Next Article in Journal
Desulfonylative Functionalization of Organosulfones via Inert (Hetero)Aryl C(sp2)–SO2 Bond Cleavage
Previous Article in Journal
Investigation of the Structural Properties and Antioxidant Potency of Pectic Polysaccharides Derived from Rohdea japonica (Thunb.) Roth
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Facile Synthesis of Red-Shifted Bis-Quinoline (BisQ) Surrogate Base

School of Pharmacy, Institute for Drug Research, The Hebrew University of Jerusalem, Hadassah Ein-Kerem, Jerusalem 91120, Israel
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2024, 29(17), 4136; https://doi.org/10.3390/molecules29174136
Submission received: 1 August 2024 / Revised: 28 August 2024 / Accepted: 30 August 2024 / Published: 31 August 2024
(This article belongs to the Section Organic Chemistry)

Abstract

:
Forced intercalation peptide nucleic acids (FIT-PNAs) are DNA mimics that act as RNA sensors. The sensing event occurs due to sequence-specific RNA hybridization, leading to a substantial increase in fluorescence. The fluorophore in the FIT-PNA is termed a surrogate base. This molecule typically replaces a purine in the PNA sequence. BisQ is a surrogate base that connects two quinolines via a monomethine bond. BisQ-based FIT-PNAs have excellent biophysical features that include high brightness and red-shifted emission (λem, max = 613 nm). In this report, we detail two chemical approaches that allow for the facile synthesis of the BisQ PNA monomer. In both cases, the key compound used for the synthesis of BisQ-CH2COOH is the tBu-ester-modified quinoline synthon (compound 5). Subsequently, one method uses the Alloc acid-protected PNA backbone, whereas the other uses the tBu ester-protected PNA backbone. In the latter case, the overall yield for BisQ acid (compound 7) and BisQ PNA monomer syntheses was 61% in six synthetic steps. This is a substantial improvement to the published procedures to date (7% total yield). Lastly, we have prepared an 11-mer FIT-PNA with either BisQ or thiazole orange (TO) and studied their photophysical properties. We find superior photophysical properties for the BisQ FIT-PNA in terms of the brightness and selectivity, highlighting the added value of using this surrogate base for RNA sensing.

1. Introduction

Peptide nucleic acids (PNAs), DNA analogs developed in the early 90s by Nielsen and co-workers [1,2], have been used for the past three decades as potent antisense molecules for the down-regulation of RNA in vitro and in vivo [3,4,5,6,7,8,9,10,11,12]. In addition, in the early 2000s, PNAs were designed as RNA-sensing molecules by a concept devised by the Seitz group termed forced intercalation PNA (FIT-PNA) [13]. FIT-PNAs are designed by replacing one of the PNA monomers in a given PNA sequence with a fluorophore that is a mono-methine cyanine dye (also known as a surrogate base). In solution, the fluorescence of FIT-PNA, as a single strand, is quenched due to the free rotation of the monomethine bond [14]. However, once hybridized to its complementary RNA/DNA strand, the FIT-PNA fluoresces at the appropriate wavelength of the cyanine dye. The most used surrogate base is thiazole orange (TO). This fluorophore emits at (λem,max ≈ 535 nm). Other cyanine dyes used in FIT-PNAs include Benzothiazole Orange (BO), Oxazole Yellow (YO), and Oxazolopyridine (JO) [15,16]. All three fluorophores are blue-shifted in comparison to TO. BisQ, developed by our lab for FIT-PNAs [17] and by the Seitz lab for RNA/DNA-based FIT probes (QB, Quinoline Blue) [18], is a red-shifted surrogate base (λem,max ≈ 602 nm) that has high brightness and quantum yields [19,20,21]. In this report, we describe two alternative synthetic routes to prepare the BisQ PNA monomer; one leading to a 10-fold improvement in the overall yield. We then prepared a model 11-mer FIT-PNA with either TO or BisQ as the surrogate base and compared their sensing properties in terms of the brightness (BR), mismatch discrimination, and binding affinity to complementary RNA.

2. Results and Discussion

2.1. Synthesis of PNA Backbone

The aeg (aminoethyl glycine) PNA backbone is an achiral backbone that may be prepared by several synthetic routes [22,23]. The reported procedure for BisQ is based on eight synthetic steps where the PNA backbone is first coupled to BisQ acid, followed by the removal of both tBOC and tBu ester-protecting groups. The final step is the instillation of the Fmoc group on the primary amine. The overall yield for this synthetic route is ca. 7% [17]. Alternatively, Wickstrom and co-workers [24], reported an alternative route for PNA backbone synthesis (for TO) that is based on the use of the Alloc-protecting group on the C-terminal acid (Scheme 1, Compound 2). In the newly designed synthesis of BisQ, we have used either the Wickstrom route or a simpler route that is depicted in Scheme 2 that avoids the re-instillation of the Alloc-protecting group. Previously, the motivation for changing the protecting group on the carboxylic acid from tBu ester to Alloc was mainly due to difficulties in the purification of the final compound [24].

2.2. Synthesis of BisQ-CH2-COOH

Previously, BisQ-CH2-COOH was synthesized by using the free acid form as bromo acetic acid (BrCH2COOH). In our subsequent syntheses of BisQ-CH2-COOH, we observed low yields in this step, presumably due to the polymerization of the starting material (BrCH2COOH). As an alternative approach, we used the protected acid, namely, BrCH2COOtBu. Using this starting material significantly improved the synthesis of the BisQ acid by both methods. BrCH2COOtBu was added to 4-methyl quinoline (compound 5, Scheme 3) and the tBu ester was removed either prior to BisQ acid formation (compound 7, Scheme 3) or after (compound 8, Scheme 3). In both methods, high yields were obtained for all synthetic steps.

2.3. Synthesis of BisQ PNA Monomer

In the first synthetic route (Scheme 4), BisQ acid (compound 8) was directly coupled to the allyl-protected PNA monomer (compound 2, Scheme 4). After coupling, compound 9 was purified by normal phase column chromatography, and the purified material was subjected to Pd (Ph3)4 for the final removal of the Alloc group. The overall yield for this synthetic route (excluding the preparation of the backbone, compound 1, Scheme 1) was 23%.
In the second synthetic route (Scheme 3), compounds 5 and 6 were added to a DCM solution with triethylamine, affording compound 7 that was purified by normal phase column chromatography. After the tBu ester removal, the TFA salt of BisQ acid was directly coupled to the tBu ester-protected PNA backbone (compound 1, Scheme 1). Subsequently, compound 4 (Scheme 2) was purified by normal phase column chromatography, followed by the final removal of the tBu ester with TFA. The overall yield for this synthetic route (excluding the preparation of the PNA backbone, compound 1, Scheme 1) was 61%.

2.4. Synthesis and Biophysical Studies on a Model 11-Mer FIT-PNA

We selected an 11-mer FIT-PNA sequence (Table 1) in order to compare the biophysical properties of BisQ with a well-documented cyanine dye, TO. These FIT-PNAs were prepared with a short cationic peptide (dK4) as a means to improve their water solubility [25]. Both FIT-PNAs were prepared on the solid support as previously reported [17].
Fluorescence spectra were recorded for both FIT-PNAs before and after annealing to a fully complementary RNA (Figure 1). As opposed to BisQ FIT-PNA, the fluorescence enhancement for TO-FIT-PNA is only ca. 2-fold (I/I0 = 1.85, Table 1).
We next examined the sequence selectivity of both FIT-PNAs by introducing a single mismatch at the flanking base (5′) to the surrogate base (Figure 2 and Figure 3).
BisQ-FIT-PNA shows a robust selectivity for all mismatches tested. However, TO-FIT-PNA has no selectivity to these mismatches. Moreover, for the TU mismatch, the fluorescence of this duplex is higher by almost 2-fold in comparison to the fully matched duplex (Figure 3). One may speculate that the uridine base that is opposite and adjacent to TO does not quench TO fluorescence due to the lack of a primary amine on this base. This may be verified, in the future, by testing other TO-FIT-PNA sequences.
From Table 1, it is clear that BisQ-FIT-PNA is a brighter and more selective RNA probe. The brightness (BR) is ca. 4.3-fold higher for BisQ-FIT-PNA compared to TO-FIT-PNA. This is manifested by a lower extinction coefficient and a lower quantum yield. It is important to note that this comparison was studied for a single 11-mer FIT-PNA sequence and one may not conclude that these dramatic differences will be valid for any given FIT-PNA.
Given the broad use of FIT-PNAs for diagnostics in various diseases, including cancer [17,24,26], RNA editing [20], and infectious diseases [16,21,27], the ease of BisQ synthesis reported in this paper will allow research groups from a variety of fields to readily synthesize and explore BisQ FIT-PNAs and other BisQ-based dyes [28].

3. Materials and Methods

3.1. Chemicals and Reagents

RNA oligos were purchased from IDT (Coralville, IA, USA). Fmoc/Bhoc-protected PNA monomers were purchased from PolyOrg Inc. (Leominster, MA, USA). Fmoc-protected amino acids and reagents for solid-phase synthesis were purchased from Merck (Darmstadt, Germany), Chem-Impex (Wood Dale, IL, USA), and Bio-Lab LTD (Jerusalem, Israel). Tertbutyl bromoacetate and 4-methylquinoline were purchased from Thermo Fischer (Waltham MA, USA). 4-chloroquinoline was purchased from AABlocks (San Diego, CA, USA).

3.2. Chemical Synthesis of BisQ Monomer

Tert-butyl N-[2-(N-9-fluorenylmethoxycarbonyl) aminoethyl] glycinate hydrochloride (compound 1). Compound 1 was synthesized as previously reported [29]. 1H NMR (500 MHz, DMSO-d6) δ 9.36 (s, 2H), 7.89 (d, J = 7.5 Hz, 2H), 7.70 (d, J = 7.5 Hz, 2H), 7.59 (t, J = 5.8 Hz, 1H), 7.42 (td, J = 7.4, 1.1 Hz, 2H), 7.33 (td, J = 7.4, 1.1 Hz, 2H), 4.33 (d, J = 6.8 Hz, 2H), 4.22 (t, J = 6.9 Hz, 1H), 3.86 (s, 2H), 3.34 (q, J = 6.2 Hz, 2H), 3.00 (t, J = 6.4 Hz, 2H), 1.45 (s, 9H); 13C NMR (500 MHz, DMSO) δ 165.65, 156.27, 143.80, 140.74, 127.64, 127.08, 125.17, 120.14, 83.01, 65.63, 47.23, 46.65, 46.38, 36.59, 27.62; HRMS m/z calculated for C23H29N2O4 [M + H]+ 397.212; 397.211.
Compound 2 was synthesized as previously described [24].
Bis-quinoline-OtBu ester (compound 3). A combination of compound 5 (1.0 g, 3.87 mmol), compound 6 (691 mg, 3.87 mmol), and triethylamine (862 mg, 8.51 mmol) in 60 mL of dry DCM was stirred for 3 h at room temperature, resulting in the formation of a blue solution. After 3 h, 30 mL of water was added, and the aqueous layer was extracted with 3 × 30 mL of DCM. The organic solvent mixture was evaporated, dehydrated using anhydrous Na2SO4, filtered, and concentrated. The crude product was purified using silica gel column chromatography (Eluent 4–6% MeOH in DCM) to afford a blue compound 3 (1.36 g, 3.41 mmol, 85–90.0%). 1H NMR (300 MHz, cdcl3) δ 8.26–8.11 (m, 3H), 7.93 (d, J = 5.9 Hz, 1H), 7.72 (t, J = 7.6 Hz, 1H), 7.67–7.56 (m, 2H), 7.55–7.41 (m, 3H), 7.34 (d, J = 7.6 Hz, 1H), 7.20 (dd, J = 8.9, 2.9 Hz, 1H), 6.79 (s, 1H), 4.95 (s, 2H), 3.97 (s, 3H), 1.45 (s, 9H); 13C NMR (300 MHz, cdcl3) δ 166.1, 150.3, 148.2, 143.8, 142.8, 138.4, 138.1, 133.0, 132.4, 126.8, 126.0, 125.8–125.4 (m), 124.9, 117.1, 115.7, 110.7, 108.5, 97.9, 84.2, 55.5, 43.0, 28.1 ppm; HRMS m/z calculated for C26H27N2O2+ 399.206; 399.205.
Fmoc-Aeg-PNA-BisQ-tert-butyl-ester (4). Compound 3 (500 mg, 1.0 mmol) was dissolved in a mixture of 10 mL of TFA:DCM (1:1) at room temperature, and the reaction mixture was stirred for 2.5 h. A volume of 10 mL of toluene was added, and the solvent was co-evaporated under vacuum. Then, 10 mL of diethyl ether was added, and the solvent mixture was evaporated twice under vacuum. This resulted in the TFA salt of compound 3, which was used in the next steps without further purification.
TFA salt of compound 3 was dissolved in 10 mL of dry DMF. Then, HATU (381 mg, 1.0 mmol), HOBt (135 mg, 1.0 mmol), and DIPEA (149 mg, 1.15 mmol) were added. The mixture was stirred at room temperature for 10 min. This reaction mixture was slowly added to a mixture of the HCl salt of Fmoc-aeg-OtBu (compound 1, 338 mg, 0.85 mmol) and DIPEA (149 mg, 1.15 mmol) in dry DMF (5 mL). The reaction mixture was stirred under an inert atmosphere (argon) for 12 h. Once the reaction was finished (followed by TLC), 10 mL of water was added to the reaction mixture and extracted with 3 × 40 mL DCM. The combined organic layers were then washed with NaHCO3 (30 mL) and saturated brine (3 × 30 mL). The organic layers were dried using anhydrous Na2SO4, filtered, and then concentrated under vacuum. After obtaining the crude product, it was purified using silica gel chromatography (5–6% MeOH in DCM). This resulted in the formation of a blue sticky compound 4 (550 mg, 0.83 mmol, 77–83% yield). 1H NMR (500 MHz, DMSO-d6); δ 8.73 (d, J = 8.6 Hz, 1H), 8.65 (d, J = 8.6 Hz, 1H), 8.34 (d, J = 7.2 Hz, 1H), 8.01–7.94 (m, 1H), 7.89 (dt, J = 9.4, 4.7 Hz, 3H), 7.79–7.63 (m, 5H), 7.61–7.46 (m, 2H), 7.44–7.38 (m, 3H), 7.35–7.26 (m, 4H), 5.55 (s, 1H), 5.28 (s, 1H), 4.39 (d, J = 6.9 Hz, 1H), 4.35 (d, J = 6.5 Hz, 1H), 4.31 (d, J = 7.0 Hz, 1H), 4.23 (tt, J = 12.8, 7.4 Hz, 2H), 4.13 (s, 2H), 3.97 (s, 1H), 3.74 (s, 1H), 3.57 (s, 1H), 3.23 (d, J = 6.2 Hz, 1H), 3.15 (dd, J = 14.4, 5.2 Hz, 1H), 2.90 (d, J = 7.9 Hz, 1H), 1.45 (s, 6H), 1.38 (s, 3H); 13C NMR (500 MHz, DMSO) δ 168.9, 166.7, 166.3, 156.5, 156.3, 156.2, 150.0, 148.2, 144.1, 143.8, 143.3, 143.1, 140.7, 138.6, 132.9, 132.2, 128.9, 127.6, 127.3, 127.1, 126.6, 125.9, 125.6, 125.4, 125.1, 124.6, 121.4, 120.1, 117.9, 117.2, 109.7, 107.6, 96.9, 82.6, 82.4, 81.0, 65.5, 65.3, 54.4, 53.8, 49.9, 48.8, 48.0, 47.3, 46.9, 42.0, 37.3, 27.7 ppm. HRMS m/z calculated for C45H45N4O5+ 721.338; 721.337.
1-[2-(1,1-Dimethylethoxy)-2-oxoethyl]-4-methyl quinolinium (compound 5). 4-methylquinoline (715.95 mg, 5 mmol) and tert-butyl bromoacetate (975.3 mg, 5 mmol) in 12.5 mL of acetonitrile were stirred for 24 h at 50 °C. The solvent was evaporated, and the brownish residue was recrystallized in acetone and vacuum dried. The solid was further washed with acetone affording compound 5 as a white solid (1.226 g, 4.75 mmol, 95%). 1H NMR (500 MHz, Deuterium Oxide) δ 8.96 (d, J = 6.1 Hz, 1H), 8.34 (dd, J = 8.5, 1.3 Hz, 1H), 8.13–8.02 (m, 2H), 7.90–7.83 (m, 2H), 5.71 (s, 2H), 2.91 (d, J = 0.8 Hz, 3H), 2.12 (s, 1H), 1.35 (s, 9H). 13C NMR (500 MHz, Deuterium Oxide) δ 166.09, 161.62, 148.63, 137.59, 135.92, 129.84, 129.11, 127.09, 122.56, 117.95, 86.36, 58.29, 27.02, 19.85. Calculated M.W. for C16H20NO2+ 258.14886; M.W. 258.14814.
1-Methyl-chloroquinolinium iodide (compound 6). Compound 6 was synthesized as previously described [17], with a 92% yield.
1-Carboxymethyl-4-methylquinolinium (compound 7). Compound 5 (645.3 mg, 2.5 mmol) and HCl (22 mL, 4M in 1,4-dioxane) were stirred for 3 h at room temperature. The solvent mixture was evaporated under reduced pressure and the resulting solid was washed with diethyl ether and collected by filtration, affording compound 7 (483 mg, 2.39 mmol, 92–96%). 1H NMR (500 MHz, Deuterium Oxide) δ 8.92 (d, J = 6.1 Hz, 1H), 8.29–8.23 (m, 1H), 8.07–7.97 (m, 2H), 7.83–7.75 (m, 2H), 5.70 (s, 2H), 2.85 (d, J = 0.8 Hz, 3H). 13C NMR (500 MHz, Deuterium Oxide) δ 169.31, 161.33, 148.51, 137.57, 135.71, 129.71, 129.09, 126.90, 122.49, 118.06, 57.91, 19.73. Calculated M.W. for C12H12NO2+ 202.08626; M.W. 202.08521.
4-[[1-Carboxymethyl-4(1H)-quinolinylidene] methyl]-1methyl quinolinium (compound 8). Compound 8 was synthesized as previously described [17].
Compound 9 was synthesized as previously described, by utilizing Compound 2 and 8 [26].
Final BisQ monomer (via allyl ester deprotection). This reaction was synthesized as previously described [27].
Final BisQ monomer (via tBu ester deprotection). Compound 4 (300 mg, 0.42 mmol) was dissolved in a 12 mL mixture of TFA and DCM in a 1:1 ratio. The solution was stirred for 3 h at room temperature. The solvent was then co-evaporated with toluene (2 × 12 mL) and further with diethyl ether (2 × 12 mL) to obtain a dark blue solid (260 mg, 0.39 mmol, 90–94% yield). 1H NMR (500 MHz, DMSO-d6) δ 8.90 (s, 1H), 8.74 (dd, J = 8.6, 4.3 Hz, 1H), 8.66 (d, J = 8.6 Hz, 1H), 8.34 (dd, J = 7.2, 5.2 Hz, 1H), 8.01–7.94 (m, 1H), 7.90 (dd, J = 7.6, 6.2 Hz, 3H), 7.78–7.66 (m, 4H), 7.61–7.53 (m, 2H), 7.51–7.39 (m, 4H), 7.37–7.26 (m, 4H), 5.55 (s, 1H), 5.33 (s, 1H), 4.40–4.35 (m, 2H), 4.31 (d, J = 7.0 Hz, 1H), 4.23 (dq, J = 14.5, 6.8 Hz, 1H), 4.14 (d, J = 2.5 Hz, 2H), 4.03 (s, 1H), 3.90 (s, 1H), 3.59 (d, J = 6.7 Hz, 1H), 3.44–3.36 (m, 2H), 3.18–3.12 (m, 1H), 3.02 (td, J = 6.7, 4.4 Hz, 1H); 13C NMR (500 MHz, DMSO-d6): major rotamer: δ 171.2, 170.3, 168.2, 166.7, 166.3, 156.4, 149.9, 148.3, 144.1, 143.8, 143.4, 143.1, 140.8, 138.6, 132.9, 129.5, 127.7, 127.1, 126.6, 125.9, 125.6, 125.4, 125.1, 124.7, 120.2, 117.9, 109.7, 109.6, 107.6, 96.9, 65.6, 47.0, 46.7, 42.1, 36.6; minor rotamer: 156.6, 156.2, 150.0, 148.3, 144.1, 143.9, 140.7, 138.7, 132.2, 129.9, 127.9,117.2, 46.7, 46.5, 42.0 ppm. HRMS calculated for C41H37N4O5+ 665.275; 665.273.

3.3. FIT-PNA Synthesis

Manual solid-phase synthesis was performed in 5 mL polyethylene syringe reactors (Phenomenex, Torrance, CA, USA) that were equipped with a fritted disk. HPLC purifications and analysis were performed on a Dionex UltiMate 3000 HPLC system (ThermoFisher Scientific, Waltham, MA, USA) using a semi-preparative C18 reversed-phase column (Jupiter C18, 5 μm, 300 Å, 250 × 10 μm, Phenomenex, Torrance, CA, USA). Eluents A (0.1% TFA in water) and B (Acetonitrile) were used in a linear gradient with a flow rate of 4 mL/min. Mass analysis of FIT-PNAs was performed by Maldi-TOF MS (Microflex, silver edition, Brucker, Billerica, MA, USA).

3.4. Tm Analysis

Melting curves (Tm) of the FIT-PNA/RNA duplexes were estimated from UV melting curves measured on an Evolution One Plus UV-Vis Spectrophotometer (Thermo Fisher Scientific, Waltham, MA, USA). FIT-PNAs and their complementary RNAs (1:1.5 ratio) were prepared in PBS buffer (pH 7.0) and adjusted to a final duplex concentration of 2.5 µM. Prior to analysis, the samples were heated from 20 °C to 90 °C at a rate of 5 °C/min and then cooled to the starting temperature at a rate of 2 °C/min. The change in absorbance was monitored at 260 nm by increasing the temperature to 90 °C at a rate of 1 °C/min.

3.5. UV-Vis Titrations and Fluorescence Measurements for Determining Quantum Yields

The measurement of the fluorescence quantum yield of FIT PNA:RNA duplexes followed the same protocol as previously reported [30]. The absorbance range of the FIT-PNA duplex determined the choice of reference dye. For example, cresyl violet was used as a reference dye for the BisQ PNA/RNA (1:1) duplex, while fluorescein was employed as a reference dye for the TO PNA/RNA (1:1) duplex.
A serial dilution of FIT-PNA/RNA duplex solutions with an OD smaller than 0.1 at the maximum wavelength was measured to verify that the absorbance maxima remained constant. FIT-PNA duplexes with similar heights of absorbance were used to determine the cut point with respect to the reference dye, and this information was used to gather the fluorescence emission spectra (see Supplementary Materials, Figures S9–S12).
The fluorescence quantum yield was calculated by considering the refractive index (RI) value of the solvents, using the integrated emission value and the OD value at the absorbance maxima of the identical samples. To mitigate the impacts of the inner filter, we employed an optical density (OD) value below 0.1. The quantum yield was calculated using the equation, as detailed in the literature [25].

3.6. Fluorescence Measurements

Fluorescence spectra of FIT-PNAs with/out synthetic RNA were measured using a Jasco FP-6500 Spectrofluorometer (Jasco Inc. Co., Ltd, Tokyo, Japan) with a band width set to 2.5 nm and a response rate of 0.2 s. For BisQ FIT-PNA, excitation was set to 580 nm and the emission spectra were recorded from 588 nm to 750 nm. For TO FIT-PNA (excitation at 499 nm), emission was recorded between 509 and 750 nm.
Solutions of FIT-PNA/FIT-PNAs (1 μM) and synthetic RNA (1 μM) were prepared in 1× PBS. The hybridization of FIT-PNA/FIT-PNAs with synthetic RNA was performed by annealing at 65 °C for 2 min followed by incubation at 37 °C for 2–3 h at a ratio of 1:1. The RNA sequence used was 5′-UAUGUAUGUUG. RNAs with mismatches were TU mm: 5′-UAUGUUUGUUG; TC mm: 5′-UAUGUCUGUUG; TG mm: 5′-UAUGUGUGUUG.

4. Conclusions

In this report, we detail two improved synthetic routes for the synthesis of the BisQ monomer, a red-shifted surrogate base used in the design of RNA/DNA sensors based on the FIT-PNA approach [13]. In comparison to published protocols, one of these synthetic routes that uses the tBu ester aeg-PNA backbone (compound 7) results in an overall yield of 61.4% (compared to 7%). In addition, a direct comparison of a model FIT-PNA with either BisQ or TO highlights the superior biophysical properties of BisQ FIT-PNA in terms of the brightness, quantum yield, and sequence selectivity.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules29174136/s1. Figures S1–S6: HPLC chromatograms and Maldi-TOF MS for FIT-PNAs; Figures S7 and S8: Tm profiles for FIT-PNA/RNA duplexes; Figures S9–S12: UV-Vis spectra for determining ε values for Fit-PNA/RNA duplexes; Figures S13–S18: 1H NMR and 13C NMR of synthesized compounds.

Author Contributions

Conceptualization, H.N. and M.K.G.; methodology, H.N., A.F. and M.K.G.; writing—original draft preparation, E.Y.; writing—review and editing, H.N. and M.K.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Len & Susan Mark Initiative for Ovarian and Uterine/MMMT Cancers—Phase III grant from the Israel Cancer Research Fund (grant No. 21-305-MI) and by the Israel Science Foundation (grant No. 572/21). H.N. is grateful for the financial support from the Neubauer Fellowship for Ph.D. students.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are presented within the submitted manuscript.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of the data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Egholm, M.; Buchardt, O.; Christensen, L.; Behrens, C.; Freier, S.M.; Driver, D.A.; Berg, R.H.; Kim, S.K.; Norden, B.; Nielsen, P.E. PNA hybridizes to complementary oligonucleotides obeying the Watson-Crick hydrogen-bonding rules. Nature 1993, 365, 566–568. [Google Scholar] [CrossRef]
  2. Nielsen, P.E.; Egholm, M.; Berg, R.H.; Buchardt, O. Sequence-selective recognition of DNA by strand displacement with a thymine-substituted polyamide. Science 1991, 254, 1497–1500. [Google Scholar] [CrossRef]
  3. Brodyagin, N.; Katkevics, M.; Kotikam, V.; Ryan, C.A.; Rozners, E. Chemical approaches to discover the full potential of peptide nucleic acids in biomedical applications. Beilstein J. Org. Chem. 2021, 17, 1641–1688. [Google Scholar] [CrossRef] [PubMed]
  4. Gupta, A.; Mishra, A.; Puri, N. Peptide nucleic acids: Advanced tools for biomedical applications. J. Biotechnol. 2017, 259, 148–159. [Google Scholar] [CrossRef] [PubMed]
  5. MacLelland, V.; Kravitz, M.; Gupta, A. Therapeutic and diagnostic applications of antisense peptide nucleic acids. Mol. Ther. Nucleic Acids 2024, 35, 102086. [Google Scholar] [CrossRef] [PubMed]
  6. Montazersaheb, S.; Hejazi, M.S.; Nozad Charoudeh, H. Potential of Peptide Nucleic Acids in Future Therapeutic Applications. Adv. Pharm. Bull. 2018, 8, 551–563. [Google Scholar] [CrossRef]
  7. Papi, C.; Gasparello, J.; Zurlo, M.; Manicardi, A.; Corradini, R.; Cabrini, G.; Gambari, R.; Finotti, A. Combined Treatment of Bronchial Epithelial Calu-3 Cells with Peptide Nucleic Acids Targeting miR-145-5p and miR-101-3p: Synergistic Enhancement of the Expression of the Cystic Fibrosis Transmembrane Conductance Regulator (CFTR) Gene. Int. J. Mol. Sci. 2022, 23, 9348. [Google Scholar] [CrossRef]
  8. Popella, L.; Jung, J.; Do, P.T.; Hayward, R.J.; Barquist, L.; Vogel, J. Comprehensive analysis of PNA-based antisense antibiotics targeting various essential genes in uropathogenic Escherichia coli. Nucleic Acids Res. 2022, 50, 6435–6452. [Google Scholar] [CrossRef] [PubMed]
  9. Pradeep, S.P.; Malik, S.; Slack, F.J.; Bahal, R. Unlocking the potential of chemically modified peptide nucleic acids for RNA-based therapeutics. RNA 2023, 29, 434–445. [Google Scholar] [CrossRef]
  10. Saarbach, J.; Sabale, P.M.; Winssinger, N. Peptide nucleic acid (PNA) and its applications in chemical biology, diagnostics, and therapeutics. Curr. Opin. Chem. Biol. 2019, 52, 112–124. [Google Scholar] [CrossRef]
  11. Tsylents, U.; Siekierska, I.; Trylska, J. Peptide nucleic acid conjugates and their antimicrobial applications—A mini-review. Eur. Biophys. J. 2023, 52, 533–544. [Google Scholar] [CrossRef] [PubMed]
  12. Zheng, H.; Clausse, V.; Amarasekara, H.; Mazur, S.J.; Botos, I.; Appella, D.H. Variation of Tetrahydrofurans in Thyclotides Enhances Oligonucleotide Binding and Cellular Uptake of Peptide Nucleic Acids. JACS Au 2023, 3, 1952–1964. [Google Scholar] [CrossRef] [PubMed]
  13. Koehler, O.; Jarikote, D.V.; Seitz, O. Forced intercalation probes (FIT probes): Thiazole orange as a fluorescent base in peptide nucleic acids for homogeneous single-nucleotide-polymorphism detection. ChemBioChem 2005, 6, 69–77. [Google Scholar] [CrossRef]
  14. Karunakaran, V.; Pérez Lustres, J.L.; Zhao, L.; Ernsting, N.P.; Seitz, O. Large dynamic Stokes shift of DNA intercalation dye Thiazole Orange has contribution from a high-frequency mode. J. Am. Chem. Soc. 2006, 128, 2954–2962. [Google Scholar] [CrossRef] [PubMed]
  15. Bethge, L.; Jarikote, D.V.; Seitz, O. New cyanine dyes as base surrogates in PNA: Forced intercalation probes (FIT-probes) for homogeneous SNP detection. Bioorganic Med. Chem. 2008, 16, 114–125. [Google Scholar] [CrossRef] [PubMed]
  16. Kummer, S.; Knoll, A.; Socher, E.; Bethge, L.; Herrmann, A.; Seitz, O. PNA FIT-Probes for the Dual Color Imaging of Two Viral mRNA Targets in Influenza H1N1 Infected Live Cells. Bioconjugate Chem. 2012, 23, 2051–2060. [Google Scholar] [CrossRef] [PubMed]
  17. Kolevzon, N.; Hashoul, D.; Naik, S.; Rubinstein, A.; Yavin, E. Single point mutation detection in living cancer cells by far-red emitting PNA-FIT probes. Chem. Commun. 2016, 52, 2405–2407. [Google Scholar] [CrossRef]
  18. Hovelmann, F.; Seitz, O. DNA Stains as Surrogate Nucleobases in Fluorogenic Hybridization Probes. Acc. Chem. Res. 2016, 49, 714–723. [Google Scholar] [CrossRef]
  19. Hoevelmann, F.; Gaspar, I.; Chamiolo, J.; Kasper, M.; Steffen, J.; Ephrussi, A.; Seitz, O. LNA-enhanced DNA FIT-probes for multicolour RNA imaging. Chem. Sci. 2016, 7, 128–135. [Google Scholar] [CrossRef]
  20. Fang, G.-m.; Chamiolo, J.; Kankowski, S.; Hövelmann, F.; Friedrich, D.; Löwer, A.; Meier, J.C.; Seitz, O. A bright FIT-PNA hybridization probe for the hybridization state specific analysis of a C → U RNA edit via FRET in a binary system. Chem. Sci. 2018, 9, 4794–4800. [Google Scholar] [CrossRef]
  21. Tepper, O.; Peled, I.; Fastman, Y.; Heinberg, A.; Mitesser, V.; Dzikowski, R.; Yavin, E. FIT-PNAs as RNA-Sensing Probes for Drug-Resistant Plasmodium falciparum. ACS Sens. 2022, 7, 50–59. [Google Scholar] [CrossRef]
  22. Stafforst, T.; Diederichsen, U. Solid-phase synthesis of acid-sensitive N-(2-aminoethyl)glycine-PNA oligomers by the Fmoc/Bhoc strategy. Eur. J. Org. Chem. 2007, 2007, 681–688. [Google Scholar] [CrossRef]
  23. Wojciechowski, F.; Hudson, R.H.E. A convenient route to N-[2-(Fmoc)aminoethyl]glycine esters and PNA oligomerization using a bis-N-boc nucleobase protecting group strategy. J. Org. Chem. 2008, 73, 3807–3816. [Google Scholar] [CrossRef]
  24. Sonar, M.V.; Wampole, M.E.; Jin, Y.-Y.; Chen, C.-P.; Thakur, M.L.; Wickstrom, E. Fluorescence Detection of KRAS2 mRNA Hybridization in Lung Cancer Cells with PNA-Peptides Containing an Internal Thiazole Orange. Bioconjugate Chem. 2014, 25, 1697–1708. [Google Scholar] [CrossRef]
  25. Tepper, O.; Zheng, H.C.; Appella, D.H.; Yavin, E. Cyclopentane FIT-PNAs: Bright RNA sensors. Chem. Commun. 2021, 57, 540–543, Erratum in: Chem. Commun. 2023, 59, 11593. [Google Scholar] [CrossRef] [PubMed]
  26. Hashoul, D.; Shapira, R.; Falchenko, M.; Tepper, O.; Paviov, V.; Nissan, A.; Yavin, E. Red-emitting FIT-PNAs: “On site” detection of RNA biomarkers in fresh human cancer tissues. Biosens. Bioelectron. 2019, 137, 271–278. [Google Scholar] [CrossRef]
  27. Sato, Y.; Miura, H.; Tanabe, T.; Okeke, C.U.; Kikuchi, A.; Nishizawa, S. Fluorescence Sensing of the Panhandle Structure of the Influenza A Virus RNA Promoter by Thiazole Orange Base Surrogate-Carrying Peptide Nucleic Acid Conjugated with Small Molecule. Anal. Chem. 2022, 94, 7814–7822. [Google Scholar] [CrossRef] [PubMed]
  28. Insuasty, D.; Cabrera, L.; Ortiz, A.; Insuasty, B.; Quiroga, J.; Abonia, R. Synthesis, photophysical properties and theoretical studies of new bis-quinolin curcuminoid BF2-complexes and their decomplexed derivatives. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2020, 230, 118065. [Google Scholar] [CrossRef] [PubMed]
  29. Jain, P.; Badger, D.B.; Liang, Y.; Gebhard, A.W.; Santiago, D.; Murray, P.; Kaulagari, S.R.; Gauthier, T.J.; Nair, R.; Kumar, M.; et al. Bioactivity improvement via display of the hydrophobic core of HYD1 in a cyclic β-hairpin-like scaffold, MTI-101. Pept. Sci. 2021, 113, e24199. [Google Scholar] [CrossRef]
  30. Würth, C.; Grabolle, M.; Pauli, J.; Spieles, M.; Resch-Genger, U. Relative and absolute determination of fluorescence quantum yields of transparent samples. Nat. Protoc. 2013, 8, 1535–1550. [Google Scholar] [CrossRef]
Scheme 1. Chemical synthesis of the Allyl–protected PNA backbone as reported by Wickstrom and co-workers [24].
Scheme 1. Chemical synthesis of the Allyl–protected PNA backbone as reported by Wickstrom and co-workers [24].
Molecules 29 04136 sch001
Scheme 2. New route for the synthesis of BisQ monomer using the tert–butyl protected PNA backbone (compound 4, Scheme 1).
Scheme 2. New route for the synthesis of BisQ monomer using the tert–butyl protected PNA backbone (compound 4, Scheme 1).
Molecules 29 04136 sch002
Scheme 3. Synthesis of BisQ–COOH (compound 8) or its tert–butyl ester (compound 3) via two synthetic routes using the tButyl–protected 4–methyl quinoline (compound 5).
Scheme 3. Synthesis of BisQ–COOH (compound 8) or its tert–butyl ester (compound 3) via two synthetic routes using the tButyl–protected 4–methyl quinoline (compound 5).
Molecules 29 04136 sch003
Scheme 4. BisQ monomer synthesis via the allyl–protected PNA backbone.
Scheme 4. BisQ monomer synthesis via the allyl–protected PNA backbone.
Molecules 29 04136 sch004
Figure 1. Fluorescence emission spectrum (normalized values) for BisQ-FIT-PNA (left) and TO-FIT-PNAs (right) towards synthetic RNA. For BisQ-FIT-PNA, λex = 580 nm; λem = 588 nm. For TO-FIT-PNA, λex = 499 nm; λem = 509 nm. [TO-FIT-PNA] = [RNA] = 0.5 μM and [BisQ-FIT-PNA] = [RNA] = 1 μM.
Figure 1. Fluorescence emission spectrum (normalized values) for BisQ-FIT-PNA (left) and TO-FIT-PNAs (right) towards synthetic RNA. For BisQ-FIT-PNA, λex = 580 nm; λem = 588 nm. For TO-FIT-PNA, λex = 499 nm; λem = 509 nm. [TO-FIT-PNA] = [RNA] = 0.5 μM and [BisQ-FIT-PNA] = [RNA] = 1 μM.
Molecules 29 04136 g001
Figure 2. Mismatch discrimination for BisQ-FIT-PNA. FIT-PNA was hybridized to fully complementary RNA (perfect match) or to RNA with a single mismatch (TG, TC, and TU). (Right)—Full fluorescence spectrum. (Left)—Maximum fluorescence for each duplex (SD based on 3 repeats). [BisQ-FIT-PNA] = [RNA] = 1 μM.
Figure 2. Mismatch discrimination for BisQ-FIT-PNA. FIT-PNA was hybridized to fully complementary RNA (perfect match) or to RNA with a single mismatch (TG, TC, and TU). (Right)—Full fluorescence spectrum. (Left)—Maximum fluorescence for each duplex (SD based on 3 repeats). [BisQ-FIT-PNA] = [RNA] = 1 μM.
Molecules 29 04136 g002
Figure 3. Mismatch discrimination for TO-FIT-PNA. FIT-PNA was hybridized to fully complementary RNA (perfect match) or to RNA with a single mismatch (TG, TC, and TU). (Right)—Full fluorescence spectrum. (Left)—Maximum fluorescence for each duplex (SD based on 3 repeats). [TO-FIT-PNA] = [RNA] = 0.5 μM.
Figure 3. Mismatch discrimination for TO-FIT-PNA. FIT-PNA was hybridized to fully complementary RNA (perfect match) or to RNA with a single mismatch (TG, TC, and TU). (Right)—Full fluorescence spectrum. (Left)—Maximum fluorescence for each duplex (SD based on 3 repeats). [TO-FIT-PNA] = [RNA] = 0.5 μM.
Molecules 29 04136 g003
Table 1. The photophysical properties and binding affinities of these FIT-PNAs are summarized.
Table 1. The photophysical properties and binding affinities of these FIT-PNAs are summarized.
FIT-PNA Duplex (1:1)λmaxƐ [mM−1 cm−1]Ex. (nm)ɸF [%]BR [mM−1 cm−1]I/I0Tm [oC]
11-mer BisQ PNA/RNA58753.625809.3352445.2 (±0.2)
11-mer TO PNA/RNA51621.734995.411.171.8549 (±0.2)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Nazzal, H.; Gupta, M.K.; Fadila, A.; Yavin, E. A Facile Synthesis of Red-Shifted Bis-Quinoline (BisQ) Surrogate Base. Molecules 2024, 29, 4136. https://doi.org/10.3390/molecules29174136

AMA Style

Nazzal H, Gupta MK, Fadila A, Yavin E. A Facile Synthesis of Red-Shifted Bis-Quinoline (BisQ) Surrogate Base. Molecules. 2024; 29(17):4136. https://doi.org/10.3390/molecules29174136

Chicago/Turabian Style

Nazzal, Huda, Manoj Kumar Gupta, Amer Fadila, and Eylon Yavin. 2024. "A Facile Synthesis of Red-Shifted Bis-Quinoline (BisQ) Surrogate Base" Molecules 29, no. 17: 4136. https://doi.org/10.3390/molecules29174136

Article Metrics

Back to TopTop