Abstract
The majority of current drugs against diseases, such as cancer, can bind to one or more sites in a protein and inhibit its activity. There are, however, well known limits on the number of druggable proteins and complementing current drugs with compounds that could selectively target DNA or RNA would greatly enhance therapeutic progress and options. We are focusing on the design of sequence-specific DNA minor groove binders that, for example, target the promoter sites of transcription factors involved in a disease. We have started with AT specific minor groove binders that are known to enter human cells and have entered clinical trials. To broaden the sequence-specific recognition of these compounds, we have identified several modules that have H-bond acceptors that strongly and specifically recognize G•C base pairs. A lead module is a thiophene-N-alkyl-benzimidazole σ-hole based system with terminal phenyl-amidines where the optimum compounds have excellent affinity and selectivity for a G•C base pair in the minor groove. Efforts are now focused on optimizing this module. We have previously optimized the alkyl group. In the work described here we are evaluating modifications to the compound aromatic system with the goal of improving GC selectivity and affinity as with the N-alkyl modifications. The lead compounds from these studies retain the thiophene-N-alkyl-BI module but have halogen substituents adjacent to an amidine group on the terminal phenyl-amidine. Other improved compounds in this set have modified amidines and conversion of the amidine to an imidazoline, for example, resulted in a strong binding compound with good specificity.
Graphical Abstract

Introduction
There are intense research efforts underway to design new synthetic compounds that have excellent selectivity for targeting RNA or DNA sequences and/or structures1–11. Such compounds could have a major impact on the treatment of important for diseases that are the result of aberrant behavior of transcription factors. Because nucleic acids are at an early stage in the disease development process, targeting them has many advantages, especially since many transcription factors are classified as “undruggable”12. Relative to drugs that bind to protein sites, however, drugs that are specific for DNA are relatively rare and except for aminoglycosides, are even more rare with RNA. Small molecule binding sites in proteins are relatively well-understood but the number of druggable proteins is only a very small part of the genome13. With this limit on protein-drug binding sites, the interest in therapeutic agents that are specific for RNA or DNA has rapidly increased.
Duplex DNA sequences and sequence-dependent local microstructures offer useful and specific targets for binding small molecules3,5,9,10. Major non-duplex structural forms of nucleic acids, such as quadruplexes, also offer the possibility of selective binding interactions and a targeted therapeutic effect14,15. With RNA an enormous variety of different structural forms are known and being discovered, but the selection of specific target regions and structures that will have a therapeutic effect and finding compounds specific for the structure remains a major challenge1,2,4,6,13.
Our development efforts have focused on novel compounds to bind with high selectivity properties in the DNA minor groove. Results so far in several laboratories have found the strongest binding, optimum recognition of DNA and most specific synthetic compounds bind in the DNA minor groove. Compounds that have served as models for AT sequence-specific minor groove binding over an extended time period include pentamidine, Hoechst 33258, netropsin, DAPI, furamidine, and analogs of all of these compounds. Such compounds are classical minor groove agents, they are all highly AT sequence-specific binders, and they have been used in therapeutic applications or for biotechnology purposes such as cell staining16–26. The compounds generally have H-bond donor groups, such as benzimidazoles, amides, and amidines/guanidiniums, that interact with the A-N3 and T=O groups on the edges of A·T base pairs (bps) at the floor of the minor groove. The G-NH2 that projects into the groove interferes with the binding of these compounds both sterically and electronically. To recognize a G·C bp then it is necessary to incorporate H-bond accepter groups into designed, synthetic minor groove binders27,28. The GC recognizing compounds should have a shape that allows space for the G-NH2 H-bond, follows the curvature of the minor groove and can also recognize AT bps in the target sequence. Needless to say, the design and preparation of such compounds is a challenge and to date has only rarely been accomplished9,10,27–36.
To initiate a project to systematically develop compounds that could specifically recognize G·C bps, a focused library of approximately 200 heterocyclic, cationic minor-groove binders with H-bond acceptor groups was evaluated. This analysis produced several initial compounds that could bind to a single G·C bp in an AT context. Three new compounds that bound strongly and specifically to a single G·C bp flanked by A·T bps (Figure S1 in Supporting Information) were selected for the initial study. The accepter groups in these compounds are azabenzimidazole, pyridine and a thiophene-N-methylbenzimidazole (N-MeBI) σ-hole motif which is preorganized into a conformation to match the curvature of the minor groove7. We have recently explored changing the methyl substituent on the thiophene-N-MeBI unit in a successful project to enhance the binding selectivity of that module9. The idea behind these substituent changes was to better recognize microstructural variations in the minor groove in addition to sequence recognition9,37,38. The results with the modified compounds were quite successful, with a modest loss in affinity to single G•C sequences, but with a major gain in binding selectivity over closely related sequences. The sequences that were tested, for example, were a single G•C sequence with an AAAGTTT binding site, and for selectivity a pure AT-binding sequence with an AAATTT site and a GC containing the binding site, AAAGCTTT (Figure 1B). These binding sites were incorporated in hairpin DNA duplexes for binding analysis.
Figure 1.

A) Chemical structure of the modified compounds from DB2457 used in this study. B) The DNA sequences used in this study; DNA sequences used for SPR studies were labeled with 5’-biotin.
In this report, we have built on the results with the modified thiophene-N-MeBI module, which illustrated that modified structures could have significantly improved recognition properties. We have addressed three specific, important questions in this work: what is the effect on binding affinity and selectivity of (i) modifying the terminal amidine group, (ii) incorporating different halogen substituents at different positions in the aromatic systems, and (iii) changing the basic structure of the thiophene-N-MeBI aromatic system? Modifications to include different halogens are based on successful, previous results with a -Cl substituent9. The amidine structure modifications as well as replacing the N-alkyl-BI-thiophene group by other chemically and stereochemically similar structures (Figure 1) are exploratory in order to establish the limits on the thiophene-N-alkyl-BI chemical space. The results with these new compounds illustrate new methods of design in the thiophene-σ-hole system with excellent specificity for recognition of G•C bps can be obtained even without the thiophene or alkyl-BI groups.
Experimental
Materials
In the DNA thermal melting (Tm) and circular dichroism (CD) experiments, the hairpin oligomer sequences were used (Figure 1B). In SPR experiments, 5′-biotin-labeled hairpin DNA oligomers were used. All DNA oligomers were obtained from Integrated DNA Technologies, Inc. (IDT, Coralville, IA) with reverse-phase HPLC purification and mass spectrometry characterization.
The buffer used in Tm and CD experiments was 50 mM Tris-HCl, 100 mM NaCl, 1 mM EDTA, pH 7.4 (TNE 100). The biosensor-surface plasmon resonance (SPR) experiments were performed in filtered, degassed TNE 100 with 0.05% (v/v) surfactant P20.
UV-vis Thermal Melting (Tm)
DNA thermal melting experiments were performed on a Cary 300 Bio UV-vis spectrophotometer (Varian). The concentration of each hairpin DNA sequence was 3 μM in TNE 100 using 1 cm quartz cuvettes. The solutions of DNA and ligands were tested with a ratio of 2:1 [ligand] / [DNA]. All samples were increased to 95 °C and cooled down to 25 °C slowly before each experiment. The spectrophotometer was set at 260 nm with a 0.5 °C/min increase beginning at 25 °C, which is below the DNA melting temperature and ending above it at 95 °C. The absorbance of the buffer was subtracted, and a graph of normalized absorbance versus temperature was created using KaleidaGraph 4.0 software. The ΔTm values were calculated using a combination of the derivative function and estimation from the normalized graphs.
Biosensor-Surface Plasmon Resonance (SPR)
SPR measurements were performed with a four-channel Biacore T200 optical biosensor system (GE Healthcare, Inc., Piscataway, NJ). A streptavidin-derivatized (SA) CM5 sensor chip was prepared for use by conditioning with a series of 180 s injections of 1 M NaCl in 50 mM NaOH (activation buffer) followed by extensive washing with HBS buffer (10 mM HEPES, 150 mM NaCl, 3 mM EDTA, and 0.05% P20, pH 7.4). Biotinylated-DNA samples (AAATTT: 5’-biotin-CCAAATTTGCCTCTGCAAATTTGG-3’; AAAGTTT: 5’-biotin-CCAAAGTTTGCTCTCAAACTTTGG-3’; AAAGCTTT: 5’-biotin-CCAAAGCTTTGCTCTCAAAGCTTTGG-3’) of 25–30 nM were prepared in HBS buffer and immobilized on the flow cell surface by noncovalent capture. Flow cell 1 was left blank as a reference, while flow cells 2–4 were immobilized separately by manual injection of biotinylated-DNA stock solutions (flow rate of 1 μL/min) until the desired amount of DNA response units (RU) was obtained (150–250 RU). Ligand solutions were prepared with degassed and filtered TNE 100 with 0.05% (v/v) surfactant P20 by serial dilutions from a concentrated stock solution. Typically, a series of different ligand concentrations (2 nM to 500 nM) were injected over the DNA sensor chip at a flow rate of 100 μL/min for 180 s, followed by buffer flow for ligand dissociation (600–1800 s). After each cycle, the sensor chip surface was regenerated with a 10 mM glycine solution (pH 2.5) for 30 s followed by multiple buffer injections to yield a stable baseline for the following cycles. The reference response from the blank cell was subtracted from the response in each flow cell containing DNA to give a signal (RU, response units) that is directly proportional to the amount of bound compound. The predicted maximum response per bound compound in the steady-state region (RUmax) was determined from the DNA molecular weight, the amount of DNA on the flow cell, the compound molecular weight, and the refractive index gradient ratio of the compound and DNA. RU was plotted as a function of free ligand concentration (Cfree), and the equilibrium binding constants were determined with a one-site binding model (K2 = 0).
| (1) |
where r represents the moles of bound compound per mole of DNA hairpin duplex, K1 and K2 are macroscopic binding constants (for a single-site or 1:1 model K2 = 0), and Cfree is the free compound concentration in equilibrium with the complex. RUmax in the equation was used as a fitting parameter, and the obtained value was compared to the predicted maximal response per bound ligand to evaluate the stoichiometry. Kinetic analysis was performed by globally fitting the binding results for the entire concentration series using a standard 1:1 kinetic model with integrated mass transport-limited binding parameters as described previously39,40.
Circular Dichroism (CD)
Circular dichroism experiments were performed on a Jasco J-1500 CD spectrometer in 1 cm quartz cuvette at 25 °C. A buffer scan as a baseline was collected first in the same cuvette and subtracted from the scan of the following samples. The hairpin DNA sequence AAATTT: 5’-CCAAATTTGCCTCTGCAAATTTGG-3’; AAAGTTT: 5’-CCAAAGTTTGCTCTCAAACTTTGG-3’ (5 μM), Figure 1B, in TNE 100 was added to the cuvette prior to the titration experiments, and then the compound was added to the DNA solution and incubated for 10 min to achieve equilibrium binding for the DNA-ligand complex formation. For each titration point, four spectra were averaged from 500 to 230 nm wavelength with a scan speed of 50 nm/min, with a response time of 1 s. Baseline-subtracted graphs were created using the KaleidaGraph 4.0 software.
Competition Electrospray Ionization Mass Spectrometry (ESI-MS)
Electrospray Ionization Mass Spectrometry (ESI-MS) analyses were performed on a Waters Q-TOF micro Mass Spectrometer (Waters Corporate, Milford, MA) equipped with an electrospray ionization source (ESI) in a negative ion mode. DNA sequences AAATTT (5’-CCAAATTTGCCTCTCGAAATTTGG-3’), AAAGTTT (5’-GCCAAAGTTTGCCTCTGCAAACTTTGGC-3’) and AAAGCTTT (5’-CCAAAGCTTTGCTCTCAAAGCTTTGG-3’), for ESI-MS experiments were purified by dialyzing it in 50 mM ammonium acetate buffer (pH 6.7) at 4 °C with 3x buffer exchange. Test samples were prepared in 50 mM ammonium acetate with 10% v/v methanol at pH 6.7 and introduced into the ion source through a direct infusion at 5 μl/min flow rate. The competitive experiments were done by mixing a ligand and DNAs with different sequences at different ratios. The instrument parameters were typically as follows: capillary voltage of 2800 V, sample cone voltage of 30 V, extraction cone voltage of 1.0 V, desolvation temperature of 70 °C, and source temperature of 100 °C. Nitrogen was used as nebulizing and drying gas. A multiply charged spectra were acquired through a full scan analysis at mass range from 300–2500 Da and then deconvoluted to the spectra presented. MassLynx 4.1 software was used for data acquisition and deconvolution.
Fluorescence Spectroscopy Binding Determinations
Fluorescence spectra were recorded on a Cary Eclipse Spectrophotometer, with excitation and emission slit width as determined depending on the concentrations of ligands. The free compound solutions at different concentrations were prepared in an appropriate buffer in TNE 100, and DNA sequence (AAATTT: 5’-CCAAATTTGCCTCTGCAAATTTGG-3’; AAAGTTT: 5’-CCAAAGTTTGCTCTCAAACTTTGG-3’; AAAGCTTT: 5’-CCAAAGCTTTGCTCTCAAAGCTTTGG-3’) aliquots were added from a concentrated stock. All titration spectra were collected after allowing an incubation time of 10 min. DB2802 was excited at 342 nm and DB2803 was excited at 336 nm based on molecular absorbance from UV-vis spectroscopy. Emission spectra of these compounds were monitored from 200 nm wavelength range. All the fluorescence titrations were performed at 25 °C. Then the Fluorescence titration spectrums and fitting plots were made in Kaleidagraph 4.0 software to determine the KD value.
Molecular Dynamics (MD) Simulations.
Structure optimization of DB2789 was performed by using DFT/B3LYP theory with the 6–31+G* basis set in Gaussian 09 (Gaussian, Inc., 2009, Wallingford, CT) with Gauss-view 5.0941. Partial charges were derived using the RESP fitting method (Restrained Electrostatic potential)42,43. AMBER 16 (Assisted Model Building with Energy Refinement) software suite was used to perform molecular dynamics (MD) simulations44. Canonical B-form ds[5’-CGAAAGTTTCG-3’][5’-CGAAACTTTCG-3’] DNA was built in Nucleic Acid Builder (NAB) tool in AMBER. AMBER preparation and force field parameter files required conducting molecular dynamics simulations for the DB2789 molecule by using ANTECHAMBER45. Specific atom types assigned for DB2789 molecule were adapted from the ff99 force field. Most of the force field parameters for DB2789 molecule were derived from the existing set of bonds, angles and dihedrals for similar atom types in parm99 and GAFF force fields46. Some dihedral angle parameters were obtained from previously reported parametrized data47,48. The molecular structure with specific atom types used for the DB2789 molecule is shown in Figure S3. Parameters of DB2789 in frcmod file are listed at the Table S1.
The AutoDock Vina program was used to dock the DB2692 in the minor groove of DNA to obtain the initial structure for the DB2789-DNA complex49. MD simulations were performed in explicit solvation conditions where the DNA-DB2789 complex was placed in a truncated octahedron box filled with TIP3P water using xleap program in AMBER. Sodium ions were used to neutralize the system. A 10 Å cutoff was applied on all van der Waals interactions. The MD simulation was carried out using the Sander module with SHAKE algorithm applied to constrain all bonds. Initially, the system was relaxed with 500 steps of steepest-descent energy minimization. The temperature of the system was then increased from 0 K to 310 K over 10 ps under constant-volume conditions. In the final step, the production run on the system was subsequently performed for 500 ns under NPT (constant-pressure) conditions.
Results
Chemistry
Scheme 1 outlines the synthesis of the final diamidines 6a-h. The nitro derivatives 2a-e were obtained by reaction of the bromo derivatives 1a-c with different boronic acid esters under the Suzuki coupling conditions using palladium tetrakis triphenylphosphine as a catalyst in dioxane and potassium carbonate solution. These nitro compounds were reduced to the corresponding amines 3a-e using tin dichloride dihydrate in ethanol. The diamines were converted to the corresponding benzimidazole dinitriles 5a-g on reaction with different aldehydes with the aid of sodium metabisulphite in DMSO. The intermediate dinitriles were converted to the final diamidines through reaction with either lithium bis(trimethylsilyl)amide in THF followed by deprotection of the silylated intermediate using HCl gas in ethanol, or applying Pinner methodology where the dinitriles are suspended and stirred in ethanolic HCl to form the imidate ester hydrochloride that was allowed to react with different amines to afford the final diamidines 6a-h. The synthesis of the diamidine 11 is described in Scheme S1 (supplemental information). 4-Ethynylbenzonitrile 7 was allowed to react with the bromo derivative 8 under Sonogoshira coupling conditions using palladium (II)bis(triphenylphosphine) dichloride and copper iodide as catalysts in a mixture of THF and triethylamine to afford the aldehyde 9. This intermediate aldehyde was converted to the benzimidazole 10 and then to the final diamidine 11, as described in Scheme 1. Compound 18 was prepared as described in Scheme S2. The intermediate bromo derivative 14 was prepared by reaction of the iodo derivative 12 and the tin compound 13 utilizing Stille cross-coupling conditions using palladium tetrakis triphenylphosphine as a catalyst in dioxane. The bromo derivative 14 was allowed to couple with cyanophenylboronic acid to afford the mono nitrile intermediate 15 by applying Suzuki conditions. This nitrile was brominated using NBS in DMF to furnish the intermediate bromo compound 16. This bromo derivative was converted to the dinitrile intermediate 17 using Suzuki coupling conditions. The final diamidine 18 was prepared after reaction lithium bis(trimethylsilyl)amide with the dinitrile as described before. Scheme S3 describes the synthesis of the final diamidine 25, starting from the bromoisatin 19. 2-Acetylthiophene 20 and 19 were heated in ethanolic potassium hydroxide to furnish the quinoline carboxylic acid derivative 21. The carboxylic acid 21 and copper cyanide in NMP heated at reflux led to cyanation and decarboxylation, furnishing the cyanoquinoline derivative 22. Compound 22 was subjected to bromination using NBS, followed by Suzuki coupling to produce the dinitrile 24. This dinitrile was converted to the final amidine 25 using the lithium bis(trimethylsilyl)amide as described before. The synthesis of the diamidines 29a,b is described in Scheme S4. Suzuki coupling of the bromo derivative 26 with cyanophenylboronic acid followed by condensation with different amines to produce the dinitriles 28a,b. Pinner conditions were applied to this dinitrile to produce the final compounds 29a,b.
Scheme 1. Reagents and conditions:

a) 4-cyanophenylboronic acid derivatives, Pd(PPh3)4, Na2CO3/H2O, dioxane, reflux; b) SnCl2/EtOH, reflux; c) Na2S2O5/DMSO, 130 °C; d) i- LiN(TMS)2/THF, ii- HCl/EtOH.OR i- HCl/EtOH. ii- appropriate diamine.
DNA Thermal Melting: Screening for Relative Binding Affinity
Changes in thermal melting temperature (Tm) of DNA provide a rapid way for initial screening of ligands for binding affinity with different DNA sequences (Table 1). Three related DNA sequences that we have successfully used in comparative studies9 were selected for testing (Figure 1B). In order to determine the binding selectivity of our compounds, we chose two quite similar sequences, AAATTT and AAAGCTTT, as the control sequences for the AAAGTTT target site.
Table 1.
Thermal Melting Studies (ΔTm, ℃) of All Test Compounds (Figure 1A) with Pure AT and Mixed DNA Sequences.a
| ΔTm (℃) | AAATTT 70 ℃ | AAAGTTT 66 ℃ | AAAGCTTT 67 ℃ |
|---|---|---|---|
| DB2457b | 6 | 14 | 5 |
| DB2708b | 4 | 14 | 5 |
| DB2798 (6a) | 3 | 6 | 2 |
| DB2759b | 3 | 12 | 3 |
| DB2801 (6b) | 3 | 10 | 3 |
| DB2789 (6c) | 3 | 12 | 4 |
| DB2740b | 5 | 17 | 7 |
| DB2788 (6d) | 5 | 14 | 6 |
| DB2647 (6e) | 2 | 4 | 1 |
| DB2650 (6f) | 5 | 3 | 2 |
| DB2786 (6g) | 4 | 15 | 5 |
| DB2787 (6h) | 3 | 11 | 2 |
| DB2770 (11) | 2 | 7 | 4 |
| DB2655 (18) | 9 | 13 | 6 |
| DB2654 (25) | 6 | 10 | 2 |
| DB2802 (29a) | 9 | 8 | 3 |
| DB2803 (29b) | 3 | 7 | 2 |
ΔTm = Tm (the complex) - Tm (the free DNA). 3 μM DNA sequences were studied in TNE100 with the ratio of 2:1 [ligand]/[DNA]. An average of two independent experiments with a reproducibility of 0.5 ℃. Full DNA sequences: AAATTT: 5’-CCAAATTTGCCTCTGCAAATTTGG-3’; AAAGTTT: 5’-CCAAAGTTTGCTCTCAAACTTTGG-3’; AAAGCTTT: 5’-CCAAAGCTTTGCTCTCAAAGCTTTGG-3’.
Reported compounds.
The parent compound, DB2457, with an N-methyl substituent was previously reported7 and has preferential binding to the single G·C bp sequence and weaker binding to the pure AT and two G·C bps sequences (Table 1). A compound with N-isopropyl substituents, DB2708, has similar binding to AAAGTTT as DB2457, and significantly weaker binding to the pure AT sequence for improved selectivity as desired9. DB2759 with a -Cl substituted on the phenyl ring keeps strong binding to AAAGTTT while binding with the two control sequences drops to an undetectable amount under the standard conditions.
Based on the successful results with modified N-substituents in improving the binding selectivity, focused modifications of the thiophene compounds in other parts of the structure is attractive to search for improved affinity and selectivity. First, considering the improved selectivity of the -Cl derivative, DB2759, -F, -Br halogen modifications were made as well as changing the position of the -Cl substituent to the opposite phenyl ring (Table 1). Both -Cl and the -Br compounds bind with the AAAGTTT sequence strongly and with good selectivity compared to AAATTT and AAAGCTTT. Surprisingly, the -F substituent binds much weaker to AAAGTTT compared to other halogen compounds. We also tested two compounds with modifications on the amidine group, DB2786, and DB2787. The isopropyl substituent obviously improves the binding to AAAGTTT, while the imidazoline compound has similar binding to all three sequences as DB27089. In addition to N-isopropyl substituents, we also tested an N-Ph substituted compound, DB2740, and found that it has strong binding and good selectivity for the single G•C sequence. DB2788 with a -Cl substituent was also prepared based on the success of the substituent in previous studies. In this compound, however, the -Cl modification did not significantly increase the binding selectivity for the N-Ph substituted compound.
In order to explore the effects on binding from variations of the basic aromatic structure, we also introduced a C≡C bond linkage into the structure (DB2770) to change the connection between phenyl rings. The acetylene linkage, however, reduces the binding by a large amount. Changing the N-RBI into a pyridine group produces DB2655, which binds strongly to AAAGTTT but with lower selectivity than the BI compounds. Changing the N-RBI into a quinoline (DB2654) decreases the single G·C bp sequence binding affinity but maintains a high level of selectivity. In addition to the N-RBI module, the thiophene was modified to a thiazole ring (DB2647 and DB2650) and, surprisingly, the simple change to a thiazole causes the compounds to entirely lose the function of the thiophene σ–hole system with the consequence of very poor binding. In addition to the thiazole, the thiophene was replaced with benzothiophene, DB2802 and DB2803. For both methyl and isopropyl BI substituted compounds, the binding affinity to the target AAAGTTT sequence decreases with a decrease in binding selectivity. The N-Me derivative, DB2802, binds to the pure AT test sequence in a similar manner to the single G·C bp sequence.
Biosensor-SPR: Methods for Quantitative Binding
A biacore sensorchip (CM5) functionalized with streptavidin was used to immobilize 5’-biotin labeled hairpin duplex DNA sequences (Figure 1B) in flow cells 2–4 and flow cell 1 was left as a blank, for background subtraction. With different compounds in the flow solutions, we were able to determine comparative binding constants for all the derivatives (Table 2). Sensorgrams were obtained and are shown for representative compounds with the different DNAs (Figure 2). With the AAATTT and AAAGCTTT sequences, most compounds showed significantly weaker binding than to AAAGTTT, with faster on and off rates. No kinetics and limited KD values could be accurately determined for the pure AT and the two G•C bps DNA sequences with most compounds. However, with the single G•C bp containing AAAGTTT sequence, excellent sensorgrams were obtained at low concentrations for most derivatives and representative SPR sensorgrams are shown in Figure 2. Kinetics fits with a one-site model were excellent and allowed a determination of KD values. Kinetics fits were generally required to determine KD values since many sensorgrams did not reach a steady-state level, especially at the lower concentrations.
Table 2.
Summary of Binding Affinity (KD, nM) for the Interaction of All Test Compounds with 5’-Biotin-labeled DNA Sequences using Biosensor-SPR Method.a
| KD (nM) | AAATTT | AAAGTTT | AAAGCTTT |
|---|---|---|---|
| DB2457b | 222 | 4 | 192 |
| DB2708b | 1020 | 4 | 223 |
| DB2798 (6a) | NB | 632 | NB |
| DB2759b | NB | 14 | NB |
| DB2801 (6b) | NB | 14 | NB |
| DB2789 (6c) | NB | 11 | NB |
| DB2740b | NB | 3 | 107 |
| DB2788 (6d) | NB | 10 | 957 |
| DB2647 (6e) | NB | NB | NB |
| DB2650 (6f) | NB | NB | NB |
| DB2786 (6g) | 713 | 1 | 55 |
| DB2787 (6h) | NB | 11 | 1890 |
| DB2770 (11) | NB | 354 | NB |
| DB2655 (18) | 108 | 12 | 315 |
| DB2654 (25) | NB | 62 | NB |
All the results in this table were investigated in Tris-HCl buffer (50 mM Tris-HCl, 100 mM NaCl, 1 mM EDTA, 0.05% P20, pH 7.4) at a 100 μL min–1 flow rate. NB means no detectible binding. The listed binding affinities are an average of two independent experiments carried out with two different sensor chips and the values are reproducible within 10% experimental errors. Full DNA sequences: AAATTT: 5’-biotin-CCAAATTTGCCTCTGCAAATTTGG-3’; AAAGTTT: 5’-biotin-CCAAAGTTTGCTCTCAAACTTTGG-3’; AAAGCTTT: 5’-biotin-CCAAAGCTTTGCTCTCAAAGCTTTGG-3’.
Reported compounds.
Figure 2.

Representative SPR sensorgrams for (A–C), DB2801, (D–F), DB2787 in the presence of AAATTT, AAAGTTT, and AAAGCTTT hairpin DNAs. In (A, C, D, and F), the concentrations of sensorgrams are 2–100 nM of each compound from bottom to top. In (B), the concentrations of DB2801 from bottom to top are 15, 30, 50, 70, and 100 nM; In (E), the concentrations of DB2787 from bottom to top are 15, 30, 50, 90, 100 nM. In (B and E), the solid grey lines are best-fit values for the global kinetic fitting of the results with a single site function. Full DNA sequences: AAATTT: 5’-biotin-CCAAATTTGCCTCTGCAAATTTGG-3’; AAAGTTT: 5’-biotin-CCAAAGTTTGCTCTCAAACTTTGG-3’; AAAGCTTT: 5’-biotin-CCAAAGCTTTGCTCTCAAAGCTTTGG-3’.
While the parent compounds, DB2457 and DB2708, have strong, selective binding to the AAAGTTT sequence, it is necessary to increase the selectivity of new compounds to a higher level for biological applications. As previously reported, DB2759 with -Cl on the aromatic system binds to AAAGTTT with a 14 nM KD and is very selective with no detectable binding to AAATTT and AAAGCTTT9. To find the optimum halogen substituent, different halogen substituents were incorporated and tested in the aromatic systems: DB2798 (-F), DB2801 (-Br) and DB2789 with -Cl on the other phenyl of the aromatic system. The different modification positions of -Cl do not make a major difference in binding affinity when DB2789 and DB2759 are compared. The -Br modification maintains similar binding affinity and specificity to AAAGTTT as with the -Cl derivatives (Figure 2). It is interesting that the smaller, electron-withdrawing atom, F, causes the binding affinity to become 40 times weaker than the -Cl modified compound (KD = 632 nM vs 14 nM). The potential intramolecular hydrogen bonding between the adjacent -F and -H of the amidine group limits the rotation of amidine on the phenyl ring to adopt an appropriate angle for binding to DNA minor groove and significantly reduces the binding affinity of the amidine group with the adjacent T=O of DNA.
As previously reported, DB2740 with an N-Ph substituent binds strongly to AAAGTTT but with limited selectivity versus AAAGCTTT9. The above results suggested that a -Cl substituent would improve the selectivity of the N-phenyl compound. To evaluate this point, DB2788 was synthesized and it binds to AAAGTTT slightly weaker than DB2740, but with 10-fold better binding selectivity to AAAGCTTT. This result shows that a -Cl on the aromatic systems, adjacent to an amidine group, can increase the AAAGTTT binding selectivity of different thiophene-N-RBI type compounds.
To evaluate the effects of changes in the chemical space of the compounds, two groups were also introduced into the amidines of DB2708. DB2786 with isopropyl amidine groups represents the strongest binding (KD = 1 nM) to AAAGTTT in this series. Unfortunately, DB2786 loses binding selectivity with a 55 nM KD for AAAGCTTT. DB2787 with amidine replaced by imidazoline binds to AAAGTTT similar to the -Cl compound but with slightly worse selectivity (Figure 2). In order to confirm the importance of the thiophene-N-RBI in the core system of the original compounds. Five compounds with basic structure variations of the thiophene σ-hole system were prepared. With the introduction of an alkyne between thiophene and phenyl ring of DB2708, the binding to AAAGTTT dramatically decreases. The replacement of N-RBI with pyridine, DB2655 gives a strong binding compound but with poor selectivity. The replacement of N-RBI-phenyl with quinoline, DB2654, gives a compound with good selectivity but weaker binding affinity. Surprisingly, the thiazole-N-RBI substituents destroyed the binding ability for all three DNA sequences. This is a surprising and puzzling result given the success of thiazole in other minor groove binders50. These results indicate the thiophene, or related derivative such as benzothiophene, is necessary but can be combined with the other kinds of N containing heterocyclic groups for a 1, 4 N···S interaction to produce a σ-hole system.
Circular dichroism (CD): Determination of DNA Binding Mode
CD titration experiments are an effective and convenient method of evaluating the binding mode and the saturation limit for compounds binding with DNA sequences. CD spectra monitor the asymmetric environment of the binding of the compound to DNA and therefore, can be used to obtain information on the binding mode51,52. No CD signals have been observed at 300–450 nm wavelength for the free compounds. However, in Figure 3 (left, middle), on the addition of the compounds into DNA, substantial positive induced CD signals (ICD) are seen in the absorption region between 300 and 450 nm. These positive ICD signals in this region indicate a minor groove binding mode by these ligands, as expected from their structures. When the Ligand/DNA ratio is titrated to 1:1, the saturated spectra indicate 1:1 binding between ligand and AAAGTTT. However, in Figure 3 (right), when we change sequence from AAAGTTT to AAATTT, the ICD is low and indicates weak binding between DB2788 and the pure AT sequence. In summary, as can be seen from Figure 3, DB2786 and DB2788 give CD spectra typical of complexes in the minor groove of the AAGTTT sequences with a 1:1 stoichiometry, in agreement with SPR results and modeling.
Figure 3.

Circular dichroism spectra for the titration of representative compounds, DB2786, DB2788, with a 5 μM AAAGTTT or AAATTT sequence in TNE100 at 25℃. Arrows indicate the changes. Full DNA sequences: AAATTT: 5’-CCAAATTTGCCTCTGCAAATTTGG-3’; AAAGTTT: 5’-CCAAAGTTTGCTCTCAAACTTTGG-3’
Fluorescence Titrations: Binding affinity
Fluorescence spectroscopy is an effective method to quantitatively measure the interactions between biomolecules and small molecules, such as a small organic molecule binding with DNA. If a change in the fluorescence intensity accompanies the binding of two species, this can be used to monitor the binding interaction and determine the stoichiometry of binding using equilibrium titration methods53,54.
In order to study the binding specificity of DB2802 and DB2803 with benzothiophene, which have non-optimum properties for SPR studies, fluorescence titration experiments were conducted. The fluorescence titration spectra and binding affinity fitting plots are shown in Figure S2 (Supporting Information). The fluorescence intensity of DB2802 increases on titration with AAATTT. For AAAGTTT and AAAGCTTT, the fluorescence intensities both decrease in the titration. From DNA thermal melting results, the binding affinity between DB2802 with AAATTT and AAAGTTT are similar and stronger than AAAGCTTT. DB2803 with an N-isopropyl group binds to AAAGTTT strongly compared to the pure AT and AAAGCTTT sequences. The results agree with our previous analysis of the N-isopropyl benzimidazole that showed strong binding to AAAGTTT with good selectivity. The above results indicate that the isopropyl group on benzimidazole matches the microstructure of AAAGTTT better than methyl group and increases the binding specificity for both the thiophene and benzothiophene groups.
Competition electrospray ionization mass spectrometry (ESI–MS): A Direct Determination of Binding Stoichiometry and Binding Specificity with Relative Binding Affinity
Mass spectrometry is an excellent method for the evaluation of relative binding affinity and specificity as well as stoichiometry. It provides a good test of experimental results obtained from other methods, such as SPR, where macromolecule-ligand stoichiometry is obtained by fitting the signal at different concentrations8,55. For this series of compounds, we chose two representative thiophene-BI substituents, DB2787, and the parent compound DB2708, for the competition ESI-MS experiments. DB2787 binds to sequence AAAGTTT with 11 nM KD value and good selectivity. DB2708 binds to AAAGTTT with 4 nM KD value but the binding selectivity is less than DB2787. In Figure 4A, three free DNA peaks are shown for AAATTT (m/z = 7302), AAAGCTTT (m/z = 7921), and AAAGTTT (m/z = 8540). On addition of DB2787, the peak for AAAGTTT (m/z = 8540) decreases with the simultaneous appearance of a new peak at m/z = 9070 that is characteristic of a 1:1 AAAGTTT–DB2787 complex (Figure 4B). There is no appearance of any complex peak with the other DNA sequences. As shown in Figure 4C, the same three DNA sequences are used to test DB2708. On the addition of DB2708, the peak of AAAGTTT almost disappeared with a new peak at m/z = 9019 that is characteristic of a 1:1 AAAGTTT–DB2708 complex (Figure 4D). However, another small new peak at m/z = 8540 appeared and is consistent with a 1:1 AAAGCTTT-DB2708 complex. Since the binding affinity between the DNA sequence AAAGCTTT and DB2708 is 223 nM KD (AAAGTTT/AAAGCTTT binding ratio is 56), with the increasing titration of DB2708 (up to 4:1 ratio), the binding between DB2708 and AAAGCTTT can be observed. Moreover, DB2787 with excellent selectivity (AAAGTTT/AAAGCTTT binding ratio is 172) only shows the AAAGTTT complex under these conditions. When excess ligands were added in the DNA mixture solution, only 1:1 Ligand/DNA complex peaks were observed. These results confirm the 1:1 stoichiometry between ligands and DNA, which agrees with CD results.
Figure 4.

ESI-MS negative mode spectra of the competition binding of sequences AAATTT (5’-CCAAATTTGCCTCTCGAAATTTGG-3’), AAAGTTT (5’-GCCAAAGTTTGCCTCTGCAAACTTTGGC-3’) and AAAGCTTT (5’-CCAAAGCTTTGCTCTCAAAGCTTTGG-3’) (10 μM each); with 40 mM DB2787 or DB2708 in buffer (50 mM ammonium acetate with 10% methanol (v/v), pH 6.8). (A and C) The ESI-MS spectra of free DNA mixture. (B and D) The ESI-MS spectra of DNA mixture with compounds. The ESI-MS results shown here are deconvoluted spectra and molecular weights are shown with each peak.
Molecular Structure: Molecular Dynamics
To help better understand the structural basis of molecular recognition of DNA sequences with a single G·C bp in an AT context, molecular dynamics (MD) simulations for the -Cl substituted N-isopropyl compound, DB2789, with the [5’-CGAAAGTTTCG-3’][5’-CGAAACTTTCG–-3’] duplex sequence were conducted. The 500 ns MD simulation has been performed by using the Amber 16 software package in the presence of 0.15 M NaCl44. Force constants for DB2789 were determined as described previously and in the Methods Section and added to the force field for the simulations46,47. Three optimum H-bonds in the complex have been observed. Both amidine groups form -N–H to T=O H-bonds (Figure 5A) that are an average of 2.9 Å in length. The amidines also form frequent highly dynamic H-bonds to terminal water molecules that move in and out of the minor groove. These water molecules frequently also form H-bonds with A·T bps at the floor of the minor groove and help link the compound to the specific binding site in the groove and stabilize the complex. The third strong H-bond is from the central G-NH that projects into the minor groove to the unsubstituted imidazole N in the BI group of DB2789, Figure 5A, to account for much of the binding selectivity of DB2789. Additional selectivity in binding is provided by the -CH group of the six-member ring of BI that points into the minor groove. This -CH forms a dynamic close interaction with the -N3 of the dG base of the central G·C bp. Additional direct stabilizing interactions are formed by other -CH groups that point to the floor of the minor groove from the two phenyls of DB2789. These -CH groups form significant dynamic interactions with A-N3 and T=O groups on the bases at the floor of the minor groove. Although DB2789 is optimally oriented along the minor groove with appropriate twist to match the minor groove curvature, the bulky -Cl group at the -ortho position of the bottom amidine group restricted the degree of freedom of the rotation of the -amidine group (Figure 5B, D). On the other hand, the top amidine group shows dynamic amidine group rotations due to the constraint reduction (Figure 5C, E).
Figure 5.

Molecular Dynamics (MD) model of DB2789 bound to an AAAGTTT site: (A) A wire-ball and stick model viewed into the minor groove of the AAAGTTT binding site with bound DB2789. The DNA bases are represented in tan-white-red-blue-yellow(C–H–O–N–P) color scheme and DB2789 is magenta-white-blue-yellow-green (C–H–N–S-Cl) color scheme. The important interactions between different sections of the DB2789-DNA complex are illustrated in yellow dashed lines. The terminal amidine groups form strong hydrogen bonds with the carbonyl groups of dT=O. The imidazole-N makes a strong hydrogen bond interaction with the exocyclic NH of dG. The BI–C–H that points to the floor of the minor groove forms a strong interaction with the N3 of dG in the minor groove. (B) and (C) represent the atoms of two amidine groups that were selected for the dihedral angle plots in (D) and (E).
Discussion
We are engaged in a project that first time is converting strictly AT specific heterocyclic diamidine DNA minor groove binders into compounds that can bind to mixed AT and GC sequences of DNA11. This involves the creation of a set of modules that can recognize a single G•C bp and combining the modules or truncated versions of them to recognize a range of complex sequences, such as promoter sites of transcription factors. The ability to target such sites could have a major impact on understanding of transcription factor function as well as therapeutic development. An example is the PU.1 transcription factor that is involved in the development of a number of cancers such as AML. There are no drugs effective against PU.1, and it has been classified as “undruggable”. In cases like PU.1 it seems more effective to target the promoter sequence rather than the transcription factor for the treatment of AML56.
The first step in this project involved the design and synthesis of a broad array of heterocyclic diamidines that could recognize a single G•C bp in an AT context7,9,57 (Figure S1). This effort produced several compounds that were selective for the PU. 1 promotor target sequence. One of our lead compounds was based on a thiophene-N-methyl-benzimidazole σ-hole system that is preorganized to bind to the DNA minor groove (Figure 1).
Our goal in the research described in this report is to probe the limits on the thiophene-N-MeBI module to determine if there are improved structures for strong and specific binding to G•C bps. As noted in the Introduction, three specific, important questions are addressed: what is the effect on binding affinity and selectivity of (i) modifying the amidine groups, (ii) incorporating different halogen substituents in the aromatic systems, and (iii) modifying the basic structure of the thiophene-N-MeBI aromatic system? In the halogen series, our initial compound had a -Cl substituent substituted adjacent to the amidine on the phenyl connected to the N-alkyl-BI group. To determine if this is the optimum halogen, -Br and -F were substituted in the same position. In quantitative SPR studies, the -Br group performed identically to the -Cl with good binding and excellent selectivity for the single G•C bp sequence. This substituent diversity may be helpful in targeting different cell types. Somewhat surprisingly, the -F substituent gave very poor binding, which we attribute to its locking the adjacent amidine into an unfavorable orientation for H-bonding to DNA through H-bonding to an adjacent amidine -NH. Putting the -Cl on the phenyl at the other end of the compound gave both stronger binding and excellent specificity for the single G•C bp sequence. The benzimidazole was also modified with an N-phenyl substituent, which gave quite strong binding but poor selectivity for the two G•C bps DNA sequence. A -Cl was added to determine if its enhanced selectivity would be available in this system. The N-phenyl-Cl compound has somewhat weaker binding, as with other -Cl derivatives, but with improved selectivity, although not as much improvement as desired.
Modifying the amidines with isopropyl groups gave the strongest binding of any compound in the series to the single G•C bp sequence, but unfortunately, the binding to the two G•C bps DNA was also enhanced. Converting the amidine to an imidazoline slightly reduced the binding to the single G•C bp sequence but the selectivity for the single G•C bp sequence over 150. A major question was whether the thiophene-N-alkyl-BI module was required or whether other σ-hole structures with H-bond acceptors would recognize the single G•C bp sequence. The first compounds in these experiments were what we thought was a conservative substitution of thiazole for thiophene to make two isomeric compounds, DB2740 and DB2788. It was a surprise that neither of these compounds gave any detectible binding to any of the three test DNAs, especially, since the triazole group has performed well in other minor groove binders50. To maintain the σ-hole the N-alkyl-BI group was replaced with a pyridine, DB2655, and quinoline, DB2654. The pyridine had satisfactory binding to the single G•C bp sequence but poor selectivity for the two G•C bps sequence and especially for the pure AT sequence. The quinoline had significantly reduced binding to the single G•C bp sequence but no detectible binding to the AT or GC test sequences.
In conclusion, there are three lead compounds from this design, synthesis and biophysical studies of GC recognition agents. All three compounds have halogen substituents adjacent to one amidine. Either amidine seems to give excellent results for the single G•C bp sequence. All three compounds have KD values in the quite strong 10 – 15 nM range for the single G•C bp sequence with no detectible binding to the AT and GC test sequences. The twist of the amidine caused by the halogen reduces binding to all sequences, but the reduction with AT and GC sequences is much larger. We have previously noted the microstructural variations in the minor groove among the three sequences and this difference obviously works to significantly enhance the selectivity of compounds with the extra twist on the amidines or size of the N-BI substituent9. The next compound to consider and perhaps test in cells is the imidazoline, DB2787. It also has good binding to the single G•C bp sequence, excellent selectivity over the AT sequence, which is the natural binding sequence for these type heterocycle diamidines and related compounds. Relative to the halogen derivatives, the imidazoline has slightly reduced selectivity for the two G•C bps sequence. This project has now shifted toward an emphasis on recognizing two G•C bps in the same sequences as will be necessary for strong binding with selectivity to other binding sites on the PU.1 promoter as well as recognizing promoters for other transition factors such as HOX A958.
Supplementary Material
Acknowledgments
We thank the National Institutes of Health Grant GM111749 (W.D.W. and D.W.B.) for financial support. We also thank the Leukemia & Lymphoma Society (LLS) award: Leukemia & Lymphoma Society 6504-17 for support. The assistance of Dr. Siming Wang is much appreciated for the data collection in electrospray ionization mass spectrometry.
Footnotes
Supporting Information
Chemistry and synthesis procedure; Scheme S1-S4 show synthesis routes of DB2770, DB2655, DB2654, DB2802, and DB2803; Figure S1 shows the structure of three initial single G•C bp binders; Figure S2 shows Fluorescence titration results of DB2802 and DB2803 with three DNA sequences; Figure S3 and Table S1 show DB2789 molecular structure and Frcmod file; 1H NMR spectrums for new compounds.
The authors declare no competing financial interest.
References
- (1).Disney MD, Dwyer BG, and Childs-Disney JL (2018) Drugging the RNA World. Cold Spring Harb. Perspect. Biol 10, a034769. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (2).Juru AU, Patwardhan NN, and Hargrove AE (2019) Understanding the Contributions of Conformational Changes, Thermodynamics, and Kinetics of RNA–Small Molecule Interactions. ACS Chem. Biol 14, 824–838. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (3).Bhaduri S, Ranjan N, and Dev P. Arya. (2018) An overview of recent advances in duplex DNA recognition by small molecules. Beilstein J. Org. Chem 14, 1051–1086. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (4).Connelly CM, Moon MH, and Schneekloth JS Jr. (2016) The Emerging Role of RNA as a Therapeutic Target for Small Molecules. Cell Chem. Biol 23, 1077–1090. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (5).Pazos E, Mosquera J, Vázquez ME, and Mascareñas JL (2011) DNA Recognition by Synthetic Constructs. ChemBioChem 12, 1958–1973. [DOI] [PubMed] [Google Scholar]
- (6).Lee J, Bai Y, Chembazhi UV, Peng S, Yum K, Luu LM, Hagler LD, Serrano JF, Chan HYE, Kalsotra A, and Zimmerman SC (2019) Intrinsically cell-penetrating multivalent and multitargeting ligands for myotonic dystrophy type 1. Proc. Natl. Acad. Sci. U. S. A 116, 8709–8714. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (7).Guo P, Paul A, Kumar A, Farahat AA, Kumar D, Wang S, Boykin DW, and Wilson WD (2016) The Thiophene “Sigma-Hole” as a Concept for Preorganized, Specific Recognition of G·C Base Pairs in the DNA Minor Groove. Chem. Eur. J 22, 15404–15412. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (8).Guo P, Paul A, Kumar A, Harika NK, Wang S, Farahat AA, Boykin DW, and Wilson WD (2017) A modular design for minor groove binding and recognition of mixed base pair sequences of DNA. Chem. Commun 53, 10406–10409. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (9).Guo P, Farahat AA, Paul A, Harika NK, Boykin DW, and Wilson WD (2018) Compound Shape Effects in Minor Groove Binding Affinity and Specificity for Mixed Sequence DNA. J. Am. Chem. Soc 140, 14761–14769. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (10).Kawamoto Y, Bando T, and Sugiyama H (2018) Sequence-specific DNA binding Pyrrole–imidazole polyamides and their applications. Bioorg. Med. Chem 26, 1393–1411. [DOI] [PubMed] [Google Scholar]
- (11).Paul A, Guo P, Boykin DW, and Wilson WD (2019) A New Generation of Minor-Groove-Binding—Heterocyclic Diamidines That Recognize G•C Base Pairs in an AT Sequence Context. Molecules 24, 946. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (12).Koehler AN (2010) A complex task? Direct modulation of transcription factors with small molecules. Curr. Opin. Chem. Biol 14, 331–340. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (13).Warner KD, Hajdin CE, and Weeks KM (2018) Principles for targeting RNA with drug-like small molecules. Nat. Rev. Drug Discovery 17, 547–558. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (14).Wang KB, Elsayed MSA, Wu G, Deng N, Cushman M, and Yang D (2019) Indenoisoquinoline Topoisomerase Inhibitors Strongly Bind and Stabilize the MYC Promoter G‑Quadruplex and Downregulate MYC. J. Am. Chem. Soc 141, 11059–11070. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (15).Gunaratnam M, Collie GW, Reszka AP, Todd AK, Parkinson GN, and Neidle S (2018) A naphthalene diimide G-quadruplex ligand inhibits cell growth and down-regulates BCL-2 expression in an imatinib-resistant gastrointestinal cancer cell line. Bioorg. Med. Chem 26, 2958–2964. [DOI] [PubMed] [Google Scholar]
- (16).Nozeret K, Loll F, Cardoso GM, Escudé C, and Boutorine AS (2018) Interaction of fluorescently labeled pyrrole-imidazole polyamide probes with fixed and living murine and human cells. Biochimie 149, 122–134. [DOI] [PubMed] [Google Scholar]
- (17).Chazotte B (2011) Labeling lysosomes in live cells with LysoTracker. Cold Spring Harb. Protoc 2, pdb.prot5557. [DOI] [PubMed] [Google Scholar]
- (18).Giordani F, Munde M, Wilson WD, Ismail MA, Kumar A, Boykin DW, and Barrett M. P. Green (2014) Fluorescent Diamidines as Diagnostic Probes for Trypanosomes. Antimicrob. Agents Chemother 58, 1793–796. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (19).Wilson WD, Nguyen B, Tanious FA, Mathis A, Hall JE, Stephens CE, and Boykin DW (2005) Dications That Target the DNA Minor Groove: Compound Design and Preparation, DNA Interactions, Cellular Distribution and Biological Activity. Curr. Med. Chem.: Anti-Cancer Agents 5, 389–408. [DOI] [PubMed] [Google Scholar]
- (20).Paine MF, Wang MZ, Generaux CN, Boykin DW, Wilson WD, De Koning HP, Olson CA, Pohlig G, Burri C, Brun R, Murilla GA, Thuita JK, Barrett MP, and Tidwell RR (2010) Diamidines for human African trypanosomiasis. Curr. Opin. Investig. Drugs 11, 876–883. [PubMed] [Google Scholar]
- (21).Soeiro MNC, Werbovetz K, Boykin DW, Wilson WD, Wang MZ, and Hemphill A (2013) Novel amidines and analogues as promising agents against intracellular parasites: a systematic review. Parasitology 140, 929–951. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (22).Ming X, Ju W, Wu H, Tidwell RR, Hall JE, and Thakker DR (2009) Transport of dicationic drugs pentamidine and furamidine by human organic cation transporters. Drug Metab. Dispos 37, 424–430. [DOI] [PubMed] [Google Scholar]
- (23).Wong C-H, Nguyen L, Peh J, Luu LM, Sanchez JS, Richardson SL, Tuccinardi T, Tsoi H, Chan WY, Chan HYE, Baranger AM, Hergenrother PJ, and Zimmerman SC (2014) Targeting Toxic RNAs that Cause Myotonic Dystrophy Type 1 (DM1) with a Bisamidinium Inhibitor. J. Am. Chem. Soc 136, 6355–6361. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (24).Siboni RB, Bodner MJ, Khalifa MM, Docter AG, Choi JY, Nakamori M, Haley MM, and Berglund JA (2015) Biological Efficacy and Toxicity of Diamidines in Myotonic Dystrophy Type 1 Models. J. Med. Chem 58, 5770–5780. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (25).Sánchez MI, Mosquera J, Vázquez EM, and Mascareñas JL (2014) Reversible supramolecular assembly at specific DNA sites: Nickel-promoted bivalent DNA binding with designed peptide and bipyridyl–bis(benzamidine) components. Angew. Chem. Int. Ed 53, 9917–9921. [DOI] [PubMed] [Google Scholar]
- (26).Rodríguez J, Mosquera J, Couceiro JR, Vázquez ME, and Mascareñas JL (2015) The AT-Hook motif as a versatile minor groove anchor for promoting DNA binding of transcription factor fragments. Chem. Sci 6, 4767–4771. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (27).Kopka ML, Yoon C, Goodsell D, Pjura P, and Dickerson RE (1985) Binding of an antitumor drug to DNA, Netropsin and C-G-C-G-A-A-T-T-BrC-G-C-G. J. Mol. Biol 183, 553–563. [DOI] [PubMed] [Google Scholar]
- (28).Lown JW, Krowicki K, Bhat UG, Skorobogaty A, Ward B, and Dabrowiak JC (1986) Molecular recognition between oligopeptides and nucleic acids: novel imidazole-containing oligopeptides related to netropsin that exhibit altered DNA sequence specificity. Biochemistry, 25, 7408–7416. [DOI] [PubMed] [Google Scholar]
- (29).Kielkopf CL, Baird EE, Dervan PB, and Rees DC (1998) Structural basis for G.C recognition in the DNA minor groove. Nat. Struct. Biol 5, 104–109. [DOI] [PubMed] [Google Scholar]
- (30).Dervan PB, and Edelson BS (2003) Recognition of the DNA minor groove by pyrrole-imidazole polyamides. Curr. Opin. Struct. Biol 3, 284–299. [DOI] [PubMed] [Google Scholar]
- (31).Nozeret K, Loll F, Cardoso GM, Escude C, and Boutorine AS (2018) Interaction of fluorescently labeled pyrrole-imidazole polyamide probes with fixed and livingmurine and human cells. Biochimie 149, 122–134. [DOI] [PubMed] [Google Scholar]
- (32).Kiakos K, Pett L, Satam V, Patil P, Hochhauser D, Lee M, and Hartley JA (2015) Nuclear localization and gene expression modulation by a fluorescent sequence-selective p-anisyl-benzimidazolecarboxamido Imidazole-Pyrrole polyamide. Chem. Biol 22, 862–875. [DOI] [PubMed] [Google Scholar]
- (33).James PL, Merkina EE, Khalaf AI, Suckling CJ, Waigh RD, Brown T, and Fox KR (2004) DNA sequence recognition by an isopropyl substituted thiazole polyamide. Nucleic Acids Res 32, 3410–3417. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (34).Mosquera J, Jiménez-Balsa A, Dodero VI, Vázquez EM, and Mascareñas JL (2013) Stimuli-responsive selection of target DNA sequences by synthetic bZIP peptides. Nat. Commun 4, 1874. [DOI] [PubMed] [Google Scholar]
- (35).Paul A, Kumar A, Nanjunda R, Farahat AA, Boykin DW, and Wilson WD (2016) Systematic synthetic and biophysical development of mixed sequence DNA binding agents. Org. Biomol. Chem 15, 827–835. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (36).Chai Y, Paul A, Rettig M, Wilson WD, and Boykin DW (2014) Design and synthesis of heterocyclic cations for specific DNA recognition: From AT-rich to mixed-base-pair DNA sequences. J. Org. Chem 79, 852–866. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (37).Zhou T, Yang L, Lu Y, Dror I, Dantas Machado AC, Ghane T, Di Felice R, and Rohs R (2013) DNAshape: a method for the high-throughput prediction of DNA structural features on a genomic scale. Nucleic Acids Res 41, W56–W62. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (38).Rohs R, West SM, Sosinsky A, Liu P, Mann RS, and Honig B (2009) The role of DNA shape in protein-DNA recognition. Nature 461, 1248–1253. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (39).Liu Y, Kumar A, Depauw S, Nhili R, David-Cordonnier M-H, Lee MP, Ismail MA, Farahat AA, Say M, Chackal-Catoen S, Batista-Parra A, Neidle S, Boykin DW, and Wilson WD (2011) Water-Mediated Binding of Agents that Target the DNA Minor Groove. J. Am. Chem. Soc 133, 10171–10183. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (40).Munde M, Kumar A, Nhili R, Depauw S, David-Cordonnier M-H, Ismail MA, Stephens CE, Farahat AA, Batista-Parra A, Boykin DW, and Wilson WD (2010) DNA minor groove induced dimerization of heterocyclic cations: compound structure, binding affinity, and specificity for a TTAA site. J. Mol. Biol 402, 847–964. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (41).Frisch MJ, Trucks GW, and Schlegel HB; Scuseria GE … Gaussian, Inc. Wallingford; CT; 2010. Gaussian 09, revision b. 01, Gaussian [Google Scholar]
- (42).Bayly CI, Cieplak P, Cornell W, and Kollman PA (1993) A well-behaved electrostatic potential based method using charge restraints for deriving atomic charges: the RESP model. J. Phys. Chem 97, 10269–10280. [Google Scholar]
- (43).Singh UC, and Kollman PA (1984) An approach to computing electrostatic charges for molecules. J. Comput. Chem 5, 129–145. [Google Scholar]
- (44).Case DA, Babin V, Berryman J, Betz RM, Cai Q, Cerutti DS, Cheatham TE III, Darden TA, Duke RE, Gohlke H, Goetz AW, Gusarov S, Homeyer N, Janowski P, Kaus J, Kolossváry I, Kovalenko A, Lee TS, LeGrand S, Luchko T, Luo R, Madej B, Merz KM, Paesani F, Roe DR, Roitberg A, Sagui C, Salomon-Ferrer R, Seabra G, Simmerling CL, Smith W, Swails J, Walker RC, Wang J, Wolf RM, Wu X, and Kollman PA (2014) AMBER 14. [Google Scholar]
- (45).Wang J, Wang W, Kollman PA, and Case DA (2006) Automatic atom type and bond type perception in molecular mechanical calculations. J. Mol. Graph. Model 25, 247–260. [DOI] [PubMed] [Google Scholar]
- (46).Harika NK, Germann MW, and Wilson WD (2017) First Structure of a Designed Minor Groove Binding Heterocyclic Cation that Specifically Recognizes Mixed DNA base pair Sequences. Chem. Eur. J 23, 17612–17620. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (47).Athri P, and Wilson WD (2009) Molecular Dynamics of Water-Mediated Interactions of a Linear Benzimidazole–Biphenyl Diamidine with the DNA Minor Groove. J. Am. Chem. Soc 131, 7618–7625. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (48).Špačková N, Cheatham TE, Ryjáček F, Lankaš F, van Meervelt L, Hobza P, and Šponer J (2003) Molecular dynamics simulations and thermodynamics analysis of DNA-drug complexes. Minor groove binding between 4’,6-diamidino-2-phenylindole and DNA duplexes in solution. J. Am. Chem. Soc 125, 1759–1769. [DOI] [PubMed] [Google Scholar]
- (49).Trott O, and Olson AJ (2009) AutoDock Vina: improving the speed and accuracy of docking with a new scoring function, efficient optimization and multithreading. J. Comput. Chem 65, 455–461. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (50).Salvia M-V, Addison F, Alniss HY, Buurma NJ, Khalaf AI, Mackay SP, Anthony NG, Suckling CJ, Evstigneev MP, Santiago AH, Waigh RD and Parkinson JA (2013) Thiazotropsin aggregation and its relationship to molecular recognition in the DNA minor groove. Biophys. Chem 179, 1–11. [DOI] [PubMed] [Google Scholar]
- (51).Fornander LH, Wu L, Billeter M, Lincoln P, and Nordén B (2013) Minor-groove binding drugs: where is the second Hoechst 33258 molecule? J. Phys. Chem. B 117, 5820–5830. [DOI] [PubMed] [Google Scholar]
- (52).Eriksson M, and Nordén B (2001) Drug-Nucleic Acid Interactions, Vol. 340, Elsevier, Amsterdam, pp. 68–98. [DOI] [PubMed] [Google Scholar]
- (53).Munde M, Wang S, Kumar A, Stephens CE, Farahat AA, Boykin DW, Wilson WD, and Poon GMK (2014) Structure-dependent inhibition of the ETS-family transcription factor PU.1 by novel heterocyclic diamidines. Nucleic Acids Res. 42, 1379–1390. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (54).Wilson WD, Tanious FA, Mathis A, Tevis D, Hall JE, and Boykin DW (2008) Antiparasitic Compounds That Target DNA. Biochimie 90, 999–1014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (55).Laughlin S, Wang S, Kumar A, Farahat AA, Boykin DW, and Wilson WD (2015) Resolution of Mixed Site DNA Complexes with Dimer-Forming Minor-Groove Binders by Using Electrospray Ionization Mass Spectrometry: Compound Structure and DNA Sequence Effects. Chem. -Eur. J 21, 5528–5539. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (56).Antony-Debré I, Paul A, Leite J, Mitchell K, Kim HM, Carvajal LA, Todorova TI, Huang K, Kumar A, Farahat AA, Bartholdy B, Narayanagari S-R, Chen J, Ambesi-Impiombato A, Ferrando AA, Mantzaris I, Gavathiotis E, Verma A, Will B, Boykin DW, Wilson WD, Poon GMK, and Steidl U (2017) Pharmacological inhibition of the transcription factor PU.1 in leukemia. J. Clin. Invest 127, 4297–4313. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (57).Farahat AA, Guo P, Shoeib H, Paul A, Boykin DW, and Wilson WD (2019) Small Size Sequence-sensitive Compounds for Specific Recognition of G·C Base Pair in DNA Minor Groove. Chem. Eur. J 10.1002/chem.201904396 [DOI] [PMC free article] [PubMed] [Google Scholar]
- (58).Depauw S, Lambert M, Jambon S, Paul A, Peixoto P, Nhili R, Marongiu L, Figeac M, Dassi C, Paul-Constant C, Billoré B, Kumar A, Farahat AA, Ismail MA, Mineva E, Sweat DP, Stephens CE, Boykin DW, Wilson WD and David-Cordonnier M-H (2019) Heterocyclic Diamidine DNA Ligands as HOXA9 Transcription Factor Inhibitors: Design, Molecular Evaluation, and Cellular Consequences in a HOXA9-Dependant Leukemia Cell Model. J. Med. Chem 62, 1306–1329. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
