Preferred Phosphodiester Conformations in Nucleic Acids. A Virtual Bond Torsion Potential to Estimate Lone-Pair Interactions in a Phosphodiester

Similar documents
Classical Potential Energy Calculations for ApA, CpC, GpG, and UpU. The Influence of the Bases on RNA Subunit Conformations

The A and B Conformations of DNA and RNA Subunits. Potential Energy Calculations for dgpdc

Influence of Ribose 2-0-Methylation on GpC Conformation by Classical Potential Energy Calculations

Effect of variations in peptide parameters and charges on the unperturbed dimensions of polypeptide chains

Structure of Guanosine-S,5 -cytidine Monophosphate. I. Semi-empirical Potential Energy Calculations and Model-Building

Yeast phenylalanine transfer RNA: atomic coordinates and torsion angles. G.J.Quigley, N.C.Seeman, A.H.-J.Wang, F.L.Suddath* and Alexander Rich

Department of Chemistry, Rutgers, The State University of New Jersey, New Brunswick, NJ 08903, USA

Figs. 1 and 2 show views of the ac and bc planes of both the. axes are shown, and inspection of the figure suggests that the

Crystal Structures of the Two Isomorphous A-DNA Decamers d(gtacgcgtac) and d(ggccgcggcc)

The Flexibility in the Proline Ring is Coupled

A CONFORMATIONAL STUDY OF PROLINE DERIVATIVES. M.E. Kamwaya * and F.N. Ngassapa

Hydrogen and hydration of DNA and RNA oligonucleotides

The flexibility in the proline ring couples to the protein backbone

Theoretical Studies on Peptidoglycans. 11. Conformations of the Disaccharide-Peptide Subunit and the Three-Dimensional Structure of Peptidoglycan*

4. Stereochemistry of Alkanes and Cycloalkanes

The Flexibility in the Proline Ring Couples to

Conformational Analysis of Macromolecules in Ground and Excited States

The typical end scenario for those who try to predict protein

POTENTIAL FUNCTIONS FOR HYDROGEN BOND INTERACTIONS

Ramachandran and his Map

Alkanes. Introduction

(A-Cm-U-Gm-A-A-Y-A--m C-U-Gp) excised from the anticodon

Junction-Explorer Help File

Conformational aspects and functions of trna

On the Mechanism of Ribonuclease A

STEREOCHEMISTRY OF ALKANES AND CYCLOALKANES CONFORMATIONAL ISOMERS

Zwitterionic character of nucleotides: possible significance in the evolution of nucleic acids

of yeast phenylalanine transfer RNA by Fourier transform NMR (conformational stability/tertiary hydrogen bonding/proton-deuteron replacement)

Introduction to Comparative Protein Modeling. Chapter 4 Part I

Nucleic acid constituent interactions

74 these states cannot be reliably obtained from experiments. In addition, the barriers between the local minima can also not be obtained reliably fro

Useful background reading

Figure 1. Molecules geometries of 5021 and Each neutral group in CHARMM topology was grouped in dash circle.

Correlation With Experimental Results

Organic Chemistry, Fifth Edition

FT-Raman Spectroscopy of Heavy Metal-Nucleotide Complexes

NMR of Nucleic Acids. K.V.R. Chary Workshop on NMR and it s applications in Biological Systems November 26, 2009

G. N. RAMACHANDRAN. and. (Department of Biophysics and Theoretical Biology, University of Chicago, Chicago, Illinois 60637, U.S.A.

The Conformation of Alkyl Sulfoxides

Stereochemistry Terminology

Stereochemical Considerations in Planning Synthesis

Chem 1140; Molecular Modeling

Molecular dynamics simulations of anti-aggregation effect of ibuprofen. Wenling E. Chang, Takako Takeda, E. Prabhu Raman, and Dmitri Klimov

Enzyme-Bound Conformations of Nucleotide Substrates. X-ray Structure and Absolute Configuration of 8,S-Cycloadenosine Monohydrate?

Interdependence of Conformational Variables in Double-Helical DNA


Structural and mechanistic insight into the substrate. binding from the conformational dynamics in apo. and substrate-bound DapE enzyme

Base pairing in DNA.

Chemistry 210 Organic Chemistry I Winter Semester 2002 Dr. Rainer Glaser

Supplementary Information

Catalysis by Entropic Guidance from Enzymes

The ideal fiber pattern exhibits 4-quadrant symmetry. In the ideal pattern the fiber axis is called the meridian, the perpendicular direction is

Organic Chemistry, Second Edition. Janice Gorzynski Smith University of Hawai i. Chapter 4 Alkanes

Chemistry 335 Supplemental Slides: Interlude 2

Chem 2100, Organic Chemistry I WS07, Dr. Rainer Glaser

Lecture 11: Potential Energy Functions

Nucleic Acids Research

Stereochemistry ofpolypeptide Chain Configurations

Only by constructing a model does one at first appreciate fully how. cyclohexane can exist in a non-planar, beautifully symmetrical, and apparently

Chapter 14: Conjugated Dienes

Other Cells. Hormones. Viruses. Toxins. Cell. Bacteria

Hydrogen bonding in oxalic acid and its complexes: A database study of neutron structures

Conformational Analysis

Abbreviations and Symbols for the Description of Conformations of Polynucleotide Chains

Advanced Cell Biology. Lecture 7

hand and delocalization on the other, can be instructively exemplified and extended

M. IQBAL and P. BALARAM, Molecular Biophysics Unit, Indian Institute of Science, Bangalore , India. Synopsis

Chemistry 2030 Introduction to Organic Chemistry Fall Semester 2013 Dr. Rainer Glaser

SUPPLEMENTARY INFORMATION

Editorial Manager(tm) for Journal of Molecular Biology Manuscript Draft. Title: The Flexibility in the Proline Ring is Coupled to the Protein Backbone

Nucleic Acid Reactivity: Challenges for Next-Generation Semiempirical Quantum Models

(1) Recall the different isomers mentioned in this tutorial.

Intrinsic DNA Curvature of Double-Crossover Tiles

Hydration of Nucleotides Thomas Wyttenbach, Dengfeng Liu, and Michael T. Bowers

Solutions to Assignment #4 Getting Started with HyperChem

DNA Structure. Voet & Voet: Chapter 29 Pages Slide 1

1. The barrier to rotation around the C-C bonds for 2-methylpropane and 2,2-dimethylpropane are shown below.

Supporting Text Z = 2Γ 2+ + Γ + Γ [1]

Secondary Structure. Bioch/BIMS 503 Lecture 2. Structure and Function of Proteins. Further Reading. Φ, Ψ angles alone determine protein structure

Supporting Information. Copyright Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, 2009

Crystal, Molecular, and Electronic Structure of 1-Acetylindoline and Derivatives

Thermodynamic parameters for the helix-coil transition of oligopeptides: Molecular dynamics simulation with the peptide growth method

The 63 rd Annual Pittsburgh Diffraction Conference

Why am I learning this, Dr. P?

Empirical calculations of interactions between two mesogenic molecules

Helical Structure and Circular Dichroism Spectra of DNA: A Theoretical Study

arxiv:cond-mat/ v1 2 Feb 94

Protein dynamics. Folding/unfolding dynamics. Fluctuations near the folded state

Alanine is helix-stabilizing in both template-nucleated and standard peptide helices

Molecular Simulation II. Classical Mechanical Treatment

First crystallographic evidence for the formation of -D-ribopyranosylamine from the reaction of ammonia with of

Algorithm for Rapid Reconstruction of Protein Backbone from Alpha Carbon Coordinates

Conformational Analysis of Cyclopropyl Methyl Ketone

Proteins polymer molecules, folded in complex structures. Konstantin Popov Department of Biochemistry and Biophysics

Proton conpling constants -

Derivation of 13 C chemical shift surfaces for the anomeric carbons of polysaccharides using ab initio methodology

SUPPLEMENTARY FIGURES

Conformational Isomers. Isomers that differ as a result of sigma bond rotation of C-C bond in alkanes

HOMOLOGY MODELING. The sequence alignment and template structure are then used to produce a structural model of the target.

Close agreement between the orientation dependence of hydrogen bonds observed in protein structures and quantum mechanical calculations

Transcription:

Preferred Phosphodiester Conformations in Nucleic Acids. A Virtual Bond Torsion Potential to Estimate Lone-Pair Interactions in a Phosphodiester A. R. SRINIVASAN and N. YATHINDRA, Department of Crystallography and Biophysics, University of Madras, Madras; and V. S. R. RAO and S. PRAKASH, Molecular Biophysics Unit, Indian Institute of Science, Bangalore, India Synopsis The lone-pair orbital interactions arising in a phosphodiester are incorporated into semiempirical conformational energy calculations using a unifold torsional potential around the virtual bond linking the ester oxygen atoms. The results explain the observed experimental data better than other methods. INTRODUCTION The preferred orientations around the internucleotide P-0(3 )(0 ) and P-O(5)(w) bonds of the phosphodiester which link the successive nucleotide residues are crucial in discerning the helical as well as nonhelical bend conformations of nucleic acids and polynucleotides. Experimental investigations other than the single-crystal x-ray diffraction methods are unable to provide directly information about P-0 bond conformations, except possibly for the stacked conformation, although the results of these studies have also conclusively substantiated the predominance of the preferred nucleotide conformations in di- and oligonucleotides. Calculations using semiempirical potential fun~tionsl-~ and semiempirical quantum mechanical method^^,^ on model sugar-phosphate-sugar backbone segments have aided in delineating the stereochemically possible and the preferred conformations around the P-0 bonds in di-, oligo-, and polynucleotides. However, some of these remained less satisfactory to the extent that a low-energy minimum occurred corresponding to the fully extended phosphodiester conformation (w,w) ( 180,1800) (tt), which was argued6 to be a considerably high energy conformation owing to the repulsive interactions between the lone-pair orbitals associated with the backbone ester oxygen O(5 ) and O(3 ) atoms. Also, no such conformation has been experimentally observed. Recent quantum chemical calculation~~--~ have shown that the fully extended phosphodiester conformation (tt) is indeed energetically unfavorable, in agreement with the earlier Biopolymers, Vol. 19, 165-171 (1980) 1c 1980 John Wiley & Sons, Inc. 0006-3525/80/0019-0165~01.OO

166 SRINIVASAN ET AL. chemical reasonings of Sundaralingam6 and with the estimated repulsive interaction energy due to eclipsed lone-pair interactions (which are of the order of 4-7 kcal/mol). It is important that this aspect be introduced into the semiempirical potential functions to enhance their predictive abilities. A straightforward way of rendering the (tt) conformation a high-energy one is to introduce a twofold torsion potential around both the ester P-0 bonds with a barrier height equivalent to half the repulsive lone-pair interaction energy.1 The use of such torsional potentials around the P-0 bonds renders the extended (tt) as a high-energy conformation due to the very nature of the potential function; however, this fails to convey the physical picture of the directional dependence of the lone-pair interaction. We propose here a potential function which avoids the use of an unrealistically high torsion potential around the P-0 bonds but directly estimates the lone-pair interaction energy in a phosphodiester bond in relation to directional properties of the bonds. The interactions between the lone-pair orbitals in a phosphodiester is regarded as arising from a torsion potential around the virtual bond linking the ester 05 and 03 oxygen atoms (Fig. l), akin to the torsion potential around the chemical bonds. The lone-pair interaction potential is expressed as VA,, = Va/2(l + cos B), where B is the dihedral angle about the virtual bond joining the two ester oxygen atoms (Fig. 1). The torsion angle B is taken as zero for the eclipsed orientations of the lone pairs. The directions 0(3 )-A and O(5 )-B in Fig. 1 represent the perpendicular bisectors of the two lone pairs. The characteristics of this potential are such that the energy is maximum for the phosphodiester conformation ( w,~) = (18Oo,18O0) at which the two lone-pair orbitals are eclipsed or in a cis conformation, i.e., 0 = 0, and the lone-pair interaction energy is minimum for phosphodiester conformations (w,w) = (Oo,18O0), (180,0 ),(6Oo,6O0), and (3OO0,30O0), when the lone-pair orbitals are in a trans conformation, i.e., B = 180. This is equivalent to a singlefold torsional potential around the virtual bond, and a simple correlation between B and w and w does not exist. The value of 2.8 kcal/mol A B? t Fig. 1. Schematic representation of lone-pair orbitals associated with O(3 ) and O(5 ) ester oxygen atoms in a phosphodiester. The dotted line joining the O(3 ) and O(5 ) represents the virtual bond and the directions O(3 )-A and 0(5 )-B are the directions of the lone-pair axes. 0 defines the torsion angle around the virtual bond O(3 )-O(5 ) and is defined as zero for the cis orientation of the O(39-A and O(5 )-B bonds.

PREFERRED PHOSPHODIESTER CONFORMATIONS 167 has been assigned for Vu from studied1 of anomeric effect in sugars. This anomeric torsional potential around the virtual bond affords a direct visualization of the correlation of the lone-pair orbital orientations with their energies while at the same time avoiding the necessity of using twofold torsion potentials around each of the P-0 bonds. Further, we find that a threefold torsion potential with the usual low barrier height of 1 kcal/mol is adequate to explain the experimental observations. The use of this anomeric potential in conjunction with the conventional semiempirical potential functions produces better agreement between theoretical predictions and experimental observations. Similar potential functions used for estimating the lone-pair interactions in polysaccharidesl have resulted in remarkable agreement between the theoretically computed and experimentally observed unperturbed end-to-end dimensions. METHOD The total potential energy in a ribodinucleoside monophosphate and ribodinucleoside triphosphate sugar-phosphate-sugar backbone is computed as a function of the internucleotide P-0(3 ) (w ) and P-0(5 ) (w) torsions, using the semiempirical potential functions and also the above anomeric torsional potential around the virtual bond (which joins the ester oxygen atoms to take into account the lone-pair interactions). The interactions due to the bases are excluded with the intention of probing the inherent-property characteristics of the polynucleotide backbone. The potential functions and the constants used here are the same as those adopted earlier.3 The (w,w) energy surfaces are shown in Figs. 2 and 3. Fig. 2. The (w,w) conformational energy map obtained for a dinucleoside monophosphate with the C(3 )-endo geometry for the nucleotides. The lone-pair orbital interaction energies are included by the use of virtual bond torsion potential. A threefold torsion potential around the P-0 bonds with a harrier height of 1 kcal/mol is also included.

168 SRINIVASAN ET AL. 300 - (3 240 - iao - 120-00 ILU Fig. 3. The (w,w) conformational energy map obtained for a dinucleoside triphosphate with C(3 )-endo geometry for the nucleotides. The lone-pair orbital interaction energies are included by the use of the virtual bond torsion potential. A threefold torsion potential around the P-0 bonds with a barrier height of 1 kcal/mol is also included. L-. RESULTS AND DISCUSSION Figure 2 shows the (o,w) energy map obtained for a C(3 )-endo ribonucleoside monophosphate backbone that also takes into account the lone-pair interaction energy through a virtual bond torsion potential of barrier height equal to 2.8 kcal/mol. The most important feature in the map is that contrary to earlier results obtained using semiempirical potential function~~-~ and some earlier molecular orbital computation^^.^ but in agreement with the refined calculations of this t ~pe,~-~ the fully extended tt conformation is now rendered a considerably high energy conformation. Similarly, the phosphodiester c.onformation g,-t = (w,o) = (300,180 ), g+t (w,w) = (6Oo,18O0) is about 2 kcal/mol higher in energy than the most stable g+g+ and g-g- phosphodiesters. The tg- = (18O0,3OO0) and the tg+ (18Oo,6O0) phosphodiesters are about 2.5 kcal/mol higher in energy than the global minimum found earlier. Except for one of the molecules of the rest of the crystal structures of ribodinucleoside monophosphates favor either the g-g- or the g+g+ phosphodiesters. w

PREFERRED PHOSPHODIESTER CONFORMATIONS 169 Similarly, the (w,w) conformational energy map (Fig. 3) obtained for a dinucleoside triphosphate backbone, which is a better representation of a polynucleotide backbone chain, also shows that the fully extended tt phosphodiester conformation is now rendered energetically highly unfavorable. This indicates that the explicit inclusion of lone-pair orbital interactions through a torsion potential around the virtual bond satisfactorily accounts for the high energy associated with the fully extended phosphodiester conformation, in conformity with experimental observations.6 The global minimum in Fig. 3 occurs at (w,w) E (300,300 ) and is now stable over the other stereochemically possible g-t, tg-, and tg+ conformations by about 2.5 kcal/mol. Values greater than 2.8 kcal/mol for the barrier height do not alter any of the above conclusions except that the energies associated with the tt, tg+, tg-, and g-t conformations become correspondingly higher than the stable g -g- conformation. We find that when only the twofold torsion potentials around the chemical P-0 bonds are used (without the threefold potential), two more minima become important in the (w,~) map (not shown) for ribodinu- cleoside monophosphate: one around (o,w) = (9Oo,27O0) having an energy equivalent to g+g+ and g-g- and another around (w,w) = (30Oo,12O0), which is about 2 kcal/mol higher than the global minimum. Only the minimum around (9Oo,27O0) becomes important for ribodinucleoside triphosphate. However, inclusion of a threefold torsion potential with an unrealistically high barrier height of 3 kcal/mol readily renders these minima high-energy conformations because of the nature of the (w,w) values at which the minima occur and because of the torsion potential. A low barrier height of the order of 1 kcal/mol is found to be insufficient to make them energetically less important. Comparison of the shape and the general features of the (w,w) map obtained here (Fig. 3) with the (w,w) map of Ref. 10 suggests that the use of twofold torsional potential around the chemical P-0 bonds considerably restricts their rotational freedom. A comparison of the above results with the phosphodiester conformations in the crystal structures of the two polymorphs of yeast trnaph would provide a measure of agreement between theory and experiment. While there are only minor differences in the conformational angles, probably due to alternative interpretation and degree of refinement, the overall conformations in these structures are very similar. Hence phosphodiestsr conformations found in the monoclinic form14j5 alone are plotted in Fig. 3. The dots represent phosphodiesters which are flanked on either side by the C(3 )-endo nucleotide geometry with the approximate gauche+ (60 ) conformation around the C(4 )-C(5 ) bond. The dots enclosed in the circles correspond to the phosphodiesters flanked at least on one side by a nucleotide having a trans (180 ) or agauche- (300 ) geometry around the C(4 )-C(5 ) bond or a C(Z )-endo pucker or both. It is noteworthy that none of the phosphodiesters occur in the tt region forbidden by lone-pair repulsive interactions, and a large number of points cluster around the most favored g-g- domain. The three dots which correspond

170 SRJNIVASAN ET AL. to phosphodiesters in the bend regions lie in the high-energy g-t domain. Thus all the points lie within the boundary prescribed by the 5 kcal/mol energy contour. Other points, some of which lie outside the energy contours, should be compared with the (o,w) maps obtained with appropriate geometry for the nucleotides. CONCLUSIONS A unifold torsional potential around the virtual bond linking the ester oxygen atoms in a phosphodiester bond is suggested to account for the lone-pair orbital interactions in semiempirical calculations. Consideration of this virtual bond torsion potential along with the conventional potential functions satisfactorily accounts for the high energy associated with the fully extended tt phosphodiester conformation in nucleic acids and polynucleotides and also the observed experimental data. The important feature of this approach is that it affords a clear visualization of the correlation between the energetics and the orientations of the lone-pair orbitals. The use of the simple virtual bond torsion potential to account for the lone-pair interactions in a phosphodiester bond should also be useful in the analysis of conformations of coenzymes like NAD, FAD, nucleoside diand triphosphates, and phosphoglycerides. An estimate of the lone-pair interactions arising as a result of conformational and configurational changes in nucleosides could also be obtained by the use of such virtual bond torsion potentials. Contribution No. 494 from the Department of Crystallography and Biophysics and Contribution No. 138 from the Molecular Biophysics Unit, Indian Institute of Science, Bangalore. References 1. Sasisekharan, A. V. & Lakshminarayanan, A. V. (1969) Biopolyrners 8,505-514. 2. Olson, W. K. & Flory, P. J. (1972) Biopolyrners 11,25-26. 3. Yathindra, N. & Sundaralingam, M. (1974) Proc. Natl. Acad. Sci. USA 71, 3325-3328. 4. Saran, A. & Govil, G. (1971) J. Theor. Biol. 33,407-418. 5. Pullman, B., Perahia, D. & Saran, A. (1972) Biochirn. Biophys. Acta 269,l-14. 6. Sundaralingam, M. (1969) Biopolyrners 7,821-869. 7. Newton, M. (1973) J. Am. Chern. Soc. 95,256-258. 8. Perahia, D., Pullman, B. & Saran, A. (1973) C. R. Acad. Sci. 277, 2257-2260. 9. Perahia, D., Pullman, B., Vasilescu, n., Cornillon, R. & Broch, H. (1977) Biochirn. Hiophys. Acta 478,244-258. 10. Govil, G. (1976) Biopolyrners 15,2303-2307. 11. Prakash, S., Srinivasan, A. R. & Rao, V. S. R. (1978) in Collected Abstracts, Proceedings of the International Symposium on Biomolecular Structure, Conformation, Function and Evolution, January 4-7, 1978, Madras, India.

PREFERRED PHOSPHODIESTER CONFORMATIONS 171 12. Rubin, J., Brennan, T. & Sundaralingam, M. (1972) Biochemistry 11,3112-3127. 13. Sussman, J. L., Seeman, N. C., Kim, S. H. & Berman, H. M. (1972) J. Mol. Biol. 66, 403-421. 14. Jack, A,, Ladner, J. E. & Klug, A. (1976) J. Mol. Biol. 108,619-650. 15. Stout, C. D., Mizuno, H., Rubin, T., Brennan, T., Rao, S. T. & Sundaralingam, M. (1976) Nucleic Acids Res. 3,1111-1123. 16. Quigley, G. J., Seeman, N. C., Wang, A. H. J., Suddath, F. L. & Rich, A. (1975) Nucleic Acids Res. 2,2329-2334. 17. Sussman, J. L. & Kim, S. H. (1976) Biochem. Biophys Res. Commun. 68,89-96. Received November 21,1978 Accepted July 20,1979