Skip to main content
Nucleic Acids Research logoLink to Nucleic Acids Research
. 2003 Dec 1;31(23):6986–6995. doi: 10.1093/nar/gkg893

L-nucleotides and 8-methylguanine of d(C1m8G2C3G4C5LG6LC7G8C9G10)2 act cooperatively to promote a left-handed helix under physiological salt conditions

Ilham Cherrak, Olivier Mauffret, Fanny Santamaria 1, Alexandre Hocquet 2, Mahmoud Ghomi 2, Bernard Rayner 3, Serge Fermandjian *
PMCID: PMC290261  PMID: 14627831

Abstract

The structure and thermal stability of a hetero chiral decaoligodeoxyribonucleotide duplex d(C1m8 G2C3G4C5LG6LC7G8C9G10)d(C11m8G12C13G14C15LG16LC17G18C19G20) (O1) with two contiguous pairs of enantiomeric 2′-deoxy-l-ribonucleotides (C5LG6L/C15LG16L) at its centre and an 8-methylguanine at position 2/12 was analysed by circular dichroism, NMR and molecular modelling. O1 resolves in a left-handed helical structure already at low salt concentration (0.1 M NaCl). The central l2-sugar portion assumes a B* left-handed conformation (mirror-image of right-handed B-DNA) while its flanking d4-sugar portions adopt the known Z left-handed conformation. The resulting Z4–B2*–Z4 structure (left-handed helix) is the reverse of that of B4–Z2*–B4 (right-handed helix) displayed by the nearly related decaoligodeoxyribonucleotide d(mC1G2mC3G4C5L G6LmC7G8mC9G10)2, at the same low salt concentration (0.1 M NaCl). In the same experimental conditions, d(C1m8G2C3G4C5G6C7G8C9G10)2 (O2), the stereoregular version of O1, resolves into a right-handed B-DNA helix. Thus, both the 8-methylguanine and the enantiomeric step CLpGL at the centre of the molecule are needed to induce left-handed helicity. Remarkably, in the various heterochiral decaoligodeoxyribonucleotides so far analysed by us, when the central CLpGL adopts the B* (respectively Z*) conformation, then the adjacent steps automatically resolves in the Z (respectively B) conformation. This allows a good optimisation of the base–base stackings and base–sugar van der Waals interactions at the ZB*/B*Z (respectively BZ*/Z*B) junctions so that the Z4–B2*–Z4 (respectively B4–Z2*–B4) helix displays a Tm (∼65°C) that is only 5°C lower than the one of its homochiral counterpart. Here we anticipate that a large variety of DNA helices can be generated at low salt concentration by manipulating internal factors such as sugar configuration, duplex length, nucleotide composition and base methylation. These helices can constitute powerful tools for structural and biological investigations, especially as they can be used in physiological conditions.

INTRODUCTION

The dl reversal of deoxyribose chirality (Fig. 1) represents an interesting modification of DNAs that can be useful for the improvement of our understanding of the structure and dynamics of DNA molecules. It can also help in the design of molecules more resistant to nucleases in the context of antisense and aptamer strategies. Previous studies have shown that homochiral l-oligonucleotides could assume reverted helicity (left-handed helices) relative to natural DNAs (1). However, the properties of heterochiral l-d oligonucleotides still remain largely unknown, although enantiomeric deoxy-l-nucleotides have been used to gain insight into the relationship between the local DNA conformation and the helical sense of a double helix (2). For instance, double-stranded oligonucleotides bearing a single l-residue are still able to form stable Watson–Crick base pairs (1,3). This is also true for the DNA duplex d(mC1G2mC3G4C5LG6LmC7G8mC9G10)2 with two contiguous l-residues in the middle of each strand (4). At neutral pH and in mild salt conditions, similar to physiological conditions, this molecule adopts an all-right-handed helix resulting from d-residues assuming the B-DNA form and l-residues the Z*-DNA form (mirror-image of Z-DNA). The practical conclusion of this study was that l-residues can cooperate with d-residues to form a heterochiral right-handed helix.

Figure 1.

Figure 1

Structures of the d- and l-deoxyribose sugars.

The stability of heterochiral right-handed helices is intimately related to an efficient base–base stacking and sugar–base van der Waals interactions. The absence of disruption in the chain of interactions contributes to maintain the helix stability at the BZ*/Z*B junctions (4). l-Enantioresidues in the Z* conformation allow such interactions to continue within the same groove of the DNA helix. Actually, the melting temperature (Tm) of the heterochiral oligonucleotide has been found to be only 5°C lower than that of its natural counterpart (4).

To learn more about the range of structures which can be assumed by heterochiral oligonucleotides, we now focuse on the d(C18mG2C3G4C5LG6LC7G8C9G10)2 (O1) molecule. Relative to the previously studied d(mC1G2mC3G4C5L G6LmC7G8mC9G10)2 oligonucleotide (4), the new oligonucleotide, O1: (i) still presents contiguous enantiomeric CL and GL at its centre; (ii) has no methylation of its cytosines; and (iii) guanine at position 2/12 now bears a C8-methylgroup that has been shown to markedly stabilise the Z-conformation of short oligonucleotides, even at low salt concentrations (5).

The properties of O1 were analysed by circular dichroism (CD), UV melting and NMR experiments, as well as molecular modelling. In the CD and UV studies, the stereoregular d(C1m8G2C3G4C5G6C7G8C9G10)2 (O2) was used in parallel to O1, as a reference molecule. Under mild salt conditions (≤0.2 M NaCl), O1 resolves into a left-handed helix, Z4–B2*–Z4, while O2 keeps the right-handed helix (B10). Thus, the methylation of guanine at position 2/12 and the replacement of two contiguous d-residues by enantiomeric l-residues at the centre of the molecule are both needed to induce the Z-conformation in physiological salt conditions.

MATERIALS AND METHODS

Oligonucleotide synthesis

2′-Deoxy-l-nucleosides were prepared according to slightly modified procedures (6,7). Introduction of a methyl group at the C8 position of 2-N-isobutyryl-2′-deoxyguanosine was performed by a free radical methylation method as described by Sugiyama et al. (5). Base-protected deoxynucleosides were dimethoxytritylated and further converted to the corresponding β-cyanoethylphosphoramidites according to standard procedures. Oligonucleotides were synthesised by means of a Model 381A (Applied Biosystems Inc.) DNA synthesiser (10 µmol scale). After deprotection with 30% aqueous ammonia for 5 h at 55°C, the crude oligonucleotides were purified by anion-exchange chromatography on a DEAE–Sephadex A25 (Pharmacia) column using triethylammonium hydrogen carbonate buffer (pH 7.5, linear gradient from 0.5 to 1.5 M) as eluent. Fractions were analysed by C18-reverse phase or anion-exchange HPLC, and those having a purity >95% were combined, co-evaporated with water several times and converted to the sodium salt with 2× Dowex 50 W (Na+ form). The final purity was >98% as determined by HPLC analysis.

CD and UV melting experiments

CD spectra of oligonucleotide solutions (70 µM in 10 mM Tris–HCl pH 7.0) were recorded in a 1 mm path length cell on a Jasco J-810 spectrophotometer equipped with a refrigerated water bath. Spectra at different temperatures were recorded after a 10 min equilibration period.

UV melting experiments were carried out on a Uvikon 931 (Kontron) spectrophotometer. The temperature of 1 cm path length cells was controlled by a Huber PD 415 temperature programmer connected to a refrigerated ethylene glycol/water bath (Huber Ministat). Annealing was performed by heating samples (7.5 µM oligonucleotide in 2 M NaCl, 10 mM Tris–HCl pH 7.0) at 90°C for 2 h and gradually cooling them at room temperature. During experiments, the sample temperature was increased at a rate of 30°C/h and the absorbance was measured at 260 nm. The melting temperature was read at the maximum of the first derivative of the UV melting curves.

NMR spectroscopy

NMR experiments were performed on a Bruker Avance- 500 MHz spectrometer. Data were processed on a Silicon Graphics Workstation with the Felix 2000 (Accelrys) and XwinNMR (Bruker) software.

Samples were dissolved in 0.4 ml of potassium dihydrogen phosphate–disodium hydrogen phosphate buffer at ionic strength 0.1 pH 7.1, and contained 1 mM EDTA. Concentrations were nearly 1.5 mM in duplex.

For assignments of exchangeable protons, samples were dissolved in H2O buffer. Phase-sensitive NOESY experiments in TPPI mode using the WATERGATE pulse sequence (8) were recorded at the four different mixing times of 40, 80, 200 and 300 ms and acquired with 4096 complex data points in t2 and 800 points in t1, with a relaxation delay of 1.5 s, at both 20 and 30°C.

For assignments of non-exchangeable protons, samples were lyophilised several times in 2H2O buffer and finally dissolved in 99.996% 2H2O. NOESY experiments were recorded at different mixing times (40, 80, 120, 200 and 300 ms) with a relaxation delay of 2 s at 20 and 30°C. A total of 2048 complex points were acquired in t2 and 512 real points in t1.

Two-dimensional (2D) P-COSY and 2D TOCSY were also used for assignments of both exchangeable and non-exchangeable protons. 1H–31P experiments (9,10) were used for the phosphorus assignments, and PFG-PEP HSQC (11) for the carbon assignments. A table of the chemical shifts is given in the Supplementary Material available at NAR Online.

NMR restraints

Interproton distance restraints used in structure calculations were measured from NOESY experiments recorded at various mixing times in H2O and 2H2O. Adequate references for distance calibrations in H2O and 2H2O were taken into account (12,13). Distances were classified as strong (1.8–3 Å), medium (1.8–4 Å), weak (1.8–5 Å) and very weak (1.8–6 Å) according to cross-peak intensities. Rotamer domains for backbone torsion angles α, β, γ, δ, ε and ζ were determined using 1H- and 31P-NMR data, as previously described (1315).

Watson–Crick hydrogen bond restraints were added for the stem residues which were experimentally determined as canonically paired. The criteria were chemical shifts of imino protons and the network of nuclear Overhauser effects (NOEs) in the base pair. Hydrogen bond restraints are the distances from a standard Watson–Crick base pair between the two heavy atoms of the hydrogen bond with an uncertainty of ±0.2 Å. For the terminal base pairs solely, an additional restraint was added, corresponding to the distance between the proton and the hydrogen bond acceptor. During the torsion angle and Cartesian dynamics steps, planarity restraints on the base pairs were added with a weight of 25 kcal/mol/Å2 as described in Kuszewski et al. (16). These restraints permitted a significant propeller twisting of the base pairs to take place unhindered.

Molecular dynamics protocol

A molecular dynamics simulated annealing protocol driven by experimental restraints was used to solve the solution structure of the molecule. This protocol uses torsion angle molecular dynamics implementation in the CNS program (17). We started with DNA strands under extended conformations, using the CNS protocol generate_extended.inp, to avoid bias of a predetermined form (17,18).

The dynamics was carried out in three stages (18). In the first stage, we used dynamics in the torsion angle space: molecules were equilibrated at 20 000 K over 90 ps, and cooled slowly at 1000 K over 90 ps. Non-bonded interactions were described by a repulsive quadratic term and calculations were performed under NOE, hydrogen bond (50 kcal/mol/Å2) and dihedral (5 kcal/mol/°2) restraints. At the beginning of the cooling step, planarity restraints were added with a weight of 25 kcal/mol/Å2. The second stage involved Cartesian molecular dynamics consisting first of a cooling step from 1000 to 300 K in 15 ps, followed by a 15 ps equilibration period at 300 K. The non-bonded interactions were described by Lennard–Jones potentials, and electrostatic terms were turned on. The force constants on dihedral angle, NOE, hydrogen bond and planarity restraints were the same as previously. In a third step, dihedral angle force constants were slowly increased from 5 to 50 kcal/mol/°2, during a Cartesian dynamics stage where the system evolved from 1000 to 300 K in a 9 ps period; this is followed by an equilibration period of 9 ps at 300 K. A total of 3000 steps of powell minimisation was then performed. Calculations were made 50 times; for each run, a different set of initial velocities was used.

Charge calculation with ab initio method

Charges of methylated nucleosides were calculated by fitting of the electrostatic potential, through the Merz Kollman procedure (19). This involved the comparative DFT geometry optimisations of guanines, 8-methylguanine, 9-methylguanine, 8,9-dimethylguanine, guanosine and 8-methylguanosine, using the B3LYP functional set (2023) along with the 6–31G* basis set (24). Amino groups of nucleic bases were checked to be pyramidalised and not planar. The two nucleosides were optimised in South and North conformations, as in Hocquet et al. (25). All calculations were made with Gaussian 98 (26).

RESULTS AND DISCUSSION

CD analysis

CD spectra of heterochiral d(C18mG2C3G4C5LG6LC7G8C9G10) (O1) and of homochiral d(C18mG2C3G4C5G6C7G8C9G10) (O2) were recorded in buffers of varying NaCl concentrations (Fig. 2). The spectral features of O1 (Fig. 2A) are specific for Z-DNA molecules (1,27) whatever the salt concentration. Even in the absence of NaCl, the O1 spectrum displays two relatively weak bands of similar magnitudes at ∼260 nm (positive) and ∼295 nm (negative). The bands demonstrate a gradual symmetrical increase as a function of NaCl concentration reflecting an increase of Z-DNA. The spectrum of the stereoregular O2 at low salt concentrations (0–0.2 M NaCl) (Fig. 2B) contains a large negative band at ∼255 nm accompanied by a smaller positive band at ∼280 nm. These are features of the right-handed B-DNA. At high salt concentration (2 M NaCl), the O2 spectrum is typical of the left-handed Z-DNA.

Figure 2.

Figure 2

Effects of NaCl on the CD spectra of (A) d(C1m8G2C3G4C5LG6LC7G8C9G10) (O1) and (B) d(C1m8G2C3G4C5G6C7G8 C9G10)2 (O2), at a 70 µM concentration in 10 mM Tris–HCl buffer pH 7, 12°C.

Sugiyama et al. (5) have concluded that m8G is a good inducer of Z-DNA, after having determined a mid-point at 0.03 M NaCl for the B→Z transition of a m8G-containing hexa-nucleotide d(C1G2C3m8G4C5G6)2. With the homochiral O2 decaoligonucleotide, that incorporates 8mG at the 2/12 positions, the mid-transition is observed at much higher NaCl concentration (largely >0.2 M), indicating that the length of the oligonucleotide and the position of the methylation in the sequence could influence the transition. Actually, the comparison of O1 and O2 proves that in long oligonucleotides, l-nucleotides are needed in addition to m8G to produce the B→Z transition at low NaCl concentration.

Temperature effects

Temperature melting effects on O1 and O2 were studied by combining CD and UV spectroscopies. We have seen above (Fig. 2) that both oligonucleotides present CD spectra typical of left-handed DNAs in 2 M NaCl, 12°C. In Figure 3A and B, the DNA thermal denaturation is reflected by a gradual reduction of the CD signal at ∼295 nm. The UV melting curves recorded at ∼260 nm for oligonucleotide O1 and O2 at concentrations nearly 10 times lower than those used in CD experiments provide Tms of 64.7 and 69.7°C, respectively (Fig. 3C). Actually, the same 5°C difference has already been found between the Tm of the heterochiral molecule d(mC1G2mC3G4C5LG6LmC7G8mC9G10)2 and that of its homochiral counterpart d(mC1G2mC3G4C5G6mC7G8mC9G10)2 (4), indicating that the heterochirality does not disturb the helix stability too much.

Figure 3.

Figure 3

CD spectra of (A) O1 and (B) O2 at 70 µM concentration in a 10 mM Tris–HCl buffer pH 7, 2 M NaCl recorded as a function of temperature. (C) UV melting curves at 260 nm of O1 and O2 at 7.5 µM concentration in the same buffer as in CD.

Thus, a global picture of the DNA double helix highly sensitive to punctual changes in the DNA chemistry and to experimental conditions can already be deduced from CD experiments. More information on the shape of the helix and on its local structures, including the ZB*/B*Z junctions, can be gleaned from NMR and molecular modelling studies.

NMR analysis

Particular NOEs and chemical shifts observed in 2H2O. We used NOESY, COSY and TOCSY spectra to assign the resonances of the non-exchangeable protons of O1 in 2H2O. These spectra are not shown. Identification of the m8G methyl group, the only methyl present in the oligonucleotide, permits in turn the assignments of the m8G deoxyribose-H1′ proton, and using through-bond experiments for the other sugar protons. Chemical shifts are listed in the Supplementary Material.

In the NOESY base–sugar H6/H8–H3′ fingerprint region, all the H6C–H3′G connectivities are strong, while those of H8G–H3′C, except H8G6–H3′C5, are very weak or absent. In the base–sugar H6/H8–H1′ fingerprint region, the GpC and the CpG steps display very distinct patterns: base–sugar H1′ NOEs are found only for GpCs and inter-residue base–sugar H2′/H2″ NOEs are systematically present in GpCs and absent in CpGs (except in C5LpG6L). The only interesidue NOE observable in the non-central CpGs concerns H8Gi–H4′C(i-1).

A noticeable feature in the NOESY spectra is the appearance of very strong intranucleotide base–sugar cross-peaks for the guanine residues G4, G8, G10 (H8–H1′) and m8G2 (Me8–H1′). Intensities are similar to those exhibited by H5–H6 in the cytosine aromatic rings. Such cross-peaks are not observed for the guanine with inverted chirality, G6L. This residue is the only guanine to adopt the glycosyl trans conformation in O1, all the other guanines assuming the glycosyl syn conformation (12,13).

Another remarkable feature concerns the unusually high-field shifts displayed by the H2′ and H5′ resonances (∼1.7 and ∼2.6 p.p.m., respectively) of all the cytosines (except C5L). The high-field shift probably results from the particular van der Waals interactions taking place between the sugars of cytosines and the bases of the succeeding guanines in O1 (5).

Particular NOEs and base pairings observed in H2O. The fingerprint region of imino protons connected to amino, aromatic and sugar H1′ protons is presented in Figure 4. Chemical shifts are typical for imino protons involved in Watson–Crick base pairs (28) at 13.29 (GL6), 13.15 (G8), 13.09 (m8G2) and 12.82 p.p.m. (G4). At the same time, imino–imino interbase NOEs are detected for GpC steps, but not for the CpG steps.

Figure 4.

Figure 4

Expanded NOESY contour plot at 500 MHz of O1 at 20°C in H2O solution, showing the connectivities between the imino and the amino, the aromatic and the sugar 1′ proton resonances. Vertical lines provide the chemical shifts and the cross-peaks of the four imino protons with the other protons, the names of which are generally noted.

Together with the NOEs typical of Watson–Crick base pairs, we observe the three following inter-residue NOEs: H1(G8)–H1′C7, H1G2–H1′C1 and H1G4–H1′C3. Their presence features the existence of large deviations from standard B-DNA at the corresponding steps.

Cross-peaks arising between the C19/C9 geminal amino protons at 8.24 and 6.88 p.p.m. and the m8G2/m8G12 imino proton at 13.09 p.p.m. are also detected. Amino protons of m8G give rise to two separate resonances, probably because the rotation of the amino group around the C2–N2 bond is considerably slowed down by the adjacent bulky methyl on the ring.

All together, the CpG and GpC steps are involved in very distinct networks of connectivities, and the enantio-CLpGL step displays spectral features which clearly deviate from those of the other CpG steps.

A sketch of the main NOE connectivities in oligonucleotide O1 is presented in Figure 5. These illustrate the differences between the CpG and the GpC steps and the particular behaviour of the C5LpG6L/G16LpC15L step.

Figure 5.

Figure 5

Main NOE connectivities in the helix O1. Arrows indicate NOEs between the non-exchangeable protons (intra-strand, black); imino and sugar/base protons (intra-strand, red); imino–imino and imino–amino protons (inter-strand, blue); amino–amino protons (inter-strand, green); and base–base protons (inter-strand, pink). See also the text.

Phosphodiester and sugar conformations in the backbone. Analysis of P-COSY and NOESY spectra provides information about sugar conformation (12,28). In the COSY spectra, we found that 3J12′ > 3J12″ (implying a C2′-endo sugar conformation) for cytosine sugars (except C5) and the reverse for guanine sugars G2, G4, G8 and G10 (implying an unusual C3′-endo conformation). This observation is confirmed by the presence of large J3′4′ and J2″3′ cross-peaks for the G2, G4, G8 and G10 residues, while these cross-peaks are absent or weak for pyrimidines (except C5). The NOESY data are in full agreements with these data. Accordingly, G2, G4, G8 and G10 were restrained to the C3′-endo conformation and the G6, C1, C3, C7 and C9 residues to the C2′-endo conformation. The C5 residue presents signs of conformational averaging and thus was not restrained.

According to the 1H–31P correlation spectra, the differences detected on the base structures extend to the backbone phosphates. A fingerprint region (31P–H3′, H4′, H5′5″) is presented in Figure 6. The chemical shifts of phosphorus atoms are listed in the Supplementary Material. Among the nine phosphorus resonances, those six arising from the five CpG steps and the G4pC5L step are found in a rather classical range (between –3.5 and –4.5 p.p.m.). The three remaining phosphorus resonances (G2pC3, G8pC9 and G6LpC7) are at low field (at about –2.5 p.p.m.), which is usually explained by an α or a ζ angle assuming the trans conformation (12,13). However, the observation of H4′–31P cross-peaks (besides the common H3′–31P cross-peak) for the GpC steps (except that of G4pC5L) and the C5LpG6L step indicates that the β angles and γ angles of these steps occupy the trans domain and the gauche + domain, respectively (12,29).

Figure 6.

Figure 6

A HETCOR 1H–31P spectrum of O1 at 500 MHz in deuterated phosphate buffer pH 7 at 20°C. H3′–P correlations are found in the left part of the spectrum and those of H4′–P in the right part.

Molecular modelling

Structure calculations. Starting with the two oligonucleotide strands in extended conformation, we performed 50 simulations applying different initial velocities. The 10 best structures selected on energetic grounds are presented in Table 1. Seventy-two percent (36 out of 50) of the structures have an overall energy better than –600 kcal/mol and do not present NOE violations >0.2 Å. The r.m.s.d. for the 10 base pairs in the 10 best structures was 0.83 Å (relative to the average structure, Table 1), which represents a rather good structural definition given the conservative restraints used in the calculations.

Table 1. Statistics of the NMR restraints and structures.

NMR restraints  
Intra-residue distances 166
Inter-residue distances 186
Hydrogen bond restraints 36
Dihedral angle restraints 84
Planarity restraints 10
Total restraints 482
Restraints/residue 24.1
Structure analysis of the 10 best conformers  
Average deviation from ideal covalent geometry  
Bond length 0.0018 ± 0.00006
Bond angles 0.4359 ± 0.0073
Improper angles 0.263 ± 0.01136
NOE violations  
 Number 0
 r.m.s.d of the violation (Å) 0.0193 ± 0.00106
Dihedral angle violations  
 Number 0.5
 r.m.s.d of the violations (°) 0.02 ± 0.0269
r.m.s.d of the structuresa 0.83 ± 0.06

ar.m.s.d relative to the average structure.

The minimised averaged helix taking into account the 10 best O1 structures combines a Z-DNA portion (base pairs 1–20 to 4–17), a B*-DNA portion (base pairs 5–16 and 6–15) and again a Z-DNA portion (base pairs 7–14 to 10–11). We obtain, therefore, a left-handed helix with a Z4–B2*–Z4 alternation (Fig. 7). In the two Z portions, the sugars alternate the directions of their O4′ atoms, thus generating a zigzag characteristic of the Z-DNA (27). In contrast, the sugars of G4/C17 and of the enantiomeric C5L/G16L and G6L/C15L base pairs belonging to the B* portion, have their sugar O4′ atoms pointing in a unique and same direction, a feature belonging to B-DNA (here, B*-DNA).

Figure 7.

Figure 7

Stereo view of the mean minimised structure of O1 obtained by molecular dynamics using the NMR data. Sugars of strand 1 are presented in a stick mode with the O4′ atoms indicated in red.

In the Z4–B2*–Z4 helix, the Z and B* portions can also be distinguished from the differences of base–base stackings and base–sugar interactions in CpG steps. The Z-DNA allows the interstrand stacking of the cytosine bases and the intrastrand van der Waals interaction of cytosine bases and guanine sugars. For the central enantio-CLpGL step, a B-type (B*-type) stacking of bases is obtained.

The torsion angle values of the best structures are given in Table 2. The guanines G2, G4, G8 and G10 adopt the C3′-endo syn conformation, while the other residues are in the C2′-endo-anti range.

Table 2. Mean phosphodiester angles for the 10 best conformers of O1 obtained with NMR data.

Mean γ (°) δ (°) ε (°) ζ (°) α (°) β (°) χ (°)
C1 –177 ± 58 148 ± 1 –141 ± 17 123 ± 22 174 ± 99 –166 ± 53 –144 ± 10
G2 176 ± 88 91 ± 1 –164 ± 16 –80 ± 21 –150 ± 10 –121 ± 5 93 ± 4
C3 48 ± 5 145 ± 1 –145 ± 8 108 ± 12 162 ± 108 –165 ± 60 –145 ± 3
G4 165 ± 81 97 ± 3 –151 ± 21 –142 ± 66 133 ± 73 –179 ± 83 82 ± 17
C5 –176 ± 32 –136 ± 13 137 ± 20 103 ± 13 164 ±143 162 ± 25 131 ± 7
G6 –96 ± 111 –149 ± 3 –158 ± 27 122 ± 50 152 ± 47 –179 ± 16 103 ± 8
C7 78 ± 30 141 ± 4 –167 ± 13 139 ± 27 157 ± 93 –141 ± 49 –159 ± 2
G8 143 ± 37 90 ± 3 –151 ± 12 –97 ± 66 –157 ± 15 –128 ± 12 70 ± 7
C9 42 ± 4 146 ± 1 –154 ± 14 136 ± 31 147 ±160 –170 ± 74 –146 ± 4
G10 –124 ± 59 91 ± 1         76 ± 25

On each line, the α and β angles are those implicating the phosphorus of the following residue (i.e. for C1, α and β angles are related to the C1pG2 step).

Description of the left-handed Z4-B2*-Z4 helix and its comparison with the right-handed B4–Z2*–B4 helix. The previously reported chimeric molecule d(mC1G2mC3G4C5L G6LmC7G8mC9G10)2 has been found to resolve in a right-handed helix B4–Z2*–B4 at low salt concentration (4). The B4–Z2*–B4 helix is the reverse image of the left-handed helix Z4–B2*–Z4 assumed by O1 in the same experimental conditions. In these oligonucleotides, the conformation of the central enantio-CLpGL step is dictated by the conformation adopted by flanking DNA portions, the conformation of this latter being itself highly sensitive to subtle chemical modifications, such as guanine methylation (O1). In the B4–Z2*–B4 helix, the base–base stacking and base–sugar interactions taking place in the B4 portions propagate to the whole helix thanks to the inversion of the sugar polarity at C5L and the passage of the G6L glycosyl bond into a syn orientation in the central enantio-CLpGL portion. These two operations that generate the Z* helix are needed to compensate the effects introduced by the chirality inversion at the centre of the molecule and to keep the base orientation in the grooves. In O1, the methylated G2/G12 residue and the enantio-CLpGL step act together to induce the left-handed helix Z4–B2*–Z4, at low salt concentration (Fig. 8). In the Z-DNA portion, the base positions are inverted relative to the right-handed helix. This occurs in response to both the inversion of polarity of the cytosine sugars and the adoption of the syn form by the guanine residues, the latter transition being strongly favoured by methylation of the guanine at position 8 (1). These two operations are needed to preserve the base stacking in the convex major groove of Z-DNA; they are not necessary in the central B2* portion, as the change of the sugar chiralities results in an effect similar to that produced by the inversion of the sugar polarity at cytosine and the anti to syn transition of the glycosyl bond at guanine. Thus, at the Z–B* junction, the C5L residue can adjust its base stacking with the preceding 5′G4 residue without changing its sugar polarity, while the G6L residue keeps the glycosyl anti conformation in order to optimise its interactions with its succeeding C7 residue. The arrangement of the G4pC5L step at the Z–B* junction of the Z4–B2*–Z4 helix is presented in Figure 9, together with the structure of the GpC step in B-DNA and Z-DNA helices. The particular arrangement adopted by the G4pC5L step in O1 allows the continuation of the base–base stackings and of the van der Waals interactions along the whole left-handed helix. This continuation does not need the inversion of the sugar polarity and the glycosyl anti to syn conversion at the succeeding C5LpG6L/G16LpC15L step.

Figure 8.

Figure 8

Stereo views of the mean minimised structures of (A) O1 and (B) d[mC1G2mC3G4C5LG6LmC7G8mC9G10]2 (4). The backbone is drawn as a ribbon.

Figure 9.

Figure 9

Structure of the GpC step in three different DNA conformations: (A) Z-DNA; (B) B-DNA; (C) at the ZB* junction of the O1 structure. Bases usually in the major groove are coloured in red, as are the sugar O4′ atoms. The base edges in the major groove are oriented forward in (B) and are opposite in (A) and (C).

CONCLUSION

The most salient feature of these studies is that heterochiral oligonucleotides can easily stabilise into right-handed or left-handed helices under physiological pH and salt conditions. Contiguous enantiomeric l-residues at the centre of self-complementary oligonucleotides can resolve into Z*-DNA (right-handed helix) or B*-DNA (left-handed helix), the result depending on the conformation of their flanking d-residues. If these assume the B (respectively Z) conformation, then the central CLpGL portion will adopt the Z* (respectively B*) conformation.

Although results demonstrate that in our chimeric oligonucleotides, the conformation of the flanking portions dictates the conformation of the central enantio-CLpGL, and therefore the whole helical conformation, the problem of the helical sense of such molecules at low salt concentration is not yet solved. We have seen, for instance, that a simple guanine methylation is sufficient to induce a left-handed helix, thus highlighting the high efficiency of subtle chemical modifications in the conformational change of chimeric oligonucleotides. Yet, how the effects are transmitted from the chemically changed position to the distant l-residues requires a detailed study with numerous heterochiral duplexes combining l- and d-DNA domains of different lengths and bearing guanine methylations at various positions.

Finally, we believe that chimeric oligonucleotides can constitute good tools in the search for molecules targeting DNAs and RNAs and to decipher the local structures and binding mechanisms of DNAs. The left-handed strands formed at low salt concentrations can be used as aptamers to detect proteins that bind tightly to Z-regions in DNAs. More importantly, they can also be useful in the preparation of antibodies that specifically recognise left-handed helices.

SUPPLEMENTARY MATERIAL

Chemical shifts of the O1 compound are provided in a table in the Supplementary Material available at NAR Online.

[Supplementary Material]

Acknowledgments

ACKNOWLEDGEMENTS

This research was supported by grants from Agence National de Recherche sur le Sida (ANRS, France) and SIDACTION (France).

REFERENCES

  • 1.Urata H., Shinohara,K., Oruga,E., Ueda,Y. and Akagi,M. (1991) Mirror-image DNA. J. Am. Chem. Soc., 113, 8174–8175. [Google Scholar]
  • 2.Urata H., Miyagoshi H., Kumashiro T., Mori K., Shoji K. and Akagi M. (2001) Basis for the right-handed helical sense of double-stranded DNA: formation of the right-handed helix by l-oligonucleotides fixed in low-anti glycosyl conformation. J. Am. Chem. Soc., 123, 4845–4846. [DOI] [PubMed] [Google Scholar]
  • 3.Blommers M.J.J., Tondelli,L. and Garbesi,A. (1994) Effects of the introduction of l-nucleotides into DNA. Solution structure of the heterochiral duplex d(G-C-G-(L)T-G-C-G)·d(C-G-C-A-C-G-C) studied by NMR spectroscopy. Biochemistry, 33, 7886–7896. [DOI] [PubMed] [Google Scholar]
  • 4.Mauffret O., ElAmri,C., Santamaria,F., Tevanian,G., Rayner,B. and Fermandjian,S. (2000) A two B–Z junction containing DNA resolves into an all right-handed double-helix. Nucleic Acids. Res., 28, 4403–4409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Sugiyama H., Kawai,K., Matsunaga,A., Fujimoto,K., Saito,I., Robinson,H. and Wang,A.H.-J. (1996) Synthesis, structure and thermodynamic properties of 8-methylguanine-containing oligonucleotides: Z-DNA under physiological salt conditions. Nucleic Acids Res., 24, 1272–1278. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Garbesi A., Capobianco,M., Colonna,F., Tondelli,L., Arcamone,F., Monzini,G., Hilbers,C.W., Aleen,J.M.E. and Blommers,M.J.J. (1993) l-DNAs as potential antimessenger oligonucleotides: a reassessment. Nucleic Acids Res., 21, 4159–4165. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Urata H., Oruga,E., Shinohara,K., Ueda,Y. and Akagi,M. (1992) Synthesis and properties of mirror-image DNA. Nucleic Acids Res., 20, 3325–3332. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Piotto M., Saudek,V. and Sklenar,V. (1992) Gradient-tailored excitation for single-quantum NMR spectroscopy of aqueous solutions. J. Biomol. NMR., 6, 661–665. [DOI] [PubMed] [Google Scholar]
  • 9.Sklenar V., Miyashiro,H., Zon,G. and Bax,A. (1986) Assignment of the 31P and 1H resonances in oligonucleotides by two-dimensional NMR spectroscopy. FEBS Lett., 208, 94–98. [DOI] [PubMed] [Google Scholar]
  • 10.Gorenstein D.G (1994). Conformation and dynamics of DNA and protein–DNA complexes by 31P NMR. Chem. Rev., 94, 1315–1338. [Google Scholar]
  • 11.Cavanagh J., Fairbrother,W.J., Palmer,A.G. and Skelton,N.J. (1996) Protein NMR Spectroscopy: Principles and Practice. Academic Press, San Diego, CA. [Google Scholar]
  • 12.Wijmenga S.S. and Van Buuren,B.N.M (1998) The use of NMR methods for conformational studies of nucleic acids. Prog. Nucl. Magn. Reson. Spectrosc., 32, 287–387. [Google Scholar]
  • 13.Varani G., Aboul-Ela,F. and Allain,F.H-T. (1996) NMR investigation of RNA structure. Prog. Nucl. Magn. Reson. Spectrosc., 29, 51–127. [Google Scholar]
  • 14.Mauffret O., Amir-Aslani., Maroun.,R.G., Monnot,M., Lescot,E. and Fermandjian,S. (1998) Comparative structural analysis by [1H,31P]-NMR and restrained molecular dynamics of two DNA hairpins from a strong DNA topoisomerase II cleavage site. J. Mol. Biol., 283, 643–655. [DOI] [PubMed] [Google Scholar]
  • 15.Chou S.-H., Tseng,Y. and Chu,B. (2000) Natural abundance heteronuclear NMR studies of the T3 mini-loop hairpin in the terminal repeat of the adenoassociated virus 2. J. Biomol. NMR, 17, 1–16. [DOI] [PubMed] [Google Scholar]
  • 16.Kuszewski J., Schwieters,C. and Clore,G.M. (2001) Improving the accuracy of NMR structures of DNA by means of a database potential of mean force describing base–base positional interaction. J. Am. Chem. Soc., 123, 3903–3918. [DOI] [PubMed] [Google Scholar]
  • 17.Brünger A.T., Adams,P.D., Clore,G.M., Delano,W.L., Gros,P., Grosse-kunstleve,R.W., Jiang,J.S., Kuswenski,J., Nilges,M., Pannu,N.S., Read,R.J., Rice,L.M., Simonson,T. and Warren,G.L. (1998) Crystallography and NMR system: a new software suite for macromolecular structure determination. Acta Crystallogr. D, 54, 905–921. [DOI] [PubMed] [Google Scholar]
  • 18.Stein E., Rice,L. and Brünger,A. (1997) Torsion-angle molecular dynamics as a new efficient tool for NMR structure calculation. J. Magn. Reson., 124, 154–164. [DOI] [PubMed] [Google Scholar]
  • 19.Besler B.H., Merz,K.M.,Jr and Kollman,P.A. (1990) Atomic charges derived from semiempirical methods. J. Comp. Chem., 11, 431–439. [Google Scholar]
  • 20.Becke A.D. (1993) Density-functional thermochemistry III. The role of exact exchange. J. Chem. Phys., 98, 5648–5652. [Google Scholar]
  • 21.Miehlich B., Savin,A., Stoll,H. and Preus,H. (1989) Results abstained with the correlation energy density functionals of Becke and Lee, Yang and Parr. Chem. Phys. Lett., 157, 200. [Google Scholar]
  • 22.Lee C., Yang,W. and Parr,R.G. (1998) Local softness and chemical reactivity in the molecules CO, SCN and H2CO. THEOCHEM, 37, 305–321. [Google Scholar]
  • 23.Vosko S.H., Wilk,L. and Nusair,M. (1980) Influence of an improved local-spin density correlation-energy functional on the cohesive energy of alkali metals. Phys. Rev. B. (Condensed Matter), 22, 3812–3843. [Google Scholar]
  • 24.Ditchfield R., Hehre,W.J. and Pople,J. (1971) A self-consistent molecular-orbital method. IX. An extended Gaussian-type basis for molecular-orbital studies of organic molecules. J. Chem. Phys., 54, 724–728. [Google Scholar]
  • 25.Hocquet N., Leulliot,M. and Ghomi,M. (2000) Ground state properties of the nucleic acid constituents studied by density functional calculations III. Role of the sugar puckering and base orientation on the energetics and geometry of 2′-deoxyribonucleosides and ribonucleosides. J. Phys. Chem. B, 104, 4560–4578. [Google Scholar]
  • 26.Frisch M.J., Trucks,G.W., Schiegel,H.B., Scueseria,G.E., Robb,M.A., Burant,J.C., Dapprich,S., Millam,J.M., Daniels,A.D., Kudin,K.N., Strain,M.C., Farkas,O., Tomasi,J., Barone,V., Cossi,M., Cammi,R., Mennucci,B., Pomelli,C., Adamo,C., Clifford,S., Octerski,J., Petersson,G.A., Ayala,P.Y., Cui,Q., Morokuma,K., Malick,D.K., Rabuck,A.D., Raghavachari,K., Foresman,J.B., Cioslowski,J., Ortiz,J.V., Baboul,A.G., Stefanov,B.B., Liu,G., Liashenko,A., Piskorz,P., Komaromi,I., Gomperts,R., Martin,R L., Fox,D.J., Keith,T., AL-Laham,M.A., Peng,C.Y., Nanayakkara,A., Gonzalez,C., Challacombe,M., Gill,P.M.W., Johnson,B.G., Chen,W., Wong,M.W., Andres,J.L., Head-Gordon,M., Replogle,E.S. and Pople,J.A. Gaussian 98 (Revision A. 3). Gaussian, Inc., Pittsburgh, 1998. [Google Scholar]
  • 27.Saenger W. (1984) Principles of Nucleic Acids Structure. Springer Verlag, New York, NY. [Google Scholar]
  • 28.Wijmenga S.S., Mooren,M.M.W. and Hilbers,C.W. (1993) NMR of nucleic acids: from spectrum to structure. In Roberts,G.C.K. (ed.), NMR of Macromolecules: A Practical Approach. Oxford University Press, Oxford, UK. pp. 217–288. [Google Scholar]
  • 29.Varani G. (1995). Exceptionally stable nucleic acid hairpins. Annu. Rev. Biophys. Biomol. Struct., 24, 379–404. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

[Supplementary Material]

Articles from Nucleic Acids Research are provided here courtesy of Oxford University Press

RESOURCES