, Miroslav Krepl, Pavel Banáš, Jiří Šponer, Giovanni Bussi. Automatic Learning of Hydrogen-Bond Fixes in the AMBER RNA Force Field. Journal of Chemical Theory and Computation 2022, 18
(7)
, 4490-4502. https://doi.org/10.1021/acs.jctc.2c00200
- Adan Gallardo, Brandon M. Bogart, Bercem Dutagaci. Protein–Nucleic Acid Interactions for RNA Polymerase II Elongation Factors by Molecular Dynamics Simulations. Journal of Chemical Information and Modeling 2022, 62
(12)
, 3079-3089. https://doi.org/10.1021/acs.jcim.2c00121
- Maxwell R. Tucker, Stefano Piana, Dazhi Tan, Michael V. LeVine, David E. Shaw. Development of Force Field Parameters for the Simulation of Single- and Double-Stranded DNA Molecules and DNA–Protein Complexes. The Journal of Physical Chemistry B 2022, 126
(24)
, 4442-4457. https://doi.org/10.1021/acs.jpcb.1c10971
- Kenneth Lam, Koushik Kasavajhala, Sarah Gunasekera, Carlos Simmerling. Accelerating the Ensemble Convergence of RNA Hairpin Simulations with a Replica Exchange Structure Reservoir. Journal of Chemical Theory and Computation 2022, 18
(6)
, 3930-3947. https://doi.org/10.1021/acs.jctc.2c00065
- Amirhossein Taghavi, Ivan Riveros, David J. Wales, Ilyas Yildirim. Evaluating Geometric Definitions of Stacking for RNA Dinucleoside Monophosphates Using Molecular Mechanics Calculations. Journal of Chemical Theory and Computation 2022, 18
(6)
, 3637-3653. https://doi.org/10.1021/acs.jctc.2c00178
- Sehr Naseem-Khan, Madison B. Berger, Emmett M. Leddin, Yazdan Maghsoud, G. Andrés Cisneros. Impact of Remdesivir Incorporation along the Primer Strand on SARS-CoV-2 RNA-Dependent RNA Polymerase. Journal of Chemical Information and Modeling 2022, 62
(10)
, 2456-2465. https://doi.org/10.1021/acs.jcim.2c00201
- Nicy, Debayan Chakraborty, David J. Wales. Energy Landscapes for Base-Flipping in a Model DNA Duplex. The Journal of Physical Chemistry B 2022, 126
(16)
, 3012-3028. https://doi.org/10.1021/acs.jpcb.2c00340
- Vojtěch Mlýnský, Michal Janeček, Petra Kührová, Thorben Fröhlking, Michal Otyepka, Giovanni Bussi, Pavel Banáš, Jiří Šponer. Toward Convergence in Folding Simulations of RNA Tetraloops: Comparison of Enhanced Sampling Techniques and Effects of Force Field Modifications. Journal of Chemical Theory and Computation 2022, 18
(4)
, 2642-2656. https://doi.org/10.1021/acs.jctc.1c01222
- Francesco Di Palma, Sergio Decherchi, Fátima Pardo-Avila, Sauro Succi, Michael Levitt, Gunnar von Heijne, Andrea Cavalli. Probing Interplays between Human XBP1u Translational Arrest Peptide and 80S Ribosome. Journal of Chemical Theory and Computation 2022, 18
(3)
, 1905-1914. https://doi.org/10.1021/acs.jctc.1c00796
- Sunit Kumar Jana, S. Harikrishna, Sruthi Sudhakar, Roberto El-Khoury, P.I. Pradeepkumar, Masad J. Damha. Nucleoside Analogues with a Seven-Membered Sugar Ring: Synthesis and Structural Compatibility in DNA–RNA Hybrids. The Journal of Organic Chemistry 2022, 87
(5)
, 2367-2379. https://doi.org/10.1021/acs.joc.1c02254
- Ana Dreab, Craig A. Bayse. Molecular Dynamics Simulations of Reduced and Oxidized TFIIIA Zinc Fingers Free and Interacting with 5S RNA. Journal of Chemical Information and Modeling 2022, 62
(4)
, 903-913. https://doi.org/10.1021/acs.jcim.1c01272
- Jianbo Zhao, Scott D. Kennedy, Douglas H. Turner. Nuclear Magnetic Resonance Spectra and AMBER OL3 and ROC-RNA Simulations of UCUCGU Reveal Force Field Strengths and Weaknesses for Single-Stranded RNA. Journal of Chemical Theory and Computation 2022, 18
(2)
, 1241-1254. https://doi.org/10.1021/acs.jctc.1c00643
- Jan Neumann, Nadine Schwierz. Artificial Intelligence Resolves Kinetic Pathways of Magnesium Binding to RNA. Journal of Chemical Theory and Computation 2022, 18
(2)
, 1202-1212. https://doi.org/10.1021/acs.jctc.1c00752
- Tamilselvan Rajasekaran, Graeme C. Freestone, Rodrigo Galindo-Murillo, Barbara Lugato, Lorena Rico, Juan C. Salinas, Hans Gaus, Michael T. Migawa, Eric E. Swayze, Thomas E. Cheatham, III, Stephen Hanessian, Punit P. Seth. Backbone Hydrocarbon-Constrained Nucleic Acids Modulate Hybridization Kinetics for RNA. Journal of the American Chemical Society 2022, 144
(4)
, 1941-1950. https://doi.org/10.1021/jacs.1c12323
- Jun Chen, Hao Liu, Xiaochen Cui, Zhengxin Li, Hai-Feng Chen. RNA-Specific Force Field Optimization with CMAP and Reweighting. Journal of Chemical Information and Modeling 2022, 62
(2)
, 372-385. https://doi.org/10.1021/acs.jcim.1c01148
- Susanta Haldar, Yashu Zhang, Ying Xia, Barira Islam, Sisi Liu, Francesco L. Gervasio, Adrian J. Mulholland, Zoë A. E. Waller, Dengguo Wei, Shozeb Haider. Mechanistic Insights into the Ligand-Induced Unfolding of an RNA G-Quadruplex. Journal of the American Chemical Society 2022, 144
(2)
, 935-950. https://doi.org/10.1021/jacs.1c11248
- Kara K. Grotz, Nadine Schwierz. Optimized Magnesium Force Field Parameters for Biomolecular Simulations with Accurate Solvation, Ion-Binding, and Water-Exchange Properties in SPC/E, TIP3P-fb, TIP4P/2005, TIP4P-Ew, and TIP4P-D. Journal of Chemical Theory and Computation 2022, 18
(1)
, 526-537. https://doi.org/10.1021/acs.jctc.1c00791
- Samanta Makurat, Zoe Cournia, Janusz Rak. Inactive-to-Active Transition of Human Thymidine Kinase 1 Revealed by Molecular Dynamics Simulations. Journal of Chemical Information and Modeling 2022, 62
(1)
, 142-149. https://doi.org/10.1021/acs.jcim.1c01157
- Gül H. Zerze, Pablo M. Piaggi, Pablo G. Debenedetti. A Computational Study of RNA Tetraloop Thermodynamics, Including Misfolded States. The Journal of Physical Chemistry B 2021, 125
(50)
, 13685-13695. https://doi.org/10.1021/acs.jpcb.1c08038
- Sangram Prusty, Raju Sarkar, Amrita Chakraborty, Susmita Roy. Correlation in Domain Fluctuations Navigates Target Search of a Viral Peptide along RNA. The Journal of Physical Chemistry B 2021, 125
(46)
, 12678-12689. https://doi.org/10.1021/acs.jpcb.1c07699
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
References
This article references 93 other publications.
- 1Weiner, S. J.; Kollman, P. A.; Case, D. A.; Singh, U. C.; Ghio, C.; Alagona, G.; Profeta, S.; Weiner, P. J. Am. Chem. Soc. 1984, 106 (3) 765– 784There is no corresponding record for this reference.
- 2Weiner, S. J.; Kollman, P. A.; Nguyen, D. T.; Case, D. A. J. Comput. Chem. 1986, 7 (2) 230– 252There is no corresponding record for this reference.
- 3Mackerell, A. D. J. Comput. Chem. 2004, 25 (13) 1584– 1604There is no corresponding record for this reference.
- 4Orozco, M.; Noy, A.; Perez, A. Curr. Opin. Struct. Biol. 2008, 18 (2) 185– 1934https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXkvFCls74%253D&md5=81ed332ed9f1903d8e13b42d1d1106bfRecent advances in the study of nucleic acid flexibility by molecular dynamicsOrozco, Modesto; Noy, Agnes; Perez, AlbertoCurrent Opinion in Structural Biology (2008), 18 (2), 185-193CODEN: COSBEF; ISSN:0959-440X. (Elsevier B.V.)A review. The recent use of mol. dynamics (MD) simulations to study flexibility of nucleic acids has been reviewed from an anal. of the publications appearing in the past two years (from 2005 till date). Despite the existence of some unsolved problems in the methodologies, these years have been witness to major advances in the field. Based on a crit. review of the most recent contributions, excitement exists on the expected evolution of the field in the next years.
- 5Fulle, S.; Gohlke, H. J. Mol. Recognit. 2010, 23 (2) 220– 2315https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXitVShsb0%253D&md5=1f0f7c926ad5672ed3765328a2ea0888Molecular recognition of RNA: challenges for modeling interactions and plasticityFulle, Simone; Gohlke, HolgerJournal of Molecular Recognition (2010), 23 (2), 220-231CODEN: JMORE4; ISSN:0952-3499. (John Wiley & Sons Ltd.)A review. There is growing interest in mol. recognition processes of RNA because of RNA's widespread involvement in biol. processes. Computational approaches are increasingly used for analyzing and predicting binding to RNA, fuelled by encouraging progress in developing simulation, free energy, and docking methods for nucleic acids. These developments take into account challenges regarding the energetics of RNA-ligand binding, RNA plasticity, and the presence of water mols. and ions in the binding interface. Accordingly, the authors detail advances in force field and scoring function development for mol. dynamics (MD) simulations, free energy computations, and docking calcns. of nucleic acid complexes. Furthermore, the authors present methods that can detect moving parts within RNA structures based on graph-theor. approaches or normal mode anal. (NMA). As an example of the successful use of these developments, the authors discuss recent structure-based drug design approaches that focus on the bacterial ribosomal A-site RNA as a drug target.
- 6MacKerell, A. D. J. Phys. Chem. B 2009, 113 (10) 3235– 3244There is no corresponding record for this reference.
- 7Foloppe, N.; Hartmann, B.; Nilsson, L.; MacKerell, A. D. Biophys. J. 2002, 82 (3) 1554– 1569There is no corresponding record for this reference.
- 8Cornell, W. D.; Cieplak, P.; Bayly, C. I.; Gould, I. R.; Merz, K. M.; Ferguson, D. M.; Spellmeyer, D. C.; Fox, T.; Caldwell, J. W.; Kollman, P. A. J. Am. Chem. Soc. 1995, 117 (19) 5179– 5197There is no corresponding record for this reference.
- 9Mackerell, A. D.; Wiorkiewiczkuczera, J.; Karplus, M. J. Am. Chem. Soc. 1995, 117 (48) 11946– 11975There is no corresponding record for this reference.
- 10Cheatham, T. E.; Kollman, P. A. J. Mol. Biol. 1996, 259 (3) 434– 444There is no corresponding record for this reference.
- 11Yang, L. Q.; Pettitt, B. M. J. Phys. Chem. 1996, 100 (7) 2564– 2566There is no corresponding record for this reference.
- 12Cheatham, T. E.; Kollman, P. A. Structure 1997, 5 (10) 1297– 1311There is no corresponding record for this reference.
- 13Cheatham, T. E.; Crowley, M. F.; Fox, T.; Kollman, P. A. Proc. Natl. Acad. Sci. U.S.A. 1997, 94 (18) 9626– 9630There is no corresponding record for this reference.
- 14Langley, D. R. J. Biomol. Struct. Dyn. 1998, 16 (3) 487– 50914https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXhtlanu7g%253D&md5=73908d8477291848c077e630d87273bfMolecular dynamic simulations of environment and sequence dependent DNA conformations: the development of the BMS nucleic acid force field and comparison with experimental resultsLangley, David R.Journal of Biomolecular Structure & Dynamics (1998), 16 (3), 487-509CODEN: JBSDD6; ISSN:0739-1102. (Adenine Press)Mol. dynamic (MD) simulations using the BMS nucleic acid force field produce environment and sequence dependent DNA conformations that closely mimic exptl. derived structures. The parameters were initially developed to reproduce the potential energy surface, as defined by quantum mechanics, for a set of small mols. that can be used as the building blocks for nucleic acid macromols. (di-Me phosphate, cyclopentane, THF, etc.). Then the dihedral parameters were fine tuned using a series of condensed phase MD simulations of DNA and RNA (in zero added salt, 4M NaCl, and 75% ethanol solns.). In the tuning process the free energy surface for each dihedral was derived from the MD ensemble and fitted to the conformational distributions and populations obsd. in 87 A- and B-DNA x-ray and 17 B-DNA NMR structures. Over 41 ns of MD simulations are presented which demonstrate that the force field is capable of producing stable trajectories, in the correct environments, of A-DNA, double stranded A- form RNA, B-DNA, Z-DNA, and a netropsin-DNA complex that closely reproduce the exptl. detd. and/or canonical DNA conformations. Frequently the MD averaged structure is closer to the exptl. detd. structure than to the canonical DNA conformation. MD simulations of A- to B- and B- to A-DNA transitions are also shown. A-DNA simulations in a low salt environment cleanly convert into the B-DNA conformation and converge into the RMS space sampled by a low salt simulation of the same sequence starting from B-DNA. In MD simulations using the BMS force field the B-form of d(GGGCCC)2 in a 75% ethanol soln. converts into the A-form. Using the same methodol., parameters, and conditions the A-form of d(AAATTT)2 correctly converts into the B-DNA conformation. These studies demonstrate that the force field is capable of reproducing both environment and sequence dependent DNA structures. The 41 ns of MD simulations presented in this paper paint a global picture which suggests that the DNA structures obsd. in low salt solns. are largely due to the favorable internal energy brought about by the nearly uniform screening of the DNA electrostatics. While the conformations sampled in high salt or mixed solvent environments occur from selective and asym. screening of the phosphate groups and DNA grooves, resp., brought about by sequence induced ion and solvent packing.
- 15MacKerell, A. D. In Molecular Modeling of Nucleic Acids; Leontis, N. B.; SantaLucia, J., Eds.; American Chemical Society: Washington, DC, 1998; Vol. 682, p 304– 311.There is no corresponding record for this reference.
- 16Foloppe, N.; MacKerell, A. D. J. Phys. Chem. B 1998, 102 (34) 6669– 6678There is no corresponding record for this reference.
- 17Cheatham, T. E.; Cieplak, P.; Kollman, P. A. J. Biomol. Struct. Dyn. 1999, 16 (4) 845– 86217https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXitF2jsb8%253D&md5=d3789449f1dd0d2ed769a22316469d4bA modified version of the Cornell et al. force field with improved sugar pucker phases and helical repeatCheatham, Thomas E., III; Cieplak, Piotr; Kollman, Peter A.Journal of Biomolecular Structure & Dynamics (1999), 16 (4), 845-862CODEN: JBSDD6; ISSN:0739-1102. (Adenine Press)We have examd. some subtle parameter modifications to the Cornell et al. force field, which has proven quite successful in reproducing nucleic acid properties, but whose C2'-endo sugar pucker phase and helical repeat for B DNA appear to be somewhat underestimated. Encouragingly, the addn. of a single V2 term involving the atoms C(sp3)-O-(sp3)-C(sp3)-N(sp2), which can be nicely rationalized because of the anomeric effect (lone pairs on oxygen are preferentially oriented relative to the electron withdrawing N), brings the sugar pucker phase of C2'-endo sugars to near perfect agreement with ab initio calcns. (W near 162°). Secondly, the use of high level ab initio calcns. on entire nucleosides (in contrast to smaller model systems necessitated in 1994-95 by computer limitations) lets one improve the χ torsional potential for nucleic acids. Finally, the O(sp3)-C(sp3)-C(sp3)-O(sp3) V2 torsional potential has been empirically adjusted to reproduce the ab initio calcd. relative energy of C2'-endo and C3'-endo nucleosides. These modifications are tested in mol. dynamics simulations of mononucleosides (to assess sugar pucker percentages) and double helixes of DNA and RNA (to assess helical and sequence specific structural properties). In both areas, the modified force field leads to improved agreement with exptl. data.
- 18MacKerell, A. D.; Banavali, N. K. J. Comput. Chem. 2000, 21 (2) 105– 120There is no corresponding record for this reference.
- 19Foloppe, N.; MacKerell, A. D. J. Comput. Chem. 2000, 21 (2) 86– 104There is no corresponding record for this reference.
- 20Wang, J. M.; Cieplak, P.; Kollman, P. A. J. Comput. Chem. 2000, 21 (12) 1049– 1074There is no corresponding record for this reference.
- 21Cheatham, T. E.; Young, M. A. Biopolymers 2000, 56 (4) 232– 256There is no corresponding record for this reference.
- 22Varnai, P.; Djuranovic, D.; Lavery, R.; Hartmann, B. Nucleic Acids Res. 2002, 30 (24) 5398– 5406There is no corresponding record for this reference.
- 23Dixit, S. B.; Beveridge, D. L.; Case, D. A.; Cheatham, T. E.; Giudice, E.; Lankas, F.; Lavery, R.; Maddocks, J. H.; Osman, R.; Sklenar, H.; Thayer, K. M.; Varnai, P. Biophys. J. 2005, 89 (6) 3721– 3740There is no corresponding record for this reference.
- 24Perez, A.; Marchan, I.; Svozil, D.; Sponer, J.; Cheatham, T. E.; Laughton, C. A.; Orozco, M. Biophys. J. 2007, 92 (11) 3817– 3829There is no corresponding record for this reference.
- 25Svozil, D.; Sponer, J. E.; Marchan, I.; Perez, A.; Cheatham, T. E.; Forti, F.; Luque, F. J.; Orozco, M.; Sponer, J. J. Phys. Chem. B 2008, 112 (27) 8188– 8197There is no corresponding record for this reference.
- 26Perez, A.; Luque, F. J.; Orozco, M. J. Am. Chem. Soc. 2007, 129 (47) 14739– 14745There is no corresponding record for this reference.
- 27Lavery, R.; Zakrzewska, K.; Beveridge, D.; Bishop, T. C.; Case, D. A.; Cheatham, T.; Dixit, S.; Jayaram, B.; Lankas, F.; Laughton, C.; Maddocks, J. H.; Michon, A.; Osman, R.; Orozco, M.; Perez, A.; Singh, T.; Spackova, N.; Sponer, J. Nucleic Acids Res. 2010, 38 (1) 299– 313There is no corresponding record for this reference.
- 28Lankas, F.; Spackova, N.; Moakher, M.; Enkhbayar, P.; Sponer, J. Nucleic Acids Res. 2010, 38 (10) 3414– 3422There is no corresponding record for this reference.
- 29Fadrna, E.; Spackova, N.; Sarzynska, J.; Koca, J.; Orozco, M.; Cheatham, T. E.; Kulinski, T.; Sponer, J. J. Chem. Theory Comput. 2009, 5 (9) 2514– 253029https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhtVWku7vN&md5=7ae168c141e60affa2fcda871a3b3057Single Stranded Loops of Quadruplex DNA As Key Benchmark for Testing Nucleic Acids Force FieldsFadrna, Eva; Spackova, Nada; Sarzynska, Joanna; Koca, Jaroslav; Orozco, Modesto; Cheatham, Thomas E., III; Kulinski, Tadeusz; Sponer, JiriJournal of Chemical Theory and Computation (2009), 5 (9), 2514-2530CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)We have carried out a set of explicit solvent mol. dynamics (MD) simulations on two DNA quadruplex (G-DNA) mols., namely the antiparallel d(G4T4G4)2 dimeric quadruplex with diagonal loops and the parallel-stranded human telomeric monomol. quadruplex d[AGGG(TTAGGG)3] with three propeller loops. The main purpose of the paper was testing of the capability of the MD simulation technique to describe single-stranded topologies of G-DNA loops, which represent a very challenging task for computational methods. The total amt. of conventional and locally enhanced sampling (LES) simulations analyzed in this study exceeds 1.5 μs, while we tested several versions of the AMBER force field (parm99, parmbsc0, and a version with modified glycosidic χ torsion profile) and the CHARMM27 force field. Further, we compared minimal salt and excess salt simulations. Post-processing MM-PBSA (Mol. Mechanics, Poisson-Boltzmann, Surface Area) free energy calcns. are also reported. None of the presently available force fields is accurate enough in describing the G-DNA loops. The imbalance is best seen for the propeller loops, as their exptl. structure is lost within a few ns of std. simulations with all force fields. Among them, parmbsc0 provides results that are clearly closest to the exptl. target values but still not in full agreement. This confirms that the improvement of the γ torsional profile penalizing the γ trans substates in the parmbsc0 parametrization was a step in the right direction, albeit not sufficient to treat all imbalances. The modified χ parameterization appears to rigidify the studied systems but does not change the ultimate outcome of the present simulations. The structures obtained in simulations with the modified χ profile are predetd. by its combination with either parm99 or parmbsc0. Exptl. geometries of diagonal loops of d(G4T4G4)2 are stable in std. simulations on the ∼10 ns time scale but are becoming progressively lost in longer and LES simulations. In addn., the d(G4T4G4)2 quadruplex contains, besides the three genuine binding sites for cations in the channel of its stem, also an ion binding site at each stem-loop junction. This arrangement of five cations in the quadruplex core region is entirely unstable in all 24 simulations that we attempted. Overall, our results confirm that G-DNA loops represent one of the most difficult targets for mol. modeling approaches and should be considered as ref. structures in any future studies aiming to develop or tune nucleic acids force fields.
- 30Reblova, K.; Lankas, F.; Razga, F.; Krasovska, M. V.; Koca, J.; Sponer, J. Biopolymers 2006, 82 (5) 504– 520There is no corresponding record for this reference.
- 31Blount, K. F.; Breaker, R. R. Nat. Biotechnol. 2006, 24 (12) 1558– 1564There is no corresponding record for this reference.
- 32Strobel, S. A.; Cochrane, J. C. Curr. Opin. Chem. Biol. 2007, 11 (6) 636– 64332https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXhtlyrtbrP&md5=4fcdfa52636d776969f6c5d939991fb5RNA catalysis: ribozymes, ribosomes, and riboswitchesStrobel, Scott A.; Cochrane, Jesse C.Current Opinion in Chemical Biology (2007), 11 (6), 636-643CODEN: COCBF4; ISSN:1367-5931. (Elsevier B.V.)A review. The catalytic mechanisms employed by RNA are chem. more diverse than initially suspected. Divalent metal ions, nucleobases, ribosyl hydroxyl groups, and even functional groups on metabolic cofactors all contribute to the various strategies employed by RNA enzymes. This catalytic breadth raises intriguing evolutionary questions about how RNA lost its biol. role in some cases, but not in others, and what catalytic roles RNA might still be playing in biol.
- 33Montange, R. K.; Batey, R. T. Annu. Rev. Biophys. 2008, 37, 117– 133There is no corresponding record for this reference.
- 34Steitz, T. A. Nat. Rev. Mol. Cell Biol. 2008, 9 (3) 242– 25334https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXitlGnu7w%253D&md5=e82078dbacb84a56b48507e6b703df8dA structural understanding of the dynamic ribosome machineSteitz, Thomas A.Nature Reviews Molecular Cell Biology (2008), 9 (3), 242-253CODEN: NRMCBP; ISSN:1471-0072. (Nature Publishing Group)A review. Ribosomes, which are central to protein synthesis and convert transcribed mRNAs into polypeptide chains, have been the focus of structural and biochem. studies for >50 yr. The structure of its larger subunit has revealed that the ribosome is a ribozyme with RNA at the heart of its enzymic activity that catalyzes peptide bond formation. Numerous initiation, elongation, and release factors ensure that protein synthesis occurs progressively and with high specificity. In the past few years, high-resoln. structures have provided mol. snapshots of different intermediates in ribosome-mediated translation in at. detail. Together, these studies have revolutionized the understanding of the mechanism of protein synthesis.
- 35Paulsen, R. B.; Seth, P. P.; Swayze, E. E.; Griffey, R. H.; Skalicky, J. J.; Cheatham, T. E., 3rd; Davis, D. R. Proc. Natl. Acad. Sci. U.S.A. 2010, 107 (16) 7263– 7268There is no corresponding record for this reference.
- 36Reddy, S. Y.; Leclerc, F.; Karplus, M. Biophys. J. 2003, 84 (3) 1421– 1449There is no corresponding record for this reference.
- 37Besseova, I.; Otyepka, M.; Reblova, K.; Sponer, J. Phys. Chem. Chem. Phys. 2009, 11 (45) 10701– 10711There is no corresponding record for this reference.
- 38Deng, N. J.; Cieplak, P. Biophys. J. 2010, 98 (4) 627– 636There is no corresponding record for this reference.
- 39Ricci, C. G.; de Andrade, A. S. C.; Mottin, M.; Netz, P. A. J. Phys. Chem. B 2010, 114 (30) 9882– 9893There is no corresponding record for this reference.
- 40Auffinger, P.; Westhof, E. Curr. Opin. Struct. Biol. 1998, 8 (2) 227– 23640https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXjt1Gktr0%253D&md5=b13df97d5a7f78d19b8dad6f28b68eceSimulations of the molecular dynamics of nucleic acidsAuffinger, Pascal; Westhof, EricCurrent Opinion in Structural Biology (1998), 8 (2), 227-236CODEN: COSBEF; ISSN:0959-440X. (Current Biology Ltd.)A review with 91 refs. The growing amt. of high quality mol. dynamics simulations generated using the latest methodol. developments and force fields has led to a sharper understanding of the forces underlying the dynamics of biomol. systems, as well as to stimulating insights into the structure and catalysis of nucleic acids. It is now clear that inclusion of long-range electrostatic interactions and of the aq. and ionic environment is necessary for producing realistic and accurate simulations. Yet, many papers hint at a force field and protocol dependence of the results and thus contain the seeds for the future improvements that will e necessary for deepening our understanding of recognition phenomena and folding of nucleic acids.
- 41Bosch, D.; Foloppe, N.; Pastor, N.; Pardo, L.; Campillo, M. J. Mol. Struc.: THEOCHEM 2001, 537, 283– 305There is no corresponding record for this reference.
- 42Foloppe, N.; MacKerell, A. D. J. Phys. Chem. B 1999, 103 (49) 10955– 10964There is no corresponding record for this reference.
- 43Mlynsky, V.; Banas, P.; Hollas, D.; Reblova, K.; Walter, N. G.; Sponer, J.; Otyepka, M. J. Phys. Chem. B 2010, 114 (19) 6642– 6652There is no corresponding record for this reference.
- 44Ode, H.; Matsuo, Y.; Neya, S.; Hoshino, T. J. Comput. Chem. 2008, 29 (15) 2531– 2542There is no corresponding record for this reference.
- 45Yildirim, I.; Stern, H. A.; Kennedy, S. D.; Tubbs, J. D.; Turner, D. H. J. Chem. Theory Comput. 2010, 6 (5) 1520– 153145https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXkvV2lurg%253D&md5=43f9653ecb996aecc3274434b4a24c3eReparameterization of RNA χ Torsion Parameters for the AMBER Force Field and Comparison to NMR Spectra for Cytidine and UridineYildirim, Ilyas; Stern, Harry A.; Kennedy, Scott D.; Tubbs, Jason D.; Turner, Douglas H.Journal of Chemical Theory and Computation (2010), 6 (5), 1520-1531CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)A reparameterization of the torsional parameters for the glycosidic dihedral angle, χ, for the AMBER99 force field in RNA nucleosides is used to provide a modified force field, AMBER99χ. Mol. dynamics simulations of cytidine, uridine, adenosine, and guanosine in aq. soln. using the AMBER99 and AMBER99χ force fields are compared with NMR results. For each nucleoside and force field, 10 individual mol. dynamics simulations of 30 ns each were run. For cytidine with AMBER99χ force field, each mol. dynamics simulation time was extended to 120 ns for convergence purposes. NMR spectroscopy, including one-dimensional (1D) 1H, steady-state 1D 1H nuclear Overhauser effect (NOE), and transient 1D 1H NOE, was used to det. the sugar puckering and preferred base orientation with respect to the ribose of cytidine and uridine. The AMBER99 force field overestimates the population of syn conformations of the base orientation and of C2'-endo sugar puckering of the pyrimidines, while the AMBER99χ force field's predictions are more consistent with NMR results. Moreover, the AMBER99 force field prefers high anti conformations with glycosidic dihedral angles around 310° for the base orientation of purines. The AMBER99χ force field prefers anti conformations around 185°, which is more consistent with the quantum mech. calcns. and known 3D structures of folded ribonucleic acids (RNAs). Evidently, the AMBER99χ force field predicts the structural characteristics of ribonucleosides better than the AMBER99 force field and should improve structural and thermodn. predictions of RNA structures.
- 46Banáš, P.; Hollas, D.; Zgarbova, M.; Jurecka, P.; Orozco, M.; Cheatham, T.; Sponer, J.; Otyepka, M. J. Chem. Theory Comput. 2010, 6 (12) 3836– 384946https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhsVSisLnK&md5=b02fa19c2fb52e1d36476274c86aaeeaPerformance of Molecular Mechanics Force Fields for RNA Simulations: Stability of UUCG and GNRA HairpinsBanas, Pavel; Hollas, Daniel; Zgarbova, Marie; Jurecka, Petr; Orozco, Modesto; Cheatham, Thomas E., III; Sponer, Jiri; Otyepka, MichalJournal of Chemical Theory and Computation (2010), 6 (12), 3836-3849CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The RNA hairpin loops represent important RNA topologies with indispensable biol. functions in RNA folding and tertiary interactions. 5'-UNCG-3' and 5'-GNRA-3' RNA tetraloops are the most important classes of RNA hairpin loops. Both tetraloops are highly structured with characteristic signature three-dimensional features and are recurrently seen in functional RNAs and ribonucleoprotein particles. Explicit solvent mol. dynamics (MD) simulation is a computational technique which can efficiently complement the exptl. data and provide unique structural dynamics information on the at. scale. Nevertheless, the outcome of simulations is often compromised by imperfections in the parametrization of simplified pairwise additive empirical potentials referred to also as force fields. We have pointed out in several recent studies that a force field description of single-stranded hairpin segments of nucleic acids may be particularly challenging for the force fields. In this paper, we report a crit. assessment of a broad set of MD simulations of UUCG, GAGA, and GAAA tetraloops using various force fields. First, we utilized the three widely used variants of Cornell et al. (AMBER) force fields known as ff94, ff99, and ff99bsc0. Some simulations were also carried out with CHARMM27. The simulations reveal several problems which show that these force fields are not able to retain all characteristic structural features (structural signature) of the studied tetraloops. Then we tested four recent reparameterizations of glycosidic torsion of the Cornell et al. force field (two of them being currently parametrized in our labs.). We show that at least some of the new versions show an improved description of the tetraloops, mainly in the syn glycosidic torsion region of the UNCG tetraloop. The best performance is achieved in combination with the bsc0 parametrization of the α/γ angles. Another critically important region to properly describe RNA mols. is the anti/high-anti region of the glycosidic torsion, where there are significant differences among the tested force fields. The tetraloop simulations are complemented by simulations of short A-RNA stems, which are esp. sensitive to an appropriate description of the anti/high-anti region. While excessive accessibility of the high-anti region converts the A-RNA into a senseless "ladder-like" geometry, excessive penalization of the high-anti region shifts the simulated structures away from typical A-RNA geometry to structures with a visibly underestimated inclination of base pairs with respect to the helical axis.
- 47Fadrna, E.; Spackova, N.; Stefl, R.; Koca, J.; Cheatham, T. E.; Sponer, J. Biophys. J. 2004, 87 (1) 227– 242There is no corresponding record for this reference.
- 48Reblova, K.; Fadrna, E.; Sarzynska, J.; Kulinski, T.; Kulhanek, P.; Ennifar, E.; Koca, J.; Sponer, J. Biophys. J. 2007, 93 (11) 3932– 3949There is no corresponding record for this reference.
- 49Faustino, I.; Perez, A.; Orozco, M. Biophys. J. 2010, 99 (6) 1876– 1885There is no corresponding record for this reference.
- 50Foloppe, N.; MacKerell, A. D. Biophys. J. 1999, 76 (6) 3206– 3218There is no corresponding record for this reference.
- 51Jurecka, P.; Hobza, P. J. Am. Chem. Soc. 2003, 125 (50) 15608– 15613There is no corresponding record for this reference.
- 52Mladek, A.; Sponer, J. E.; Jurecka, P.; Banas, P.; Otyepka, M.; Svozil, D.; Sponer, J. J. Chem. Theory Comput. 2010, 6 (12) 3817– 383552https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhsVegurfK&md5=25b44243e0c15ce1315e85b277cd5aa0Conformational Energies of DNA Sugar-Phosphate Backbone: Reference QM Calculations and a Comparison with Density Functional Theory and Molecular MechanicsMladek, Arnost; Sponer, Judit E.; Jurecka, Petr; Banas, Pavel; Otyepka, Michal; Svozil, Daniel; Sponer, JiriJournal of Chemical Theory and Computation (2010), 6 (12), 3817-3835CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The study investigates electronic structure and gas-phase energetics of the DNA sugar-phosphate backbone via advanced quantum chem. (QM) methods. The anal. has been carried out on biol. relevant backbone conformations composed of 11 canonical BI-DNA structures, 8 pathol. structures with α/γ torsion angles in the g+/t region, and 3 real noncanonical γ-trans structures occurring in the loop region of guanine quadruplex DNA. The influence of backbone conformation on the intrinsic energetics was primarily studied using a model system consisting of two sugar moieties linked together via a phosphodiester bond (SPSOM model). To get the conformation of the studied system fully under control, for each calcn. we have frozen majority of the dihedral angles to their target values. CCSD(T) energies extrapolated to the complete basis set were utilized as ref. values. However, the calcns. show that inclusion of higher-order electron correlation effects for this system is not crucial and complete basis set second-order perturbation calcns. are sufficiently accurate. The ref. QM data are used to assess performance of 10 contemporary d. functionals with the best performance delivered by the PBE-D/TZVPP combination along with the Grimme's dispersion correction, and by the TPSS-D/6-311++G(3df,3pd) augmented by Jurecka's dispersion term. In addn., the QM calcns. are compared to mol. mechanics (MM) model based on the Cornell et al. force field. The destabilization of the pathol. g+/t conformers with respect to the ref. canonical structure and the network of intramol. CH···O interactions were investigated by means of natural bond orbital anal. (NBO) and atoms-in-mols. (AIM) Bader anal. Finally, four addnl. model systems of different sizes were assessed by comparing their energetics to that of the SPSOM system. Energetics of smaller MOSPM model consisting of a sugar moiety linked to a phosphate group and capped with Me and methoxy group on the 5'- and 3'-ends, resp., is fairly similar to that of SPSOM, while the role of undesired intramol. interactions is diminished.
- 53Halkier, A.; Helgaker, T.; Jorgensen, P.; Klopper, W.; Koch, H.; Olsen, J.; Wilson, A. K. Chem. Phys. Lett. 1998, 286 (3–4) 243– 252There is no corresponding record for this reference.
- 54Halkier, A.; Helgaker, T.; Jorgensen, P.; Klopper, W.; Olsen, J. Chem. Phys. Lett. 1999, 302 (5–6) 437– 446There is no corresponding record for this reference.
- 55Ahlrichs, R.; Bar, M.; Haser, M.; Horn, H.; Kolmel, C. Chem. Phys. Lett. 1989, 162 (3) 165– 16955https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3cXkt1yrtg%253D%253D&md5=b6aa32e6226a8e11b511e4d09cc60dc8Electronic structure calculations on workstation computers: the program system TURBOMOLEAhlrichs, Reinhart; Baer, Michael; Haeser, Marco; Horn, Hans; Koelmel, ChristophChemical Physics Letters (1989), 162 (3), 165-9CODEN: CHPLBC; ISSN:0009-2614.The basic structure of the program system TURBOMOLE for SCF - including first and second anal. derivs. with respect to nuclear coordinates - and MP2 calcns. is briefly described. The program takes full advantage of all discrete point group symmetries and has only modest - and (partially) adjustable - I/O and background storage requirements. The performance of TURBOMOLE is documented for demonstrative applications.
- 56Weigend, F.; Häser, M. Theor. Chem. Acc. 1997, 97 (1–4) 331– 34056https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXmvFCitL8%253D&md5=14ae7a8718188931367e3a192be50351RI-MP2. First derivatives and global consistencyWeigend, Florian; Haser, MarcoTheoretical Chemistry Accounts (1997), 97 (1-4), 331-340CODEN: TCACFW; ISSN:1432-881X. (Springer)The evaluation of RI-MP2 first derivs. with respect to nuclear coordinates or with respect to an external elec. field is described. The prefix RI indicates the use of an approx. resoln. of identity in the Hilbert space of interacting charge distributions (Coulomb metric), i.e., the use of an auxiliary basis set to approx. charge distributions. The RI technique is applied to first derivs. of the MP2 correlation energy expression while the (restricted) Hartree-Fock ref. is treated in the usual way. Computational savings by a factor of 10 over conventional approaches are demonstrated in an application to porphyrin. The RI approxn. to MP2 derivs. does not entail any significant loss in accuracy. Finally, the relative energetic stabilities of a representative sample of closed-shell mols. built from first and second row elements were investigated by the RI-MP2 approach, and thus it is tested whether such properties that refer to potential energy hypersurfaces in a more global way can be described with similar consistency to the more locally defined derivs.
- 57Werner, H. J.; Knowles, P. J.; Lindh, R.; Manby, F. R.; Schütz, M.; Celani, P.; Korona, T.; Rauhut, G.; Amos, R. D.; Bernhardsson, A.; Berning, A.; Cooper, D. L.; Deegan, M. J. O.; Dobbyn, A. J.; Eckert, F.; Hampel, C.; Hetzer, G.; Lloyd, A. W.; McNicholas, S. J.; Meyer, W.; Mura, M. E.; Nicklass, A.; Palmieri, P.; Pitzer, R.; Schumann, U.; Stoll, H.; Stone, A. J.; Tarroni, R.; Thorsteinsson, T. Molpro Version 2006.1, a package of ab initio programs; 2006; http://www.molpro.net (accessed July 2011).There is no corresponding record for this reference.
- 58Krishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A. J. Chem. Phys. 1980, 72 (1) 650– 654There is no corresponding record for this reference.
- 59Clark, T.; Chandrasekhar, J.; Spitznagel, G. W.; Schleyer, P. V. J. Comput. Chem. 1983, 4 (3) 294– 301There is no corresponding record for this reference.
- 60Gill, P. M. W.; Johnson, B. G.; Pople, J. A. J. Chem. Phys. 1992, 96 (9) 7178– 7179There is no corresponding record for this reference.
- 61Frisch, M. J.; Pople, J. A.; Binkley, J. S. J. Chem. Phys. 1984, 80 (7) 3265– 3269There is no corresponding record for this reference.
- 62Jurecka, P.; Cerny, J.; Hobza, P.; Salahub, D. R. J. Comput. Chem. 2007, 28 (2) 555– 569There is no corresponding record for this reference.
- 63Schafer, A.; Huber, C.; Ahlrichs, R. J. Chem. Phys. 1994, 100 (8) 5829– 5835There is no corresponding record for this reference.
- 64Klamt, A.; Schuurmann, G. J. Chem. Soc., Perkin Trans. 2 1993, 5) 799– 80564https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXkvVSku7o%253D&md5=a8e9e41514d60ca7f11212cf5c281424COSMO: a new approach to dielectric screening in solvents with explicit expressions for the screening energy and its gradientKlamt, A.; Schueuermann, G.Journal of the Chemical Society, Perkin Transactions 2: Physical Organic Chemistry (1972-1999) (1993), (5), 799-805CODEN: JCPKBH; ISSN:0300-9580.Starting from the screening in conductors, an algorithm for the accurate calcn. of dielec. screening effects in solvents is presented, which leads to rather simple explicit expressions for the screening energy and its analytic gradient with respect to the solute coordinates. Thus geometry optimization of a solute within a realistic dielec. continuum model becomes practicable for the first time. The algorithm is suited for mol. mechanics as well as for any MO algorithm. The implementation into MOPAC and some example applications are reported.
- 65Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, J.A.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A. Gaussian 03, revision D.02; Gaussian, Inc.: Wallingford, CT, 2004.There is no corresponding record for this reference.
- 66Case, D. A.; Cheatham, T. E.; Darden, T.; Gohlke, H.; Luo, R.; Merz, K. M.; Onufriev, A.; Simmerling, C.; Wang, B.; Woods, R. J. J. Comput. Chem. 2005, 26 (16) 1668– 1688There is no corresponding record for this reference.
- 67Lu, Q.; Luo, R. J. Chem. Phys. 2003, 119 (21) 11035– 11047There is no corresponding record for this reference.
- 68Luo, R.; David, L.; Gilson, M. K. J. Comput. Chem. 2002, 23 (13) 1244– 1253There is no corresponding record for this reference.
- 69Guvench, O.; MacKerell, A. D., Jr. J. Mol. Model. 2008, 14 (8) 667– 67969https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhtVKgtLo%253D&md5=b5bd1c35ed00665a83be39345f88a6feAutomated conformational energy fitting for force-field developmentGuvench, Olgun; MacKerell, Alexander D., Jr.Journal of Molecular Modeling (2008), 14 (8), 667-679CODEN: JMMOFK; ISSN:0948-5023. (Springer GmbH)We present a general conformational-energy fitting procedure based on Monte Carlo simulated annealing (MCSA) for application in the development of mol. mechanics force fields. Starting with a target potential energy surface and an unparametrized mol. mechanics potential energy surface, an optimized set of either dihedral or grid-based correction map (CMAP) parameters is produced that minimizes the root mean squared error RMSE between the parametrized and targeted energies. The fitting is done using an MCSA search in parameter space and consistently converges to the same RMSE irresp. of the randomized parameters used to seed the search. Any no. of dihedral parameters can be simultaneously parametrized, allowing for fitting to multi-dimensional potential energy scans. Fitting options for dihedral parameters include non-uniform weighting of the target data, constraining multiple optimized parameters to the same value, constraining parameters to be no greater than a user-specified max. value, including all or only a subset of multiplicities defining the dihedral Fourier series, and optimization of phase angles in addn. to force consts. The dihedral parameter fitting algorithm's performance is characterized through multi-dimensional fitting of cyclohexane, tetrahydropyran, and hexopyranose monosaccharide energetics, with the latter case having a 30-dimensional parameter space. The CMAP fitting is applied in the context of polypeptides, and is used to develop a parametrization that simultaneously captures the .vphi.,ψ energetics of the alanine dipeptide and the alanine tetrapeptide. Because the dihedral energy term is common to many force fields, we have implemented the dihedral-fitting algorithm in the portable Python scripting language and have made it freely available as "fit_dihedral.py" for download at http://mackerell.umaryland.edu.
- 70Duan, Y.; Wu, C.; Chowdhury, S.; Lee, M. C.; Xiong, G. M.; Zhang, W.; Yang, R.; Cieplak, P.; Luo, R.; Lee, T.; Caldwell, J.; Wang, J. M.; Kollman, P. J. Comput. Chem. 2003, 24 (16) 1999– 201270https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXovVygsbc%253D&md5=b4531cbf2d90deb9a173ec314dfd7b5dA point-charge force field for molecular mechanics simulations of proteins based on condensed-phase quantum mechanical calculationsDuan, Yong; Wu, Chun; Chowdhury, Shibasish; Lee, Mathew C.; Xiong, Guoming; Zhang, Wei; Yang, Rong; Cieplak, Piotr; Luo, Ray; Lee, Taisung; Caldwell, James; Wang, Junmei; Kollman, PeterJournal of Computational Chemistry (2003), 24 (16), 1999-2012CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)Mol. mechanics models have been applied extensively to study the dynamics of proteins and nucleic acids. Here we report the development of a third-generation point-charge all-atom force field for proteins. Following the earlier approach of Cornell et al., the charge set was obtained by fitting to the electrostatic potentials of dipeptides calcd. using B3LYP/cc-pVTZ/HF/6-31G** quantum mech. methods. The main-chain torsion parameters were obtained by fitting to the energy profiles of Ace-Ala-Nme and Ace-Gly-Nme di-peptides calcd. using MP2/cc-pVTZ//HF/6-31G** quantum mech. methods. All other parameters were taken from the existing AMBER data base. The major departure from previous force fields is that all quantum mech. calcns. were done in the condensed phase with continuum solvent models and an effective dielec. const. of ε = 4. We anticipate that this force field parameter set will address certain crit. short comings of previous force fields in condensed-phase simulations of proteins. Initial tests on peptides demonstrated a high-degree of similarity between the calcd. and the statistically measured Ramanchandran maps for both Ace-Gly-Nme and Ace-Ala-Nme di-peptides. Some highlights of our results include (1) well-preserved balance between the extended and helical region distributions, and (2) favorable type-II poly-proline helical region in agreement with recent expts. Backward compatibility between the new and Cornell et al. charge sets, as judged by overall agreement between dipole moments, allows a smooth transition to the new force field in the area of ligand-binding calcns. Test simulations on a large set of proteins are also discussed.
- 71Dockbregeon, A. C.; Chevrier, B.; Podjarny, A.; Johnson, J.; Debear, J. S.; Gough, G. R.; Gilham, P. T.; Moras, D. J. Mol. Biol. 1989, 209 (3) 459– 474There is no corresponding record for this reference.
- 72Drew, H. R.; Wing, R. M.; Takano, T.; Broka, C.; Tanaka, S.; Itakura, K.; Dickerson, R. E. Proc. Natl. Acad. Sci.: Biol. 1981, 78 (4) 2179– 218372https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3MXktVajt74%253D&md5=1278977a90de8b049e2577294d569c45Structure of a B-DNA dodecamer. I. Conformation and dynamicsDrew, Horace R.; Wing, Richard M.; Takano, Tsunehiro; Broka, Christopher; Tanaka, Shoji; Itakura, Keiichi; Dickerson, Richard E.Proceedings of the National Academy of Sciences of the United States of America (1981), 78 (4), 2179-83CODEN: PNASA6; ISSN:0027-8424.The crystal structure of the synthetic DNA dodecamer d(CpGpCpGpApApTpTpCpGpCpG) was refined to a residual error of R = 17.8% at 1.9-Å resoln. (2-σ data). The mol. forms slightly >1 complete turn of right-handed double-stranded B helix. The 2 ends of the helix overlap and interlock minor grooves with neighboring mols. up and down a 21 screw axis, producing a 19° bend in helix axis over the 11-base-pair steps of the dodecamer. In the center of the mol., where perturbation is least, the helix has a mean rotation of 36.9° per step, or 9.8 base pairs per turn. The mean propeller twist (total dihedral angle between base planes) between A·T base pairs in the center of the mol. is 17.3°, and that between C·G pairs on the 2 ends avs. 11.5°. Individual deoxyribose ring conformations, measured by the C5'-C4'-C3'-O3' torsion angle δ, exhibit an approx. Gaussian distribution centered around the C1'-exo position with δav. = 123° and a range of 79-157°. Purine sugars cluster at high δ values, and pyrimidine sugars cluster at lower δ. A tendency toward 2-fold symmetry in sugar conformation about the center of the mol. is detectable in spite of the destruction of ideal 2-fold symmetry by the mol. bending. More strikingly, sugar conformations of paired bases appear to follow a principle of anticorrelation, with δ values lying approx. the same distance to either side of the center value, δ = 123°. This same anticorrelation is also obsd. in other DNA and DNA·RNA structures.
- 73Timsit, Y.; Bombard, S. RNA 2007, 13 (12) 2098– 2107There is no corresponding record for this reference.
- 74Klosterman, P. S.; Shah, S. A.; Steitz, T. A. Biochemistry 1999, 38 (45) 14784– 14792There is no corresponding record for this reference.
- 75Aqvist, J. J. Phys. Chem. 1990, 94 (21) 8021– 8024There is no corresponding record for this reference.
- 76Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R. W.; Klein, M. L. J. Chem. Phys. 1983, 79 (2) 926– 93576https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3sXksF2htL4%253D&md5=a1161334e381746be8c9b15a5e56f704Comparison of simple potential functions for simulating liquid waterJorgensen, William L.; Chandrasekhar, Jayaraman; Madura, Jeffry D.; Impey, Roger W.; Klein, Michael L.Journal of Chemical Physics (1983), 79 (2), 926-35CODEN: JCPSA6; ISSN:0021-9606.Classical Monte Carlo simulations were carried out for liq. H2O in the NPT ensemble at 25° and 1 atm using 6 of the simpler intermol. potential functions for the dimer. Comparisons were made with exptl. thermodn. and structural data including the neutron diffraction results of Thiessen and Narten (1982). The computed densities and potential energies agree with expt. except for the original Bernal-Fowler model, which yields an 18% overest. of the d. and poor structural results. The discrepancy may be due to the correction terms needed in processing the neutron data or to an effect uniformly neglected in the computations. Comparisons were made for the self-diffusion coeffs. obtained from mol. dynamics simulations.
- 77Lu, X. J.; Olson, W. K. Nucleic Acids Res. 2003, 31 (17) 5108– 5121There is no corresponding record for this reference.
- 78Holroyd, L. F.; van Mourik, T. Chem. Phys. Lett. 2007, 442 (1–3) 42– 46There is no corresponding record for this reference.
- 79Valdes, H.; Klusak, V.; Pitonak, M.; Exner, O.; Stary, I.; Hobza, P.; Rulisek, L. J. Comput. Chem. 2008, 29 (6) 861– 870There is no corresponding record for this reference.
- 80Jensen, F. J. Chem. Theory Comput. 2010, 6 (1) 100– 10680https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhtlOksrzM&md5=dc6efcef8f646babf9de2136568222c5An Atomic Counterpoise Method for Estimating Inter- and Intramolecular Basis Set Superposition ErrorsJensen, FrankJournal of Chemical Theory and Computation (2010), 6 (1), 100-106CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)An at. counterpoise method is proposed to calc. ests. of inter- and intramol. basis set superposition errors. The method ests. the basis set superposition error as a sum of at. contributions and can be applied for both independent particle and electron correlation models. The at. counterpoise method provides results very similar to the mol. counterpoise method for intermol. basis set superposition errors at both the HF and MP2 levels of theory with a sequence of increasingly larger basis sets. The advantage of the at. counterpoise method is that it can be applied with equal ease to est. intramol. basis set superposition errors, for which few other methods exist. The at. counterpoise method is computationally quite efficient, requiring typically double the amt. of computer time as required for calcg. the uncorrected energy.
- 81Zhao, Y.; Truhlar, D. G. Theor. Chem. Acc. 2008, 120 (1–3) 215– 24181https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXltFyltbY%253D&md5=c31d6f319d7c7a45aa9b716220e4a422The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionalsZhao, Yan; Truhlar, Donald G.Theoretical Chemistry Accounts (2008), 120 (1-3), 215-241CODEN: TCACFW; ISSN:1432-881X. (Springer GmbH)We present two new hybrid meta exchange-correlation functionals, called M06 and M06-2X. The M06 functional is parametrized including both transition metals and nonmetals, whereas the M06-2X functional is a high-nonlocality functional with double the amt. of nonlocal exchange (2X), and it is parametrized only for nonmetals. The functionals, along with the previously published M06-L local functional and the M06-HF full-Hartree-Fock functionals, constitute the M06 suite of complementary functionals. We assess these four functionals by comparing their performance to that of 12 other functionals and Hartree-Fock theory for 403 energetic data in 29 diverse databases, including ten databases for thermochem., four databases for kinetics, eight databases for noncovalent interactions, three databases for transition metal bonding, one database for metal atom excitation energies, and three databases for mol. excitation energies. We also illustrate the performance of these 17 methods for three databases contg. 40 bond lengths and for databases contg. 38 vibrational frequencies and 15 vibrational zero point energies. We recommend the M06-2X functional for applications involving main-group thermochem., kinetics, noncovalent interactions, and electronic excitation energies to valence and Rydberg states. We recommend the M06 functional for application in organometallic and inorganometallic chem. and for noncovalent interactions.
- 82Zgarbova, M.; Otyepka, M.; Sponer, J.; Hobza, P.; Jurecka, P. Phys. Chem. Chem. Phys. 2010, 12, 10476− 10493.There is no corresponding record for this reference.
- 83Neidle, S. In Nucleic Acid Structure and Recognition;Oxford University Press Inc.: Oxford, 2002.There is no corresponding record for this reference.
- 84Sklenovsky, P.; Florova, P.; Banas, P.; Reblova, K.; Lankas, F.; Otyepka, M.; Sponer, J. J. Chem. Theory Comput. accepted for publication.There is no corresponding record for this reference.
- 85Merz, K. M. J. Chem. Theory Comput. 2010, 6 (4) 1018– 1027There is no corresponding record for this reference.
- 86Simmerling, C.; Strockbine, B.; Roitberg, A. E. J. Am. Chem. Soc. 2002, 124 (38) 11258– 11259There is no corresponding record for this reference.
- 87Feig, M.; Pettitt, B. M. Biophys. J. 1998, 75 (1) 134– 149There is no corresponding record for this reference.
- 88Altona, C.; Sundaralingam, M. J. Am. Chem. Soc. 1972, 94 (23) 8205– 8212There is no corresponding record for this reference.
- 89Sponer, J.; Kypr, J. J. Mol. Biol. 1991, 221 (3) 761– 764There is no corresponding record for this reference.
- 90Bhattacharyya, D.; Bansal, M. J. Biomol. Struct. Dyn. 1989, 6 (4) 635– 65390https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXhsFSqsro%253D&md5=9df52ea7d56288b5e508bad5858e4a08A self-consistent formulation for analysis and generation of non-uniform DNA structuresBhattacharyya, Dhananjay; Bansal, ManjuJournal of Biomolecular Structure & Dynamics (1989), 6 (4), 635-53CODEN: JBSDD6; ISSN:0739-1102.A fully self-consistent formulation is described here for the anal. and generation of base-pairs in nonuniform DNA structures, in terms of various local parameters. The internal wedge parameters are math. related to the parameters describing the base-pair orientation with respect to an external helix axis. Hence any one set of three translation and three rotation parameters are necessary and sufficient to completely describe the relative orientation of the base-pairs comprising a step (or doublet). A general procedure is outlined for obtaining an av. or global helix axis from the local helix axes for each step. A graphical representation of the local helix axes in the form of a polar plot is also shown and its application for estg. the curvature of oligonucleotide structures is illustrated, with examples of both A and B type structures.
- 91Besseova, I.; Reblova, K.; Leontis, N. B.; Sponer, J. Nucleic Acids Res. 2010, 38 (18) 6247– 6264There is no corresponding record for this reference.
- 92Tolbert, B. S.; Miyazaki, Y.; Barton, S.; Kinde, B.; Starck, P.; Singh, R.; Bax, A.; Case, D. A.; Summers, M. F. J. Biomol. NMR 2010, 47 (3) 205– 219There is no corresponding record for this reference.
- 93Tjandra, N.; Bax, A. Science 1997, 278 (5344) 1697– 1697There is no corresponding record for this reference.
Supporting Information
Supporting Information
Torsion profiles of the studied ribo- and deoxyribonucleosides in vacuo and in COSMO and PB solvent models, χOL-DFT torsion parameters, and tables of structural parameters for 1RNA and 2R20′ structures. This material is available free of charge via the Internet at http://pubs.acs.org.
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.