Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2018 Jun 19.
Published in final edited form as: FEBS Lett. 2017 Jun 19;591(12):1770–1784. doi: 10.1002/1873-3468.12681

Robust IR Based Detection of Stable and Fractionally Populated G-C+ and A-T Hoogsteen Base Pairs in Duplex DNA

Allison L Stelling 1, Yu Xu 2, Huiqing Zhou 1, Seung H Choi 1, Mary C Clay 1, Dawn K Merriman 2, Hashim M Al-Hashimi 1,2,*
PMCID: PMC5584567  NIHMSID: NIHMS877959  PMID: 28524232

Abstract

Non-canonical G-C+ and A-T Hoogsteen base pairs can form in duplex DNA and play roles in recognition, damage repair, and replication. Identifying Hoogsteen base pairs in DNA duplexes remains challenging due to difficulties in resolving syn versus anti purine bases with X-ray crystallography; and size limitations and line broadening can make them difficult to characterize by NMR spectroscopy. Here, we show how infrared (IR) spectroscopy can identify G-C+ and A-T Hoogsteen base pairs in duplex DNA across a range of different structural contexts. The utility of IR-based detection of Hoogsteen base pairs is demonstrated by characterizing the first example of adjacent A-T and G-C+ Hoogsteen base pairs in a DNA duplex where severe broadening complicates detection with NMR.

Introduction

NMR relaxation dispersion studies1--6 have shown that in canonical DNA duplexes, Watson-Crick A-T and G-C base pairs (bps) exist in dynamic equilibrium with their Hoogsteen7,8 counterparts, which have populations on the order of 0.1% to 1% (Fig. 1). A-T and G-C+ Hoogsteen bps can form by flipping the purine base in a Watson-Crick bp and creating a new set of hydrogen bonds that require closer proximity of the two bases and protonation of cytosine N3 in the case of G-C. (Fig. 1). While Hoogsteen bps are a minor species in naked DNA duplexes, they can become the dominant species in DNA-drug9--16 and DNA-protein17--21 complexes where they are thought to play roles in DNA recognition,22 in damaged DNA where they are thought to play roles in damage induction, accommodation and repair;23,24 and in the active site of Y-family polymerases such as human polymerase iota25,26 where they are proposed to play roles in damage-bypass during replication.

Figure 1. Chemical structures for Watson-Crick and Hoogsteen bps.

Figure 1

Hydrogen bond acceptors are shown in red, and donors in blue. Green arrows denote the C1′ to C1′ distances between the Watson-Crick and Hoogsteen bps.

The detection and characterization of Hoogsteen bps can pose unique challenges to existing structural characterization techniques. There are now several studies documenting difficulties in distinguishing Hoogsteen from Watson-Crick bps by X-ray crystallography particularly for low-to-medium resolution maps. For example, a crystal structure of the Y-family DNA polymerase iota revealing Hoogsteen pairing during replication25,26 was challenged27 on the grounds that the weak electron density for the active site A-T bp makes it difficult to resolve a Watson-Crick from a Hoogsteen geometry. There have been other studies documenting difficulties in resolving Hoogsteen versus Watson-Crick bps in X-ray structures of DNA-homeodomain17 and DNA-p7320 complexes. Because of these ambigu-ities,17 it is conceivable that Hoogsteen bps may have been mismodeled as Watson-Crick in existing X-ray structures of DNA complexes. The characterization of Hoogsteen bps by X-ray crystallography can also be complicated by potential effects arising due to crystal packing forces. For example, several AT-rich sequences form DNA duplexes composed of Hoogsteen bps by X-ray crystallography28--32 but predominantly form canonical duplexes composed of Watson-Crick bps when examined using solution NMR. While solution NMR can be used to unambiguously characterize Hoogsteen bps, size limitations hinder application to large protein-DNA complexes. There have also been documented difficulties4,6,33 in detecting Hoogsteen bps by NMR due to severe line broadening of resonances arising either because a fractional population of Hoogsteen (>20%) exchanges with the Watson-Crick form on the microsecond-to-millisecond timescale or due to enhanced solvent exchange. Such broadening contributions can lead to a blind spot for NMR-based detection, even in small systems. Both NMR and X-ray also require significant amounts of nucleic acid material, and data collection and analysis can be significantly time-consuming.

IR spectroscopy can in principle provide a convenient approach for identifying Hoogsteen bps. With vibrational spectroscopy's inherently picosecond timescale,34 IR is unburdened by molecular weight or chemical exchange. DNA has been extensively studied35--37 with both Raman and IR spectroscopy. Early work comparing the Raman or IR spectrum to the X-ray structures of crystalline DNA, aided by computational methods and isotopic labeling,38--50 allowed precise correlations between the geometry of short duplexes and their wavenumber positions;51--53 as well as yielding marker bands diagnostic54--60 for a number of DNA conformations and environments.

IR and Raman have previously been used to examine the Hoogsteen bps in 1-methyl thymine and 9-methyladenine co-crystals,61,62 in triple helices,63--66 G-quadraplexes,67--70 and an all-Hoogsteen parallel duplex DNA.71 While there have also been a few Raman and IR studies of Hoogsteen bps in DNA,72--74 these have primarily focused on G-C+ bps under non-physiological low pH conditions in which other features of the DNA structure could be affected due to protonation. More recently, IR studies have examined75 the electronic excited state of G-C+ Hoogsteen bps in D2O in a duplex containing nine alternating G-C bps GC9. Fewer studies have examined A-T Hoogsteen bps within DNA duplexes with most studies focusing on non-duplex environments such as co-crystals of 9-methyladenine and 1-methylthymine monomers,61, 62 triplexes63--65, 76--78 and parallel duplexes.71 This is in part due to challenges in obtaining inducible or trapped A-T Hoogsteen bps in DNA duplexes. Thus, IR marker bands for G-C+ Hoogsteen bps in duplex DNA in H2O under physiological pH conditions remain to be fully characterized, and IR marker bands for A-T Hoogsteen bps in anti-parallel DNA duplexes remain to be established. The robustness of these Hoogsteen specific IR marker bands under a variety of DNA duplex conditions has yet to be assessed and is key for establishing IR as a method for Hoogsteen detection.

Given the growing interest in Hoogsteen bps within the duplex environment, we developed an IR-based approach for the robust detection of G-C+ and A-T Hoogsteen bps under physiological conditions. We used NMR verified model duplexes containing various compositions of G-C+ and A-T Hoogsteen bps to identify and validate robust IR marker bands for G-C+ and A-T Hoogsteen bps. We show that these IR marker bands can be used to detect Hoogsteen bps in duplex DNA even in cases where they are NMR-invisible due to chemical exchange. The utility of IR-based detection of Hoogsteen bps is demonstrated by characterizing the first example of adjacent A-T and G-C+ Hoogsteen base pairs in a DNA duplex where severe line broadening complicates NMR-based detection under physiological pH values.

Methods

Sample preparation

All chemicals (NaCl, sodium phosphate, HCl, KOH, DCl, KOD and D2O) used in both ddH2O and D2O buffer preparation were purchased from Sigma-Aldrich. Unmodified DNA samples were purchased as single-stranded oligonucleotides from Integrated DNA Technologies (IDT) with standard desalting purification. The m1G and m1A modified DNA constructs were purchased from Yale-Keck Oligo Synthesis Resource (W.M. Keck Foundation, New Haven, CT).

Samples were prepared by annealing at 95 °C for 5 min followed by 40 min of cooling at ambient temperature in pure ddH2O, after which they were exchanged 3× into their respective buffers with Amicon Ultra-15 or Ultra-0.5 centrifugal filters with a 3 kDa MW cutoff (Millipore). The samples with N1-methyladenine were annealed at 65 °C for 5 min. Concentrations were determined via UV-vis. with a NanoDrop 2000 (ThermoFisher) using the supplier provided extinction coefficients at 260 nm. Purity was assessed with 20% denaturing PAGE (Fig. S1A).

For the D2O experiments, samples were exchanged 3× into the D2O buffers at pD 6.8 (15 mM sodium phosphate, 25 mM NaCl) or pD 5.4. The pD on all D2O buffers was determined by adding 0.4 to the pH value after incubating the meter in D2O. The D8 GC9 isotope construct was prepared using a protocol described for RNA.79,80 Approximately 1 μmole of GC9 was re-solubilized in 99.9% D2O to a concentration of 1 mM. Triethylamine (TEA) was then added to a concentration of 75 mM, incubated at 95 °C for 24 hours, and lyophilized overnight to remove D2O and TEA. Finally, the D8 DNA was exchanged 3X into either the pH 3.1 buffer (1 mM glycine, 1 mM NaCl, pH 3.1) or a low NaCl, pH 7 buffer (1 mM Tris, 1 mM NaCl, pH 7).

DNA:echinomycin complexes were formed as previously described13 by adding 3× molar excess of echinomycin in methanol to the DNA duplex and shaking at 4 °C for 3 hours. The sample was then placed under a dry air flow until it was completely dehydrated, after which it was re-suspended in ddH2O.

CD and UV-visible spectroscopy

CD spectra were measured on an Aviv 202 CD spectropolarimeter equipped with a Peltier temperature control unit and a recirculating, cooling water bath. The temperature was set at 15 °C and the sensitivity was 100 mdeg. Concentrations of the GC9 construct were 30 μM for each buffer condition (25 mM NaCl, 15 mM sodium phosphate, pH 6.8; and 1 mM glycine 1 mM NaCl, pH 3.1). The average of three measurements were taken and the buffer blank subtracted out with a custom Python script. UV-vis. measurements were performed on a ThermoFisher NanoDrop 2000 using a 300 μL quartz cuvette from Starna Cells Inc.

NMR spectroscopy

NMR experiments were performed on a Bruker Avance III 600MHz spectrometer or 700-MHz spectrometer equipped with 5 mm triple-resonance cryogenic probes. Assignments of GC9 and CA-DNA:echinomycin were determined using two dimensional (2D) 1H-13H Heteronuclear Single Quantum Coherence (HSQC), 2D SO-FAST 1H-15N Heteronuclear Multiple Quantum Coherence (HMQC) imino,81 and 2D 1H-1H Nuclear Overhauser Effect Spectroscopy (NOESY) experiments. GC9 experiments were performed at 10 °C with 1 to 2 mM GC9 at 10 °C, and CA-DNA:echinomycin experiments were performed at 25 °C at concentrations of 1 mM. Data were processed with NMRpipe82 and analyzed in SPARKY (http://www.cgl.ucsf.edu/home/sparky/).

IR spectroscopy

The IR measurements were performed with a PerkinElmer Frontier FTIR with a Pike Technologies MIRacle ATR (attenuated total reflectance) KSR5 coated diamond optic and a narrow band MCT detector. Measurements on buffer exchanged DNA were performed by taking the background scans with the flow-through buffer on the crystal. Typically, 128 scans and 4 cm−1 resolution using approximately 1 to 4 μL of 3 to 4 mM sample was used; and measurements were performed at ambient temperature. CO2 and water vapor was subtracted from the spectrum using the provided automatic background removal function. For the direct titration of GC9, 32 scans and 6 μL was employed for the initial measurement, after which 0.5 μL aliquots of 48 mM HCl in ddH2O was added to the droplet, aspirated with a pipette three times to ensure mixing, followed by acquisition. These were found to be comparable to results from GC9 exchanged into buffers of known pH (Fig. S1B).

IR data processing was performed with the PerkinElmer Spectrum software. Spectra were normalized to the 970 cm−1 band, which have been found in previous work to be invariant to multiple conformational changes and were insensitive to D2O exchange;37 and all IR intensities are reported in normalized ATR units. The interactive baselining function was used to add a multipoint baseline prior to normalization. Subtraction spectra were calculated using the difference function on the normalized and baselined spectra. For the high salt GC9 IR experiments, the construct was exchanged into 1 mM glycine and 300 mM NaCl, pH 6.9, at approximately 8 mM GC9. The Z-form IR spectra of GC9 were taken at pH 7.0 in buffers with 4 M NaCl and 15 mM sodium phosphate. The IR measurement of TA-DNA and CA-DNA echinomycin complexes was performed in a 15 mM sodium phosphate, 25 mM NaCl buffer at either pH 6.8 or 5.4. To reduce spectral congestion from the echinomycin IR bands, the spectrum of echinomycin suspended in water was first subtracted from the spectrum of the DNA:echinomycin complex to remove the strong amide bands present in the echinomycin spectrum and to isolate the IR spectrum of the DNA itself.

Results

Characterizing a pH-induced Watson-Crick to Hoogsteen transition in GC9 by UV-visible, CD and NMR spectroscopy

Prior Raman, UV-vis, and CD studies72,75 showed that under low salt concentrations (1 mM NaCl), GC9 undergoes a transition from a Watson-Crick duplex to a Hoogsteen duplex when the pH is lowered to below 4.5. This pH inducible Hoogsteen duplex provides an ideal opportunity to identify IR bands unique to G-C+ Hoogsteen bps. We first confirmed this transition using UV-visible and CD spectroscopy (Fig. 2). Consistent with prior CD83,84 and vibrational studies,72,75 reducing the pH to below 4.5 resulted in an increase in intensity at 280 nm in the UV-vis reflecting formation of cytosine cations (Fig. 2A, top) and the appearance of a weak positive signal at 290 nm and weak negative signal at 255 nm in the CD spectrum attributed to formation of G-C+ Hoogsteen bps (Fig. 2A, bottom).

Figure 2. UV, CD, and NMR spectra of GC9.

Figure 2

(A) UV-visible (top) and CD spectra (bottom) of GC9 at low (red) and neutral (blue) pH. (B) SOFAST 1H 1D NMR spectra of GC9 at pH 6.8 (blue), 3.9 (black) and 3.1 (red), (all in 1 mM glycine, 1 mM NaCl, 0.1 mM EDTA, and 10% D2O buffers). (C-E) Aromatic (C), aliased C1′-H1′ (D) and C5-H5 (E) HSQC spectra. (F) NOESY of GC9 at pH 3.1 (buffer conditions as in (B)). Labels for resonances belonging to Hoogsteen and Watson-Crick forms are shown in red and blue, respectively.

We used NMR spectroscopy to further verify formation of G-C+ Hoogsteen bps at low pH in GC9. We observed unique NMR signatures of G-C+ Hoogsteen bps85 at pH 3.1, including the imino resonances of C+ at approximately 15 p.p.m. and a resonance at approximately 11 p.p.m. reflecting the now solvent-exposed G N3 (Fig. 2B). We also observe two sets of resonances in the 2D C-H HSQC spectra in slow exchange on the NMR timescale; a major Watson-Crick species and minor (approximately 30% based on volume analysis) Hoogsteen species with characteristic1,3,4,6 approximately 2 to 3 p.p.m. downfield shifted G C1′, G C8, and C C6 and approximately 1 p.p.m. upfield shifted C C5 resonances (Fig. 2C-E). The Hoogsteen resonances also exhibit the expected Hoogsteen-specific NOE-based distance connectivities,1,3,4,6,24,86,87 including an enhanced NOE cross peak between G-H8 and G-H1′ due to formation of a syn base (Fig. 2F). While the highly repetitive sequence makes it difficult to assign NOEs to specific nucleotide positions along the DNA, we do observe sequential NOEs between the Hoogsteen species, and between Watson-Crick species, but not between Watson-Crick and Hoogsteen species, consistent with cooperative formation of all Hoogsteen or all Watson-Crick bp helices. However, we cannot rule out the existence of species with mixed Watson-Crick and Hoogsteen populations that are not readily detectable or assignable based on the NOESY data.

IR marker bands allow detection of stable and fractional G-C+ Hoogsteen base pairs in duplex DNA

We used the pH inducible GC9 duplex to identify unique IR bands that could potentially be used in the identification of G-C+ Hoogsteen bps in DNA duplexes in H2O. We observe significant changes in the IR spectra of GC9 when lowering the pH from 7.0 to 3.1 (Fig.s3 and S1B and C). The difference IR spectra of GC9 show the all the expected variations due to formation of Hoogsteen base pairs63--65,67--70,77,78 when lowering the pH from 7.0 to 3.1 (Fig. 3). These include increased intensities in Hoogsteen-specific IR bands due to formation of a syn G (increases at 1730 cm−1, between 1350 and 1365 cm−1; and between 1315 and 1325 cm−1), protonated C+ (increases at 1730, 1540, and 1282 cm−1), and a decrease at 1498 cm−1 due to Hoogsteen type H-bonding at G N7 (Fig. 3B and Table 1).

Figure 3. Hoogsteen IR marker bands in GC9.

Figure 3

(A) IR spectra for a direct pH titration of GC9, with the initial Watson-Crick GC9 signal (blue) converting to Hoogsteen (red) through intermediates (grey). Initial buffer conditions were pH 7.0, 1 mM NaCl, 1 mM glycine. (B) Low - neutral pH IR difference spectra of the pH titration spectra in (A), red line is approximately pH 3.1. Hoogsteen IR marker bands are in red, Watson-Crick are in blue, and backbone are in black. Bands whose assignments are preliminary are in italics. Inset: Structures of Watson-Crick (blue) and Hoogsteen G-C+ (red) bps with highlights on the atoms which are major contributors to the IR marker bands. All spectra are normalized to the 970 cm−1 band.

Table 1.

Watson-Crick and Hoogsteen IR G-C+ signals in H2O.57,72,110 Positions are in wavenumbers (cm−1). Normal mode composition for the Watson-Crick form is given in parenthesis listed with the highest contributors first.

Watson-Crick Hoogsteen Notes
1708 1730 C+ and GC6=O (C6=O, C5-C6, C5-C4, N1-H, N1-C6 for G35)
1530 1540 C+ (skeletal out of phase ring mode37)
1498 1498 Decrease in intensity upon increased H-bonding to N7 (N7=C8, C8-N9, C4=C5, C4-C9, N-C8H of G37)
1375 1350 syn G (C1′-N9, N9-C1′H, N7=C8 of G111)
1330 1320 syn G (C2′-H, C1′-C2′-H, C3′-C2′-H of G111)
1296 1282 C+ (C4-NH2 of C)

Control pH titrations on a mixed sequence with GC steps did not yield these spectral changes (Fig. S1D) indicating that the IR markers assigned to Hoogsteen bps do not arise from more generalized protonation of DNA duplexes. We confirmed assignment of the 1730, 1540, and 1282 cm−1 bands to C+ based on observation of isotope shifts when exchanging the sample into D2O and collecting IR data at two pDs (Fig. S2A and B). The 1498 cm−1 band is an important marker of Hoogsteen type H-bond at G N7.60,88--94 Based on studies of triplexes and quadruplexes37 this band is expected to appear between 1480 to 1490 cm−1. We confirmed that the 1498 cm−1 band in the DNA duplex has significant contributions from a mode involving the G N7 via deuterium labeling at the nonexchange-able G H8 sites (Fig. S2C and D); as previously described.38,56, 95

The above Hoogsteen IR signals in GC9 were assigned based primarily on prior Raman work.72,75 We located two other IR markers which may be tentatively assigned to Hoogsteen in plane C=O stretching and NH2 modes. The G NH2 and C C2=O each lose a H-bond in the Hoogsteen bp, and thus bands with contributions from these functional groups are expected to increase in wavenumber. The negative band at 1620 cm−1 and the positive band at 1640 cm−1 are consistent with loss of H-bonding to a G NH2,35 while the loss at 1655 cm−1 and gain between 1680 and 1670 cm−1 are consistent with a similar loss of an H-bond to the C C=O.35 Supporting this assignment is the large isotope shift these bands undergo in D2O, particularly for the 1620 cm−1 band (Fig. S2A). These spectral changes may also arise from alterations in the G N1-C6 and G N1-H ring modes35,37 when in Hoogsteen bps, and which also have IR bands between 1600 and 1700 cm−1.

Formation of Z-DNA under low pH conditions is not expected based on the low ionic strength conditions used and could be ruled out based on the NMR and CD studies above, and by lack of an increase in the IR spectrum at the IR marker bands for Z-DNA37 at 1690 cm−1 and 1215 cm−1 (Fig. S3). Exchanging GC9 into a 300 mM NaCl pH 3.1 buffer also resulted in the expected suppression72 of the Hoogsteen-specific IR signals (Fig. S3, bottom). Formation of non-duplex structures such as G-quadraplexes was additionally ruled out by native PAGE gel electrophoresis (Fig. S1A). Finally, base open states were ruled out since they are expected to blue shift the N7 H-bond marker (1498 cm−1) band due to decreased H-bonding and a red shift is observed consistent with increased H-bonding to a G N7 (1498 cm−1 in pH 3.1 GC9).

Interestingly, during the above studies of GC9, Hoogsteen bps could be detected using IR under conditions where they were undetectable by NMR due to extensive line broadening. In the NMR spectrum of GC9, no Hoogsteen iminos were detected at pH 3.9, with weak imino signals only appearing at pH 3.1 (Fig. 2B, red line). In contrast, the IR G-C+ Hoogsteen signature is clearly visible in the difference spectrum over a full range of pH values, including pH 3.9 (Fig. 3B, red line; and Fig. S1B). This highlights one of the advantages of using IR in the detection of fractional Hoogsteen bps in DNA duplexes.

IR detection of a single m1G-C+ Hoogsteen base pair in a Watson-Crick duplex under physiological pH

In studies of Hoogsteen bps in duplex DNA, one will often be interested in the detection of single Hoogsteen bps surrounded by Watson-Crick bps under physiological pH conditions. Such single Hoogsteen bps have been observed in DNA:drug9,12,33, 96--99 and DNA:protein complexes17--19, 21 and in DNA containing damaged nucleotides by both X-ray crystallography17,20, 25,26 and NMR2--4,6,33 spectroscopy. We therefore examined the feasibility of using the IR markers of G-C+ Hoogsteen bps obtained for GC9 at low pH to characterize single G-C+ Hoogsteen bps surrounded by Watson-Crick under physiological pH. This is also an important test of the sensitivity of the IR-based method and ability to resolve a few Hoogsteen bps from a background of Watson-Crick bps. For these studies, we trapped a single G-C+ Hoogsteen bp by incorporating N1-methylated guanine (m1G) in the well characterized1,4,6,100 d(GCCAAAAAATCG) duplex (A6-DNAm1G, Fig. 4A). Prior NMR studies showed3,6,24,86 that m1G-C+ forms a stable Hoogsteen bp surrounded by minimally perturbed Watson-Crick bps.

Figure 4. Hoogsteen IR marker bands in m1G-C+ base pairs.

Figure 4

(A) Structure of the m1G-C+ bp and the A6-DNA sequence; m1G site in red. (B) IR spectra of A6-DNAm1G (red) and A6-DNA IR spectra (black) at pH 5.4 (top) and 6.8 (bottom). (C) Comparison of the A6-DNAm1G - A6-DNA (red) and the GC9 low - neutral pH (grey) difference spectra. Shaded regions and red labels indicate the positions of Hoogsteen G-C+ marker bands present in both spectra; black labels indicate backbone IR bands. (D) IR difference spectra of A6-DNAm1G at pH 5.4 (light grey), pH 6.8 (black) and pH 9 (blue), all subtracted from A6-DNA at pH 6.8. Red labels indicate marker bands for Hoogsteen that are lost as the pH is increased. Buffers consisted of 25 mM NaCl and 15 mM sodium phosphate. All spectra are normalized to the 970 cm−1 band.

The IR spectrum of A6-DNAm1G and A6-DNA (Fig. 4B) were subtracted, and their difference spectra compared to the difference spectra from low versus neutral pH GC9 (Fig. 4C). The 1730 cm−1 signal and other assigned signature G-C+ Hoogsteen bands are clearly observed in the A6-DNAm1G difference spectra (Fig. 4C, Table 1). Exchange into pH 9.0, 6.8, and 5.4 buffers resulted in the expected increases/decreases under acidic or basic H2O or D2O conditions (Fig.s 4D and S4) with controls demonstrating these bands to be distinct from IR signals due to generalized duplex protonation (Fig. S4). These results show that the IR G-C+ Hoogsteen bp marker bands can be used to detect single G-C+ Hoogsteen bps surrounded by Watson-Crick bps under physiological pH.

IR marker bands of A-T Hoogsteen base pairs

Studies have shown1,3,5,28--32 that A-T Hoogsteen bps are energetically more favorable than G-C+ Hoogsteen bps and are more likely to form in duplex DNA. Fewer studies have examined IR and Raman bands of A-T Hoogsteen bps as compared to G-C+ in part due to the absence of model duplexes for inducing Hoogsteen such as GC9. While formation A-T Hoogsteen bps involves smaller changes in H-bonding and charge as compared to G-C+ Hoogsteen bps, a more subtle A-T Hoogsteen IR signature is still expected. Based on calculations and by analogy to the syn adenines in Z-DNA,39,42, 55,101 A-T Hoogsteen IR marker bands are expected to lie at approximately 1337 cm−1 (shifted from 1344 cm−1 for anti A), between 1315 and 1325 cm−1 (shifted from approximately 1335 cm−1 for anti A), between 1365 and 1355 cm−1 (shifted from approximately 1375 cm−1 for anti A), and between approximately 1437 and 1442 cm−1 for a syn A (shifted from 1426 cm−1 anti A bands). Other bands for Hoogsteen A-T bps are expected35,40 to lie between 1650 and 1708 cm−1 (H-bond alterations at A N7, T C4=O, and A N6-H2), and approximately 1309 cm−1 (A N9-C8 and N3=C2).

We used m1A to trap an A-T Hoogsteen bp in two previously studied1 DNA duplexes; d(GCCAAAAAATCG) (A6-DNA) and d(GCCAATTGATCG) (A2-DNA) (Fig. 5A). We also investigated IR marker bands for A-T Hoogsteen bps using a d(ACACCTACGTGT) (TA-DNA, Fig. 5A) duplex in complex with the antibiotic echinomycin12--16 that forms A-T Hoogsteen bps. Echinomycin binds preferentially to GC steps9 separated by spacer bps (Fig. 5A, blue). NMR,12,13 X-ray,14,15 and footprinting16 have shown that for spacers with a TA step, complex formation results in tandem A-T Hoogsteen bps.

Figure 5. Persistent IR Signals for A-T Hoogsteen base pairs in different contexts.

Figure 5

(A) Chemical structure of m1A (top), and A6-DNAm1A, A2-DNAm1A, and TA-DNA sequences. Hoogsteen base pairs are in red and echinomycin binding site in blue. (B) Overlay of A6-DNAm1A - A6-DNA (top), A2-DNAm1A - A2-DNA (middle), and TA-DNA:drug - drug - free TA-DNA (bottom) difference IR spectra. Bands due to Hoogsteen G-C+ present in both spectra are shown in red; shaded regions indicate the range of positions seen for these marker bands. Buffers consisted of 25 mM NaCl and 15 mM sodium phosphate. All spectra are normalized to the 970 cm−1 band.

Significant changes are apparent in the difference spectra when comparing A6-DNAm1A and A2-DNAm1A with their respective unmethylated duplexes (Fig. 5B and Table 2, and Fig. S5A and B) that are distinct from those seen for m1G (Fig. S5A). This was also the case when comparing IR spectra of the TA-DNA with and without the echinomycin (Fig. S5C and D). We observe consistent changes across the three systems including the expected decrease in intensity in the N7 band (1490 cm−1 in the TA-DNA:drug complex) due to Hoogsteen-type H-bonding to A N737 and the expected increases between 1325 and 1315 cm−1, approximately 1337 cm−1, between 1365 and 1350 cm−1, and an increase at approximately 1437 cm−1 due to formation of a syn A (Fig. 5B and Table 2). This agreement, particularly in these conformationally sensitive regions of the IR, across three separate Hoogsteen A-T bp contexts supports these bands as markers for A-T Hoogsteen formation.

Table 2.

Preliminary assignments for Hoogsteen IR A-T marker bands in H2O observed in A6-DNAm1A - A6-DNA, A2-DNAm1A - A2-DNA, and the TA-DNA:drug complex - drug - free TA-DNA difference IR spectra. Positions are in wavenumbers (cm−1). Normal mode composition for the Watson-Crick form is given in parenthesis listed with the highest contributors first.

Watson-Crick Hoogsteen Notes
1498 1498 Decrease in intensity upon increased H-bonding to N7 (N-C8H, N7=C8, C2=N3 of A37)
1455 1436 to 1444 syn A (N1-C6, C6-N12 of A35)
1375 1350 to 1360 syn A (C1′-N9, C6-N6, N9-C1′H of A42)
1344 1335 syn A (N9-C8-H, C2-N3-H, C2=N3 of A42)
1344 1320 syn A (N9-C8-H, C2-N3-H, C2=N3 of A42)
1300 to 1308 1300 to 1308 A-N9=C8, decrease in intensity (N9-C8, N3-C2, C8-H, C2-H of A35)

Interestingly, the A6-DNAm1A and A2-DNAm1A difference IR spectra do not overlay perfectly with each other, with changes from the A6-DNAm1A duplex significantly more intense in the double bond region than in A6-DNAm1A (Fig. S5B). This increase may reflect formation of a more stable Hoogsteen m1A-T bp in A6-DNAm1A than in A2-DNAm1A, as reported previously by NMR.1 The two m1A-T Hoogsteen containing duplexes and the complex do show differences between 1700 to 1550 cm−1 region which may reflect more globally altered base pairing in the complex as compared to the m1A duplexes (Fig. S5D).

Application of IR in detecting tandem G-C+/A-T Hoogsteen bps

Thus far, echinomycin33, 96--99 and the related molecule triostin A9 are the only two molecules that have been shown to induce Hoogsteen bps in duplex DNA. These compounds have been shown to induce Hoogsteen bps in a manner strongly dependent on DNA sequence. For example, when the spacer sequences between two CG binding sites are TA/AT, they form tandem Hoogsteen bps, however Watson-Crick bps are observed99 when the spacer sequences are AT/TA. NMR studies99 of DNA:echinomycin and DNA:triostin complexes have encountered difficulties in the detection of Hoogsteen bps due to severe line broadening. These challenges could potentially be overcome with the IR-based method which combined with the lower cost and ease of data collection and analysis might provide an efficient method to screen for compounds that induce Hoogsteen.

Hoogsteen bps often occur in tandem as adjacent TA/AT and CG/GC Hoogsteen bps in DNA-drug,9--16 and DNA-protein17--19, 21 complexes. Interestingly, adjacent A-T and G-C+ Hoogsteen bps have not been observed to date. As an application of the IR-based detection of Hoogsteen bps, we examined whether tandem A-T and G-C+ Hoogsteen bps could be formed by placing a CA spacer sequence between the two echinomycin binding sites in the d(ACGTGCGT) duplex (CA-DNA, Fig. 6A). Due to line broadening, NMR spectra of the CA-DNA:drug complex at pH 6.8 were inconclusive as to whether Hoogsteen bps were present (Fig. 6A). However, at pH 5.4 we observed NMR signatures characteristic of both G-C+ and A-T Hoogsteen bps (Fig. 6A). These included resonances from C+ iminos; as well as downfield shifted purine C8 (Fig. S6A) and C1′ (Fig. S6B) chemical shifts. To our knowledge, this represents the first observation of a G-C+ and A-T Hoogsteen bps adjacent to one another.

Figure 6. IR evidence for tandem G-C+/A-T echinomycin-induced Hoogsteen base pairs under physiological pH.

Figure 6

(A) CA-DNA sequence and SOFAST 1H 1D NMR spectra of the pH 6.8 (blue) and pH 5.4 (red) CA-DNA:drug complex in 15 mM sodium phosphate, 25 mM NaCl, 0.1 mM EDTA, and 10% D2O buffers. (B) The 1700 to 1400 cm−1 region for A6-DNAm1A - A6-DNA (top), A6-DNAm1G - A6-DNA (bottom), and the CA-DNA:drug - drug - free CA-DNA spectra at pH 6.8 (middle, red). (C-F) Enhanced views of the 1480 to 1460 cm−1 (C), 1370 to 1345 cm−1 (D), 1330 to 1300 cm−1 (E), and 1300 to 1270 cm−1 (F) IR difference spectra; labels as in (B). Bands due to the tandem formation of G-C+ and A-T Hoogsteen bps are in red; shaded regions indicate the range of positions seen for these marker bands. Buffers for the IR experiments consisted of 25 mM NaCl and 15 mM sodium phosphate. All spectra are normalized to the 970 cm−1 band.

We used IR to examine whether the tandem G-C+ and A-T Hoogsteen bps persist under physiological pH. IR difference spectra of the CA-DNA:drug complex was compared to the G-C+ Hoogsteen bp in A6-DNAm1G as well as A6-DNAm1A (Fig. 6B - F). Unlike NMR, the CA-DNA:drug IR difference spectrum shows changes consistent with both A-T and G-C+ Hoogsteen bps at neutral pH (Fig. 6B). Weak bands due to G-C+ appear at 1730 cm−1, 1540 cm−1, and 1280 cm−1 (C+), between 1436 and 1441 cm−1 (syn A), between 1360 and 1355 cm−1 (syn A and G), approximately 1335 cm−1(syn A), and between 1325 and 1315 cm−1 (syn A and G). The CA-DNA:drug complex shows both sets of purine syn bands, indicating the presence of both A-T and G-C+ Hoogsteen bps. As expected, the G-C+ Hoogsteen signals increased upon lowering the pH (Fig. S6C), consistent with the NMR data. These results again highlight how IR can be applied to identify Hoogsteen bps under conditions where detection via NMR is significantly hampered due to line broadening from solvent exchange.

Discussion

The results presented in this work indicate that IR marker bands that are unique to Hoogsteen bps can be used to robustly detect Hoogsteen bps in duplex DNA under a wide range of conditions and contexts; including low pH duplexes with all G-C+ Hoogsteen bps, single m1A-T and m1G-C+ Hoogsteen bps surrounded by Watson-Crick bps; and AT/TA and AT/GC Hoogsteen bps in DNA:drug complexes. Our study also shows that IR can be used to identify Hoogsteen bps that are only fractionally populated (>10%). Unlike NMR, IR is not burdened by chemical/solvent exchange line broadening. There are now several cases suspected to involve exchange with Hoogsteen in which the key resonances are broadened precluding NMR-based assessment of Watson-Crick versus Hoogsteen.4,6,33 The added convenience of IR and lower sample requirements makes it an ideal method for exploring Hoogsteen bps in a variety of contexts such as in different damaged bases, when bound to different small molecules, under different physiochemical conditions (crowding, pH, salt concentration, etc.).

Our IR studies of these different systems also provide new insights into the structure of Hoogsteen bps. It is clear from the IR difference spectra of both GC9 and m1G that backbone signals near 1220 and 1088 cm−1 are altered between the G-C+ Watson-Crick and Hoogsteen. The approximately 1088 cm−1 region in particular has contributions from complex modes involving C-O and P-O bonds,48 and is sensitive to backbone parameters such as sugar pucker and torsion angles. GC9 shows near-complete overlap with Z-DNA at 1060 cm−1, indicating similarities between Z and Hoogsteen backbone geometries. A6-DNAm1G shows a large intensity increase at 1075 cm−1, an intermediate value between 1088 cm−1 (B-form) and 1060 cm−1 (Z-form), which is likely48 due to ζ and α torsion angles away from away from 0° and towards 180°. This is consistent with a recent study.102 which found similar alterations to these angles in a structure-based survey of the PDB.

The P-O backbone band at approximately 1225 cm−1 is also altered in A6-DNAm1G. In A6-DNAm1G, this band exhibits a marked decrease compared to the unmethylated duplex, with a decrease at 1216 cm−1; which may indicate a transition either towards a BII-like form,103 or a decrease in A-from-like backbone character.37 This is in line with the observed103 decrease at 1050 cm−1, which is also a B-DNA backbone marker. This points to Hoogsteen G-C+ bps exerting an influence over the distribution of DNA conformational substrates103 and may have implications for protein recognition of DNA by indirect readout. As with A6-DNAm1G, there are large changes to the backbone bands at approximately 1225 cm−1 and 1060 cm−1 in both m1A constructs. This increase at 1060 cm−1 is observed in Z-DNA91 and m1A, and thus m1A containing DNA may contain some Z-DNA backbone character.

Previous studies have performed theoretical investigations of the Raman104--107 and IR spectra106,107 of G-C+ Hoogsteen bps.107 Early calculations104 typically found shifts in the opposite direction to those observed in experiment. This prompted a few studies105--107 that investigated the effects of anharmonicity on the vibrational density of states of Watson-Crick and Hoogsteen bps. These anharmonic corrections105,106 to the potential provided better predictions of the Raman experimental value for the Hoogsteen protonated cytosine at 1260 cm−1 and which we observe at 1280 cm−1 in the IR measurements. The anharmonic corrections were particularly important for the G-C+ Hoogsteen bp modes involving the extra proton. Additionally, these studies105,106 list several high frequency IR modes (between 1740 and 1770 cm−1, depending on the potential used) due to carbonyl stretching and N-H bending modes that are not observed in our experimental spectra. In addition, these theoretical studies did not predict the change in position of the cytosine ring mode which we observe at around 1540 cm−1 in Hoogsteen (1530 cm−1 in Watson-Crick). The data provided in this work may help guide future theoretical studies of Hoogsteen IR bands.

As IR is unaffected by molecular weight, it holds great promise in the identification of Hoogsteen bps within DNA-protein complexes. The spectral congestion in larger complexes will present challenges due to overlap with Hoogsteen marker IR bands; however these problems may be overcome using isotope editing strategies108,109 already being used for RNA and nucleic acid:protein complexes.

Conclusions

In conclusion, we have identified robust Hoogsteen IR bands that can be used in the identification of G-C+ in duplex DNA and made preliminary assignments of bands for A-T Hoogsteen bps. The IR spectra of pH titrations on mixed sequence content DNA (A6-DNA and A2-DNA) showed some changes in the IR, but were much smaller in magnitude than those seen for the Hoogsteen GC9. Significant changes were also observed for A-T Hoogsteen base pairs in both N1-methylated adenines and in DNA-drug complexes. We used these marker bands to identify the first example of tandem A-T and Hoogsteen bps in a DNA:drug complex. The persistence of these different IR marker bands across a range of different sequence contexts and environments is a promising sign for future studies examining larger DNA complexes for the presence of Hoogsteen bps, as well as analysis of structural perturbations that occur in large DNAs and RNAs containing modified bases.

Supplementary Material

Supp FigS1
Supp FigS2
Supp FigS3
Supp FigS4
Supp FigS5
Supp FigS6
Supp info

Acknowledgments

We thank members of the Al-Hashimi laboratory for assistance and critical input. We acknowledge technical support and resources from the Duke Magnetic Resonance Spectroscopy Center at Duke University. This work was supported by NIH grants (R01GM089846 to H.M.A).

References

  • 1.Nikolova EN, Kim E, Wise AA, O'Brien PJ, Andricioaei I, Al-Hashimi HM. Transient Hoogsteen base pairs in canonical duplex DNA. Nature. 2011 Feb;470:498–502. doi: 10.1038/nature09775. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Nikolova EN, Gottardo FL, Al-Hashimi HM. Probing transient Hoogsteen hydrogen bonds in canonical duplex DNA using NMR relaxation dispersion and single-atom substitution. Journal of the American Chemical Society. 2012 Feb;134:3667–3670. doi: 10.1021/ja2117816. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Nikolova EN, Goh GB, Brooks CL, Al-Hashimi HM. Characterizing the protonation state of cytosine in transient GC Hoogsteen base pairs in duplex DNA. Journal of the American Chemical Society. 2013;135(18):6766–6769. doi: 10.1021/ja400994e. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Nikolova EN, Stull F, Al-Hashimi HM. Guanine to inosine substitution leads to large increases in the population of a transient GC Hoogsteen base pair. Biochemistry. 2014;53(46):7145–7147. doi: 10.1021/bi5011909. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Alvey HS, Gottardo FL, Nikolova EN, Al-Hashimi HM. Widespread transient Hoogsteen base-pairs in canonical duplex DNA with variable energetics. Nature Communications. 2014;5:4786. doi: 10.1038/ncomms5786. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Zhou H, Kimsey IJ, Nikolova EN, Sathyamoorthy B, Grazioli G, McSally J, Bai T, Wunderlich CH, Kreutz C, Andricioaei I, Al-Hashimi HM. m1A and m1G disrupt A-RNA structure through the intrinsic instability of Hoogsteen base pairs. Nature Structural & Molecular Biology. 2016 Sep;23:803–810. doi: 10.1038/nsmb.3270. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Hoogsteen K. The crystal and molecular structure of a hydrogen-bonded complex between 1-methylthymine and 9-methyladenine. Acta Crystallographica. 1963 Sep;16:907–916. [Google Scholar]
  • 8.Hoogsteen K. The structure of crystals containing a hydrogen-bonded complex of 1-methylthymine and 9-methyladenine. Acta Crystallographica. 1959;12:822. [Google Scholar]
  • 9.Wang A, Ughetto G, Quigley G, Hakoshima T, van der Marel G, van Boom J, Rich A. The molecular structure of a DNA-triostin A complex. Science. 1984;225(4667):1115–1121. doi: 10.1126/science.6474168. [DOI] [PubMed] [Google Scholar]
  • 10.Quigley GJ, Ughetto G, van der Marel GA, van Boom JH, Wang AH, Rich A. Non-Watson-Crick G-C and A-T base pairs in a DNA-antibiotic complex. Science. 1986 Jun;232:1255–1258. doi: 10.1126/science.3704650. [DOI] [PubMed] [Google Scholar]
  • 11.Singh UC, Pattabiraman N, Langridge R, Kollman PA. Molecular mechanical studies of d(CGTACG)2: complex of triostin A with the middle A-T base pairs in either Hoogsteen or Watson-Crick pairing. Proceedings of the National Academy of Sciences. 1986;83(17):6402–6406. doi: 10.1073/pnas.83.17.6402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Gao X, Patel DJ. NMR studies of echinomycin bisintercalation complexes with d(A1-C2-G3-T4) and d(T1-C2-G3-A4) duplexes in aqueous solution: sequence-dependent formation of Hoogsteen A1-T4 and Watson-Crick T1-A4 base pairs flanking the bisintercalation site. Biochemistry. 1988;27(5):1744–1751. doi: 10.1021/bi00405a054. [DOI] [PubMed] [Google Scholar]
  • 13.Gilbert DE, van der Marel GA, van Boom JH, Feigon J. Unstable Hoogsteen base pairs adjacent to echinomycin binding sites within a DNA duplex. Proceedings of the National Academy of Sciences. 1989;86(9):3006–3010. doi: 10.1073/pnas.86.9.3006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Cuesta-Seijo JA, Sheldrick GM. Structures of complexes between echinomycin and duplex DNA. Acta Crystallogr D Biol Crystallogr. 2005;61(4):442–448. doi: 10.1107/S090744490500137X. [DOI] [PubMed] [Google Scholar]
  • 15.Cuesta-Seijo JA, Weiss MS, Sheldrick GM. Serendipitous SAD phasing of an echinomycin-(ACGTACGT)2 bisintercalation complex. Acta Crystallogr D Biol Crystallogr. 2006;62(4):417–424. doi: 10.1107/S0907444906003763. [DOI] [PubMed] [Google Scholar]
  • 16.Pfoh R, Cuesta-Seijo JA, Sheldrick GM. Interaction of an echinomycin-DNA complex with manganese ions. Acta Crystallogr Sect F Struct Biol Cryst Commun. 2009;65(7):660–664. doi: 10.1107/S1744309109019654. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Aishima J, Gitti RK, Noah JE, Gan HH, Schlick T, Wolberger C. A Hoogsteen base pair embedded in undistorted B-DNA. Nucleic Acids Research. 2002;30(23):5244–5252. doi: 10.1093/nar/gkf661. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Patikoglou GA, Kim JL, Sun L, Yang SH, Kodadek T, Burley SK. TATA element recognition by the TATA box-binding protein has been conserved throughout evolution. Genes & Development. 1999;13(24):3217–3230. doi: 10.1101/gad.13.24.3217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Kitayner M, Rozenberg H, Rohs R, Suad O, Rabinovich D, Honig B, Shakked Z. Diversity in DNA recognition by p53 revealed by crystal structures with Hoogsteen base pairs. Nature Structural & Molecular Biology. 2010 Apr;17:423–429. doi: 10.1038/nsmb.1800. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Ethayathulla AS, Tse P, Monti P, Nguyen S, Inga A, Fronza G, Viadiu H. Structure of p73 DNA-binding domain tetramer modulates p73 transactivation. Proceedings of the National Academy of Sciences. 2012;109(16):6066–6071. doi: 10.1073/pnas.1115463109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Vainer R, Cohen S, Shahar A, Zarivach R, Arbely E. Structural basis for p53 Lys120 acetylation dependent DNA binding mode. Journal of Molecular Biology. 2016;428(15):3013–2025. doi: 10.1016/j.jmb.2016.06.009. [DOI] [PubMed] [Google Scholar]
  • 22.Rohs R, Jin X, West SM, Joshi R, Honig B, Mann RS. Origins of specificity in protein-DNA recognition. Annual Review of Biochemistry. 2010;79:233–269. doi: 10.1146/annurev-biochem-060408-091030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Shanmugam G, Kozekov ID, Guengerich FP, Rizzo CJ, Stone MP. Structure of the 1,N2-ethenodeoxyguanosine adduct opposite cytosine in duplex DNA: Hoogsteen base pairing at pH 5.2. Chemical Research in Toxicology. 2008;21(9):1795–1805. doi: 10.1021/tx8001466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Yang H, Zhan Y, Fenn D, Chi LM, Lam SL. Effect of 1-methyladenine on double-helical DNA structures. FEBS Letters. 2008;582(11):1629–1633. doi: 10.1016/j.febslet.2008.04.013. [DOI] [PubMed] [Google Scholar]
  • 25.Nair DT, Johnson RE, Prakash S, Prakash L, Aggarwal AK. Replication by human DNA polymerase-iota occurs by Hoogsteen base-pairing. Nature. 2004;430:377–380. doi: 10.1038/nature02692. [DOI] [PubMed] [Google Scholar]
  • 26.Johnson RE, Prakash L, Prakash S. Biochemical evidence for the requirement of Hoogsteen base pairing for replication by human DNA polymerase iota. Proceedings of the National Academy of Sciences. 2005;102(30):10466–10471. doi: 10.1073/pnas.0503859102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Wang J. DNA polymerases: Hoogsteen base-pairing in DNA replication? Nature. 2005;437(7057):E6–E7. doi: 10.1038/nature04199. [DOI] [PubMed] [Google Scholar]
  • 28.De Luchi D, Tereshko V, Gouyette C, Subirana JA. Structure of the DNA coiled coil formed by d(CGATATATATAT) ChemBioChem. 2006;7(4):585–587. doi: 10.1002/cbic.200500449. [DOI] [PubMed] [Google Scholar]
  • 29.Acosta-Reyes FJ, Alechaga E, Subirana JA, Campos JL. Structure of the DNA duplex d(ATTAAT)2 with Hoogsteen hydrogen bonds. PLOS ONE. 2015 Mar;10:1–9. doi: 10.1371/journal.pone.0120241. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Pous J, Urpí L, Subirana JA, Gouyette C, Navaza J, Campos JL. Stabilization by extra-helical thymines of a DNA duplex with Hoogsteen base pairs. Journal of the American Chemical Society. 2008;130(21):6755–6760. doi: 10.1021/ja078022+. [DOI] [PubMed] [Google Scholar]
  • 31.Abrescia NGA, Thompson A, Huynh-Dinh T, Subirana JA. Crystal structure of an antiparallel DNA fragment with Hoogsteen base pairing. Proceedings of the National Academy of Sciences. 2002;99(5):2806–2811. doi: 10.1073/pnas.052675499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Abrescia NGA, Gonzalez C, Gouyette C, Subirana JA. X-ray and NMR Studies of the DNA oligomer d(ATATAT): Hoogsteen base pairing in duplex DNA. Biochemistry. 2004;43(14):4092–4100. doi: 10.1021/bi0355140. [DOI] [PubMed] [Google Scholar]
  • 33.Gilbert DE, Feigon J. Proton NMR study of the d(ACGTATACGT)]2-2echinomycin complex: conformational changes between echinomycin binding sites. Nucleic Acids Research. 1992;20(10):2411–2420. doi: 10.1093/nar/20.10.2411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Carey PR. Spectroscopic characterization of distortion in enzyme complexes. Chemical Reviews. 2006;106(8):3043–3054. doi: 10.1021/cr0502854. [DOI] [PubMed] [Google Scholar]
  • 35.Tsuboi M, Takahashi S, Harada I. Infrared and Raman spectra of nucleic acids: vibrations in the base-residues. In: Duchesne J, editor. Structural Studies on Nucleic Acids and Other Biopolymers. Ch. 11. Academic Press; 1973. pp. 91–145. [Google Scholar]
  • 36.Tsuboi M. Application of infrared spectroscopy to structure studies of nucleic acids. Applied Spectroscopy Reviews. 1970;3(1):45–90. [Google Scholar]
  • 37.Banyay M, Sarkar M, Gräslund A. A library of IR bands of nucleic acids in solution. Biophysical Chemistry. 2003;104(2):477–488. doi: 10.1016/s0301-4622(03)00035-8. [DOI] [PubMed] [Google Scholar]
  • 38.Delabar JM, Majoube M. Infrared and Raman spectroscopic study of 15N and D-substituted guanines. Spectrochimica Acta, Part A: Molecular and Biomolecular Spectroscopy. 1978;34(2):129–140. [Google Scholar]
  • 39.Ghomi M, Taboury J, Taillandier E. Experimental and calculated study of the vibrational modes of poly(dG-dC)·poly(dG-dC) in B and Z conformations. Biochimie. 1984;66(2):87–92. doi: 10.1016/0300-9084(84)90184-6. [DOI] [PubMed] [Google Scholar]
  • 40.Hirakawa AY, Okada H, Sasagawa S, Tsuboi M. Infrared and Raman spectra of adenine and its 15N and 13C substitution products. Spectrochimica Acta, Part A: Molecular and Biomolecular Spectroscopy. 1985;41(1):209–216. [Google Scholar]
  • 41.Nishimura Y, Tsuboi M, Kubasek WL, Bajdor K, Peticolas WL. Ultraviolet resonance Raman bands of guanosine and adenosine residues useful for the determination of nucleic acid conformation. Journal of Raman Spectroscopy. 1987;18(3):221–227. [Google Scholar]
  • 42.Letellier R, Ghomi M, Taillandier E. Interpretation of dna vibration modes II- the adenosine and thymidine residues involved in oligonucleotides and polynucleotides. Journal of Biomolecular Structure and Dynamics. 1987 Feb;4:663–683. doi: 10.1080/07391102.1987.10507667. [DOI] [PubMed] [Google Scholar]
  • 43.Dohy D, Ghomi M, Taillandier E. Interpretation of DNA vibration modes: III - the behaviour of the sugar pucker vibration modes as a function of its pseudorotation parameters. Journal of Biomolecular Structure & Dynamics. 1989;6(4):741–754. doi: 10.1080/07391102.1989.10507734. [DOI] [PubMed] [Google Scholar]
  • 44.Ghomi M, Letellier R, Liquier J, Taillandier E. Interpretation of DNA vibrational spectra by normal coordinate analysis. International Journal of Biochemistry. 1990;22(7):691–699. doi: 10.1016/0020-711x(90)90003-l. [DOI] [PubMed] [Google Scholar]
  • 45.Toyama A, Takino Y, Takeuchi H, Harada I. Ultraviolet resonance Raman spectra of ribosyl C(1′)-deuterated purine nucleosides: evidence of vibrational coupling between purine and ribose rings. Journal of the American Chemical Society. 1993;115(24):11092–11098. [Google Scholar]
  • 46.Toyama A, Hanada N, Abe Y, Takeuchi H, Harada I. Assignment of adenine ring in-plane vibrations in adenosine on the basis of 14N and 13C isotopic frequency shifts and TUV resonance Raman enhancement. Journal of Raman Spectroscopy. 1994;25(7-8):623–630. [Google Scholar]
  • 47.Tsuboi M, Ueda T, Ushizawa K, Sasatake Y, Ono A, Kainosho M, Ishido Y. Assignments of the deoxyribose vibrations: Isotopic thymidine. Bulletin of the Chemical Society of Japan. 1994;67(5):1483–1484. [Google Scholar]
  • 48.Guan Y, Thomas GJ. Vibrational analysis of nucleic acids. III. Conformation-dependent Raman markers of the phosphodiester backbone modeled by dimethyl phosphate. Journal of Molecular Structure. 1996;379(1):31–41. doi: 10.1016/S0006-3495(96)79474-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Tsuboi M, Komatsu M, Hoshi J, Kawashima E, Sekine T, Ishido Y, Russell MP, Benevides JM, George J, Thomas J. Raman and infrared spectra of (2′S)-2′-2H]thymidine: vibrational coupling between deoxyribosyl and thymine moieties and structural implications. Journal of the American Chemical Society. 1997;119(8):2025–2032. [Google Scholar]
  • 50.Toyama A, Fujimoto N, Hanada N, Ono J, Yoshimitsu E, Matsubuchi A, Takeuchi H. Assignments and hydrogen bond sensitivities of UV resonance Raman bands of the C8-deuterated guanine ring. Journal of Raman Spectroscopy. 2002;33(9):699–708. [Google Scholar]
  • 51.Thamann TJ, Lord RC, Wang AH, Rich A. The high salt form of poly(dG-dC)/poly(dG-dC) is left-handed Z-DNA: Raman spectra of crystals and solutions. Nucleic Acids Research. 1981;9(20):5443–5458. doi: 10.1093/nar/9.20.5443. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Nishimura Y, Tsuboi M, Nakano T, Higuchi S, Sato T, Shida T, Uesugi S, Ohtsuka E, Ikehara M. Raman diagnosis of nucleic acid structure: sugar-puckering and glycosidic conformation in the guanosine moiety. Nucleic Acids Research. 1983;11(5):1579–1588. doi: 10.1093/nar/11.5.1579. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Benevides JM, Wang AHJ, Rich A, Kyogoku Y, Van der Marel GA, Van Boom JH, Thomas GJ. Raman spectra of single crystals of r(GCG)d(CGC) and d(CCCCGGGG) as models for A DNA, their structure transitions in aqueous solution, and comparison with double-helical poly(dG)·poly(dC) Biochemistry. 1986;25(1):41–50. doi: 10.1021/bi00349a007. [DOI] [PubMed] [Google Scholar]
  • 54.Benevides J, Thomas G. Characterization of DNA structures by Raman spectroscopy: high-salt and low-salt forms of double helical poly(dG-dC) in H2O and D2O solutions and application to B, Z and A-DNA. Nucleic Acids Research. 1983;11(16):5747–5761. doi: 10.1093/nar/11.16.5747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Benevides JM, Wang AH, van der Marel GA, van Boom JH, Rich A, Thomas GJ. The Raman spectra of left-handed DNA oligomers incorporating adenine-thymine base pairs. Nucleic Acids Research. 1984;12(14):5913–5925. doi: 10.1093/nar/12.14.5913. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Taboury JA, Liquier J, Taillandier E. Characterization of DNA structures by infrared spectroscopy: double helical forms of poly(dG-dC)·poly(dG-dC), poly(dD8G-dC)·poly(dD8G-dC), and poly(dG-dm5C)·poly(dG-dm5C) Canadian Journal of Chemistry. 1985;63(7):1904–1909. [Google Scholar]
  • 57.Nishimura Y, Tsuboi M, Sato T, Aoki K. Conformation-sensitive Raman lines of mononucleotides and their use in a structure analysis of polynucleotides: guanine and cytosine nucleotides. Journal of Molecular Structure. 1986;146(C):123–153. [Google Scholar]
  • 58.Nishimura Y, Torigoe C, Tsuboi M. Salt induced B - A transition of poly(dG)·poly(dC) and the stabilization of A form by its methylation. Nucleic Acids Research. 1986;14(6):2737–2748. doi: 10.1093/nar/14.6.2737. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Taillandier E, Ridoux JP, Liquier J, Leupin W, Denny WA, Wang Y, Thomas GA, Peticolas WL. Infrared and Raman studies show that poly(dA)·poly(dT) and d(AAAAATTTTT)2 exhibit a heteronomous conformation in films at 75% relative humidity and a B-type conformation at high humidities and in solution. Biochemistry. 1987;26(12):3361–3368. doi: 10.1021/bi00386a017. [DOI] [PubMed] [Google Scholar]
  • 60.Taillandier E, Liquier J, Ghomi M. Conformational transitions of nucleic acids studied by IR and Raman spectroscopies. Journal of Molecular Structure. 1989;214:185–211. [Google Scholar]
  • 61.Kyogoku Y, Higuchi S, Tsuboi M. Infrared absorption spectra of the single crystals of 1-methyl-thymine, 9-methyladenine and their 1:1 complex. Spectrochimica Acta, Part A: Molecular and Biomolecular Spectroscopy. 1967;23(4):969–983. [Google Scholar]
  • 62.Strobel JL, Scovell WM. Laser Raman spectroscopy of a complementary base pair in the Hoogsteen configuration. Biochimica et Biophysica Acta. 1980;608(2):201–214. doi: 10.1016/0005-2787(80)90167-7. [DOI] [PubMed] [Google Scholar]
  • 63.Liquier J, Letellier R, Dagneaux C, Ouali M, Morvan F, Raynier B, Imbach JL, Taillandier E. Triple helix formation by alpha-oligodeoxynucleotides: A vibrational spectroscopy and molecular modeling study. Biochemistry. 1993;32(40):10591–10598. doi: 10.1021/bi00091a008. [DOI] [PubMed] [Google Scholar]
  • 64.Ouali M, Letellier R, Sun JS, Akhebat A, Adnet F, Liquier J, Taillandier E. Determination of G*G·C triple-helix structure by molecular modeling and vibrational spectroscopy. Journal of the American Chemical Society. 1993;115(10):4264–4270. [Google Scholar]
  • 65.Dagneaux C, Liquier J, Taillandier E. FTIR study of a nonclassical dT10*dA10-dT10 intramolecular triple helix. Biochemistry. 1995;34(45):14815–14818. doi: 10.1021/bi00045a023. [DOI] [PubMed] [Google Scholar]
  • 66.Escudé C, Mohammadi S, Sun JS, Nguyen CH, Bisagni E, Liquier J, Taillandier E, Garestier T, Héléne C. Ligand-induced formation of Hoogsteen-paired parallel DNA. Chemical Biology. 1996;3(1):57–65. doi: 10.1016/s1074-5521(96)90085-x. [DOI] [PubMed] [Google Scholar]
  • 67.Miura T, Thomas GJ. Structure and dynamics of interstrand guanine association in quadruplex telomeric DNA. Biochemistry. 1995;34(29):9645–9654. doi: 10.1021/bi00029a042. [DOI] [PubMed] [Google Scholar]
  • 68.Miura T, Benevides J, Thomas GJ. A phase-diagram for sodium and potassiumion control of polymorphism in telomeric DNA. Journal of Molecular Biology. 1995 Apr 28;248:233–238. doi: 10.1016/s0022-2836(95)80046-8. [DOI] [PubMed] [Google Scholar]
  • 69.Krafft C, Benevides JM, Thomas GJ. Secondary structure polymorphism in Oxytricha nova telomeric DNA. Nucleic Acids Research. 2002;30(18):3981–3991. doi: 10.1093/nar/gkf517. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Pagba CV, Lane SM, Wachsmann-Hogiu S. Conformational changes in quadruplex oligonucleotide structures probed by Raman spectroscopy. Biomedical Optics Express. 2011 Feb;2:207–217. doi: 10.1364/BOE.2.000207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Liu K, Miles HT, Frazier J, Sasisekharan V. A novel DNA duplex. A parallel-stranded DNA helix with Hoogsteen base pairing. Biochemistry. 1993;32(44):11802–11809. doi: 10.1021/bi00095a008. [DOI] [PubMed] [Google Scholar]
  • 72.Segers-Nolten GMJ, Sijtsema NM, Otto C. Evidence for Hoogsteen GC base pairs in the proton-induced transition from right-handed to left-handed poly(dG-dC)-poly(dG-dC) Biochemistry. 1997;36(43):13241–13247. doi: 10.1021/bi971326w. [DOI] [PubMed] [Google Scholar]
  • 73.Tajmir-Riahi H, Neault J, Naoui M. Does DNA acid fixation produce left-handed Z structure? FEBS Letters. 1995;370:105–108. doi: 10.1016/0014-5793(95)00802-g. [DOI] [PubMed] [Google Scholar]
  • 74.Puppels GJ, Otto C, Greve J, Robert-Nicoud M, Arndt-Jovin DJ, Jovin TM. Raman microspectroscopic study of low-pH-induced changes in DNA structure of polytene chromosomes. Biochemistry. 1994;33(11):3386–3395. doi: 10.1021/bi00177a032. [DOI] [PubMed] [Google Scholar]
  • 75.Zhang Y, de La Harpe K, Beckstead AA, Martínez-Fernández L, Improta R, Kohler B. Excited-state dynamics of DNA duplexes with different H-bonding motifs. Journal of Physical Chemistry Letters. 2016;7:950–954. doi: 10.1021/acs.jpclett.6b00074. [DOI] [PubMed] [Google Scholar]
  • 76.Fang Y, Wei Y, Bai C, Tang Y, Lin S, Kan L. Hydrated water molecules of pyrimidine/purine/pyrimidine DNA triple helices as revealed by FT-IR spectroscopy: a role of cytosine methylation. Journal of Biomolecular Structure & Dynamics. 1997;14(4):485–93. doi: 10.1080/07391102.1997.10508147. [DOI] [PubMed] [Google Scholar]
  • 77.White AP, Powell JW. Observation of the hydration-dependent conformation of the (dG)20(dG)20(dC)20 oligonucleotide triplex using FTIR spectroscopy. Biochemistry. 1995;34(4):1137–1142. doi: 10.1021/bi00004a006. [DOI] [PubMed] [Google Scholar]
  • 78.Akhebat A, Dagneaux C, Liquier J, Taillandier E. Triple helical polynucleotidic structures: an FTIR study of the C+·G·C triplet. Journal of Biomolecular Structure & Dynamics. 1992;10(3):577–588. doi: 10.1080/07391102.1992.10508669. [DOI] [PubMed] [Google Scholar]
  • 79.Huang X, Yu P, LeProust E, Gao X. An efficient and economic site-specific deuteration strategy for NMR studies of homologous oligonucleotide repeat sequences. Nucleic Acids Research. 1997;25(23):4758–4763. doi: 10.1093/nar/25.23.4758. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Lu K, Miyazaki Y, Summers MF. Isotope labeling strategies for NMR studies of RNA. Journal of Biomolecular NMR. 2010;46(1):113–125. doi: 10.1007/s10858-009-9375-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Sathyamoorthy B, Lee J, Kimsey I, Ganser LR, Al-Hashimi H. Development and application of aromatic [13C, 1H] SOFAST-HMQC NMR experiment for nucleic acids. Journal of Biomolecular NMR. 2014;60(2):77–83. doi: 10.1007/s10858-014-9856-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Delaglio F, Grzesiek S, Vuister GW, Zhu G, Pfeifer J, Bax A. NMRPipe: A multidimensional spectral processing system based on UNIX pipes. Journal of Biomolecular NMR. 1995;6(3):277–293. doi: 10.1007/BF00197809. [DOI] [PubMed] [Google Scholar]
  • 83.Kumar GS, Das S, Bhadra K, Maiti M. Protonated forms of poly[d(G-C)] and poly(dG).poly(dC) and their interaction with berberine. Bioorganic & Medicinal Chemistry. 2003;11(23):4861–4870. doi: 10.1016/j.bmc.2003.09.028. [DOI] [PubMed] [Google Scholar]
  • 84.Bhadra K, Kumar GS, Das S, Islam MM, Maiti M. Protonated structures of naturally occurring deoxyribonucleic acids and their interaction with berberine. Bioorganic & Medicinal Chemistry. 2005;13(16):4851–4863. doi: 10.1016/j.bmc.2005.05.010. [DOI] [PubMed] [Google Scholar]
  • 85.Guéron M, Charretier E, Hagerhorst J, Kochoyan M, Leroy JL, Moraillon A. Imino proton exchange in DNA and RNA. Structure & Methods, Vol 3: DNA & RNA. 1990:113–137. [Google Scholar]
  • 86.Yang H, Lam SL. Effect of 1-methyladenine on thermodynamic stabilities of double-helical DNA structures. FEBS Letters. 2009;583(9):1548–1553. doi: 10.1016/j.febslet.2009.04.017. [DOI] [PubMed] [Google Scholar]
  • 87.Delaney JC, Essigmann JM. Mutagenesis, genotoxicity, and repair of 1-methyladenine, 3-alkylcytosines, 1-methylguanine, and 3-methylthymine in alkB Es-cherichia coli. Proceedings of the National Academy of Sciences. 2004;101(39):14051–14056. doi: 10.1073/pnas.0403489101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Taboury J, Bourtayre P, Liquier J, Taillandier E. Interaction of Z form poly(dG-dC)·poly(dG-dC) with divalent metal ions: localization of the binding sites by I.R. spectroscopy. Nucleic Acids Research. 1984;12(10):4247–4258. doi: 10.1093/nar/12.10.4247. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Keller PB, Hartman KA. Structural forms and transitions for the complex of mercury (II) with poly(dG-dC) Journal of Biomolecular Structure & Dynamics. 1987;4(6):1013–1026. doi: 10.1080/07391102.1987.10507694. [DOI] [PubMed] [Google Scholar]
  • 90.Keller P, Loprete D, Hartman K. Structural forms and transitions of poly (dG-dC) with Cd (II), Ag (I) and NaNO3. Journal of Biomolecular Structure & Dynamics. 1988;5(6):1221–1229. doi: 10.1080/07391102.1988.10506465. [DOI] [PubMed] [Google Scholar]
  • 91.Taillandier E, Peticolas W, Adam S, Huynh-Dinh T, Igolen J. Polymorphism of the d(CCCGCGGG)2 double helix studied by FT-IR spectroscopy. Spectrochimica Acta, Part A: Molecular and Biomolecular Spectroscopy. 1990;46(1):107–112. [Google Scholar]
  • 92.Langlais M, Tajmir-Riahi HA, Savoie R. Raman spectroscopic study of the effects of Ca2+, Mg2+, Zn2+, and Cd2+ ions on calf thymus DNA: binding sites and conformational changes. Biopolymers. 1990;30(7-8):743–752. doi: 10.1002/bip.360300709. [DOI] [PubMed] [Google Scholar]
  • 93.Duguid J, Bloomfield VA, Benevides J, T GJ. Raman spectroscopy of DNA-metal complexes. I. Interactions and conformational effects of the divalent cations: Mg, Ca, Sr, Ba, Mn, Co, Ni, Cu, Pd, and Cd. Biophysical Journal. 1993;65(5):1916. doi: 10.1016/S0006-3495(93)81263-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Muntean CM, Dostál L, Misselwitz R, Welfle H. DNA structure at low pH values, in the presence of Mn2+ ions: a Raman study. Journal of Raman Spectroscopy. 2005;36(11):1047–1051. [Google Scholar]
  • 95.Ghomi M, Taillandier E. Normal coordinate analysis of 5′-dGMP and its deuterated derivatives. European Biophysics Journal. 1985;12(3):153–162. doi: 10.1007/BF00254073. [DOI] [PubMed] [Google Scholar]
  • 96.Wang AHJ, Ughetto G, Quigley GJ, Rich A. Interactions of quinoxaline antibiotic and DNA: the molecular structure of a Triostin A-d(GCGTACGC) complex. Journal of Biomolecular Structure & Dynamics. 1986;4(3):319–342. doi: 10.1080/07391102.1986.10506353. [DOI] [PubMed] [Google Scholar]
  • 97.Gao X, Patel DJ. Antitumour drug-DNA interactions: NMR studies of echinomycin and chromomycin complexes. Quarterly Reviews of Biophysics. 1989 May;22:93–138. doi: 10.1017/s0033583500003814. [DOI] [PubMed] [Google Scholar]
  • 98.Searle MS, Wickham G. Hoogsteen versus Watson-Crick A-T basepairing in DNA complexes of a new group of ‘quinomycin-like’ antibiotics. FEBS Letters. 1990;272(1-2):171–174. doi: 10.1016/0014-5793(90)80476-y. [DOI] [PubMed] [Google Scholar]
  • 99.Gilbert DE, Feigon J. The DNA sequence at echinomycin binding sites determines the structural changes induced by drug binding: NMR studies of echinomycin binding to d(ACGTACGT)2 and d(TCGATCGA)2. Biochemistry. 1991;30(9):2483–2494. doi: 10.1021/bi00223a027. [DOI] [PubMed] [Google Scholar]
  • 100.Nikolova EN, Zhou H, Gottardo FL, Alvey HS, Kimsey IJ, Al-Hashimi HM. A historical account of hoogsteen base-pairs in duplex DNA. Biopolymers. 2013;99(12):955–968. doi: 10.1002/bip.22334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Adam S, Liquier J, Taboury JA, Taillandier E. Right- and left-handed helixes of poly-d(A-T).poly-d(A-T) investigated by infrared spectroscopy. Biochemistry. 1986;25(11):3220–3225. doi: 10.1021/bi00359a021. [DOI] [PubMed] [Google Scholar]
  • 102.Zhou H, Hintze BJ, Kimsey IJ, Sathyamoorthy B, Yang S, Richardson JS, Al-Hashimi HM. New insights into Hoogsteen base pairs in DNA duplexes from a structure-based survey. Nucleic Acids Research. 2015;43(7):3420–3433. doi: 10.1093/nar/gkv241. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Khesbak H, Savchuk O, Tsushima S, Fahmy K. The role of water H-bond imbalances in B-DNA substate transitions and peptide recognition revealed by time-resolved FTIR spectroscopy. Journal of the American Chemical Society. 2011;133(15):5834–5842. doi: 10.1021/ja108863v. [DOI] [PubMed] [Google Scholar]
  • 104.Morari CI, Muntean CM. Numerical simulations of Raman spectra of guanine-cytosine Watson-Crick and protonated Hoogsteen base pairs. Biopolymers. 2003;72(5):339–344. doi: 10.1002/bip.10418. [DOI] [PubMed] [Google Scholar]
  • 105.Bende A, Bogdan D, Muntean CM, Morari C. Localization and anharmonicity of the vibrational modes for G-C Watson-Crick and Hoogsteen base pairs. Journal of Molecular Modeling. 2011;17(12):3265–3274. doi: 10.1007/s00894-011-1002-y. [DOI] [PubMed] [Google Scholar]
  • 106.Bende A, Muntean CM. The influence of anharmonic and solvent effects on the theoretical vibrational spectra of the guanine-cytosine base pairs in Watson-Crick and Hoogsteen configurations. Journal of Molecular Modeling. 2014;20(3):2113. doi: 10.1007/s00894-014-2113-z. [DOI] [PubMed] [Google Scholar]
  • 107.Morari C, Muntean CM, Tripon C, Buimaga-Iarinca L, Calborean A. DFT investigation of the vibrational properties of GC Watson-Crick and Hoogsteen base pairs in the presence of Mg2+, Ca2+, and Cu2+ ions. Journal of Molecular Modeling. 2014;20(4):2220. doi: 10.1007/s00894-014-2220-x. [DOI] [PubMed] [Google Scholar]
  • 108.Chen Y, Eldho NV, Dayie TK, Carey PR. Probing adenine rings and backbone linkages using base specific isotope-edited Raman spectroscopy: Application to group II intron ribozyme domain V. Biochemistry. 2010;49(16):3427–3435. doi: 10.1021/bi902117w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Chen Y, Basu R, Gleghorn ML, Murakami KS, Carey PR. Time-resolved events on the reaction pathway of transcript initiation by a single-subunit RNA polymerase: Raman crystallographic evidence. Journal of the American Chemical Society. 2011;133(32):12544–12555. doi: 10.1021/ja201557w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.O'Connor T, Johnson C, Scovell WM. Raman pH profiles for nucleic acid constituents I. Cytidine and uridine ribonucleosides. Biochimica et Biophysica Acta. 1976;447(4):484–494. doi: 10.1016/0005-2787(76)90085-x. [DOI] [PubMed] [Google Scholar]
  • 111.Letellier R, Ghomi M, Taillandier E. Interpretation of DNA vibration modes: I- The guanosine and cytidine residues involved in poly(dG-dC)·poly(dG-dC) and d(CG)3·d(CG)3. Journal of Biomolecular Structure and Dynamics. 1986;3(4):671–687. doi: 10.1080/07391102.1986.10508455. [DOI] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

Supp FigS1
Supp FigS2
Supp FigS3
Supp FigS4
Supp FigS5
Supp FigS6
Supp info

RESOURCES