Next Article in Journal
COF-SiO2@Fe3O4 Composite for Magnetic Solid-Phase Extraction of Pyrethroid Pesticides in Vegetables
Previous Article in Journal
Co-Adsorption of Alcohols and Water in JUK-8 Studied Using Quasi-Equilibrated Thermodesorption
Previous Article in Special Issue
Chlorine in an Organic Molecule, a Universal Promoter—Workhorse—Of Reactions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nitrogen-Centered Radicals Derived from Azidonucleosides

by
Yahaira Reyes
1,
Amitava Adhikary
2 and
Stanislaw F. Wnuk
1,*
1
Department of Chemistry and Biochemistry, Florida International University, Miami, FL 33199, USA
2
Department of Chemistry, Oakland University, Rochester, MI 48309, USA
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(10), 2310; https://doi.org/10.3390/molecules29102310
Submission received: 1 April 2024 / Revised: 4 May 2024 / Accepted: 8 May 2024 / Published: 14 May 2024
(This article belongs to the Special Issue Featured Reviews in Organic Chemistry 2024)

Abstract

:
Azido-modified nucleosides have been extensively explored as substrates for click chemistry and the metabolic labeling of DNA and RNA. These compounds are also of interest as precursors for further synthetic elaboration and as therapeutic agents. This review discusses the chemistry of azidonucleosides related to the generation of nitrogen-centered radicals (NCRs) from the azido groups that are selectively inserted into the nucleoside frame along with the subsequent chemistry and biological implications of NCRs. For instance, the critical role of the sulfinylimine radical generated during inhibition of ribonucleotide reductases by 2′-azido-2′-deoxy pyrimidine nucleotides as well as the NCRs generated from azidonucleosides by radiation-produced (prehydrated and aqueous) electrons are discussed. Regio and stereoselectivity of incorporation of an azido group (“radical arm”) into the frame of nucleoside and selective generation of NCRs under reductive conditions, which often produce the same radical species that are observed upon ionization events due to radiation and/or other oxidative conditions that are emphasized. NCRs generated from nucleoside-modified precursors other than azidonucleosides are also discussed but only with the direct relation to the same/similar NCRs derived from azidonucleosides.

1. Introduction

Nitrogen-centered radicals (NCRs) play an important role in chemical biology and cellular signaling [1,2,3] as well as in organic synthesis, as they allow access to new synthetic pathways in nonconventional ways [4,5,6,7,8]. NCRs are categorized into four main types of radicals: σ-iminyl (R=N·), π-aminyl (R1-N(·)-R2), π-amidyl (R1-CO-N(·)-R2), and π-aminium (R1-N(·)H-R2)+. They are generated via homolytic cleavage, reductive/oxidative conditions, and proton-coupled electron transfer (PCET) methods [4,5,7,8]. Precursors to NCRs include N-halogenated amines [9,10], aryloxyamides [11], sulfonylamides [12], or O-aroyloximes [13], among others [4,8]. NCRs generated from azides (R-N3) are important due to the synthetic ease of the incorporation of azido groups into the frame of complex molecules, including natural products and their versatile application to the subsequent functionalization reactions [14,15].
Azido-modified nucleosides have been of interest for over six decades, and the finding that 3′-azido-3′-deoxythymidine (AZT) is a therapeutic agent [16,17] for the treatment of acquired immunodeficiency syndrome (AIDS) has sparked attention to their chemistry. The synthesis of azidonucleosides, their reactions, and biological activities have been subject of comprehensive reviews [18,19,20]. Azidonucleosides have been explored as substrates for the (a) synthesis of amino nucleosides [21], (b) click chemistry [20], (c) bioconjugation and ligation [19], (d) the metabolic labeling of DNA and RNA and for live cell fluorescence imaging [22], and (e) radical biology including enzyme inhibitions [23,24], among others [18,25].
In the last twenty years, azido-modified nucleosides, nucleotides, and oligonucleotides have been extensively explored as substrates for click chemistry [26,27,28,29]. In general, the application of the azidonucleosides and oligonucleotides in click chemistry is more demanding than their alkyne counterparts due to the challenging chemical [30,31,32,33,34] and enzymatic [35] synthesis and their lack of compatibility with the solid-phase synthesis of DNA fragments [36,37]. Also, the triazole click products are often used for the fluorescence imaging of cancer cells [38,39,40] and for cross-linking cellular nucleic acids [41]. The azide group can often cyclize with nitrogen atoms present in the pyrimidine or purine ring to form tetrazole tautomers [42,43]. Moreover, 5-azidouracil and 8-azidoadenine have also been utilized as photoaffinity probes. Upon UV-irradiation, these aryl azides produce highly reactive nitrenes that can react with most amino acid residues, thereby resulting in nucleic acid–protein photo-crosslinking in yields above 50% [25].
Recently, 2′-deoxy-2′-β-fluoro-4′-azidocytidine (Azvudine), a clinical candidate originally developed for HIV treatment, entered clinical trials in China for evaluating its efficacy and safety and showed promise for treating coronavirus disease (COVID-19) [44,45]. Moreover, 4′-azidocytidine is a potent inhibitor of HCV (R1479) [46]. Azido-substituted nucleosides have been developed for use as prodrugs [42,47], as adenosine receptor antagonists [48], for determining protein–DNA/RNA interactions [34,49], or as anticancer [47] and anti-viral agents [50,51].
Furthermore, 3′-azido-3′-deoxythymidine (AZT) has been employed as a radiation sensitizer in the radiotherapy of tumors for HIV-positive patients [52]. AZT demonstrated significant radiosensitization in irradiated human colon cancer, larynx squamous carcinoma, and malignant glioma cells [53,54,55]. The aim of this account is to discuss the application of azidonucleosides in the selective and site-specific generation of NCRs and their biological implications. Pulse radiolysis, photolysis, and electrochemical investigations as well as enzymatic and biomimetic model studies in combination with electron paramagnetic resonance (EPR) and density functional theory investigations were employed to elucidate the formation of NCRs in azidonucleosides and their subsequent reactions. Thus, the focus of this review is not on the synthesis of azidonucleosides but rather on the importance of the chemistries of NCRs generated from the azido group. NCRs generated from non-azidonucleoside-modified precursors, by ionization, or by the one-electron oxidation of the parent nucleobase are discussed only in relation (comparison) to the azido-derived radical chemistry.

1.1. 2′-Azido-2′-Deoxy Pyrimidine Nucleosides and Nucleotides: Inhibition of Ribonucleotide Reductases and Importance of Nitrogen-Centered Radical Chemistry

2′-Azido-2′-deoxynucleoside 5′-diphosphates (e.g., 1, N3UDP) are potent inactivators of ribonucleotide reductases (RDPR). Sjöberg et al. found that the inactivation of RDPR by 2′-azido-2′-deoxynucleotides was accompanied by the appearance of new EPR signals for a nitrogen-centered radical (anisotropic triplet with a second hyperfine interaction) and the concomitant decay of peaks for the tyrosyl radical [56], which was the first direct evidence for free radical chemistry with RDPR. The structure of this elusive nitrogen radical was studied extensively and shown to be derived from the azide moiety [57,58,59]. A proposed mechanism postulated azide loss (as an anion) from the initial C3′ radical intermediate to give the ketyl radical 2 and subsequent reduction by proton-coupled electron transfer to generate the 2′-deoxy-3′-ketonucleotide 3 (Figure 1). This process leaves a thiyl radical in the active site. The reaction of hydrazoic acid with the thiyl radical generates stoichiometric N2 and a sulfinylimine radical 8. The protonated azide (pKa of 4.6) was hypothesized to be essential for that mechanism [57,59]. Conversion (2 → 3) is analogous to the proposed mechanism for the reduction of natural nucleotides that proceeds by the generation of the same 3′-keto-2′-deoxynucleotide intermediate, which makes the investigation of the inhibition of RDPR by N3UDP even more significant [23,60]. The initial NCR 8 reacts further with the oxygen or carbon atoms of a carbonyl group of the 3′-keto-2′-deoxynucleotide to generate radicals 4 or 7, respectively [58].
The inactivation of the RDPR with 3′[17O]-N3UDP [61] 1 was consistent with the formation of the radical, R-S-N·-C(3′)-OH 7, and provided the first evidence for the trapping of a 3′-ketonucleotide in the reduction process by a nitrogen-centered radical 8 [59]. Chemical requirements also favor the formation of 7 (over 4), and there is precedent in the literature for the analogous addition of aminyl radicals to carbonyl [62] and imino groups [63]. Moreover, the inactivation of the adenosylcobalamine-dependent RTPR with 2′-arabino-2′-azido-2′-deoxyadenosine-5′-triphosphate was accompanied by the detection of a paramagnetic species by EPR spectroscopy. In a tentatively proposed mechanism, perhaps due to an altered sugar pucker and the steric constraints imposed by the azido moiety on the β-face of the nucleoside, the C2′-azide was acting here as a radical trap for the initially formed protein thiyl radical to generate a sulfinylimine NCR [64].
Figure 1. Proposed structures for the nitrogen-centered radicals (NCRs) and pathways for their generation during the inactivation of RDPR by N3UDP [57,58,59,65].
Figure 1. Proposed structures for the nitrogen-centered radicals (NCRs) and pathways for their generation during the inactivation of RDPR by N3UDP [57,58,59,65].
Molecules 29 02310 g001
The theoretical modeling study of the inactivation of RDPR by N3UDP by Pereira and coworkers generated an alternative mechanistic proposal (depict by red arrow path, Figure 1) [65]. In that hypothesis, the released azide anion (N3) was proposed to add to the 2′-ketyl radical 2 first with the concomitant protonation of the ketone oxygen by E441. The resulting radical 5 is then reduced at the 2′ position by Cys225 to generate the Cys225 thiyl radical 6. A subsequent attack of the thiyl radical on an alkyl azide (instead of HN3 or N3) would lead to the same nitrogen-centered radical 7 that was detected experimentally [59]. The addition of azide to ketones has chemical precedence [66], but there are no model systems where azide adds to ketyl radicals. The reduction of alkyl azides with tin [62,67], alkoxy [68], and silyl (in the presence of thiols) [69,70] radicals have been reported [15].

1.2. Biomimetic Studies

Thao and coworkers designed model 3′-azido-3′-deoxynucleosides with thiol or vicinal dithiol substituents at C2′ or C5′ to study reactions postulated to occur during the inhibition of ribonucleotide reductases (RNRs) by N3UDP (911, Figure 2) [71]. It was presumed that the intramolecular addition of the thiyl radical in 12 (generated from 9) to the azido group via an eight-membered TS would produce the transient triazenyl radical 13. A loss of N2 would generate nitrogen-centered radical 14 (similar to sulfinylimine radical of type 7; vide supra), and an abstraction of a hydrogen atom by 14 should give a cyclic sulfenamide, which might undergo ring opening to give 3′-amino-3′-deoxy products. Density functional theory calculations predicted that intramolecular reactions between the generated thiyl radicals and azido group on the model compounds 911 will be exothermic by 33.6–41.2 kcal/mol with low energy barriers of 10.4–13.5 kcal/mol [71].
The heating of 3′-azido-3′-deoxy-5-O-(2,3-dimercaptopropanoyl)thymidine 11 in H2O with 2′-azobis-(2-methyl-2-propionamidine) dihydrochloride (AAPH) as an initiator for the production of thiyl radicals gave 3′-amino-3′-deoxythymidine, whereas the analogous treatment of 3′-azido-3′-deoxythymidine (AZT, a control substrate lacking a thiol substituent) resulted in the isolation of major quantities of unchanged AZT. Moreover, the γ-radiolysis of N2O-saturated aqueous solutions of AZT and cysteine produced 3′-amino-3′-deoxythymidine and thymine. The DFT-calculated predictions and results with radical-initiated intramolecular azido substituent reduction with model compounds, such as 9 and 11, bearing azido and thiol substituents might be in harmony with the enzymatic positioning of azido-containing substrates in close proximity with thiol functionalities that exist in the active sites of RNRs [71].
Robins and coworkers explored further biomimetic reactions to occur during the mechanism-based inhibition of RDPR by N3UDP with model 2′-substituted nucleoside derivatives [72], which can cause the elimination of radical or ionic species from C2′ upon the generation of a radical at C3′, including 2′-azido-substituted uridine and adenosine models (e.g., 15 [21]) [73]. Azide 15 was converted to 3′-O-phenoxythiocarbonyl (PTC) derivative 16. Compounds 15 and 16 were subjected to parallel treatment with tributylstannane/AIBN/toluene/Δ (Method a) and also with triphenylsilane/dibenzoyl peroxide/toluene/Δ (Method b; Figure 3). The known stannyl “radical-mediated” reduction [74] of azide 15 to amine 18 occurred with tributylstannane/AIBN. Interestingly, a reduction of azide 15 to amine 18 was not observed with triphenylsilane/dibenzoyl peroxide. It is also noteworthy that although the radical-mediated cleavage of the carbon-nitrogen bond with azides is unknown, the treatment of the ribo epimer of 15 with excess triphenylsilane and a prolonged reaction time caused some dehydrogenolytic deazidation to give small quantities of the 2′-deoxy derivative 19 in the absence of amine formation [73].
The treatment of 3′-thionocarbonate 16 with Bu3SnH resulted in radical-mediated elimination upon the generation of a radical at C3′ 17 to give 2′,3′-didehydro-2′,3′-dideoxy nucleoside 21 in moderate yields, but competing reduction of the azido group and hydrogenolysis of the thionocarbonate group also produced byproduct 20. Elimination product 21 was formed almost exclusively upon the treatment of 16 with Ph3SiH but at a lower conversion rate and with the recovery of starting material 16 (Figure 3) [73].

2. Nitrogen-Centered Radicals Generated from Azidonucleosides by Radiation-Produced Electrons

Azide anion (N3) has a low reactivity with radiation-produced electrons (e.g., with aqueous electrons, i.e., with fully solvated electrons when water is the solvent) [75]; however, the aqueous electrons react with azidonucleosides with an almost diffusion-controlled rate (~1010 M−1s−1), forming NCRs [76]. The plausible involvement of dissociative electron attachment pathway in the formation of NCRs from azidonucleosides could be inferred from the dissociative electron attachment (DEA) spectra of phenyl azide (Scheme 1). Studies establish that the major fragmentation pathway involves N2 loss [77].

2.1. From Azido-Modified Nucleoside Sugars

Sevilla and Adhikary developed a method for the generation of aminyl radicals on dissociative electron attachment (DEA) to azidonucleosides, which allows for detailed EPR spectroscopic studies of aminyl radicals and their subsequent chemistry [79]. EPR spectral studies and density functional theoretical calculations showed that the predominant site of electron capture in AZT (22) is at the azide group (ca. 80%) and not at the thymine moiety (ca. 20%), which is the most electron-affinic DNA-base [79]. In this work, the radiation-produced prehydrated electrons led to the site-specific formation of a localized π-aminyl radical (RNH·) 25 [in homogeneous glassy 7.5 M aqueous (D2O or H2O) LiCl solutions; Figure 4]. It was demonstrated that the neutral azide can capture an electron due to its high electron affinity and form a transient negative ion (TNI), an unstable azide anion radical RN3·, 23 (see also Scheme 1) when irradiated. RN3· can then facilely lose N2, leaving a nitrene anion radical (RN·, 24; neither were detected by EPR even at 77 K) that upon protonation from the surrounding solvent becomes a neutral π-aminyl radical (RND·/RNH·, 25), which was detected by EPR. Upon annealing to higher temperatures (ca 160–170 K), RNH· 25 undergoes bimolecular H-atom abstraction from the C5-methyl group of 22, generating the allyl radical 26 or from the sugar moiety yielding C5′· (27) [79].
A radiation-produced electron addition to 5′-azido-5′-deoxythymidine 28 forms C5′-NH·, which undergoes predominant (ca. 80%) bimolecular H-atom abstraction from a proximate C5-methyl group generating the allyl radical (of type 26) and ca. 20% of the σ-type iminyl radical in the sugar moiety (similar to 32; see Figure 5 and relevant discussion) [79,80]. Interestingly, electron attachment to 3′-azido-2′,3′-dideoxyguanosine 29 results in the one-electron oxidation of the guanine base via an intramolecular electron-coupled proton transfer pathway to give the G(N1-H)· radical [79,81] but not the elusive guanyl radical G(N2-H)· [82]. The latter is formed via the deprotonation of a guanyl cation radical from the exocyclic amine of the guanine base in DNA but not in monomers [79,83,84].
Mudgal et al. reported that the site of azido substitution in the sugar moiety of azidopyrimidine nucleosides influences the reactivity of aminyl radicals formed by dissociative electron attachment (Figure 5) [80]. Employing a 15N-labeled azido group and deuterations at specific sites in the sugar and base, it was shown that they initially form aminyl radicals RNH· upon annealing samples from 77 K to 170 K: (i) at a primary carbon site (e.g., 31; generated from 5′-azido-2′,5′-dideoxyuridine, 30), it is converted to a σ-type iminyl radical (R=N·) 32 via a concentration-dependent bimolecular H-atom abstraction reaction between 30 and 31 and subsequent nitrogen loss from the intermediary α-azidoalkyl radical 33 (path A); (ii) at a secondary carbon site RNH· 35, it is generated from 2′-azido-2′-deoxyuridine (34, 2′-N3dU) and underwent bimolecular electrophilic addition to the C5=C6 double bond of a proximate pyrimidine base to give C6 36a and C5 36b radicals (path B); and (iii) at tertiary alkyl carbon (e.g., in 4′-azidocytidine 37a), RNH· 37b is quite stable and undergoes little to no reaction (path C). These results show the influence of the stereochemical and electronic environment on RNH· reactivity and potentially should allow for the selection of azidonucleosides, which would be the most effective in augmenting cellular radiation damage [80].
To further test the mechanism for sugar radical formation from the π-aminyl radicals without the nucleobase interaction, Mudgal et al. investigated the formation and reactivity of the π-aminyl radicals from azidopentafuranoses [e.g., methyl 2-azido-2-deoxy-α-D-lyxofuranoside, 38 (1-Me-2-Azlyxo), and its β-D-ribo isomer 43 (1-Me-2-Azribo)] [85]. Prehydrated electron attachment to 38 and its 15N- and 2H-labeled derivatives showed unequivocal (concentration independent) intramolecular H-atom abstraction from the C5 by RNH· in 39 via a favorable six-membered transition state ([1,5]-hydrogen shift) to produce the primary α-hydroxy C5· 40. Subsequent ring opening (4041) and unimolecular conversion produced a secondary C4· 42 under the reductive environment of DEA (Figure 6) [85]. However, for 43, EPR studies established thermally-activated (concentration dependent) intermolecular H-atom abstraction by RNH· 44 from the methyl group at the C1 of 43 to generate the carbon-centered radical 45.
The general character of the site-specific generation of NCRs from azide group has been also illustrated with other natural products, such as sesquiterpene lactones [e.g., parthenolide (PTL) and dehydroleucodine (DhL)] [86]. The addition of radiation-produced electrons to azido-PTL 46 leads to the formation of highly reactive oxidizing aminyl radical 47, which transforms into a stable α-carbonyl-stabilized tertiary C-centered radical AmPTL· 48 via [1,3]-hydrogen shift (Figure 7). Remarkably, the radiation of azido-DhL produces the corresponding aminyl radical, which, after bimolecular H-atom abstraction from substrate Azido-DhL, generates α-azidoalkyl radical 49. Interestingly, no iminyl radicals have been detected upon the radiation of these sesquiterpene lactones. Azido-PTL and azido-DhL significantly suppressed proliferation rate and colony-forming ability in MCF-7 cells [86]. The azido-PTL in combination with radiation restricted its colony-forming ability to a greater extent than the PTL itself. The radiosensitization has been attributed to the increased reactive oxygen species (ROS) generated by the radicals produced from the azido group.

2.2. From Azido-Modified Nucleobases

The synthetic ability of the regioselective incorporation of the azido group at chemically distinctive positions of nucleobases (e.g., C5/C6 of pyrimidine bases or C2/C6 of pyrimidine moiety as well as C8 of imidazole moiety of purine bases [87]) provides a possibility for the selective generation of NCRs under the reductive conditions of dissociative electron attachment. It is noteworthy that often, it is challenging to selectively generate and elucidate the same specific NCR species during radiation and/or other oxidative conditions used during ionization events of non-azido nucleobases.
Wen et al. have employed 5-azidomethyl pyrimidine derivatives, such as AmdU 50 and AmdC, to study the radiation-mediated formation of RNH· and its subsequent reactions for potential anticancer properties (Figure 8) [88]. The authors hypothesized that the incorporation of azido-modified nucleosides into genomic DNA would augment radiation-induced damage in cells owing to the reactions of aminyl radicals under hypoxic conditions and therefore act as potential radiosensitizers. Their findings revealed that radiation-produced electron addition to 5-AmdU 50 generates the π-aminyl RNH· 51. Radical 51 then undergoes an intermolecular reaction with 5-AmdU, abstracting the H-atom from the CH2N3 group at the C5 of 5-AmdU, to form the α-azido carbon radical 52. The subsequent loss of N2 from 52 generates the thermodynamically more stable σ-iminyl radical [15,89] (R=N·) 53. The AmdU and AmdC phosphates incorporate into DNA fragments in polymerase-catalyzed reactions, and AmdU demonstrated effective radiosensitization in EMT6 tumor cells in the presence or absence of oxygen [88]. AmdU has been also metabolically incorporated into DNA in living cells for the click labeling of DNA [90], and its 5′-triphosphate was found to be the substrate for DNA polymerases and PCR amplification [91]. Moreover, its 5′-triphosphate prodrug was shown to enhance its incorporation into the DNA of wild-type cells and animals [22]. Thus, AmdU can serve a dual purpose of labeling tumor cells prior to, during, or after radiotherapy, and it may radiosensitize the tumor during radiotherapy.
The addition of a radiation-produced electron to 5-(1-azidovinyl)-2′-deoxyuridine (AvdU, 54) also generates π-RNH· 55, which undergoes facile tautomerization to thermodynamically more stable σ-iminyl radical 56 (Figure 9) [88]. One-electron attachment to the cytidine analog (AvdC) proved that the formation of the aminyl radical and its tautomerization to the iminyl radical has a general character and occurs independently of the nucleobase [88]. Owing to the high concentrations of various free radical scavengers in cells [92], the bimolecular conversion of π-RNH· to σ-R=N· from 5-azidomethyl nucleosides (e.g., AmdU) is unlikely to take place. However, the tautomerization of the π-RNH· to σ-R=N· from 5-azidovinyl nucleosides (e.g., AvdU) should occur even in cells. Therefore, it was expected that the π-RNH· from AmdU could augment radiation damage more effectively than the σ-R=N· from AvdU. The experiments in EMT6 tumor cells indeed showed the higher radiosensitizing effect of AmdU compared with that of AvdU [88].
Distinct pathways of dissociative electron attachment have been observed in 6-azidomethyl uracil nucleosides [93] compared to their 5-azidomethyl counterparts (Figure 10) [76]. Contrary to the results with 5-AmdU [88], where radiation-mediated prehydrated electrons in the absence of oxygen led to π-aminyl and σ-iminyl radicals (see Figure 8), radiation-produced electron addition to 6-azidomethyluridine (6-AmU, 57) leads to the unexpected loss of azide as an anion via dissociative electron attachment from the initially formed azide anion radical intermediate (58, U-6-CH2-N3·¯) or transient radical anion to generate the C6 allylic radical 59 [76].
The characterization of nitrogen-centered radicals formed via dissociative electron attachment to the azido group directly attached to a nucleobase, such as in 5-azidouridine 60, 6-azidouridine 61, and 4-azidopyrimidine analogue 62, which have demonstrated how the presence of the azido group at different positions in the pyrimidine base (meta to N1 and N3 in 60 and ortho/para to N1 and N3 in 61 and 62) distinctly affects the nature and stability of the nitrogen-centered radicals (Figure 11) [94]. The formation of (i) RNH· (64) from 60, (ii) R=N· (65) from 61, and (iii) RN3· (66) from 62, after gamma-irradiation at 77 K, was observed. Moreover, tetrazolocytidine 63 (a cylic derivative of 62), upon irradiation, also produced azide anion radical 67 [94,95].
Hocek’s group developed azidophenyl (AzP)-labeled 5-(4-azidophenyl)-2′-deoxycytidine 68 and 7-(4-azidophenyl)-7-deaza-2′-deoxyadenosine 69a triphosphates (dAAzPTP and dCAzPTP) as substrates for the enzymatic labeling of double- or single-stranded DNA (Figure 12) [96,97,98]. These novel electroactive azidophenyl-modified nucleotides and DNA gave strong signals in voltammetric studies at −0.9 V due to the reduction of the azido group, outside the potential region around −1.5 V, where natural bases are reduced. The proposed mechanism for the electrochemical reduction of dAAzP 69b on mercury surface involves the one-electron reduction of the azido-group to nitrene anion radical 70a accompanied by dinitrogen release. The nitrenium ion-radical is then protonated to form the π-aminyl radical 70b, which is stabilized by an aromatic phenylene linker bound to nucleobase. The subsequent electrochemical reduction of the aminyl radical by one electron and one proton yield the amine [98]. The nucleosides with the new AzP redox label are not only suitable for the electrochemical detection but can also be transformed to another redox label or silenced through the click reactions.

3. Nitrogen-Centered Radicals Generated on Non-Azido Nucleobases

Approaches for the generation of the similar NCRs from the selectively modified adenine, guanine, and cytosine substrates “armed with radical initiators” other than the azido group have also been developed [99,100,101]. Thus, Wagner’s group demonstrated the formation of N6-aminyl radical 72 with the photolysis of 6-N-arylhydrazones of 2′-deoxyadenosine 71a and 71b in the presence of H-donors, such as glutathione (GSH; Figure 13) [99]. Specifically, 6-N-(4-methoxyphenyl)hydrazone 71a was more efficient for NCR formation compared to phenylhydrazone 71b. Upon the photolysis of the latter, N6-aminyl radical 72 and benzylidene iminyl radical 73 were postulated to be generated via homolytic cleavage as the subsequent H-atom abstraction by 72 and 73 from GSH-produced 2′-deoxyadenosine (dAdo) and benzaldehyde. The decomposition of N6-aminyl radical 72 in the excess of dAdo revealed the generation of 6-amino-2-imino product 74 (45%) via the recombination of 72 and 73 at the C2. These studies reveal the potential new path for the selective formation of adenosyl-6-N-aminyl radical 72 under neutral conditions.
Greenberg’s group identified the aminyl radicals with the photolysis of the hydrazine modifications at the 6-N-positon of 2′-deoxyadenosine and 2-N-positon of 2′-deoxyguanosine and showed the compatibility of these hydrazine analogues for solid-phase oligonucleotide synthesis to study DNA–hole transfer processes [100]. Specifically, the photolysis of hydrazine dAdo precursor 75 produced C8-diphenylamino-substituted dAdo 76 in addition to dAdo (Figure 14). The identification of these photolyzed products indicated that N6-aminyl radical 72 is initially generated from 75 with the loss of the diphenyl aminyl radical (Ph2N·). The subsequent reaction between 72 and Ph2N· produced 76. The compatibility of these hydrazines for solid-phase oligonucleotide synthesis was proven via the generation of dodecameric duplex 77, as Tm decreased significantly compared to its dAdo duplicate [100]. Although spectroscopic evidence (UV, MS) and (in)direct synthetic proofs were presented to indicate the formation of the aminyl radical 72, there was no conclusive EPR results unequivocally characterizing this radical due to a very poor rate of fragmentation in glassy systems at low temperature. It is noteworthy to point out that in Wagner’s studies of the in situ-generated benzylidene iminyl radical, 73 was postulated to add to the C2 position of the purine ring (π-deficient pyrimidine ring) to give 74, whereas in Greenberg’s investigations, the diphenyl aminyl radical adds to the C8 position (π-excessive imidazole ring) to produce 76.
The photocleavage of ketone 78 and the β-fragmentation of the initially formed alkyl radical 79 also led to aminyl radical 72 (Figure 14) [101]. The formation of 72 was followed by laser flash photolysis (LFP), which yielded a transient with λmax ≈ 340 nm and a broader weaker absorption centered at ~560 nm. The calculations indicate that the iminyl tautomer 80 of 72 is 13.0 kcal/mol higher in energy in the gas phase and its forms must rapidly isomerize to 72. Precursor 78 has been incorporated into DNA fragments, and the site-selective generation of the neutral purine nitrogen radical 72 was shown to produce tandem lesions. The involvement of 2′-deoxyadenosin-6-N-yl radical 72 in this process could be detected because it was independently generated from the synthetic precursors [102,103,104].
The photolysis of dGuo hydrazine 81 provided strong evidence for the formation of dGuo-N2-yl radical 82 as determined by the generation of 2′-deoxyguanosine (Figure 15) [100]. The independent generation and time-resolved detection of radical 82 was developed by the photolysis of ketone radical initiator 83 [81]. The LFP experiments showed no evidence for the water-assisted tautomerization of dG(N2-H)· 82 to dG(N1-H)· 85 within hundreds of microseconds, supporting the theoretical prediction by von Sonntag [92,105]. This observation suggests that the generation of dG(N1-H)· via dG(N2-H)· following hydrogen atom abstraction from dG and subsequent tautomerization is unlikely [81]. The formation of the N1-yl radical G(N1-H)· 85 has been indirectly observed via the one-electron oxidation of N-(aryloxy)naphthalimide 84 [106]. Photolabile N-hydroxypyrid-2(1H)-one or thione 86 were also used for guanine radical generation [107,108]. It is noteworthy that preliminary results on the addition of a radiation-produced electron to 2-azido-2′-deoxyinosine produces the complex EPR signals, and subsequent UV photoexcitation leads to the mixture of radicals, including the possible formation of sugar radicals rather than the elusive guanyl aminyl radical 82 [95].
The Greenberg’s group reported the photochemical generation of σ-iminyl radical 2′-deoxycytidin-4-N-yl 88 from a nitrophenyl oxime precursor 87 and the synthesis of oligonucleotides containing 87 for DNA incorporation (Figure 16) [109]. It is worth noting that the attempted photolytic generation of 88 from oxime esters 89 and 90 was unsuccessful due to instability and poor fragmentation. The formation of the iminyl radical 88 from 87 was confirmed (in)directly via LFP and transient UV-absorption spectroscopy [109]. Photolysis studies also revealed 88 can recombine with the aryloxyl radical to regenerate 87, undergo C5-C6 addition to produce diradicals, generate deoxycytidine via reduction, or react under aerobic conditions to generate other radicals/products. Iminyl radical 88 was also independently generated (and characterized by EPR) via single-electron transfer to oxime ester 89 [110]. Like with adeninyl radical 72, tandem lesion formation from cytosinyl radical 88 is traceless because it is reduced to dC during the process. However, unlike 72, an isolated 88, which is a stronger oxidant, directly oxidizes dG. In this regard, 88 is more similar to a nucleoside alkyl aminyl radical, such as the one generated from AZT [79,109].
One-electron oxidation and ESR studies of 1-methylcytosine and 2′-deoxycytidine 91 showed that the cytidine cation radical 92 preferentially deprotonates to form an aminyl radical 94 syn to the carbonyl moiety (Figure 17). The tautomerization of syn-aminyl radical also leads to the iminyl σ-radical 88 [111,112]. Interestingly, contrary to these findings, [109,110,111] the one-electron reduction (conditions for DEA) of azide precursors at the C4 positions of pyrimidine bases (62 or 63) led to the formation of anion radicals 66 or 67 (see Figure 11) [94,95].

4. Conclusions

Nucleosides with a regioselectively inserted azido group at the sugar or base moiety are unique precursors for the generation of nitrogen-centered radicals (NCRs) under enzymatic, electrochemical, and dissociative electron attachment (reductive) conditions. The azidonucleoside substrates under reductive conditions produce the same NCR species that are observed upon radiation and/or other oxidative conditions of parent nucleosides, making them unique and desired precursors. The generation of the NCRs from azidonucleosides and their subsequent reactivity depends on the nucleophilicity/electronegativity as well as redox potentials of nucleobases, which define the fragmentation of the initially formed transient negative ions [(RN3]* and their dissociation from the release of N2 (to form π-aminyl radicals, RNH·) or loss of N3 (to form R·; Scheme 1). Due to their reactivity, the RNH· undergo a plethora of reactions, including selective hydrogen atom abstractions from the phosphate–sugar backbone or from the nucleobases as well as their addition to the double bond in the proximate base moiety.
The azidonucleosides have shown significant promise in their application as radiosensitizers for increasing the efficacy of tumor radiochemotherapy. For instance, 5-azidomethyl-2′-deoxyuridine demonstrated effective radiosensitization in EMT6 tumor cells in the presence or absence of oxygen. On the molecular level, it was found that the inactivation of RDPR by 2′-azido-2′-deoxynucleotides was accompanied by the appearance of new EPR signals for a nitrogen-centered radical resulting from the reaction of azide with a protein thiyl radical. Thus, this overview summarizes the investigations outlining the involvement and role of azidonucleoside-produced NCRs that are involved in the first signaling steps (i.e., radicals leading to stable damage products) that affect cellular functions.
Studies on the nucleoside substrates “armed with radical initiators” other than an azido group placed selectively on the exo-amino group of adenine, guanine, and cytosine have led to the generation/detection of similar NCRs under photolytic conditions. These non-azido analogues were found to be compatible for the solid-phase synthesis of DNA fragments, and photolysis was used to study DNA interaction, stability, and hole transfer.

Author Contributions

Conceptualization, S.F.W.; writing-original draft preparation, Y.R. and S.F.W.; writing—review and editing, A.A. and S.F.W.; supervision, S.F.W. All authors have read and agreed to the published version of the manuscript.

Funding

A.A. acknowledges the support from National Cancer Institute of the NIH (Grant RO1CA045424) and the NSF under Grant No. CHE-1920110.

Acknowledgments

A.A. acknowledges the travel support from IntraOp and the same from DNA (MDPI). A.A. also thanks Université Paris-Saclay for the Visiting Professorship at the Institut de Chimie Physique.

Conflicts of Interest

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

References

  1. Kerru, N.; Gummidi, L.; Maddila, S.; Gangu, K.K.; Jonnalagadda, S.B. A Review on Recent Advances in Nitrogen-Containing Molecules and Their Biological Applications. Molecules 2020, 25, 1909. [Google Scholar] [CrossRef] [PubMed]
  2. Dedon, P.C.; Tannenbaum, S.R. Reactive nitrogen species in the chemical biology of inflammation. Arch. Biochem. Biophys. 2004, 423, 12–22. [Google Scholar] [CrossRef] [PubMed]
  3. Adams, L.; Franco, M.C.; Estevez, A.G. Reactive nitrogen species in cellular signaling. Exp. Biol. Med. 2015, 240, 711–717. [Google Scholar] [CrossRef] [PubMed]
  4. Pratley, C.; Fenner, S.; Murphy, J.A. Nitrogen-Centered Radicals in Functionalization of sp2 Systems: Generation, Reactivity, and Applications in Synthesis. Chem. Rev. 2022, 122, 8181–8260. [Google Scholar] [CrossRef] [PubMed]
  5. Jiang, H.; Studer, A. Chemistry With N-Centered Radicals Generated by Single-Electron Transfer-Oxidation Using Photoredox Catalysis. CCS Chem. 2019, 1, 38–49. [Google Scholar] [CrossRef]
  6. Wang, P.; Zhao, Q.; Xiao, W.; Chen, J. Recent advances in visible-light photoredox-catalyzed nitrogen radical cyclization. Green Synth. Catal. 2020, 1, 42–51. [Google Scholar] [CrossRef]
  7. Xiong, P.; Xu, H.-C. Chemistry with Electrochemically Generated N-Centered Radicals. Acc. Chem. Res. 2019, 52, 3339–3350. [Google Scholar] [CrossRef] [PubMed]
  8. Yu, X.-Y.; Zhao, Q.-Q.; Chen, J.; Xiao, W.-J.; Chen, J.-R. When Light Meets Nitrogen-Centered Radicals: From Reagents to Catalysts. Acc. Chem. Res. 2020, 53, 1066–1083. [Google Scholar] [CrossRef] [PubMed]
  9. Cheng, C.; Chen, D.; Li, Y.; Xiang, J.-N.; Li, J.-H. Fluoroamide-driven intermolecular hydrogen-atom-transfer-enabled intermolecular 1,2-alkylarylation of alkenes. Org. Chem. Front. 2023, 10, 943–950. [Google Scholar] [CrossRef]
  10. Sirinimal, H.S.; Hebert, S.P.; Samala, G.; Chen, H.; Rosenhauer, G.J.; Schlegel, H.B.; Stockdill, J.L. Synthetic and Computational Study of Tin-Free Reductive Tandem Cyclizations of Neutral Aminyl Radicals. Org. Lett. 2018, 20, 6340–6344. [Google Scholar] [CrossRef] [PubMed]
  11. Liu, L.; Ward, R.M.; Schomaker, J.M. Regioselective Intramolecular Allene Amidation Enabled by an EDA Complex. Chem. Eur. J. 2020, 26, 13783–13787. [Google Scholar] [CrossRef]
  12. Azzi, E.; Ghigo, G.; Sarasino, L.; Parisotto, S.; Moro, R.; Renzi, P.; Deagostino, A. Photoinduced Chloroamination Cyclization Cascade with N-Chlorosuccinimide: From N-(Allenyl)sulfonylamides to 2-(1-Chlorovinyl)pyrrolidines. J. Org. Chem. 2022, 88, 6420–6433. [Google Scholar] [CrossRef]
  13. Yamada, K.; Sato, M.; Tanaka, K.; Wakabayashi, A.; Igarashi, T.; Sakurai, T. Nitrile-forming radical elimination reactions of 1-naphthaldehyde O-(4-substituted benzoyl)oximes activated by triplet benzophenone. J. Photochem. Photobiol. A Chem. 2006, 183, 205–211. [Google Scholar] [CrossRef]
  14. Shee, M.; Singh, N.D.P. Chemical versatility of azide radical: Journey from a transient species to synthetic accessibility in organic transformations. Chem. Soc. Rev. 2022, 51, 2255–2312. [Google Scholar] [CrossRef]
  15. Minozzi, M.; Nanni, D.; Spagnolo, P. From Azides to Nitrogen-Centered Radicals: Applications of Azide Radical Chemistry to Organic Synthesis. Chem. Eur. J. 2009, 15, 7830–7840. [Google Scholar] [CrossRef]
  16. Horwitz, J.P.; Chua, J.; Noel, M.; Nucleosides, V. The Monomesylates of 1-(2’-Deoxy-β-D-lyxofuranosyl)thymine1,2. J. Org. Chem. 1964, 29, 2076–2078. [Google Scholar] [CrossRef]
  17. Mitsuya, H.; Weinhold, K.J.; Furman, P.A.; St Clair, M.H.; Lehrman, S.N.; Gallo, R.C.; Bolognesi, D.; Barry, D.W.; Broder, S. 3′-Azido-3′-deoxythymidine (BW A509U): An antiviral agent that inhibits the infectivity and cytopathic effect of human T-lymphotropic virus type III/lymphadenopathy-associated virus in vitro. Proc. Natl. Acad. Sci. USA 1985, 82, 7096–7100. [Google Scholar] [CrossRef]
  18. Pathak, T. Azidonucleosides:  Synthesis, Reactions, and Biological Properties. Chem. Rev. 2002, 102, 1623–1668. [Google Scholar] [CrossRef]
  19. Müggenburg, F.; Müller, S. Azide-modified Nucleosides as Versatile Tools for Bioorthogonal Labeling and Functionalization. Chem. Rec. 2022, 22, e202100322. [Google Scholar] [CrossRef]
  20. Perrone, D.; Marchesi, E.; Preti, L.; Navacchia, M.L. Modified Nucleosides, Nucleotides and Nucleic Acids via Click Azide-Alkyne Cycloaddition for Pharmacological Applications. Molecules 2021, 26, 3100. [Google Scholar] [CrossRef]
  21. Robins, M.J.; Hawrelak, S.D.; Hernández, A.E.; Wnuk, S.F. Nucleic Acid Related Compounds. LXXXI. Efficient General Synthesis of Purine (Amino, Azido, and Triflate)-Sugar Nucleosides. Nucleosides Nucleotides 1992, 11, 821–834. [Google Scholar] [CrossRef]
  22. Tera, M.; Glasauer, S.M.K.; Luedtke, N.W. In Vivo Incorporation of Azide Groups into DNA by Using Membrane-Permeable Nucleotide Triesters. ChemBioChem 2018, 19, 1939–1943. [Google Scholar] [CrossRef]
  23. Stubbe, J.; van der Donk, W.A. Ribonucleotide reductases: Radical enzymes with suicidal tendencies. Chem. Biol. 1995, 2, 793–801. [Google Scholar] [CrossRef]
  24. Stubbe, J.; Nocera, D.G. Radicals in Biology: Your Life Is in Their Hands. J. Am. Chem. Soc. 2021, 143, 13463–13472. [Google Scholar] [CrossRef]
  25. Meisenheimer, K.M.; Koch, T.H. Photocross-Linking of Nucleic Acids to Associated Proteins. Crit. Rev. Biochem. Mol. Biol. 1997, 32, 101–140. [Google Scholar] [CrossRef]
  26. Sirivolu, V.R.; Vernekar, S.K.V.; Ilina, T.; Myshakina, N.S.; Parniak, M.A.; Wang, Z. Clicking 3′-Azidothymidine into Novel Potent Inhibitors of Human Immunodeficiency Virus. J. Med. Chem. 2013, 56, 8765–8780. [Google Scholar] [CrossRef]
  27. Wu, J.; Yu, W.; Fu, L.; He, W.; Wang, Y.; Chai, B.; Song, C.; Chang, J. Design, synthesis, and biological evaluation of new 2′-deoxy-2′-fluoro-4′-triazole cytidine nucleosides as potent antiviral agents. Eur. J. Med. Chem. 2013, 63, 739–745. [Google Scholar] [CrossRef]
  28. Kore, A.R.; Charles, I. Click Chemistry Based Functionalizations of Nucleoside, Nucleotide and Nucleic Acids. Curr. Org. Chem. 2013, 17, 2164–2191. [Google Scholar] [CrossRef]
  29. El-Sagheer, A.H.; Brown, T. Click chemistry with DNA. Chem. Soc. Rev. 2010, 39, 1388–1405. [Google Scholar] [CrossRef]
  30. Fomich, M.A.; Kvach, M.V.; Navakouski, M.J.; Weise, C.; Baranovsky, A.V.; Korshun, V.A.; Shmanai, V.V. Azide phosphoramidite in direct synthesis of azide-modified oligonucleotides. Org. Lett. 2014, 16, 4590–4593. [Google Scholar] [CrossRef]
  31. Krause, A.; Hertl, A.; Muttach, F.; Jäschke, A. Phosphine-Free Stille–Migita Chemistry for the Mild and Orthogonal Modification of DNA and RNA. Chem. Eur. J. 2014, 20, 16613–16619. [Google Scholar] [CrossRef]
  32. Fauster, K.; Hartl, M.; Santner, T.; Aigner, M.; Kreutz, C.; Bister, K.; Ennifar, E.; Micura, R. 2′-Azido RNA, a versatile tool for chemical biology: Synthesis, X-ray structure, siRNA applications, click labeling. ACS Chem. Biol. 2012, 7, 581–589. [Google Scholar] [CrossRef]
  33. Beyer, C.; Wagenknecht, H.-A. In situ azide formation and “click” reaction of nile red with DNA as an alternative postsynthetic route. Chem. Commun. 2010, 46, 2230–2231. [Google Scholar] [CrossRef]
  34. Wada, T.; Mochizuki, A.; Higashiya, S.; Tsuruoka, H.; Kawahara, S.-i.; Ishikawa, M.; Sekine, M. Synthesis and properties of 2-azidodeoxyadenosine and its incorporation into oligodeoxynucleotides. Tetrahedron Lett. 2001, 42, 9215–9219. [Google Scholar] [CrossRef]
  35. Weisbrod, S.H.; Baccaro, A.; Marx, A. DNA Conjugation by Staudinger Ligation. Nucleic Acids Symp. Ser. 2008, 52, 383–384. [Google Scholar] [CrossRef]
  36. Jawalekar, A.M.; Meeuwenoord, N.; Cremers, J.S.; Overkleeft, H.S.; van der Marel, G.A.; Rutjes, F.P.; van Delft, F.L. Conjugation of nucleosides and oligonucleotides by [3 + 2] cycloaddition. J. Org. Chem. 2008, 73, 287–290. [Google Scholar] [CrossRef]
  37. Coppola, C.; Simeone, L.; De Napoli, L.; Montesarchio, D. On the Compatibility of Azides in Phosphoramidite-Based Couplings: Synthesis of a Novel, Convertible Azido-Functionalized CyPLOS Analogue. Eur. J. Org. Chem. 2011, 2011, 1155–1165. [Google Scholar] [CrossRef]
  38. Ozols, K.; Cīrule, D.; Novosjolova, I.; Stepanovs, D.; Liepinsh, E.; Bizdēna, Ē.; Turks, M. Development of N6-methyl-2-(1,2,3-triazol-1-yl)-2′-deoxyadenosine as a novel fluorophore and its application in nucleotide synthesis. Tetrahedron Lett. 2016, 57, 1174–1178. [Google Scholar] [CrossRef]
  39. Zayas, J.; Annoual, M.; Das, J.K.; Felty, Q.; Gonzalez, W.G.; Miksovska, J.; Sharifai, N.; Chiba, A.; Wnuk, S.F. Strain Promoted Click Chemistry of 2- or 8-Azidopurine and 5-Azidopyrimidine Nucleosides and 8-Azidoadenosine Triphosphate with Cyclooctynes. Application to Living Cell Fluorescent Imaging. Bioconjugate Chem. 2015, 26, 1519–1532. [Google Scholar] [CrossRef]
  40. Wen, Z.; Tuttle, P.R.; Howlader, A.H.; Vasilyeva, A.; Gonzalez, L.; Tangar, A.; Lei, R.; Laverde, E.E.; Liu, Y.; Miksovska, J.; et al. Fluorescent 5-Pyrimidine and 8-Purine Nucleosides Modified with an N-Unsubstituted 1,2,3-Triazol-4-yl Moiety. J. Org. Chem. 2019, 84, 3624–3631. [Google Scholar] [CrossRef]
  41. Tera, M.; Luedtke, N.W. Chapter Nineteen—Cross-linking cellular nucleic acids via a target-directing double click reagent. In Methods in Enzymology; Chenoweth, D.M., Ed.; Academic Press: Cambridge, MA, USA, 2020; Volume 641, pp. 433–457. [Google Scholar]
  42. Kotra, L.P.; Wang, P.; Bartlett, M.G.; Shanmuganathan, K.; Xu, Z.; Cavalcanti, S.; Newton, M.G.; Chu, C.K. 4-Azido-2-pyrimidinone Nucleosides and Related Chemistry. J. Org. Chem. 1997, 62, 7267–7271. [Google Scholar] [CrossRef]
  43. Lakshman, M.K.; Singh, M.K.; Parrish, D.; Balachandran, R.; Day, B.W. Azide-tetrazole equilibrium of C-6 azidopurine nucleosides and their ligation reactions with alkynes. J. Org. Chem. 2010, 75, 2461–2473. [Google Scholar] [CrossRef]
  44. Ren, Z.; Luo, H.; Yu, Z.; Song, J.; Liang, L.; Wang, L.; Wang, H.; Cui, G.; Liu, Y.; Wang, J.; et al. A Randomized, Open-label, Controlled Clinical Trial of Azvudine Tablets in the Treatment of Mild and Common COVID-19, A Pilot Study. Adv. Sci. 2020, 7, 2001435. [Google Scholar] [CrossRef] [PubMed]
  45. Yu, B.; Chang, J. Azvudine (FNC): A promising clinical candidate for COVID-19 treatment. Signal transduct. target. ther. 2020, 5, 236. [Google Scholar] [CrossRef]
  46. Klumpp, K.; Lévêque, V.; Le Pogam, S.; Ma, H.; Jiang, W.R.; Kang, H.; Granycome, C.; Singer, M.; Laxton, C.; Hang, J.Q.; et al. The novel nucleoside analog R1479 (4′-azidocytidine) is a potent inhibitor of NS5B-dependent RNA synthesis and hepatitis C virus replication in cell culture. J. Biol. Chem. 2006, 281, 3793–3799. [Google Scholar] [CrossRef] [PubMed]
  47. Kotra, L.P.; Manouilov, K.K.; Cretton-Scott, E.; Sommadossi, J.P.; Boudinot, F.D.; Schinazi, R.F.; Chu, C.K. Synthesis, Biotransformation, and Pharmacokinetic Studies of 9-(β-d-Arabinofuranosyl)-6-azidopurine:  A Prodrug for Ara-A Designed To Utilize the Azide Reduction Pathway. J. Med. Chem. 1996, 39, 5202–5207. [Google Scholar] [CrossRef] [PubMed]
  48. Cosyn, L.; Palaniappan, K.K.; Kim, S.-K.; Duong, H.T.; Gao, Z.-G.; Jacobson, K.A.; Van Calenbergh, S. 2-Triazole-Substituted Adenosines:  A New Class of Selective A3 Adenosine Receptor Agonists, Partial Agonists, and Antagonists. J. Med. Chem. 2006, 49, 7373–7383. [Google Scholar] [CrossRef] [PubMed]
  49. Fabrega, C.; Garcia, R.G.; Diaz, A.R.; Eritja, R. Studies on the Synthesis of Oligonucleotides Containing Photoreactive Nucleosides: 2-Azido-2’-Deoxyinosine and 8-Azido-2′-Deoxyadenosine. Biol. Chem. 1998, 379, 527–533. [Google Scholar] [CrossRef] [PubMed]
  50. Lakshman, M.K.; Kumar, A.; Balachandran, R.; Day, B.W.; Andrei, G.; Snoeck, R.; Balzarini, J. Synthesis and Biological Properties of C-2 Triazolylinosine Derivatives. J. Org. Chem. 2012, 77, 5870–5883. [Google Scholar] [CrossRef] [PubMed]
  51. Lioux, T.; Gosselin, G.; Mathé, C. Azido/Tetrazole Tautomerism in 2-Azidoadenineβ-D-Pentofuranonucleoside Derivatives. Eur. J. Org. Chem. 2003, 2003, 3997–4002. [Google Scholar] [CrossRef]
  52. Housri, N.; Yarchoan, R.; Kaushal, A. Radiotherapy for patients with the human immunodeficiency virus: Are special precautions necessary? Cancer 2010, 116, 273–283. [Google Scholar] [CrossRef] [PubMed]
  53. Coucke, P.A.; Cottin, E.; Decosterd, L.A. Simultaneous alteration of de novo and salvage pathway to the deoxynucleoside triphosphate pool by (E)-2’-Deoxy-(fluoromethylene)cytidine (FMdC) and zidovudine (AZT) results in increased radiosensitivity in vitro. Acta Oncol. 2007, 46, 612–620. [Google Scholar] [CrossRef] [PubMed]
  54. Zhou, F.-X.; Liao, Z.-K.; Dai, J.; Xiong, J.; Xie, C.-H.; Luo, Z.-G.; Liu, S.-Q.; Zhou, Y.-F. Radiosensitization effect of zidovudine on human malignant glioma cells. Biochem. Biophys. Res. Commun. 2007, 354, 351–356. [Google Scholar] [CrossRef] [PubMed]
  55. Liao, Z.-K.; Zhou, F.-X.; Luo, Z.-G.; Zhang, W.-J.; Jie, X.; Bao, J.; Han, G.; Zhang, M.-S.; Xie, C.-H.; Zhou, Y.-F. Radioactivation of hTERT promoter in larynx squamous carcinoma cells: An ‘indirected-activator’ strategy in radio-gene-therapy. Oncol. Rep. 2008, 19, 281–286. [Google Scholar] [PubMed]
  56. Sjoberg, B.M.; Graslund, A.; Eckstein, F. A substrate radical intermediate in the reaction between ribonucleotide reductase from Escherichia coli and 2’-azido-2’-deoxynucleoside diphosphates. J. Biol. Chem. 1983, 258, 8060–8067. [Google Scholar] [CrossRef]
  57. Salowe, S.; Bollinger, J.M., Jr.; Ator, M.; Stubbe, J.; McCracken, J.; Peisach, J.; Samano, M.C.; Robins, M.J. Alternative model for mechanism-based inhibition of Escherichia coli ribonucleotide reductase by 2′-azido-2′-deoxyuridine 5′-diphosphate. Biochemistry 1993, 32, 12749–12760. [Google Scholar] [CrossRef] [PubMed]
  58. van der Donk, W.A.; Stubbe, J.; Gerfen, G.J.; Bellew, B.F.; Griffin, R.G. EPR Investigations of the Inactivation of E. coli Ribonucleotide Reductase with 2′-Azido-2′-deoxyuridine 5′-Diphosphate: Evidence for the Involvement of the Thiyl Radical of C225-R1. J. Am. Chem. Soc. 1995, 117, 8908–8916. [Google Scholar] [CrossRef]
  59. Fritscher, J.; Artin, E.; Wnuk, S.; Bar, G.; Robblee, J.H.; Kacprzak, S.; Kaupp, M.; Griffin, R.G.; Bennati, M.; Stubbe, J. Structure of the Nitrogen-Centered Radical Formed during Inactivation of E. coli Ribonucleotide Reductase by 2′-Azido-2′-deoxyuridine-5′-diphosphate: Trapping of the 3′-Ketonucleotide. J. Am. Chem. Soc. 2005, 127, 7729–7738. [Google Scholar] [CrossRef] [PubMed]
  60. Zipse, H.; Artin, E.; Wnuk, S.; Lohman, G.J.S.; Martino, D.; Griffin, R.G.; Kacprzak, S.; Kaupp, M.; Hoffman, B.; Bennati, M.; et al. Structure of the Nucleotide Radical Formed during Reaction of CDP/TTP with the E441Q-a2b2 of E. coli Ribonucleotide Reductase. J. Am. Chem. Soc. 2009, 131, 200–211. [Google Scholar] [CrossRef]
  61. Wnuk, S.F.; Chowdhury, S.M.; Garcia, P.I., Jr.; Robins, M.J. Stereo-defined Synthesis of O3’-Labeled Uracil Nucleosides. 3′-[17O]-2′-Azido-2′-deoxyuridine 5′-Diphosphate as a Probe for the Mechanism of Inactivation of Ribonucleotide Reductases. J. Org. Chem. 2002, 67, 1816–1819. [Google Scholar] [CrossRef] [PubMed]
  62. Kim, S.; Joe, G.H.; Do, J.Y. Highly efficient intramolecular addition of aminyl radicals to carbonyl groups: A new ring expansion reaction leading to lactams. J. Am. Chem. Soc. 1993, 115, 3328–3329. [Google Scholar] [CrossRef]
  63. Kim, S.; Joe, G.H.; Do, J.Y. Novel Radical Cyclizations of Alkyl Azides. A New Route to N-Heterocycles. J. Am. Chem. Soc. 1994, 116, 5521–5522. [Google Scholar] [CrossRef]
  64. Ashley, G.W.; Lawrence, C.C.; Stubbe, J.; Robins, M.J. Mechanism based inactivation of the adenosylcobalamin-dependent ribonucleotide reductase from L. leichmannii by 2′-ara-2′-azido-2′-deoxy adenosine-5′-triphosphate: Observation of paramagnetic intermediates. Tetrahedron 1997, 53, 12005–12016. [Google Scholar] [CrossRef]
  65. Pereira, S.; Fernandes, P.A.; Ramos, M.J. Theoretical study of ribonucleotide reductase mechanism-based inhibition by 2’-azido-2’-deoxyribonucleoside 5’-diphosphates. J. Comput. Chem. 2004, 25, 227–237. [Google Scholar] [CrossRef] [PubMed]
  66. Bräse, S.; Gil, C.; Knepper, K.; Zimmermann, V. Organic Azides: An Exploding Diversity of a Unique Class of Compounds. Angew. Chem. Int. Ed. 2005, 44, 5188–5240. [Google Scholar] [CrossRef] [PubMed]
  67. Benati, L.; Bencivenni, G.; Leardini, R.; Minozzi, M.; Nanni, D.; Scialpi, R.; Spagnolo, P.; Zanardi, G. Thermal Reactions of Tributyltin Hydride with a-Azido Esters: Unexpected Intervention of Tin Triazene Adducts under Both Nonradical and Radical Conditions. J. Org. Chem. 2005, 70, 3046–3053. [Google Scholar] [CrossRef] [PubMed]
  68. Roberts, B.P.; Winter, J.N. Electron spin resonance studies of radicals derived from organic azides. J. Chem. Soc. Perk. Trans. 2 1979, 1353–1361. [Google Scholar] [CrossRef]
  69. Benati, L.; Bencivenni, G.; Leardini, R.; Minozzi, M.; Nanni, D.; Scialpi, R.; Spagnolo, P.; Zanardi, G. Radical Reduction of Aromatic Azides to Amines with Triethylsilane. J. Org. Chem. 2006, 71, 5822–5825. [Google Scholar] [CrossRef] [PubMed]
  70. Postigo, A.; Kopsov, S.; Ferreri, C.; Chatgilialoglu, C. Radical Reactions in Aqueous Medium Using (Me3Si)3SiH. Org. Lett. 2007, 9, 5159–5162. [Google Scholar] [CrossRef] [PubMed]
  71. Dang, T.P.; Sobczak, A.J.; Mebel, A.M.; Chatgilialoglu, C.; Wnuk, S.F. Investigation of reactions postulated to occur during inhibition of ribonucleotide reductases by 2′-azido-2′-deoxynucleotides. Tetrahedron 2012, 68, 5655–5667. [Google Scholar] [CrossRef] [PubMed]
  72. Robins, M.J.; Guo, Z.; Samano, M.C.; Wnuk, S.F. Biomimetic Simulation of Free Radical-Initiated Cascade Reactions Postulated To Occur at the Active Site of Ribonucleotide Reductases. J. Am. Chem. Soc. 1999, 121, 1425–1433. [Google Scholar] [CrossRef]
  73. Robins, M.J.; Wnuk, S.F.; Hernandez-Thirring, A.E.; Samano, M.C. Nucleic Acid-Related Compounds. 91. Biomimetic Reactions Are in Harmony with Loss of 2’-Substituents as Free Radicals (Not Anions) during Mechanism-Based Inactivation of Ribonucleotide Reductases. Differential Interactions of Azide, Halogen, and Alkylthio Groups with Tributylstannane and Triphenylsilane. J. Am. Chem. Soc. 1996, 118, 11341–11348. [Google Scholar]
  74. Samano, M.C.; Robins, M.J. Efficient reduction of azides to amines with tributylstannane. high-yield syntheses of amino and diamino deoxynucleosides. Tetrahedron Lett. 1991, 32, 6293–6296. [Google Scholar] [CrossRef]
  75. Hayon, E.M. Formation and Reactions of Azide Radicals; Accesssion Number: AD0750330; Army Natick Laboratories, National Technical Information Service, U.S. Department of Commerce: Natick, MA, USA, 1972; pp. 45–56. [Google Scholar]
  76. Adjei, D.; Reyes, Y.; Kumar, A.; Ward, S.; Denisov, S.A.; Alahmadi, M.; Sevilla, M.D.; Wnuk, S.F.; Mostafavi, M.; Adhikary, A. Pathways of the Dissociative Electron Attachment Observed in 5- and 6-Azidomethyluracil Nucleosides: Nitrogen (N2) Elimination vs Azide Anion (N3) Elimination. J. Phys. Chem. B 2023, 127, 1563–1571. [Google Scholar] [CrossRef] [PubMed]
  77. Živanov, S.; Ibanescu, B.C.; Paech, M.; Poffet, M.; Baettig, P.; Sergenton, A.C.; Grimme, S.; Allan, M. Dissociative Electron Attachment and Electron Energy-Loss Spectra of Phenyl Azide. J. Phys. B At. Mol. Opt. Phys. 2007, 40, 101–107. [Google Scholar] [CrossRef]
  78. Narayanan, S.J.J.; Tripathi, D.; Verma, P.; Adhikary, A.; Dutta, A.K. Secondary Electron Attachment-Induced Radiation Damage to Genetic Materials. ACS Omega 2023, 8, 10669–10689. [Google Scholar] [CrossRef] [PubMed]
  79. Adhikary, A.; Khanduri, D.; Pottiboyina, V.; Rice, C.T.; Sevilla, M.D. Formation of Aminyl Radicals on Electron Attachment to AZT: Abstraction from the Sugar Phosphate Backbone versus One-Electron Oxidation of Guanine. J. Phys. Chem. B 2010, 114, 9289–9299. [Google Scholar] [CrossRef]
  80. Mudgal, M.; Dang, T.P.; Sobczak, A.J.; Lumpuy, D.A.; Dutta, P.; Ward, S.; Ward, K.; Alahmadi, M.; Kumar, A.; Sevilla, M.D.; et al. Site of Azido Substitution in the Sugar Moiety of Azidopyrimidine Nucleosides Influences the Reactivity of Aminyl Radicals Formed by Dissociative Electron Attachment. J. Phys. Chem. B 2020, 124, 11357–11370. [Google Scholar] [CrossRef] [PubMed]
  81. Zheng, L.; Dai, X.; Su, H.; Greenberg, M.M. Independent Generation and Time-Resolved Detection of 2′-Deoxyguanosin-N2-yl Radicals. Angew. Chem. Int. Ed. 2020, 59, 13406–13413. [Google Scholar] [CrossRef] [PubMed]
  82. Chatgilialoglu, C.; D’Angelantonio, M.; Guerra, M.; Kaloudis, P.; Mulazzani, Q.G. A reevaluation of the ambident reactivity of the guanine moiety towards hydroxyl radicals. Angew. Chem. Int. Ed. 2009, 48, 2214–2217. [Google Scholar] [CrossRef] [PubMed]
  83. Adhikary, A.; Kumar, A.; Becker, D.; Sevilla, M.D. The Guanine Cation Radical:  Investigation of Deprotonation States by ESR and DFT. J. Phys. Chem. B 2006, 110, 24171–24180. [Google Scholar] [CrossRef] [PubMed]
  84. Adhikary, A.; Kumar, A.; Munafo, S.A.; Khanduri, D.; Sevilla, M.D. Prototropic equilibria in DNA containing one-electron oxidized GC: Intra-duplex vs. duplex to solvent deprotonation. Phys. Chem. Chem. Phys. 2010, 12, 5353–5368. [Google Scholar] [CrossRef] [PubMed]
  85. Mudgal, M.; Rishi, S.; Lumpuy, D.A.; Curran, K.A.; Verley, K.L.; Sobczak, A.J.; Dang, T.P.; Sulimoff, N.; Kumar, A.; Sevilla, M.D.; et al. Prehydrated One-Electron Attachment to Azido-Modified Pentofuranoses: Aminyl Radical Formation, Rapid H-Atom Transfer, and Subsequent Ring Opening. J. Phys. Chem. B 2017, 121, 4968–4980. [Google Scholar] [CrossRef] [PubMed]
  86. Reyes, Y.; Barbieri, M.; Pathak, R.; Ward, S.; Sevilla, M.D.; Adhikary, A.; Wnuk, S.F. Studies of amino and azido derivatives of sesquiterpene lactones and their potential augmentation for radiation damage to cancer cells. In Proceedings of the 263rd ACS National Meeting, Medicinal Chemistry Division, Abstract 3647326, San Diego, CA, USA, 20–24 March 2022. [Google Scholar]
  87. Liang, Y.; Wen, Z.; Cabrera, M.; Howlader, A.H.; Wnuk, S.F. Purines. In Science of Synthesis: Houben-Weyl Methods of Molecular Transformations: Knowledge Updates 2020/1; Christmann, M., Huang, Z., Jiang, X., Li, J.J., Oestreich, M., Petersson, E.J., Schaumann, E., Wang, M., Eds.; Georg Thieme Verlag: Stuttgart, Germany, 2020; pp. 195–384. [Google Scholar]
  88. Wen, Z.; Peng, J.; Tuttle, P.R.; Ren, Y.; Garcia, C.; Debnath, D.; Rishi, S.; Hanson, C.; Ward, S.; Kumar, A.; et al. Electron-Mediated Aminyl and Iminyl Radicals from C5 Azido-Modified Pyrimidine Nucleosides Augment Radiation Damage to Cancer Cells. Org. Lett. 2018, 20, 7400–7404. [Google Scholar] [CrossRef] [PubMed]
  89. Hu, B.; DiMagno, S.G. Reactivities of vinyl azides and their recent applications in nitrogen heterocycle synthesis. Org. Biomol. Chem. 2015, 13, 3844–3855. [Google Scholar] [CrossRef] [PubMed]
  90. Neef, A.B.; Luedtke, N.W. An azide-modified nucleoside for metabolic labeling of DNA. ChemBioChem 2014, 15, 789–793. [Google Scholar] [CrossRef] [PubMed]
  91. Ren, X.; El-Sagheer, A.H.; Brown, T. Azide and trans-cyclooctene dUTPs: Incorporation into DNA probes and fluorescent click-labelling. Analyst 2015, 140, 2671–2678. [Google Scholar] [CrossRef] [PubMed]
  92. von Sonntag, C. Free-Radical-Induced DNA Damage and Its Repair; Springer: Berlin/Heidelberg, Germany, 2006. [Google Scholar]
  93. Reyes, Y.; Mebel, A.; Wnuk, S.F. 6-azido and 6-azidomethyl uracil nucleosides. Nucleosides Nucleotides Nucl. Acids 2023, 43, 453–471. [Google Scholar] [CrossRef] [PubMed]
  94. de Cabrera, M.; Reyes, Y.; Ward, S.; Alahmadi, M.; Barbolovici, A.; Kumar, A.; Sevilla, M.D.; Wnuk, S.F.; Adhikary, A. Characterization of nitrogen-centered radicals formed via dissociative electron attachment from 4-, 5-, and 6-azido substituted pyrimidine nucleosides. In Proceedings of the 263rd ACS National Meeting, Physical Chemistry Division, Abstract 3649777, San Diego, CA, USA, 20–24 March 2022. [Google Scholar]
  95. De Cabrera, M. Nucleoside analogues for positron emission tomography imaging and to study radiation mediated generation of radicals from azides. Ph.D. Thesis, Florida international University, Miami, FL, USA, 2019. [Google Scholar]
  96. Balintová, J.; Špaček, J.; Pohl, R.; Brázdová, M.; Havran, L.; Fojta, M.; Hocek, M. Azidophenyl as a click-transformable redox label of DNA suitable for electrochemical detection of DNA–protein interactions. Chem. Sci. 2015, 6, 575–587. [Google Scholar] [CrossRef] [PubMed]
  97. Danhel, A.; Trosanova, Z.; Balintova, J.; Havran, L.; Hocek, M.; Barek, J.; Fojta, M. Voltammetric analysis of 5-(4-Azidophenyl)-2′-deoxycytidine nucleoside and azidophenyl-labelled single- and double-stranded DNAs. Electrochim. Acta 2016, 215, 72–83. [Google Scholar] [CrossRef]
  98. Daňhel, A.; Trošanová, Z.; Balintová, J.; Simonová, A.; Pospíšil, L.; Cvačka, J.; Hocek, M.; Fojta, M. Electrochemical reduction of azidophenyl-deoxynucleoside conjugates at mercury surface. Electrochim. Acta 2018, 259, 377–385. [Google Scholar] [CrossRef]
  99. Kuttappan-Nair, V.; Samson-Thibault, F.; Wagner, J.R. Generation of 2′-Deoxyadenosine N6-Aminyl Radicals from the Photolysis of Phenylhydrazone Derivatives. Chem. Res. Toxicol. 2010, 23, 48–54. [Google Scholar] [CrossRef] [PubMed]
  100. Zheng, L.; Lin, L.; Qu, K.; Adhikary, A.; Sevilla, M.D.; Greenberg, M.M. Independent Photochemical Generation and Reactivity of Nitrogen-Centered Purine Nucleoside Radicals from Hydrazines. Org. Lett. 2017, 19, 6444–6447. [Google Scholar] [CrossRef] [PubMed]
  101. Zheng, L.; Griesser, M.; Pratt, D.A.; Greenberg, M.M. Aminyl Radical Generation via Tandem Norrish Type I Photocleavage, β-Fragmentation: Independent Generation and Reactivity of the 2’-Deoxyadenosin- N6-yl Radical. J. Org. Chem. 2017, 82, 3571–3580. [Google Scholar] [CrossRef] [PubMed]
  102. Zheng, L.; Greenberg, M.M. DNA Damage Emanating From a Neutral Purine Radical Reveals the Sequence Dependent Convergence of the Direct and Indirect Effects of γ-Radiolysis. J. Am. Chem. Soc. 2017, 139, 17751–17754. [Google Scholar] [CrossRef] [PubMed]
  103. Zheng, L.; Greenberg, M.M. Traceless Tandem Lesion Formation in DNA from a Nitrogen-Centered Purine Radical. J. Am. Chem. Soc. 2018, 140, 6400–6407. [Google Scholar] [CrossRef] [PubMed]
  104. Sun, H.; Zheng, L.; Greenberg, M.M. Independent Generation of Reactive Intermediates Leads to an Alternative Mechanism for Strand Damage Induced by Hole Transfer in Poly(dA–T) Sequences. J. Am. Chem. Soc. 2018, 140, 11308–11316. [Google Scholar] [CrossRef] [PubMed]
  105. Naumov, S.; von Sonntag, C. Guanine-derived radicals: Dielectric constant-dependent stability and UV/Vis spectral properties: A DFT study. Radiat. Res. 2008, 169, 364–372. [Google Scholar] [CrossRef] [PubMed]
  106. Zheng, L.; Greenberg, M.M. Independent Generation and Reactivity of 2′-Deoxyguanosin-N1-yl Radical. J. Org. Chem. 2020, 85, 8665–8672. [Google Scholar] [CrossRef] [PubMed]
  107. Vrantza, D.; Kaloudis, P.; Leondiadis, L.; Gimisis, T.; Vougioukalakis, G.C.; Orfanopoulos, M.; Gasparutto, D.; Cadet, J.; Encinas, S.; Paris, C.; et al. Modification of Guanine with Photolabile N-Hydroxypyridine-2(1H)-thione: Monomer Synthesis, Oligonucleotide Elaboration, and Photochemical Studies. Helv. Chim. Acta 2006, 89, 2371–2386. [Google Scholar] [CrossRef]
  108. Kaloudis, P.; Paris, C.; Vrantza, D.; Encinas, S.; Pérez-Ruiz, R.; Miranda, M.A.; Gimisis, T. Photolabile N-hydroxypyrid-2(1H)-one derivatives of guanine nucleosides: A new method for independent guanine radical generation. Org. Biomol. Chem. 2009, 7, 4965–4972. [Google Scholar] [CrossRef] [PubMed]
  109. Peng, H.; Jie, J.; Mortimer, I.P.; Ma, Z.; Su, H.; Greenberg, M.M. Reactivity and DNA Damage by Independently Generated 2′-Deoxycytidin-N4-yl Radical. J. Am. Chem. Soc. 2021, 143, 14738–14747. [Google Scholar] [CrossRef] [PubMed]
  110. Peng, H.; Vu, S.; Retes, P.; Ward, S.; Kumar, A.; Sevilla, M.D.; Adhikary, A.; Greenberg, M.M. Photochemical and Single Electron Transfer Generation of 2′-Deoxycytidin-N4-yl Radical from Oxime Esters. J. Org. Chem. 2023, 88, 7381–7390. [Google Scholar] [CrossRef]
  111. Adhikary, A.; Kumar, A.; Bishop, C.T.; Wiegand, T.J.; Hindi, R.M.; Adhikary, A.; Sevilla, M.D. π-Radical to σ-Radical Tautomerization in One-Electron-Oxidized 1-Methylcytosine and Its Analogs. J. Phys. Chem. B 2015, 119, 11496–11505. [Google Scholar] [CrossRef] [PubMed]
  112. Rodríguez, M.A. Light-Induced Iminyl Radicals: Generation and Synthetic Applications. In Molecular Photochemistry: Various Aspects; Satayen, S., Ed.; IntechOpen: London, UK, 2012; pp. 265–282. [Google Scholar]
Figure 2. Model azidonucleosides with the cysteinate or vicinal dithiol unit attached to C2′ or C5′ to study the assumed ring closure reactions between the thiyl radical and the azido group. Plausible intramolecular interactions of the thiyl radical with an azido group at C3′ for adenosine-derived substrate 9 [71].
Figure 2. Model azidonucleosides with the cysteinate or vicinal dithiol unit attached to C2′ or C5′ to study the assumed ring closure reactions between the thiyl radical and the azido group. Plausible intramolecular interactions of the thiyl radical with an azido group at C3′ for adenosine-derived substrate 9 [71].
Molecules 29 02310 g002
Figure 3. Differential interaction of azidonucleosides with tributylstannane and triphenylsilane. Reduction versus elimination of azido group from C2′ position [73].
Figure 3. Differential interaction of azidonucleosides with tributylstannane and triphenylsilane. Reduction versus elimination of azido group from C2′ position [73].
Molecules 29 02310 g003
Scheme 1. Schematic representation of the steps involved in the dissociate electron attachment process involving the neutral azide molecule, RN3. The first step, attachment, leads to the formation of a transient negative ion (TNI), [(RN3]*. This is followed by dissociation [78].
Scheme 1. Schematic representation of the steps involved in the dissociate electron attachment process involving the neutral azide molecule, RN3. The first step, attachment, leads to the formation of a transient negative ion (TNI), [(RN3]*. This is followed by dissociation [78].
Molecules 29 02310 sch001
Figure 4. The formation and reactivity of the aminyl radical generated upon the addition of radiation-produced electrons to AZT [79].
Figure 4. The formation and reactivity of the aminyl radical generated upon the addition of radiation-produced electrons to AZT [79].
Molecules 29 02310 g004
Figure 5. Types of reactions undergone by π-RNH· attached to a primary (path A), secondary (path B), or tertiary (path C) alkyl carbon in the ribose sugar of pyrimidine nucleosides.
Figure 5. Types of reactions undergone by π-RNH· attached to a primary (path A), secondary (path B), or tertiary (path C) alkyl carbon in the ribose sugar of pyrimidine nucleosides.
Molecules 29 02310 g005
Figure 6. The formation of the ring-opened C4· 42 via the ring opening of a C5· intermediate in 1-Me-2-Azlyxo (38) and thermodynamically stable -OCH2· 45 with an intact ribofuranose ring in 1-Me-2-Azribo (43).
Figure 6. The formation of the ring-opened C4· 42 via the ring opening of a C5· intermediate in 1-Me-2-Azlyxo (38) and thermodynamically stable -OCH2· 45 with an intact ribofuranose ring in 1-Me-2-Azribo (43).
Molecules 29 02310 g006
Figure 7. The intramolecular formation of the neutral tertiary carbon-centered radical 48 from azido-PTL 46. Structure of α-azidoalkyl radical 49 from azido-DhL.
Figure 7. The intramolecular formation of the neutral tertiary carbon-centered radical 48 from azido-PTL 46. Structure of α-azidoalkyl radical 49 from azido-DhL.
Molecules 29 02310 g007
Figure 8. The formation of the neutral aminyl radical 51 by radiation-produced prehydrated electron attachment to 5-AmdU and its subsequent H-atom abstraction reaction with 5-AmdU to form iminyl radical 53 via the intermediary α-azidoalkyl radical 52 [88].
Figure 8. The formation of the neutral aminyl radical 51 by radiation-produced prehydrated electron attachment to 5-AmdU and its subsequent H-atom abstraction reaction with 5-AmdU to form iminyl radical 53 via the intermediary α-azidoalkyl radical 52 [88].
Molecules 29 02310 g008
Figure 9. Tautomerization of π aminyl radical, generated from AvdU, to σ-iminyl radical [88].
Figure 9. Tautomerization of π aminyl radical, generated from AvdU, to σ-iminyl radical [88].
Molecules 29 02310 g009
Figure 10. The formation of allylic radical 59 via the unexpected loss of azide as an anion (N3¯) from the azide anion radical intermediate 58 via dissociative electron attachment [76].
Figure 10. The formation of allylic radical 59 via the unexpected loss of azide as an anion (N3¯) from the azide anion radical intermediate 58 via dissociative electron attachment [76].
Molecules 29 02310 g010
Figure 11. Nitrogen-centered radicals formed via dissociative electron attachment to 5-azidouridine 60, 6-azidouridine 61, and 4-azidopyrimidine analogue 62 and its cyclic tetrazolo derivative 63.
Figure 11. Nitrogen-centered radicals formed via dissociative electron attachment to 5-azidouridine 60, 6-azidouridine 61, and 4-azidopyrimidine analogue 62 and its cyclic tetrazolo derivative 63.
Molecules 29 02310 g011
Figure 12. Azidophenyl-modified nucleosides and their electrochemical reduction on mercury surface to the aminyl radicals.
Figure 12. Azidophenyl-modified nucleosides and their electrochemical reduction on mercury surface to the aminyl radicals.
Molecules 29 02310 g012
Figure 13. Plausible generation of dAdo-6-N-aminyl radical 72 by photolysis of dAdo-6-N-arylhydrazones 71a and 71b and its subsequent chemistry to form 2-imino-dAdo derivative 74 [99].
Figure 13. Plausible generation of dAdo-6-N-aminyl radical 72 by photolysis of dAdo-6-N-arylhydrazones 71a and 71b and its subsequent chemistry to form 2-imino-dAdo derivative 74 [99].
Molecules 29 02310 g013
Figure 14. The plausible generation of dAdo-6-N-aminyl radical 72 by the photolysis of ketone 78 as well as hydrazine 75 and the subsequent chemistry to form 8-amino-dAdo derivative 76.
Figure 14. The plausible generation of dAdo-6-N-aminyl radical 72 by the photolysis of ketone 78 as well as hydrazine 75 and the subsequent chemistry to form 8-amino-dAdo derivative 76.
Molecules 29 02310 g014
Figure 15. Photochemical generation of guanosyl radicals from synthetic precursors.
Figure 15. Photochemical generation of guanosyl radicals from synthetic precursors.
Molecules 29 02310 g015
Figure 16. The generation of 2′-deoxycytidin-N4-yl iminyl radical 88 by photolysis and radiation-produced prehydrated electron attachment approaches.
Figure 16. The generation of 2′-deoxycytidin-N4-yl iminyl radical 88 by photolysis and radiation-produced prehydrated electron attachment approaches.
Molecules 29 02310 g016
Figure 17. Formation of cytidine π-aminyl radicals 93 and 94 that tautomerize to σ-iminyl radical 88 upon one-electron oxidation of cytosine nucleobase [111].
Figure 17. Formation of cytidine π-aminyl radicals 93 and 94 that tautomerize to σ-iminyl radical 88 upon one-electron oxidation of cytosine nucleobase [111].
Molecules 29 02310 g017
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

Reyes, Y.; Adhikary, A.; Wnuk, S.F. Nitrogen-Centered Radicals Derived from Azidonucleosides. Molecules 2024, 29, 2310. https://doi.org/10.3390/molecules29102310

AMA Style

Reyes Y, Adhikary A, Wnuk SF. Nitrogen-Centered Radicals Derived from Azidonucleosides. Molecules. 2024; 29(10):2310. https://doi.org/10.3390/molecules29102310

Chicago/Turabian Style

Reyes, Yahaira, Amitava Adhikary, and Stanislaw F. Wnuk. 2024. "Nitrogen-Centered Radicals Derived from Azidonucleosides" Molecules 29, no. 10: 2310. https://doi.org/10.3390/molecules29102310

Article Metrics

Back to TopTop