Abstract
We have developed a simple method to synthesize 6-seleno-2′-deoxyguanosine (SedG) by selectively replacing the 6-oxygen atom with selenium. This selenium-atom-specific modification (SAM) alters the optical properties of the naturally occurring 2′-deoxyguanosine (dG). Unlike the native dG, the UVabsorption of SedG is significantly influenced by the pH of the aqueous solution. Moreover, SedG is fluorescent at the physiological pH and exhibits pH-dependent fluorescence in aqueous solutions. Furthermore, SedG has noticeable fluorescence in non-aqueous solutions, indicating its sensitivity to environmental changes. This is the first time a fluorescent nucleoside by single-atom alteration has been observed. Fluorescent nucleosides modified by a single atom have great potential as molecular probes with minimal perturbations to investigate nucleoside interactions with proteins, such as membrane-transporter proteins.
Keywords: selenium derivatization and fluorescence, nucleoside with single-atom modification, fluorescent DNA, RNA, nucleic acid probes
1 Introduction
Selenium derivation of nucleic acids opens a new paradigm for nucleic acids and related research, especially probing the structures and biological functions of nucleic acids and their modifications [1–12]. Se-modifications of nucleic acids can lead to novel materials and therapeutics with unique combinations of biochemical and physiochemical properties [12–15]. Selenium replacement of oxygen in the nucleobases further stimulates the investigation of structure-and-function relationships such as fluorescence, nucleobaseparing, duplex stability, structure flexibility, replication efficiency and fidelity, and polymerase recognition of the Se-modified nucleobases.
Natural nucleic acids are colorless and non-fluorescent under physiological conditions [16]. However, strong acid solution (with a pH as low as 1.0) can induce weak fluorescence in native nucleobases, such as adenine and guanine [17–19]. Such acidic conditions are not suited for probing nucleic acids and their interactions with other molecules. Numerous efforts have been made to develop fluorescent nucleosides [20–28]. Various nucleoside analogs containing extended-fluorescent nucleobases and attached fluorescent labels have been designed and used to study the native nucleic acids and their interactions with proteins. Unfortunately, the modifications are normally bulky and often result in fluorescent nucleobases that are colorless. These modifications can cause perturbations and are not ideal for probing nucleic acid structure and function. However, it is very challenging to design and discover minimally modified nucleosides that are colored and exhibit significant fluorescence.
Our previous report suggests that the 6-Se-modification on guanosine causes minimal perturbation in the duplex structure [29]. The synthesis and antitumor activities of 6-Se-2′-deoxyguanosine and analogues have been reported earlier [30–33]. Herein we report a simple synthesis and optical studies of SedG (3 in Scheme 1), achieved with our SAM strategy in which the exo-oxygen at C-6 of guanosine is replaced with selenium. The synthesized SedG is bright yellow and exhibits a strong absorption at 357 nm in water (Figure 1(a)). We also found that its UV-absorption maximum is strongly dependent on the solution pH and the polarity of the solvent. Moreover, to our surprise, we observed that SedG exhibits pH-dependent fluorescence; maximum fluorescence was observed in solution with pH 6 (λex = 305 nm; λem = 390 nm). The fluorescent behavior is appreciably influenced by solvent polarity. A noticeable fluorescence was observed in isopropanol (λem = 320 nm; λem =395 nm). With minimal perturbation, this Se-nucleoside may serve as a valuable visual and optical probe for exploring the structure and dynamics of nucleic acids and their complexes with proteins and small molecules.
Scheme 1.

Synthesis of 6-seleno-2′-deoxyguanosine. Reaction conditions: (i) 80% acetic acid, 25 °C, 1 h, 91%; (ii) 0.05 M K2CO3 in methanol, 25 °C, 2 h, 42%.
Figure 1.

(A) Absorption spectra of dG (dotted line), N2-tBPAc-6-CE-Se-dG (broken line), and SedG (solid line) in water at 25 °C; inset: SedG or 6SedG (left), yellow; dG (right), colorless. (B) RP-HPLC analysis of dG and SedG [monitored at 260 nm (red) and 360 nm (black)], (a) and (b): commercial dG (retention time: 13.9 min); (c) and (d): synthesized SedG (retention time: 14.8 min); (e) and (f): co-injection of commercial dG and synthesized SedG (retention times: 13.9 and 14.8 min, respectively).
In particular, this fluorescent Se-nucleoside would be very useful in studying the structure and dynamics of trans-membrane transporter proteins. The analysis of structure and function of transporter proteins has recently drawn widespread attention from multidisciplinary research teams [34–37]. This information on the nucleoside-protein interaction is crucial for understanding the molecular basis of the action of drugs that target the transporter proteins, which are important drug targets [38]. A variety of biochemical and biophysical techniques have been applied to elucidate the mechanisms of active transportion in biological membranes at the molecular level [39–42]. With a fluorescence emission at 390 nm, which is red-shifted by approximately 50 nm from the emission range of most proteins, the SedG nucleoside and its nucleotides could also present a valuable chromophore or fluorescent ligand to investigate proteinnucleic acid interactions and further contribute to molecular understanding of the protein dynamics [43].
2 Experimental
2.1 Materials and methods
The synthesis was carried out using commercially available reagents. The solid reagents were dried under vacuum and the reactions were performed under argon. The solvents were purged with argon before use. Water purified by ion exchange to a resistivity of 18.2 MΩ was used to prepare all buffers. Solvent mixtures are indicated as volume/volume ratios. Analytical thin layer chromatography (TLC, Dynamic Adsorbent) was performed using Merck Whatman 60 F254 plates (0.25 mm thick), and visualized under UV light. A Ce-Mo staining solution (25 g phosphomolybdate, 10 g Ce(SO4)2 • 4H2O, 60 mL conc. H2SO4, 940 mL H2O) was used to monitor the detritylation reaction. Column chromatography was performed using Fluka silica gel (mesh size 0.040–0.063 mm). NMR spectra were recorded on a Bruker 400 XWIN spectrometer. Chemical shift values are in ppm. High-resolution MS was determined by electrospray mass spectrometry.
2.2 Synthesis of 6-seleno-2′-deoxyguanosine, SedG
N2-[2-(4-tert-Butylphenoxy)acetyl]-6-(2-cyanoethyl) seleno-2′-deoxyguanosine (2)
Compound 1 (160 mg, 0.183 mmol) synthesized by previously published procedures [29] was dissolved in acetic acid (80%, 1 mL) and stirred at 25 °C for 1 h to deprotect the 5′-DMTr group and yield 2. Compound 2 was then purified by a silica gel column equilibrated with methylene chloride and eluted with a step-wise gradient of methanol/methylene chloride mixtures (1.0%, 2.0%, 3.0%, 4.0% MeOH in CH2Cl2, 200 mL each fraction) to afford 2 (95 mg, 91%; Figures S1–S3). HRMS (ESI-TOF): C25H30N6O5Se; [M + H]+: 575.1508 (calc. 575.1516), [M + Na]+: 597.1351 (calc. 597.1335); 1H NMR (400 MHz, CD3SOCD3) δ: 1.25 (s, 9H, 3 × CH3-tBu), 2.30–2.35 and 2.69–2.76 (2 × m, J1′–2′ = 6.4 Hz, 2H, H-2′), 3.19 (t, J = 6.4 Hz, 2H, Se–CH2–CH2-CN), 3.53–3.62 (m, 2H, Se–CH2–CH2–CN), 3.53–3.62 (m, 2H) and 3.87–3.88 (br, 1H) (H-3′ and H-5′), 4.43 (br, 1H, H-4′), 4.95 (s, 2H, CH2–O), 6.35 (t, J1′–2′ = 6.4 Hz, 1H, H-1′), 6.86–6.88 (d, J = 8.64 Hz, 2H, CH-arom), 7.30–7.32 (d, J = 8.72 Hz, 2H, CH-arom), 8.59 (s, 1H, H-8), 10.82 (s, 1H, NH); 13C NMR (100 MHz, CD3SOCD3) δ: 18.37 and 18.89 (Se–CH2–CH2–CN), 31.22 (CH3–tBu), 33.68 (C–tBu), 61.35 (C-5′), 67.12 (CH2–O), 70.39 (C-3′), 83.31 (C-4′), 87.85 (C-1′), 113.83 and 125.99 (CH-arom), 119.84 (CN), 130.25 (C-5), 142.51 (C-8), 143.03 and 157.24 (C-arom), 148.46 (C-6), 151.57 (C-4), 155.53 (C-2), 167.23 (C=O).
6-Seleno-2′-deoxyguano sine (3)
Compound 2 (10 mg, 0.017 mmol) was treated with 0.05 M K2CO3 (in methanol) at 25 °C for 2 h to remove the cyanoethyl and (4-tert-butyl-phenoxy)acetyl protecting groups. The crude product was neutralized with HCl and purified by reverse phase HPLC. 2 was eluted at 355 nm using a combination of two solvents at varying proportions: buffer A (2.5 mM TEAAc in water) and buffer B (2.5 mM TEAAc, 50% water and 50% acetonitrile). The fractions were lyophilized to afford bright yellow 6-seleno-2′-deoxyguanosine (3, 2.5 mg, 42%; Figures S4–S6). HRMS (ESI-TOF): C10H13N5O3Se; [M–H]−: 330.0103 (calc. 330.0111); 1H NMR (400 MHz, CD3OD) δ: 2.34–2.40 and 2.62–2.71 (2× m, J1′–2′ = 6.4 Hz, 2H, H-2′), 3.70–3.80 (quat. of d, 2H) and 3.98 (br, 1H, H-3′ and H-5′), 4.51 (br, 1H, H-4′), 6.26 (t, J1′–2′ = 6.4 Hz, 1H, H-1′), 8.23 (s, 1H, NH); 13C NMR (100 MHz, CD3OD) δ: 41.38 (C-2′), 63.28 (C-5′), 72.67 (C-3′), 85.68 (C-4′), 89.39 (C-1′), 120.91 (C-5), 133.90 (C-8), 141.37 (C-4), 147.55 (C-2), 174.51 (C-6).
2.3 UV-Vis and fluorescence measurements
UV-Vis spectra were recorded on a Varian Cary-100 Bio (UV/Vis Model 240) spectrophotometer with respect to pure solvent/buffer reference. Fluorescence measurements were performed on a Perkin-Elmer LS55 fluorescence spectrometer. The excitation and emission slit widths were fixed at 10 nm. All the solution measurements were made in 10 mM sodium phosphate buffer. The pH of the buffer was adjusted using HCl or NaOH solutions. Sample concentration was fixed at 5 μM for all fluorescence measurements (unless otherwise specified). The baseline was subtracted at all times. Both dG and SedG stock solutions were prepared in methanol for all measurements.
2.4 HPLC analysis and purification
The crude 6-seleno-2′-deoxyguanosine was purified and analyzed by reverse-phase high-performance liquid chromatography (RP-HPLC). The purification was performed on a Welchrom C18-XB, 10 μm, 21.2 mm × 250 mm column using Shimadzu SPD-10A VP liquid chromatography with a flow-rate of 6 mL/min [buffer A: 2.5 mM triethylammoniumacetate (TEAAc, pH 7.1) in water; buffer B: 2.5 mM TEAAc (pH 7.1) in 50% acetonitrile] and a linear gradient from 100% buffer A to 100% buffer B in 15 min. The desired peak was collected and the buffers were removed by lyophilization. All samples were analyzed on a Welchrom C18-XB column (4.6 mm × 250 mm) and measured at a flow rate of 1.0 mL/min and a linear gradient of 0 to 25% buffer B in 20 min [buffer A: 20 mM TEAAc (pH 7.1) in water; buffer B: 20 mM TEAAc (pH 7.1) in 50% acetonitrile].
2.5 pH titration curve of 6-seleno-2′-deoxyguanosine
The measurements were made at room temperature in 10 mM sodium phosphate buffer. The pH of the solution was adjusted using HCl and NaOH solutions. The absorption was measured every 0.1 pH unit between pH 7–9 and at every 0.5–1.0 pH unit between pH 4–7 and pH 9–12. The pH of each solution was measured before and after spectrum collection and the error range was within ±0.02 pH units.
3 Results and discussion
6-Se-2′-deoxyguanosine (Scheme 1) was synthesized using 5′-DMTr-2-(4-tert-butyl-phenoxy)acetyl-6-(2-cyanoethyl)-seleno-2′-deoxyguanosine (1). The synthesis of nucleoside 1 has been previously reported [29]. Nucleoside 1 was detritylated using acid solution. The cyanoethyl (CE) and (4-tert-butyl-phenoxy)-acetyl (t-BPAc) protecting groups were removed under ultra-mild conditions (0.05 M K2CO3 in methanol).
Native 2′-deoxyguanosine (dG) reaches an absorption maximum at 254 nm. Introduction of selenium yields an interesting absorption profile to the nucleoside. The absorption spectrum of 2-(4-tert-butyl-phenoxy)acetyl-6-(2-cyanoethyl)-seleno-2′-deoxyguanosine(N2-tBPAc-6-CE-Se-dG, 2) shows discreet bands at 256 nm and 308 nm (Figure 1(a)) in aqueous solution. When the protecting groups are completely removed, the resulting SedG(3) shows absorption maximum at 357 nm and is bright yellow in color (inset: Figure 1(a)). Thus, a single atom replacement not only imparts a strong color to the otherwise colorless nucleoside, but also results in more than 100 nm red-shift in the UV absorption. This observation was confirmed by an HPLC analysis comparison of commercial dG with synthesized SedG at 260 nm and 360 nm (Figure 1(b)). SedG absorbs at both 260 nm (slightly) and 360 nm (strongly), whereas native dG has no absorbance at 360 nm.
Earlier reports, using X-ray crystallographic analysis and theoretical calculations, have confirmed the existence of 6-thioguanine in different tautomeric forms [44–46]. Tautomerism in 6-selenoguanine has also been investigated using high-level ab initio calculations. The 6-selenol form was found to be the most stable in gas phase, whereas the 6-selenone form was predicted to be the most stable in aqueous solution (Figure 2). The presence of selenium also influences the pKa of N-1 imino proton in guanosine; the deprotonation generates the more stable 6-selenolic form [47, 48]. Interestingly, during the pKa measurements on SedG, we observed that the 6-selenolic form has a distinct UV-absorption profile. The protonation and deprotonation of SedG was monitored by UV spectrophotometry in solutions with pH values ranging from 1 to 12 (Figure 3; dG also shown for comparison). Under basic pH, the absorption maximum of SedG shifts to 330 nm, whereas under neutral and acidic conditions the absorption maximum is 357 nm. Because the UV-absorption is strongly dependent on solution pH, we recorded the UV-Vis spectra of SedG at different pH values and calculated the pKa. The pKa(7.57 ± 0.02) of SedG was calculated from the fitted titration plot (Figure 4).
Figure 2.

The tautomers, protonated, and deprotonated forms of SedG.
Figure 3.

UV-Vis spectra. Absorption profiles of (a) SedG and (b) dGas a function of pH.
Figure 4.

pKa titration plot. pH versus wavelength (nm) plot for SedG; the fitted curve yields the pKa value 7.57 (±0.02).
Guanine and its nucleosides and nucleotides are known to fluoresce under extreme pH conditions, such as pH 1 [16, 18]. This fluorescence is dictated by the electron distribution on the guanine base at low pH. Therefore, it is meaningful to investigate the impact of selenium modification on guanosine, especially SedG fluorescence as a function of pH, because the electron-rich Se atom alters the dG electron distribution. For a direct comparison, we measured the fluorescence of both dG (Figure 5) and SedG (Figure 6) in solutions with pH values ranging from 1 to 12. The pH-dependent fluorescence profile of native dG, with a fluorescence emission maximum at 395 nm, is consistent with the literature reports [16, 18]. However, the fluorescence profile of SedG, as a function of pH, shows a different pattern (Figure 6). Unlike native dG, the excitation spectrum (excitation wavelength: 305 nm; emission maximum: 390 nm) of SedG is significantly different than its UV-Vis spectrum (absorption maximum: 357 nm). Probably due to the fluorescence-quenching by the solvent (water) molecules, no fluorescence was observed by excitation of SedG at 360 nm. We also found that SedG is practically non-fluorescent under strong acid conditions (pH 1–2), probably due to the protonation of the SedG 2-amino group (Figure 2). Higher pH (> 7.57), which causes deprotonation of the Se-nucleobase, will also decrease the fluorescence. Under other pH conditions, SedG is fluorescent; here, the fluorescence reached a peak at pH 6.0. Thus, the SedG fluorescence is sensitive to charges on the Se-nucleobase and such charges alter the electron delocalization on the nucleobase, which may explain why SedG fluorescence reaches its maximum at pH 6.0. Moreover, our study indicated that SedG is fluorescent under physiological pH 7.4. Concentration-dependent fluorescence profiles are presented in Figure 7 and S7. Due to the complicated conversion between the SedG tautomers in protonated and deprotonated forms, it is challenging to figure out the effect of pH on the tautomerization. Currently, we are investigating how pH may affect both the tautomer formation and fluorescence intensity.
Figure 5.

Fluorescence spectra of dG in aqueous solutions. (a) Excitation spectra of dG as a function of pH at 25 °C; the emission wavelength was 395 nm; (b) emission spectra of dGas a function of pH at 25 °C, with excitation at 258 nm.
Figure 6.

Fluorescence spectra of SedG in aqueous solutions. (a) Excitation spectra of SedG as a function of pH at 25 °C; the emission wavelength was 390 nm; (b) emission spectra of SedG as a function of pH at 25 °C, with excitation at 305 nm.
Figure 7.

Concentration-dependent fluorescence spectra of SedG at pH 7.4. (a) Excitation spectra of SedG at pH 7.4 and 25 °C. The emission wavelength was 390 nm; (b) emission spectra of SedG at pH 7.4 and 25 °C. The excitation wavelength was 305 nm.
Solvent polarity is another significant parameter when considering any environmental effect on the absorption and emission spectra [23, 49–51]. Therefore, we measured the UV-Vis absorption of dG in different solvents and observed that SedG nucleoside exhibits interesting solvatochromicity (Figure 8(a)). In general, the UV-absorption maximum is red-shifted to approximately 370 nm in polar solvents. A detailed analysis on the fluorescence of SedG, in solvents with different polarities, was also performed (Figure 8(b, c)). We were excited to observe that SedG is noticeably emissive in isopropanol and has a little or no fluorescence in the other solvents used in this study, which suggests that the solvent polarity can significantly affect the emission profile. Such polarity-dependent emission is probably due to ground-state interactions between the nucleoside and the solvent molecules. The excitation spectra of SedG in different solvents were monitored at 405 nm. We found that the fluorescence excitation maxima were slightly red-shifted to 320 nm in the organic solvents. Concentration-dependent fluorescence profiles for SedG in isopropanol (Figure 9) and ethanol (Figure S8) were also obtained.
Figure 8.

UV-Vis and fluorescence spectra of SedG in different solvents. (a) Absorption spectra of SedG in different solvents at 25 °C; (b) excitation spectra of SedG in different solvents at 25 °C, emission wavelength 405 nm; (c) emission spectra of SedG in different solvents at 25 °C, excitation wavelength 320 nm.
Figure 9.

Concentration-dependent fluorescence spectra of SedG in iso-propanol. (a) Excitation spectra of SedG in iso-propanol at 25 °C; emission wavelength 395 nm; (b) emission spectra of SedG in iso-propanol at 25°C; excitation wavelength 320 nm.
4 Conclusions
We have developed a simple strategy for 6-seleno-2′-deoxyguanosine (SedG) synthesis. We have found that SedG exhibits distinctive optical properties although it is structurally similar to the native dG. Unlike the corresponding native, SedG is bright yellow and fluorescent under physiological pH. This is the first time that observations of a fluorescent nucleoside by single-atom alteration have been made. We also found that the UV absorption and fluorescent emission of SedG are sensitive to pH and solvent polarity. Therefore, the Se-modified nucleoside has great potential as a useful visual and fluorescent probe for nucleoside, nucleotide, and nucleic acid structure-function studies such as nucleoside transportation, nucleic acid-protein interaction, and nucleic-acid folding.
Acknowledgments
This work was financially supported by the US National Science Foundation (NSF, MCB-0824837), and the Georgia Cancer Coalition (GCC) Distinguished Cancer Clinicians and Scientists Awards.
References
- 1.Lin L, Sheng J, Huang Z. Nucleic acid X-ray crystallography via direct selenium derivatization. Chem Soc Rev. 2011;40(9):4591–4602. doi: 10.1039/c1cs15020k. [DOI] [PubMed] [Google Scholar]
- 2.Salon J, Gan J, Abdur R, Liu H, Huang Z. Synthesis of 6-Seguanosine RNAs for structural study. Org Lett. 2013;(15):3934–3937. doi: 10.1021/ol401698n. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Sheng J, Gan J, Soars AS, Salon J, Huang Z. Structural insights of non-canonical U*U pair and hoogsteen interaction probed with Se atom. Nucleic Acids Res. 2013;41:10475–10487. doi: 10.1093/nar/gkt799. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Sheng J, Zhang W, Hassan AE, Gan J, Soares AS, Geng S, Ren Y, Huang Z. Hydrogen bond formation between the naturally modified nucleobase and phosphate backbone. Nucleic Acids Res. 2012;40(16):8111–8118. doi: 10.1093/nar/gks426. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Sheng J, Huang Z. Selenium derivatization of nucleic acids for X-ray crystal-structure and function studies. Chem Biodiversity. 2010;7(4):753–785. doi: 10.1002/cbdv.200900200. [DOI] [PubMed] [Google Scholar]
- 6.Salon J, Sheng J, Jiang J, Chen G, Caton-Williams J, Huang Z. Oxygen replacement with selenium at the thymidine 4-position for the Se base pairing and crystal structure studies. J Am Chem Soc. 2007;129(16):4862–4863. doi: 10.1021/ja0680919. [DOI] [PubMed] [Google Scholar]
- 7.Salon J, Chen G, Portilla Y, Germann MW, Huang Z. Synthesis of a 2′-Se-uridine phosphoramidite and its incorporation into oligonucleotides for structural study. Org Lett. 2005;7(25):5645–5648. doi: 10.1021/ol052270y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Hassan AE, Sheng J, Zhang W, Huang Z. High fidelity of base pairing by 2-selenothymidine in DNA. J Am Chem Soc. 2010;132(7):2120–2121. doi: 10.1021/ja909330m. [DOI] [PubMed] [Google Scholar]
- 9.Hassan AE, Sheng J, Jiang J, Zhang W, Huang Z. Synthesis and crystallographic analysis of 5-Se-thymidine DNAs. Org Lett. 2009;11(12):2503–2506. doi: 10.1021/ol9004867. [DOI] [PubMed] [Google Scholar]
- 10.Sheng J, Jiang J, Salon J, Huang Z. Synthesis of a 2′-Se-thymidine phosphoramidite and its incorporation into oligonucleotides for crystal structure study. Org Lett. 2007;9(5):749–752. doi: 10.1021/ol062937w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Zhang W, Hassan EA, Huang Z. Synthesis of novel di-Se-containing thymidine and Se-DNAs for structure and function studies. Sci China Chem. 2013;56(3):273–278. doi: 10.1007/s11426-012-4800-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Sun H, Jiang S, Caton-Williams J, Liu H, Huang Z. 2-Selenouridine triphosphate synthesis and Se-RNA transcription. RNA. 2013;19:1309–1314. doi: 10.1261/rna.038075.112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Lin L, Sheng J, Momin RK, Liu H, Huang Z. Facile synthesis and anti-tumor cell activity of Se-containing nucleosides. Nucleosides, Nucleotides Nucleic Acids. 2009;28(1):56–66. doi: 10.1080/15257770802581765. [DOI] [PubMed] [Google Scholar]
- 14.Brandt G, Carrasco N, Huang Z. Efficient substrate cleavage catalyzed by hammerhead ribozymes derivatized with selenium for X-ray crystallography. Biochemistry. 2006;45(29):8972–8977. doi: 10.1021/bi060455m. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Carrasco N, Ginsburg D, Du Q, Huang Z. Synthesis of selenium-derivatized nucleosides and oligonucleotides for X-ray crystallography. Nucleosides, Nucleotides Nucleic Acids. 2001;20(9):1723–1734. doi: 10.1081/NCN-100105907. [DOI] [PubMed] [Google Scholar]
- 16.Longworth J, Rahn R, Shulman R. Luminescence of pyrimidines, purines, nucleosides, and nucleotides at 77 K. The effect of ionization and tautomerization. J Chem Phys. 1966;45:2930. doi: 10.1063/1.1728048. [DOI] [PubMed] [Google Scholar]
- 17.Walaas E. Fluorescence of adenine and inosine nucleotides. Acta Chem Scand. 1963;17:461–463. [Google Scholar]
- 18.Börresen H. On the luminescence properties of some purines and pyrimidines. Acta Chem Scand. 1963;17(4):921–929. [Google Scholar]
- 19.Cohen BJ, Goodman L. Luminescence of purines 1. J Am Chem Soc. 1965;87(23):5487–5490. [Google Scholar]
- 20.Korshun VA, Manasova EV, Balakin KV, Malakhov AD, Perepelov AV, Sokolova TA, Berlin Yu A. New fluorescent nucleoside derivatives-5-alkynylated 2-deoxyuridines. Nucleosides Nucleotides. 1998;17(9-11):1809–1812. [Google Scholar]
- 21.Kovaliov M, Segal M, Fischer B. Fluorescent p-substitutedphenyl-imidazolo-cytidine analogues. Tetrahedron. 2013;69(18):3698–3705. [Google Scholar]
- 22.Bag SS, Saito Y, Hanawa K, Kodate S, Suzuka I, Saito I. Intelligent fluorescent nucleoside in sensing cytosine base: Importance of hydrophobic nature of perylene fluorophore. Bioorg Med Chem Lett. 2006;16(24):6338–6341. doi: 10.1016/j.bmcl.2006.09.039. [DOI] [PubMed] [Google Scholar]
- 23.Ben Gaied N, Glasser N, Ramalanjaona N, Beltz H, Wolff P, Marquet R, Burger A, Mély Y. 8-Vinyl-deoxyadenosine, an alternative fluorescent nucleoside analog to 2′-deoxyribosyl-2-aminopurine with improved properties. Nucleic Acids Res. 2005;33(3):1031–1039. doi: 10.1093/nar/gki253. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Wilson JN, Gao J, Kool ET. Fluorescent nucleoside analogs: Synthesis, properties and applications. Tetrahedron. 2007;63(17):3415. [Google Scholar]
- 25.Xie Y, Maxson T, Tor Y. Fluorescent nucleoside analogue displays enhanced emission upon pairing with guanine. Org Biomol Chem. 2010;8(22):5053–5055. doi: 10.1039/c0ob00413h. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Sinkeldam RW, Greco NJ, Tor Y. Fluorescent analogs of biomolecular building blocks: Design, properties and applications. Chem Rev. 2010;110(5):2579–2619. doi: 10.1021/cr900301e. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Dodd D, Hudson R. Intrinsically fluorescent base-discriminating nucleoside analogs. Mini-Rev Org Chem. 2009;6(4):378–391. [Google Scholar]
- 28.Akimitsu O, Yoshio S, Isao S. Design of base-discriminating fluorescent nucleosides. J Photochem Photobiol, C. 2005;6:108–122. [Google Scholar]
- 29.Salon J, Jiang J, Sheng J, Gerlits OO, Huang Z. Derivatization of DNAs with selenium at 6-position of guanine for function and crystal structure studies. Nucleic Acids Res. 2008;36(22):7009–7018. doi: 10.1093/nar/gkn843. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Milne GH, Townsend LB. Synthesis and antitumor activity of alpha-and beta-2′-deoxy-6-selenoguanosine and certain related derivatives. J Med Chem. 1974;17(3):263–268. doi: 10.1021/jm00249a002. [DOI] [PubMed] [Google Scholar]
- 31.Chu S-H, Davidson DD. Potential antitumor agents. 2. Alpha- and Beta-2′-deoxy-6-selenoguanosine and related compounds. J Med Chem. 1972;15(10):1088–1089. doi: 10.1021/jm00280a031. [DOI] [PubMed] [Google Scholar]
- 32.Chu SH. Potential antitumor agents. Selenoguanosine and related compounds. J Med Chem. 1971;14(3):254–255. doi: 10.1021/jm00285a024. [DOI] [PubMed] [Google Scholar]
- 33.Beltagy Y, Waugh W, Repta A. Antioxidants in purification, stabilization, and formulation of the antineoplastic agent 6-selenoguanosine. J Pharm Sci. 1980;69(10):1168–1170. doi: 10.1002/jps.2600691012. [DOI] [PubMed] [Google Scholar]
- 34.Harris NJ, Booth PJ. Folding and stability of membrane transport proteins in vitro. Biochim Biophys Acta, Biomembr. 2012;1818(4):1055–1066. doi: 10.1016/j.bbamem.2011.11.006. [DOI] [PubMed] [Google Scholar]
- 35.Charalambous K, Booth PJ, Woscholski R, Seddon JM, Templer RH, Law RV, Barter LM, Ces O. Engineering de novo membrane-mediated protein-protein communication networks. J Am Chem Soc. 2012;134(13):5746–5749. doi: 10.1021/ja300523q. [DOI] [PubMed] [Google Scholar]
- 36.Werten P, Rémigy HW, de Groot B, Fotiadis D, Philippsen A, Stahlberg H, Grubmüller H, Engel A. Progress in the analysis of membrane protein structure and function. FEBS Lett. 2002;529(1):65–72. doi: 10.1016/s0014-5793(02)03290-8. [DOI] [PubMed] [Google Scholar]
- 37.Li F, Xia Y, Meiler J, Ferguson-Miller S. Characterization and modeling of the oligomeric state and ligand binding behavior of purified translocator protein 18 KDa from rhodobacter sphaeroides. Biochemistry. 2013;54(34):5884–5899. doi: 10.1021/bi400431t. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Rask-Andersen M, Masuram S, Fredriksson R, Schiöth HB. Solute carriers as drug targets: Current use, clinical trials and prospective. Mol Aspects Med. 2013;34(2):702–710. doi: 10.1016/j.mam.2012.07.015. [DOI] [PubMed] [Google Scholar]
- 39.Lau FW, Bowie JU. A method for assessing the stability of a membrane protein. Biochemistry. 1997;36(19):5884–5892. doi: 10.1021/bi963095j. [DOI] [PubMed] [Google Scholar]
- 40.Hebling CM, Morgan CR, Stafford DW, Jorgenson JW, Rand KD, Engen JR. Conformational analysis of membrane proteins in phospholipid bilayer nanodiscs by hydrogen exchange mass spectrometry. Anal Chem. 2010;82(13):5415–5419. doi: 10.1021/ac100962c. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Grewer C, Gameiro A, Mager T, Fendler K. Electrophysiological characterization of membrane transport proteins. Annu Rev Biophys. 2013;42:95–120. doi: 10.1146/annurev-biophys-083012-130312. [DOI] [PubMed] [Google Scholar]
- 42.Shaikh SA, Li J, Enkavi G, Wen PC, Huang Z, Tajkhorshid E. Visualizing functional motions of membrane transporters with molecular dynamics simulations. Biochemistry. 2013;52(4):569–587. doi: 10.1021/bi301086x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Teale F. The ultraviolet fluorescence of proteins in neutral solution. Biochem J. 1960;76(2):381–388. doi: 10.1042/bj0760381. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Alhambra C, Luque F, Estelrich J, Modesto Orozco. Tautomerism of neutral and protonated 6-thioguanine in the gas phase and in aqueous solution. An ab initio study. J Org Chem. 1995;60(4):969–976. [Google Scholar]
- 45.Leszczynski J. Tautomers of 6-thioguanine: Structures and properties. J Phys Chem. 1993;97(14):3520–3524. [Google Scholar]
- 46.Stewart MJ, Leszczynski J, Rubin YV, Blagoi YP. Tautomerism of thioguanine: From gas phase to DNA. J Phys Chem A. 1997;101(26):4753–4760. [Google Scholar]
- 47.Venkateswarlu D, Leszczynski J. Tautomerism and proton transfer in 6-selenoguanine: A post Hartree-Fock level ab initio SCF-MO investigation. J Phys Chem A. 1998;102(30):6161–6166. [Google Scholar]
- 48.Leszczyński J. Guanine, 6-thioguanine and 6-selenoguanine: Ab initio HF/DZP and MP2/DZP comparative studies. J Mol Struct. 1994;311:37–44. [Google Scholar]
- 49.Cho HY, Woo SK, Hwang GT. Synthesis and photophysical study of 2′-deoxyuridines labeled with fluorene derivatives. Molecules. 2012;17(10):12061–12071. doi: 10.3390/molecules171012061. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Yoshio S, Azusa S, Shinya I, Isao S. Synthesis and photophysical properties of novel push-pull-type solvatochromic 7-deaza-2′-deoxypurine nucleosides. Tetrahedron Lett. 2011;52(37):4726–4729. [Google Scholar]
- 51.Callis PR. Electronic states and luminescence of nucleic acid systems. Annu Rev Phys Chem. 1983;34(1):329–357. [Google Scholar]
