Sabrina Pantera,
Boris Illarionov
b,
Jing Chen
a,
Adelbert Bacher
c,
Markus Fischer
b and
Stefan Weber
*a
aInstitut für Physikalische Chemie, Albert-Ludwigs-Universität Freiburg, Albertstr. 21, 79104 Freiburg, Germany. E-mail: stefan.weber@pc.uni-freiburg.de
bInstitut für Lebensmittelchemie, Universität Hamburg, Grindelallee 117, 20146 Hamburg, Germany
cTUM School of Natural Sciences, Technische Universität München, Lichtenbergstr. 4, 85747 Garching, Germany
First published on 12th August 2025
6,7,8-Trimethyllumazine (TML) is a structural analog of the natural cofactor 6,7-dimethyl-8-ribityllumazine. Under basic conditions, TML undergoes a distinctive disproportionation reaction upon photoexcitation. The transiently formed radical pair can be investigated by photo-chemically induced dynamic nuclear polarization (photo-CIDNP) spectroscopy. In this contribution, the structure of the TML anion is analyzed systematically using NMR spectroscopy. Furthermore, the transiently formed TML radicals are investigated and their hyperfine structures elucidated by 1H and 13C photo-CIDNP spectroscopy. Experimental photo-CIDNP intensities are compared with isotropic hyperfine coupling constants from density functional theory (DFT) calculations. The results confirm the formation of an oxidized TML˙ radical and a reduced TMLH˙− radical, the latter potentially protonated at N1. Comparative analysis reveals a substantially different hyperfine structure of the formed radical species which is rationalized based on calculations of spin density distributions. The results provide important insights into photo-induced one-electron transfer reactions of 6,7-dimethyllumazines and their potential role in redox processes in biological systems. The detection and characterization of the oxidized TML˙ radical is of special interest as this oxidation state has not been satisfactorily described in the literature so far. Thus this contribution advances the understanding of the mechanism of formation and the structure of lumazine radicals.
![]() | ||
Fig. 1 Structure of lumazines: lumazine: R1 = R2 = R3 = H; 6,7-dimethyl-8-ribityllumazine: R1 = R2 = CH3, R3 = ribityl; 6,7,8-trimethyllumazine: R1 = R2 = R3 = CH3. |
In nature, DMRL was first identified in 1966 as a direct biosynthetic precursor of riboflavin.12 The final synthesis step, catalyzed by riboflavin synthase, involves the remarkable transfer of a four-carbon fragment between two DMRL molecules to form the riboflavin molecule.13–15 The lumazine synthase/riboflavin synthase complex has been extensively studied in the context of antibiotic development, as their inhibition disrupts riboflavin biosynthesis in microorganisms, see e.g. ref. 16–20. Since 1978, DMRL has also been known to function as a chromophore in a protein thereafter named lumazine protein (LumP) from the marine bacterium Photobacterium phosphoreum. This protein was found to form a complex with the fluorescent protein luciferase.21,22 The complex exhibits a blue-shifted bioluminescence due to Förster resonance energy transfer from luciferase to DMRL, as well as an increased quantum yield compared to unbound luciferase.23,24 Furthermore, DMRL has been identified as an additional cofactor in a recently discovered subgroup of the photolyase/cryptochrome family which is involved in e.g. DNA repair25 and various light-driven biological responses.26 In the FAD-binding protein cryptochrome B (CryB) from Rhodobacter sphaeroides, DMRL is located in the antenna-binding domain and broadens the absorbance section of the protein.27 Further investigation of another member of the photolyase/cryptochrome subgroup, the (6-4) photolyase protein B (PhrB) from Agrobacterium tumefaciens,28 indicates that DMRL plays a role surpassing the one of a simple antenna chromophore in this protein: DMRL evidently acts as a “photoprotective pigment” coupling the oxidation of the FAD cofactor with the reduction of DMRL under intense illumination.29
In general, lumazines can mediate one- and two-electron transfer reactions, as demonstrated by cyclovoltammetry with unsubstituted lumazine.30 This makes three oxidation states accessible: the oxidized lumazine, the one-electron reduced lumazine radical and the fully reduced lumazine. In this regard, they share a similar redox reactivity as flavins, which can access the same biologically relevant oxidation states.31
The first optical absorption spectra of lumazine radicals in aqueous solution were obtained using pulse radiolysis.32 Several studies have confirmed the formation of lumazine radicals in protein environment. DMRL bound to flavodoxin forms a radical upon dithionite titration.33 A 6,7-alkylated 5-ribityllumazine (6,7-(2,3-dimethylbutano)-N(8)-ribityllumazine-5′-monophosphate) generated an anionic radical bound to old-yellow enzyme under reduction with dithionite.34 Paulus et al. studied the wildtype lumazine protein from Photobacterium leiognathi as well as several mutants with DMRL and riboflavin as cofactors employing time-resolved absorption spectroscopy and derived kinetics of their photoreduction.35
Data on the lumazine radical obtained by magnetic resonance spectroscopy are scarce. Ehrenberg et al.36 conducted the first continuous-wave electron paramagnetic resonance (cw-EPR) study on DMRL and several derivatives at room temperature under acidic conditions. Amongst others, hyperfine couplings of a cationic TML radical protonated at N1 and N5 were reported.36 Westerling et al. later studied different 5-alkylated 5,6,7,8-tetrahydrolumazine radicals using cw-EPR at room temperature.37 The aforementioned study by Paulus et al. included cw-EPR and electron nuclear double resonance (ENDOR) data on the DMRL radical bound to lumazine protein. The authors were able to determine the g factor as well as several hyperfine couplings of the neutral DMRL radical protonated at N5. This study highlights that DMRL can in principle act as a redox-active cofactor.35
The study of TML by Wörner et al.11 conducted in our laboratories focused on a disproportionation reaction between neutral (TMLH) and anionic (TML−) TML molecules, a photo-induced reaction involving a one-electron transfer process. Thus, a triplet-born, spin-correlated radical pair (SCRP) comprising an oxidized radical TML˙ and a reduced radical TMLH˙− is formed. The oxidized radical TML˙ corresponds to an oxidation state of lumazines previously unknown. It is only mentioned in a contribution by Tu and coworkers,38 who give the redox potential of the one-electron oxidation of TML in acetonitrile without providing a valid reference. This oxidation state is analogous to a “superoxidized” flavin radical,39,40 that is accessible through one-electron oxidation of the flavin with strong oxidants such as tetranitromethane or sulfate radical.41 The rather harsh conditions of synthesis indicate that this oxidation state has no biological relevance. Thus, the first detection and characterization of TML˙ by Wörner et al. expand the redox chemistry of lumazines to a fourth oxidation state, which unlike the respective flavin redox state is readily accessible simply by irradiation with light. Furthermore, DFT calculations suggest that TMLH˙− may be protonated at N5 in a subsequent step following radical pair formation to yield TMLH2˙ (N5). Proton hyperfine couplings of the 6α and 8α methyl groups of both radical species have been determined in the contribution.
This study employed photo-chemically induced dynamic nuclear polarization (photo-CIDNP) spectroscopy to investigate the transiently formed TML radicals.11 Photo-CIDNP spectroscopy is a NMR technique that enables the indirect detection of short-lived SCRPs, offering an alternative to EPR techniques. This is achieved by probing the diamagnetic products of the SCRP which contain the fingerprint of the SCRP's electronic structure; for recent reviews on solution-state photo-CIDNP see ref. 42–44. Photo-CIDNP, established in 1967 by Bargon, Fischer and Johnsen45,46 as well as Ward and Lawler,47 is based on a combination of two aspects: (i) the fate of a photo-induced SCRP is multiplicity-dependent and (ii) the SCRP undergoes singlet–triplet-mixing with the mixing frequency being dependent on the difference of g factors of both radicals and the isotropic hyperfine coupling constants Aiso. This dynamic interplay leads to a spin-sorting process which manifests itself in hyperpolarized nuclear spin resonances of the diamagnetic product. When employing a time-resolved photo-CIDNP technique, the relative size of enhancement for each nucleus is proportional to its Aiso in the transient radical.
With this contribution, we aim to further characterize the radical states of TML, especially in light of the detection of a formerly unknown oxidation state in the lumazine realm. Given the emerging evidence for the role of DMRL in light-induced redox reactions within protein environments,29 a detailed investigation of one-electron reduced and oxidized 6,7-dimethyllumazine radical states will provide crucial insights into lumazine redox reactivity. We chose TML over DMRL for two reasons: its unique disproportionation reaction allows convenient access to two TML radical species. Additionally, the formation of multiple anionic structures of DMRL in basic solution complicates analogous disproportionation reactions. Previous publications mentioned an enhanced photodegradation of free DMRL in solution33,35,48 which renders this molecule an unsuitable candidate for the characterization of its radical state in solution.
This study characterizes radicals formed from TMLH and TML− using 1H and 13C photo-CIDNP spectroscopy. By comparing experimental Aiso values with values from DFT calculations, we determine the protonation states of the reduced and oxidized radical species. Together with previous work by Wörner et al.,11 we provide a more complete understanding of the photo-induced disproportionation reaction of TML and the electronic structure of the oxidized TML radical. Additionally, we conduct a systematic structural analysis of TML− in basic solution using (1H,1H)-NOESY spectroscopy to clarify the structure of the anion.
To the best of our knowledge, no 2D data of TML− detailing its structure are available in the literature. Therefore, a (1H,1H)-NOESY experiment of TML in basic solution was employed, see Fig. 2 for the resulting data and the structure of TML−. It is notable that only two cross peaks of 7α protons are visible, between H6α and H7α′ as well as between H8α and H7α′′. This clearly indicates that the structure resembles an exomethylene or that the rotation around the C7–C7α bond is very slow on the NMR time scale. For a carbanion, additional NOESY cross peaks between H6α and H7α′′ as well as between H8α and H7α′ would be expected.
To test whether temperature-dependent rotational dynamics affect the two H7α resonances, we conducted 1D 1H experiments within a temperature range of 283–333 K, see Fig. S4. The H7α′ and H7α′′ resonances are gradually shifted to higher chemical shifts until H7α′ overlaps with the not fully suppressed H2O signal. Line broadening or other indications of coalescence between both H7α signals are not observed, thus suggesting a rigid exomethylene structure.
Exomethylene and carbanion structures are expected to have different H7α′–C7α–H7α′′ angles. A relaxed potential energy surface scan of TML− was performed by varying the H7α′–C7α–H7α′′ angle from 102° to 120°. The resulting data, as depicted in Fig. S5, show one energy minimum at 117°. This value corresponds to the angle of a slightly distorted sp2-hybridized carbon, which has a typical angle of 120°. The absence of a local minimum at around 109° indicates that a carbanion structure is energetically unfavorable.
The data presented do not indicate a carbanion structure, as the H7α resonances are well separated and demonstrate no sign of temperature-dependent coalescence. The NOESY data provide clear evidence that the structure of TML− is not dynamic. Consequently, previous findings11 must be reevaluated. The discussion of the authors is based on the SOMO of the TML˙ radical which provides insight into the electronic structure with a high electron density at C7α. The binding situation of the diamagnetic TML− molecule clearly does not reflect this.
As the source of photo-CIDNP polarization, Wörner et al.11 proposed a redox cycle comprising photoexcitation of TMLH followed by single-electron transfer from the anionic TML− to TMLH. Comparison of relative photo-CIDNP intensities with DFT calculations of Aiso suggested a subsequent protonation of the reduced TMLH˙− species at N5, thereby forming a transient TMLH2˙ radical species. Protonation of N1 was ruled out. The photo-CIDNP data presented in this contribution were analyzed in a similar way. To perform a linear correlation of the photo-CIDNP intensities with absolute Aiso values, we relied on DFT calculations as experimental data on the hyperfine couplings of TML radicals are scarce. The resulting Aiso values for the relevant nuclei are listed in absolute values in Table 1. The relative photo-CIDNP intensities from Fig. 3 were determined by integration and normalization to the most intense resonance arising from H7α′. These values are listed in Table 2, along with relative Aiso as determined by DFT for better comparability.
Nucleus | Aiso(abs)/MHz | |||
---|---|---|---|---|
TML˙ | TMLH˙− | TMLH2˙ (N1) | TMLH2˙ (N5) | |
H6α | 15.10 | −5.31 | −5.37 | 1.67 |
H7α′ | 34.12 | |||
H7α′′ | −34.51 | |||
H7α | 29.15 | 30.55 | 21.73 | |
H8α | 6.20 | 15.52 | 14.48 | 19.13 |
C6 | 24.84 | −28.57 | −29.25 | −15.51 |
C6α | −8.07 | 0.69 | 1.00 | −1.60 |
C7 | −34.12 | 27.25 | 30.85 | 12.41 |
C7α | 39.43 | −14.28 | −14.94 | −9.82 |
Nucleus | Aiso(rel) (DFT) | Aiso(rel) (CIDNP) | Aiso(rel) (DFT) | Aiso(rel) (CIDNP) | ||
---|---|---|---|---|---|---|
TML˙ | TML− | TMLH˙− | TMLH2˙ (N1) | TMLH2˙ (N5) | TMLH | |
H6α | 0.44 | 0.53 | −0.16 | −0.16 | 0.05 | −0.08 |
H7α′ | −1.00 | −1.00 | ||||
H7α′′ | −1.01 | −0.87 | ||||
H7α | 0.85 | 0.90 | 0.64 | 0.69 | ||
H8α | 0.18 | 0.14 | 0.45 | 0.42 | 0.56 | 0.28 |
C6 | −0.63 | −0.61 | −0.72 | −0.74 | −0.39 | −0.44 |
C6α | 0.20 | 0.17 | 0.02 | 0.03 | −0.04 | 0.00 |
C7 | 0.87 | 0.57 | 0.69 | 0.78 | 0.31 | 0.49 |
C7α | −1.00 | −1.00 | −0.36 | −0.38 | −0.25 | −0.15 |
The linear correlation of the relative photo-CIDNP intensities of TML− with Aiso(TML˙) is demonstrated in Fig. 4(a). A high correlation with a coefficient of determination of R2 = 0.9860 as well as a slope m of −0.0278 MHz−1 was found. It is noteworthy that the correlation is marginally lower than the R2 of 0.9996 reported by Wörner et al.,11 which is to be expected with two additional data points. Still, the correlation remains remarkably high, substantiating the conclusion that this radical species is indeed formed as part of the transient SCRP. In a similar manner, linear correlations were performed of the relative photo-CIDNP intensities of TMLH with Aiso of the reduced radical species TMLH˙−, TMLH2˙ (N1) and TMLH2˙ (N5), see Fig. 4(b)–(d). For TMLH˙− a high correlation with R2 = 0.9745 was found (m = 0.0222 MHz−1). Similar values were calculated for TMLH2˙ (N1): R2 = 0.9884 and m = 0.0218 MHz−1. However, R2 = 0.7512 obtained from the correlation of TMLH2˙ (N5) is poor, which stands in contrast to the findings reported by Wörner et al.11 DFT calculation of Aiso(H6α) predicts a positive value for TMLH2˙ (N5), contrasting with the negative sign predicted for both TMLH˙− and TMLH2˙ (N1). This additional data point of H6α leads to a substantial decrease in R2. Furthermore, as visible in Fig. 4(b)–(d), H7α exhibits a higher photo-CIDNP intensity than predicted by DFT calculations for TMLH2˙ (N5). The combined contributions of these resonances result in a more refined hyperfine pattern for the reduced TMLH radical species. Consequently, we can eliminate TMLH2˙ (N5) as a potential source of photo-CIDNP polarization. The distinction between TMLH˙− and TMLH2˙ (N1) remains difficult based on 1H photo-CIDNP alone as both radicals share a comparable proton hyperfine coupling pattern and thus a similarly high correlation with DFT predictions.
![]() | ||
Fig. 4 Linear correlations of the experimentally determined relative 1H photo-CIDNP intensities of TML− and TMLH with calculated Aiso values of the corresponding radical species: (a) TML˙ (R2 = 0.9860, m = −0.0278 MHz−1), (b) TMLH˙− (R2 = 0.9745, m = 0.0222 MHz−1), (c) TMLH2˙ (N1) (R2 = 0.9884, m = 0.0218 MHz−1) and (d) TMLH2˙ (N5) (R2 = 0.7512, m = 0.0240 MHz−1). The relative photo-CIDNP intensities of TML− are negatively proportional to the respective hyperfine coupling according to Kaptein's rule.60 |
A simple rule established by Kaptein60 correlates the sign of photo-CIDNP enhancement with the sign of Aiso, the sign of Δg of the SCRP, the multiplicity of the SCRP's precursor and the reaction route of the SCRP. For two radicals of the same SCRP, Δg changes its sign. This results in a negative proportionality of Aiso and photo-CIDNP enhancement for nuclei in the oxidized TML radical while nuclei in the reduced TML radical exhibit a positive proportionality, compare Fig. 4(a)–(d).
The information obtained from 1H photo-CIDNP experiments is limited, as merely four and three distinct protons are available for the oxidized and reduced TML radical species, respectively. However, the use of 13C-labeled TML isotopologues can give access to the 13C hyperfine structure, thereby providing more detailed insights into the radical species. An isotopologue that is both readily available and inexpensive (in terms of costs and synthesis efforts) is [6,6α,7,7α-13C4]6,7,8-trimethyllumazine.
The 13C resonances of [6,6α,7,7α-13C4]TML− and [6,6α,7,7α-13C4]TMLH were assigned according to (1H,13C)-HSQC data, see Fig. S6, and the splitting pattern of the 13C resonances in the 13C spectrum, see Fig. S7 for spectra at pH 7 and pH 13. The transient photo-CIDNP spectrum of [6,6α,7,7α-13C4]TML at pH 13 is depicted in Fig. 5. TML− shows prominent emissive signals for both C7α and C6. C6α and C7 exhibit weaker absorptive resonances. From TMLH, signals attributed to C6 and C7α are clearly visible as emissive resonances. C7 shows a weaker absorptive resonance. A photo-CIDNP signal arising from C6α was not observed, which is in accordance to predictions of Aiso by DFT ranging from 0.69 MHz to 1.60 MHz, see Table 1. Comparison of the relative photo-CIDNP signals reveals that both radicals are easily distinguishable: TML˙ exhibits a strong hyperfine coupling for C7α, followed by C7 and C6. Only a weak hyperfine coupling is attributed to C6α. For reduced TML radical species, the strongest hyperfine couplings are found for C6 and C7. The hyperfine coupling of C7α is significantly weaker. This indicates a substantially different 13C hyperfine pattern, a finding consistent with the previously discussed 1H photo-CIDNP experiment.
The relation |Aiso(7α)| > |Aiso(6α)| is found for the reduced TML radical in 1H and 13C photo-CIDNP although the sign of Aiso is inverted when switching from 1H to 13C. Heller, Chesnut and McConnell elucidated that for π-based radicals, the hyperfine coupling interaction of both α- and β-standing atoms depend on the spin population at the atom.61,62 This relation accounts for the similarity in the size relations of 6α and 7α nuclei. In both cases, the transfer of spin density from the atom in the π system to the substituents occurs through polarization of σ-bonds. The mechanism of spin density transfer is direct polarization for α substituents and hyperconjugation for β substituents, which is the reason for opposite signs of hyperfine couplings for 1H and 13C nuclei.
The relative photo-CIDNP intensities were correlated with calculations of Aiso for the relevant TML radicals, see Fig. 6 and Table 2. For this purpose, the photo-CIDNP signals were fitted using Voigt line shapes. For C7α from TMLH, the relative intensity was calculated by integration, as fitting was not possible due to poor resolution of J coupling. The photo-CIDNP resonances of TML− were correlated to Aiso(TML˙), yielding a high correlation of R2 = 0.9631 and a slope of −0.0222 MHz−1, as illustrated in Fig. 6(a). R2 is lower than that calculated with 1H photo-CIDNP (R2 = 0.9860), which can be attributed to the lower S/N ratio of 13C experiments. The correlation of Aiso(TMLH˙−) with relative photo-CIDNP intensities of TMLH yielded R2 = 0.9787 and m = 0.0159 MHz−1. Comparable values of R2 = 0.9850 and 0.0149 MHz−1 were found for the correlation of TMLH2˙ (N1). The correlation of TMLH2˙ (N5) yielded a lower correlation of R2 = 0.9173 and a higher slope of 0.0291 MHz−1 compared to the other reduced radical species. Nonetheless, this experiment shows a significantly higher correlation for TMLH2˙ (N5) than the 1H photo-CIDNP experiment. This indicates that the protonation site N5 exerts a greater influence on 1H hyperfine couplings than on 13C hyperfine couplings so that the differentiation between different protonation states is easier with 1H photo-CIDNP thus yielding a lower correlation for TMLH2˙ (N5). Based on the presented photo-CIDNP data we are confident to claim that TMLH˙− is not protonated at N5 after radical pair formation. However, further differentiation between TMLH˙− and TMLH2˙ (N1) remains speculative based on photo-CIDNP data. For both experiments, a slightly higher correlation for TMLH2˙ (N1) is demonstrated. Nevertheless, the difference in R2 is not significant enough to reach a definitive conclusion. To the best of our knowledge, no information on pKA values of 6,7-dimethylated lumazine radicals are available. Therefore, we refer to the pKA value of TML in its ground state of 0.93 determined for N1.63 We think it unlikely, that the pKA for this protonation site in the reduced TML radical is elevated compared to the ground state, so that protonation at pH 13 is possible.
![]() | ||
Fig. 6 Linear correlations of the experimentally determined relative 13C photo-CIDNP intensities of [6,6α,7,7α-13C4]TML− and [6,6α,7,7α-13C4]TMLH with calculated Aiso of the relevant radical species: (a) TML˙ (R2 = 0.9631, m = −0.0222 MHz−1), (b) TMLH˙− (R2 = 0.9787, m = 0.0159 MHz−1), (c) TMLH2˙ (N1) (R2 = 0.9850, m = 0.0149 MHz−1) and (d) TMLH2˙ (N5) (R2 = 0.9173, m = 0.0291 MHz−1). The relative photo-CIDNP intensities of TML− exhibit a negative proportionality with the respective hyperfine coupling due to Kaptein's rule.60 |
It is noteworthy that the absolute slopes calculated for correlations of the reduced radical species (TMLH˙− or TMLH2˙ (N1)) are lower than the one calculated for the oxidized species (for 1H photo-CIDNP, compare 0.0222 MHz−1 and 0.0218 MHz−1 with 0.0278 MHz−1. For 13C photo-CIDNP, compare 0.0159 MHz−1 and 0.0149 MHz−1 with 0.0222 MHz−1). Without any additional polarization loss pathway on a short microsecond timescale which is the time resolution of the experiment, the slopes calculated for both parts of one SCRP should be equal in absolute magnitude.64 Several effects can be potential sources of polarization loss in TMLH: (i) assuming that TMLH2˙ (N1) is indeed generated during the photo-induced reaction, the subsequent back electron transfer would generate TMLH2+ in a primary step. This species is expected to readily deprotonate to form TMLH, which could lead to a dissipation of hyperpolarization into the solvent. (ii) A cancellation due to degenerate electron exchange,
↓TMLHred˙− + TMLH → ↓TMLH + TMLHred˙−, | (1) |
2↓TMLHred˙− → ↓TMLH2− + ↓TML−, | (2) |
The investigated TML radical species have not been previously studied and data on other 6,7-dimethyllumazine radicals are scarce. Therefore, the Aiso determined cannot be directly compared to literature data. As previously mentioned, Ehrenberg et al.36 investigated a compound, which, in the context of the present publication, can be designated TMLH3˙+. This compound corresponds to the reduced TMLH˙− radical protonated at N1 and N5. The authors determined a number of isotropic hyperfine couplings in absolute values for the nuclei N5, N8, C8α, C7α, C6α and H5. However, a direct comparison with the experimental data of the present study is not feasible as only hyperfine couplings of C6α and C7α are determined in both cases. Consequently, the literature data are tentatively compared to theoretical Aiso from DFT which correspond to the real hyperfine structure quite well, see Table 3. The Aiso of TMLH3˙+ are converted to MHz.
Nucleus | |Aiso|/MHz | Aiso/MHz | ||
---|---|---|---|---|
TMLH3˙+ (ref. 36) | TMLH˙− | TMLH2˙ (N1) | TMLH2˙ (N5) | |
C6α | 2.44 | 0.69 | 1.00 | −1.60 |
C7α | 18.87 | −14.28 | −14.94 | −9.82 |
C8α | 16.43 | −7.77 | −7.01 | −8.55 |
H5 | 22.71 | −25.34 | ||
N5 | 20.61 | 16.37 | 16.41 | 14.23 |
N8 | 16.43 | 8.88 | 7.89 | 11.39 |
Overall, the Aiso values determined for TMLH3˙+ follow a similar order of size as Aiso determined for all three reduced TML radical species: H5 shows the highest Aiso value followed by N5, C7α, N8 and C8α, and C6α. For N8 and C8α, the same value is determined by Ehrenberg et al.36 whereas in this study, |Aiso|(N8) is greater than |Aiso|(C8α) for all three radicals. It is anticipated that TMLH3˙+ would resemble more closely to one of the TML radical species. However, a direct comparison shows deviations between TMLH3˙+ and all three reduced TML radical species. The protonation at both N1 and N5 appears to have a substantial impact on the hyperfine structure of the TML radical. Nevertheless, disparate experimental conditions could account for the differences, as TMLH3˙+ was probed in CF3COOH solution, while the DFT calculations were conducted with a simulation of water solvation.
To further assess whether TMLH˙− is protonated, photo-CIDNP experiments employing a higher variety of 13C isotopologues may provide more information, see Table S1 for a list of all Aiso values of the TML radical species. For example, Aiso(C2) is expected to be affected by protonation at N1, which may be discernible by photo-CIDNP. It is not anticipated that valuable information will be obtained from 15N photo-CIDNP, as the isotropic hyperfine interactions of 15N nuclei do not exhibit significant variation between the protonation states of the TMLH˙− radical. A comparable challenge in differentiating between protonation states was encountered in studies of the 5-deazaflavin radical through photo-CIDNP employing 5-deazaflavin mononucleotide67 and demethylated 5-deazariboflavins.68 Contrary to flavin radicals, protonation at N5 is impeded by exchanging N by C–H at this position. The second protonation site of 5-deazariboflavin, N1, does not significantly alter the hyperfine structure of the radical, so that both protonation states share a similar photo-CIDNP spectrum. In this regard, flavin and 6,7-dimethyllumazine radicals demonstrate similar behavior.
![]() | ||
Fig. 7 Reaction mechanism of the photo-induced disproportionation of TMLH in basic solution. Initially, TMLH is excited into a triplet state following intersystem crossing (ISC).11 Electron transfer (ET) from TML− results in the formation of the SCRP [TMLH˙−(red)⋯TML˙(ox)]. Thereafter, two pathways are possible: either the SCRP directly decays back to TMLH and TML− or TMLH˙− is protonated at N1 by the solvent followed by the decay of the SCRP via deprotonation and ET. The available data does not allow for a clear distinction between these two pathways. |
This study provides evidence that 6,7-dimethyllumazine anions are capable of undergoing light-induced dismutations. This finding may have implications for future studies on the mechanism of riboflavin biosynthesis from DMRL69 and on the photocycle of lumazine-containing proteins. It should be noted that the photochemical reactivity of a protein-bound cofactor may be impacted by the protein environment. Therefore, the electron transfer properties of bound 6,7-dimethyllumazines may deviate from the findings of this study. The study by Paulus et al.35 on DMRL and riboflavin bound to lumazine protein exemplifies this aspect. The authors probed the kinetics of photoreduction of DMRL in solution and incorporated in different mutants of lumazine protein and found significant deviations, notably a slower photoreduction of the cofactor and a higher accumulation of the DMRL radical in the lumazine protein.
To account for the origin of the different hyperfine structures of the TML radical species, Mulliken spin populations were calculated using DFT, see Fig. 8 for a graphical representation and Fig. S2 for a bar chart visualization. When not determined experimentally, Aiso are calculated by DFT, see Table S1.
In summary, the left side of the 13C and 15N framework in TML is predominantly affected in their spin population distribution when comparing oxidized and reduced TML radical, partly to a considerable extent. The different oxidation states show sign flips and substantially different magnitudes of spin population. The right side of the TML structure is affected less significantly due to small spin populations. For this reason, these nuclei are omitted in the following detailed discussion of how the differences in hyperfine coupling result from the presented Mulliken spin populations.
In the reduced TML radical, with (TMLH2˙ (N1)) or without protonation at N1 (TMLH˙−), N5 carries the highest spin population with about 37%. In both species, Aiso(C4a) exhibits a medium-sized negative hyperfine coupling, despite its positive spin population, due to strong polarization by N5. Conversely, C6 exhibits a pronounced negative hyperfine coupling. This is due to a moderately negative spin density of C6. Additionally, the adjacent N5 and C7 exert a polarizing effect through their high positive spin densities. C7 itself exhibits a strong positive hyperfine coupling. The nucleus has a high positive spin density of 34–37%. The neighboring nuclei, C6 and N8, exhibit moderate negative and positive spin densities, so that their polarization of C7 is averaged. The methyl groups 6α, 7α and 8α are not part of the π system. Consequently, their spin population is due to direct polarization from their adjacent π nuclei as established by McConnell, Chesnut and Heller.61,62 The methyl carbons exhibit smaller and opposite hyperfine couplings compared to their neighboring π nuclei. This is particularly evident in C6α, which exhibits an Aiso value lower than 1 MHz.
The formal deprotonation of C7α of TMLH˙− results in the formation of the oxidized radical TML˙. Its spin density pattern is predominantly influenced by C7α, which is part of the π network. Consequently, this leads to a significant reduction and inversion of the spin population of N5. C4a exhibits a high amount of positive spin density. Together with the polarizing effect of the adjacent N5, an exceptionally high positive hyperfine coupling results. Similarly, C6 exhibits a high positive hyperfine coupling resulting from its positive spin population and the polarization by moderate negative spin populations of N5 and C7. Despite the relatively modest negative spin population of C7, the combined polarization by C7α, C6 and N8, each exhibiting moderate to high positive spin populations, culminate in a pronounced negative hyperfine coupling. The equally strong, positive hyperfine coupling of C7α can be attributed to its substantial spin population of 52% as well as a slight polarizing effect of C7.
Photo-CIDNP spectroscopy confirmed a transient SCRP consisting of an oxidized and reduced TML radical formed in a unique disproportionation reaction of neutral and anionic TML. Analysis of the CIDNP resonances reveals substantial differences in the oxidized and reduced TML radicals. The experimental hyperfine coupling constants are in strong agreement with DFT calculations. Thus, the TML radicals formed are the oxidized TML˙ and the reduced TMLH˙− or TMLH2˙ protonated at N1. The formation of a TMLH2˙ radical protonated at N5 as described previously11 can be excluded. The analyzed hyperfine structures are rationalized on the basis of calculations of the spin population distribution in both radical species. Overall, our results contribute to a deeper understanding of 6,7-dimethyllumazine redox chemistry, particularly in the context of one-electron transfer processes. As the oxidized TML radical corresponds to a formerly unknown oxidation state of lumazines, this study provides an in-depth investigation of its electronic structure. The results can be used to derive the hyperfine structure of DMRL radicals which are not readily accessible by photo-CIDNP spectroscopy. This contribution may have implications for further studies on lumazine-containing proteins and the role of lumazines in electron transfer reactions in biological systems.
This journal is © the Owner Societies 2025 |