Next Article in Journal
Development of a Long-Term Sampling Method for Determination of NMHCs in Indoor Air
Previous Article in Journal
Identification of Gentian-Related Species Based on Two-Dimensional Correlation Spectroscopy (2D-COS) Combined with Residual Neural Network (ResNet)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Water-Induced Regeneration of a 2,2-Diphenyl-1-picrylhydrazyl Radical after Its Scandium Ion-Promoted Electron-Transfer Disproportionation in an Aprotic Medium

1
Quantum RedOx Chemistry Team, Institute for Quantum Life Science (iQLS), Quantum Life and Medical Science Directorate, National Institutes for Quantum Science and Technology (QST), Inage-ku, Chiba 263-8555, Japan
2
Institute for Advanced Co-Creation Studies, Open and Transdisciplinary Research Initiatives, Osaka University, Suita 565-0871, Japan
3
Department of Chemistry and Nano Science, Ewha Womans University, Seoul 03760, Republic of Korea
4
Department of Chemistry, Faculty of Pure and Applied Sciences, University of Tsukuba, Tsukuba 305-8571, Japan
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(13), 5002; https://doi.org/10.3390/molecules28135002
Submission received: 17 May 2023 / Revised: 14 June 2023 / Accepted: 23 June 2023 / Published: 26 June 2023
(This article belongs to the Section Physical Chemistry)

Abstract

:
A neutral, stable radical, 2,2-diphenyl-1-picrylhydrazyl radical (DPPH), has been frequently used to estimate the activity of antioxidants for more than 60 years. However, the number of reports about the effect of metal ions on the reactivity of DPPH is quite limited. We have recently reported a unique electron-transfer disproportionation of DPPH to produce the DPPH cations (DPPH+) and anions (DPPH) upon the addition of scandium triflate [Sc(OTf)3 (OTf = OSO2CF3)] to an acetonitrile (MeCN) solution of DPPH. The driving force of this reaction is suggested to be an interaction between DPPH and Sc3+. In this study, it is demonstrated that the addition of H2O to the DPPH–Sc(OTf)3 system in MeCN resulted in an increase in the absorption band at 519 nm due to DPPH. This indicated that an electron-transfer comproportionation occurred to regenerate DPPH. The regeneration of DPPH was also confirmed by electron paramagnetic resonance (EPR) spectroscopy. The amount of DPPH increased with an increasing amount of added H2O to reach a constant value. The detailed mechanism of regeneration of DPPH was proposed based on the detailed spectroscopic and kinetic analyses, in which the reaction of DPPH+ with [(DPPH)2Sc(H2O)3]+ generated upon the addition of H2O to [(DPPH)2Sc]+ is the rate-determining step.

1. Introduction

2,2-Diphenyl-1-picrylhydrazyl radical (DPPH) is a neutral, stable radical that has been frequently used to estimate the activity of antioxidants for more than 60 years [1,2,3,4]. It is known that the radical-scavenging reactivity of antioxidants is significantly affected by the reaction environments, such as solvents [5,6], pH [7,8], the presence of metal ions [9,10,11,12,13,14,15,16,17,18], and so on. However, the number of reports about the reactivity of DPPH in the presence of metal ions is quite limited. We have demonstrated that the DPPH-scavenging reactivity of phenolic compounds, such as a vitamin E model, flavonoids, and hydroquinones, is significantly enhanced in the presence of redox-inactive metal ions with a moderate Lewis acidity, such as Mg2+ [10] and Al3+ [9]. The coordination of the metal ion to the one-electron reduced species of DPPH (DPPH) may stabilize the product, resulting in the acceleration of the electron transfer. On the other hand, DPPH is known to undergo reversible one-electron reduction and oxidation to produce DPPH and the corresponding cation (DPPH+), respectively, in organic solvents (Figure 1A) [19,20,21,22]. We have also reported that an electron-transfer disproportionation of DPPH to produce DPPH+ and DPPH occurs upon the addition of scandium triflate [Sc(OTf)3 (OTf = OSO2CF3)] to an acetonitrile (MeCN) solution of DPPH [23]. Since there is no proton sauce in this reaction system, DPPH does not undergo protonation to produce DPPH-H. Then, DPPH may significantly be stabilized by the strong Lewis acidity of Sc3+ with a formation constant of 2.3 × 103 M–1. Recently, Denzo et al. have reported the reactivity of DPPH in the presence of metal cations (Cu2+ and Zn2+) and acids (HClO4 and HNO3) in MeCN [24]. A strong Brønsted acid, such as HClO4, is required for the disproportionation of DPPH to occur. We report herein that the addition of water to the MeCN solution containing DPPH+, DPPH and Sc(OTf)3 resulted in the electron-transfer comproportionation between DPPH+ and DPPH to regenerate DPPH, demonstrating the reversibility of the Sc3+-catalyzed electron-transfer disproportionation of DPPH. The reversible redox reactivity of DPPH in the presence of the redox-inactive metal ion with strong Lewis acidity shows a unique electron-transfer redox reaction of radical species in aprotic media.

2. Results and Discussion

When Sc(OTf)3 was added to an MeCN solution of DPPH, a decrease in the absorption band of DPPH at 519 nm was observed, accompanied by an increase in the absorption band at 380 nm due to the electron-transfer disproportionation [23]. The band at 380 nm is characteristic of DPPH+. The spectral titration conducted in this study shows the Sc(OTf)3/DPPH molar ratio being 1:4 (Figure 2). Thus, two molecules of DPPH are suggested to be stabilized by one Sc3+, as shown in Figure 1B, although the [(DPPH)2Sc]+ complex has yet to be detected.
Upon the addition of H2O to this solution, the absorption band at 519 nm due to DPPH increased. The time course changes in the absorbance at 519 nm after the addition of several amounts of H2O are shown in Figure 3A,B. At all the concentrations of H2O, the reaction has completed after 1500 s. Figure 4 shows the overlapped absorption spectra at 1500 s after the addition of varying amounts of H2O. The absorption band at 380 nm due to DPPH+ decreased, accompanied by an increase in the absorption band at 519 nm due to DPPH with clear isosbestic points at 344 and 449 nm.
The regeneration of DPPH upon the addition of H2O to the DPPH–Sc(OTf)3 system in MeCN was also confirmed by the electron paramagnetic resonance (EPR) spectroscopy. The well-resolved five lines having a g value of 2.0036 were observed in the EPR spectrum of DPPH in MeCN (Figure 5A). Upon addition of Sc(OTf)3 to the MeCN solution of DPPH, the signal intensity was significantly decreased, as shown in Figure 5B. The addition of H2O to this reaction system resulted in the regeneration of DPPH, which was confirmed by the increase in the EPR signal intensity due to DPPH (Figure 5C).
Figure 3C shows [DPPH] vs. [H2O] at 1500 s after the addition of H2O to the 3 mL MeCN solution containing DPPH (7.1 × 10−5 M) and Sc(OTf)3 (2.0 × 10−5 M). The [DPPH] values were calculated using the extinction coefficient (ε) of 1.2 × 104 M−1 cm−1 at 519 nm [2] and increased with increasing [H2O] to reach a constant value. It is suggested that the complex formation of Sc3+ with H2O may weaken the interaction between DPPH and Sc3+, leading to the electron-transfer comproportionation to produce DPPH. In fact, a hexaaqua complex, Sc(H2O)63+, has been reported for the Sc3+ hydration in aqueous perchlorate solution [25].
The rise of the absorbance at 519 nm due to DPPH shown in Figure 3A,B obeyed pseudo-first-order kinetics. Figure 6 shows a double logarithmic plot of the pseudo-first-order rate constants (kobs) vs. [H2O]. The slope of this plot (dashed line in Figure 6), except for the kobs value at 9.3 × 10−1 M H2O, is about three, suggesting that a triaqua complex, [(DPPH)2Sc(H2O)3]+, may be formed by the addition of H2O to [Sc(DPPH)2]+ as shown in Figure 7A. Then, the reaction B (Figure 7) occurs to produce two molecules of DPPH, DPPH, and [Sc(H2O)3] as the rate-determining step followed by a rapid reaction between DPPH and DPPH+ to produce two molecules of DPPH (Figure 7C).

3. Materials and Methods

3.1. Materials

DPPH was commercially obtained from Tokyo Chemical Industry Co., Ltd., Tokyo, Japan. Sc(OTf)3 was purchased from Sigma–Aldrich, St. Louis, MO, USA. MeCN (spectral grade) used as a solvent was commercially obtained from Nacalai Tesque, Inc., Kyoto, Japan, and used as received. The water used in this study was freshly prepared with a Milli-Q system (Millipore Direct-Q UV3) (Merck Millipore, Burlington, MA, USA).

3.2. Spectral Measurements

Typically, a 10 µL aliquot of Sc(OTf)3 (6.1 × 10−3 M) in MeCN was added to a quartz cuvette (10 mm i.d.) which contained DPPH in MeCN (2.95, 2.90, 2.85, 2.80, 2.75 or 2.70 mL). This led to an electron-transfer disproportionation of DPPH to produce DPPH+ and [(DPPH)2Sc]+. After 1 h, water (50, 100, 150, 200, 250, or 300 µL) was added to this MeCN solution (2.95, 2.90, 2.85, 2.80, 2.75, or 2.70 mL, respectively). The molar concentrations of 50, 100, 150, 200, 250, and 300 µL H2O in 3 mL MeCN–H2O are 9.3 × 10−1, 1.9, 2.8, 3.7, 4.2, 5.6 M, respectively. The final concentrations of DPPH and Sc(OTf)3 were 7.1 × 10−5 M and 2.0 × 10−5 M, respectively, in 3 mL MeCN–H2O. UV-vis spectral changes associated with the reaction were monitored using an Agilent 8453 photodiode array spectrophotometer thermostated with a Peltier temperature control at 298 K (Agilent Technologies, Santa Clara, CA, USA). The regeneration rates of DPPH were followed by monitoring the absorbance change at 519 nm due to DPPH on the Agilent 8453 photodiode array spectrophotometer ([H2O] = 9.3 × 10−1 and 1.9 M) or on a UNISOKU RSP-1000-02NM stopped-flow spectrophotometer (UNISOKU Co., Ltd., Osaka, Japan), which was thermostated with a Thermo Scientific NESLAB RTE-7 Circulating Bath (Thermo Fisher Scientific, Inc., Waltham, MA, USA) at 298 K ([H2O] = 2.8, 3.7, 4.2, and 5.6 M). The kobs values were obtained by a least-square curve fit using an Apple MacBook Pro personal computer (Apple Inc., Cupertino, CA, USA). The first-order plots of ln(AA) vs. time (A and A are the absorbance at the reaction time and the final absorbance, respectively) were linear until three or more half-lives, with a correlation coefficient ρ > 0.999. In each case, it was confirmed that the kobs values derived from at least three independent measurements agreed within experimental error of ±5%. In all cases, solutions were normally equilibrated with air.

3.3. EPR Measurements

The EPR spectra of DPPH (7.1 × 10−5 M) in the presence or absence of Sc(OTf)3 (2.0 × 10−5 M) and/or H2O (5.6 M) in MeCN were taken using an LLC-04B ESR sample tube (LABOTEC Co., Ltd., Tokyo, Japan) on a JEOL X-band spectrometer (JES-RE1X) (JEOL Ltd., Tokyo, Japan) at room temperature under the following conditions: microwave frequency 9.43 GHz, microwave power 8 mW, center field 338 mT, sweep width 15 mT, sweep rate 3 mT min−1, modulation frequency 100 kHz, modulation amplitude 0.2 mT, and time constant 0.1 s. EPR data acquisition was controlled by the WIN-RAD ESR Sata Analyzer System (Radical Research, Inc., Tokyo, Japan). The g values were calibrated with an Mn2+ marker. In all cases, solutions were normally equilibrated with air.

4. Conclusions

The addition of H2O to the MeCN solution containing DPPH+ and [(DPPH)2Sc]+ resulted in the regeneration of DPPH. It is suggested that the complex formation of Sc3+ with H2O may weaken the interaction between DPPH and Sc3+, leading to the electron-transfer comproportionation to produce DPPH. The detailed mechanism of regeneration of DPPH was proposed based on the detailed spectroscopic and kinetic analyses, in which the reaction of DPPH+ with [(DPPH)2Sc(H2O)3]+ generated upon the addition of H2O to [(DPPH)2Sc]+ is the rate-determining step.

Author Contributions

Conceptualization, I.N.; methodology, I.N.; validation, Y.S.; formal analysis, Y.S.; investigation, Y.S.; resources, I.N.; data curation, Y.S.; writing—original draft preparation, I.N.; writing—review and editing, K.O., H.I. and S.F.; visualization, Y.S.; supervision, I.N., K.O. and S.F.; project administration, I.N.; funding acquisition, I.N. and K.O. All authors have read and agreed to the published version of the manuscript.

Funding

Please add: This work was partially supported by Grant-in-Aid (No. JP18K06620 to I.N., JP20H02779, JP20H04819, JP18H04650, JP17H03010, and JP16H02268 to K.O., and JP23K04686 to S.F.) from the Ministry of Education, Culture, Sports, Science and Technology, Japan.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Acknowledgments

We thank Tsukasa Waki (Chiba University) for this help in the experiments.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Angeli, L.; Imperiale, S.; Ding, Y.; Scampicchio, M.; Morozova, K. A novel stoichio-kinetic model for the DPPH• assay: The importance of the side reaction and application to complex mixtures. Antioxidants 2021, 10, 1019. [Google Scholar] [CrossRef] [PubMed]
  2. Foti, M.C. Use and abuse of the DPPH radical. J. Agric. Food Chem. 2015, 63, 8765–8776. [Google Scholar] [CrossRef] [PubMed]
  3. Pyrzynska, K.; Pękal, A. Application of free radical diphenylpicrylhydrazyl (DPPH) to estimate the antioxidant capacity of food samples. Anal. Methods 2013, 5, 4288–4295. [Google Scholar] [CrossRef]
  4. Blois, M.S. Antioxidant determinations by the use of a stable free radical. Nature 1958, 181, 1199–1200. [Google Scholar] [CrossRef]
  5. Litwinienko, G.; Ingold, K.U. Solvent effects on the rates and mechanisms of reaction of phenols with free radicals. Acc. Chem. Res. 2007, 40, 222–230. [Google Scholar] [CrossRef]
  6. Valgimigli, L.; Banks, J.T.; Lusztyk, J.; Ingold, K.U. Solvent effects on the antioxidant activity of vitamin E. J. Org. Chem. 1999, 64, 3381–3383. [Google Scholar] [CrossRef]
  7. Mukai, K.; Tokunaga, A.; Itoh, S.; Kanesaki, Y.; Ohara, K.; Nagaoka, S.; Abe, K. Structure–activity relationship of the free-radical-scavenging reaction by vitamin E (α-, β-, γ-, δ-tocopherols) and ubiquinol-10: pH dependence of the reaction rates. J. Phys. Chem. B 2007, 111, 652–662. [Google Scholar] [CrossRef]
  8. Mitani, S.; Ouchi, A.; Watanabe, E.; Kanesaki, Y.; Nagaoka, S.; Mukai, K. Stopped-flow kinetic study of the aroxyl radical-scavenging action of catechins and vitamin C in ethanol and micellar solutions. J. Agric. Food Chem. 2008, 56, 4406–4417. [Google Scholar] [CrossRef]
  9. Nakanishi, I.; Ohkubo, K.; Ogawa, Y.; Matsumoto, K.; Ozawa, T.; Fukuzumi, S. Aluminium ion-promoted radical-scavenging reaction of methylated hydroquinone derivatives. Org. Biomol. Chem. 2016, 14, 7956–7961. [Google Scholar] [CrossRef] [Green Version]
  10. Waki, T.; Kobayashi, S.; Ozawa, T.; Kamada, T.; Nakanishi, I. Effects of ionic radius of redox-inactive bio-related metal ions on the radical-scavenging activity of flavonoids evaluated using photometric titration. Chem. Commun. 2013, 49, 9842–9844. [Google Scholar] [CrossRef]
  11. Nakanishi, I.; Shimada, T.; Ohkubo, K.; Manda, S.; Shimizu, T.; Urano, S.; Okuda, H.; Miyata, N.; Ozawa, T.; Anzai, K.; et al. Involvement of electron transfer in the radical-scavenging reaction of resveratrol. Chem. Lett. 2007, 36, 1276–1277. [Google Scholar] [CrossRef]
  12. Nakanishi, I.; Kawaguchi, K.; Ohkubo, K.; Kawashima, T.; Manda, S.; Kanazawa, H.; Takeshita, K.; Anzai, K.; Ozawa, T.; Fukuzumi, S.; et al. Scandium-ion accelerated scavenging reaction of cumylperoxyl radical by a cyclic nitroxyl radical via electron transfer. Chem. Lett. 2007, 36, 378–379. [Google Scholar] [CrossRef]
  13. Nakanishi, I.; Kawashima, T.; Ohkubo, K.; Kanazawa, H.; Inami, K.; Mochizuki, M.; Fukuhara, K.; Okuda, H.; Ozawa, T.; Itoh, S.; et al. Electron-transfer mechanism in radical-scavenging reactions by a vitamin E model in a protic medium. Org. Biomol. Chem. 2005, 3, 626–629. [Google Scholar] [CrossRef] [PubMed]
  14. Nakanishi, I.; Ohkubo, K.; Miyazaki, K.; Hakamata, W.; Urano, S.; Ozawa, T.; Okuda, H.; Fukuzumi, S.; Ikota, N.; Fukuhara, K. A planar catechin analogue having a more negative oxidation potential than (+)-catechin as an electron transfer antioxidant against a peroxyl radical. Chem. Res. Toxicol. 2004, 7, 26–31. [Google Scholar] [CrossRef] [PubMed]
  15. Nakanishi, I.; Miyazaki, K.; Shimada, T.; Ohkubo, K.; Urano, S.; Ikota, N.; Ozawa, T.; Fukuzumi, S.; Fukuhara, K. Effects of metal ions distinguishing between one-step hydrogen- and electron-transfer mechanisms for the radical-scavenging reaction of (+)-catechin. J. Phys. Chem. A 2002, 206, 11123–11126. [Google Scholar] [CrossRef]
  16. Mukai, K.; Oi, M.; Ouchi, A.; Nagaoka, S. Kinetic study of the α-tocopherol-regeneration reaction of ubiquinol-10 in methanol and acetonitrile solutions: Notable effect of the alkali and alkaline earth metal salts on the reaction rates. J. Phys. Chem. B 2012, 116, 2615–2621. [Google Scholar] [CrossRef]
  17. Kohno, Y.; Fujii, M.; Matsuoka, C.; Hashimoto, H.; Ouchi, A.; Nagaoka, S.; Mukai, K. Notable effects of the metal salts on the formation and decay reactions of α-tocopheroxyl radical in acetonitrile solution. The complex formation between α-tocopheroxyl and metal cations. J. Phys. Chem. B 2011, 115, 9880–9888. [Google Scholar] [CrossRef]
  18. Ouchi, A.; Nagaoka, S.; Abe, K.; Mukai, K. Kinetic study of the aroxyl radical-scavenging reaction of α-tocopherol in methanol solution: Notable effect of the alkali and alkaline earth metal salts on the reaction rates. J. Phys. Chem. B 2009, 113, 13322–13331. [Google Scholar] [CrossRef]
  19. Solon, E.; Bard, A.J. The electrochemistry of diphenylpicrylhydrazyl. J. Am. Chem. Soc. 1964, 86, 1926–1928. [Google Scholar] [CrossRef]
  20. Kalinowski, M.K.; Kimkiewicz, J. Solvation effects in the electrochemistry of diphenylpicrylhydrazyl. Monatsh. Chem. 1983, 114, 1035–1043. [Google Scholar] [CrossRef]
  21. Deutchoua, A.D.D.; Siegnin, R.; Kouteu, G.K.; Denzo, G.K.; Ngameni, E. Electrochemistry of 2,2-diphenyl-1-picrylhydrazyl (DPPH) in acetonitrile in presence of ascorbic acid—Application for antioxidant properties evaluation. ChemistrySelect 2019, 4, 13746–13753. [Google Scholar] [CrossRef]
  22. Patrascu, P.; Lete, C.; Popescu, C.; Matache, M.; Paun, A.; Madalan, A.; Ionita, P. Synthesis and spectral comparison of electronic and molecular properties of some hydrazines and hydrazyl free radicals. Arkivoc 2020, vi, 1–10. [Google Scholar] [CrossRef] [Green Version]
  23. Nakanishi, I.; Kawashima, T.; Ohkubo, K.; Waki, T.; Uto, Y.; Kamada, T.; Ozawa, T.; Matsumoto, K.; Fukuzumi, S. Disproportionation of a 2,2-diphenyl-1-picrylhydrazyl as a model of reactive oxygen species catalysed by Lewis and/or Brønsted acids. Chem. Commun. 2014, 50, 814–816. [Google Scholar] [CrossRef] [PubMed]
  24. Ngueumaleu, Y.; Deuchoua, A.D.D.; Hanga, S.S.P.; Liendji, R.W.; Denzo, G.K.; Ngameni, E. Probing the reactivity of 2,2-diphenyl-1-picrylhydrazyl (DPPH) with metal cations and acids in acetonitrile by electrochemistry and UV-Vis spectroscopy. Phys. Chem. Chem. Phys. 2023, 25, 5282–5920. [Google Scholar] [CrossRef] [PubMed]
  25. Rudolph, W.W.; Pye, C.C. Raman spectroscopic measurements of scandium(III) hydration in aqueous perchlorate solution and ab intio molecular orbital studies of scandium(III) water clusters: Does Sc(III) occur as a hexaaqua complex? J. Phys. Chem. A 2000, 104, 1627–1639. [Google Scholar] [CrossRef]
Figure 1. (A) Redox behavior or DPPH. (B) Sc3+-induced disproportionation of DPPH.
Figure 1. (A) Redox behavior or DPPH. (B) Sc3+-induced disproportionation of DPPH.
Molecules 28 05002 g001
Figure 2. Plot of the absorbance at 519 nm vs. [Sc(OTf)3]/[DPPH] in MeCN. [DPPH] and [Sc(OTf)3] are concentrations of DPPH (6.6 × 10–5 M) and Sc(OTf)3 (4.2 × 10–6 M, each), respectively.
Figure 2. Plot of the absorbance at 519 nm vs. [Sc(OTf)3]/[DPPH] in MeCN. [DPPH] and [Sc(OTf)3] are concentrations of DPPH (6.6 × 10–5 M) and Sc(OTf)3 (4.2 × 10–6 M, each), respectively.
Molecules 28 05002 g002
Figure 3. (A) Time course change monitored by the Agilent 8453 photodiode array spectrophotometer in the absorbance at 519 nm after the addition of H2O (closed circles 50 µL (9.3 × 10−1 M), open circles 100 µL (1.9 M), closed triangles 150 µL (2.8 M), open squares 200 µL (3.7 M), closed squares 250 µL (4.2 M), and open triangles 300 µL (5.6 M)) to the MeCN solution containing DPPH and Sc(OTf)3 at 298 K. The final concentrations of DPPH and Sc(OTf)3 are 7.1 × 10−5 M and 2.0 × 10−5 M, respectively, in 3 mL MeCN–H2O. (B) Time course change monitored by the stopped-flow spectrophotometer in the absorbance at 519 nm after addition of H2O (closed triangles 150 µL (2.8 M), open squares 200 µL (3.7 M), closed squares 250 µL (4.2 M), and open triangles 300 µL (5.6 M)) to the MeCN solution containing DPPH and Sc(OTf)3 at 298 K. The final concentrations of DPPH and Sc(OTf)3 are 7.1 × 10–5 M and 2.0 × 10–5 M, respectively, in 3 mL MeCN–H2O. (C) Plot of [DPPH] vs. [H2O] at 1500 s after the addition of H2O to the MeCN solution containing DPPH and Sc(OTf)3. The final concentrations of DPPH and Sc(OTf)3 are 7.1 × 10–5 M and 2.0 × 10–5 M, respectively, in 3 mL MeCN–H2O.
Figure 3. (A) Time course change monitored by the Agilent 8453 photodiode array spectrophotometer in the absorbance at 519 nm after the addition of H2O (closed circles 50 µL (9.3 × 10−1 M), open circles 100 µL (1.9 M), closed triangles 150 µL (2.8 M), open squares 200 µL (3.7 M), closed squares 250 µL (4.2 M), and open triangles 300 µL (5.6 M)) to the MeCN solution containing DPPH and Sc(OTf)3 at 298 K. The final concentrations of DPPH and Sc(OTf)3 are 7.1 × 10−5 M and 2.0 × 10−5 M, respectively, in 3 mL MeCN–H2O. (B) Time course change monitored by the stopped-flow spectrophotometer in the absorbance at 519 nm after addition of H2O (closed triangles 150 µL (2.8 M), open squares 200 µL (3.7 M), closed squares 250 µL (4.2 M), and open triangles 300 µL (5.6 M)) to the MeCN solution containing DPPH and Sc(OTf)3 at 298 K. The final concentrations of DPPH and Sc(OTf)3 are 7.1 × 10–5 M and 2.0 × 10–5 M, respectively, in 3 mL MeCN–H2O. (C) Plot of [DPPH] vs. [H2O] at 1500 s after the addition of H2O to the MeCN solution containing DPPH and Sc(OTf)3. The final concentrations of DPPH and Sc(OTf)3 are 7.1 × 10–5 M and 2.0 × 10–5 M, respectively, in 3 mL MeCN–H2O.
Molecules 28 05002 g003
Figure 4. Spectral change observed 1500 s after the addition of H2O (9.3 × 10–1, 1.9, 2.8, 3.7, 4.2, and 5.6 M) to the MeCN solution containing DPPH and Sc(OTf)3. The final concentrations of DPPH and Sc(OTf)3 are 7.1 × 10–5 M and 2.0 × 10–5 M, respectively, in 3 mL MeCN–H2O.
Figure 4. Spectral change observed 1500 s after the addition of H2O (9.3 × 10–1, 1.9, 2.8, 3.7, 4.2, and 5.6 M) to the MeCN solution containing DPPH and Sc(OTf)3. The final concentrations of DPPH and Sc(OTf)3 are 7.1 × 10–5 M and 2.0 × 10–5 M, respectively, in 3 mL MeCN–H2O.
Molecules 28 05002 g004
Figure 5. EPR spectra of (A) DPPH (7.1 × 10−5 M) in MeCN, (B) DPPH (7.1 × 10−5 M) in the presence of Sc(OTf)3 (2.0 × 10−5 M) in MeCN, and (C) DPPH (7.1 × 10−5 M)in the presence of Sc(OTf)3 (1.8 × 10−5 M) and H2O (5.6 M) in MeCN.
Figure 5. EPR spectra of (A) DPPH (7.1 × 10−5 M) in MeCN, (B) DPPH (7.1 × 10−5 M) in the presence of Sc(OTf)3 (2.0 × 10−5 M) in MeCN, and (C) DPPH (7.1 × 10−5 M)in the presence of Sc(OTf)3 (1.8 × 10−5 M) and H2O (5.6 M) in MeCN.
Molecules 28 05002 g005
Figure 6. Plot of log kobs vs. log[H2O].
Figure 6. Plot of log kobs vs. log[H2O].
Molecules 28 05002 g006
Figure 7. Proposed mechanism of the DPPH regeneration. (A) Addition of H2O to form [(DPPH)2Sc(H2O)3]+. (B) Rate-determining electron transfer from [(DPPH)2Sc(H2O)3]+ to DPPH+. (C) Rapid comproportionation between DPPH and DPPH+.
Figure 7. Proposed mechanism of the DPPH regeneration. (A) Addition of H2O to form [(DPPH)2Sc(H2O)3]+. (B) Rate-determining electron transfer from [(DPPH)2Sc(H2O)3]+ to DPPH+. (C) Rapid comproportionation between DPPH and DPPH+.
Molecules 28 05002 g007
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

Nakanishi, I.; Shoji, Y.; Ohkubo, K.; Ito, H.; Fukuzumi, S. Water-Induced Regeneration of a 2,2-Diphenyl-1-picrylhydrazyl Radical after Its Scandium Ion-Promoted Electron-Transfer Disproportionation in an Aprotic Medium. Molecules 2023, 28, 5002. https://doi.org/10.3390/molecules28135002

AMA Style

Nakanishi I, Shoji Y, Ohkubo K, Ito H, Fukuzumi S. Water-Induced Regeneration of a 2,2-Diphenyl-1-picrylhydrazyl Radical after Its Scandium Ion-Promoted Electron-Transfer Disproportionation in an Aprotic Medium. Molecules. 2023; 28(13):5002. https://doi.org/10.3390/molecules28135002

Chicago/Turabian Style

Nakanishi, Ikuo, Yoshimi Shoji, Kei Ohkubo, Hiromu Ito, and Shunichi Fukuzumi. 2023. "Water-Induced Regeneration of a 2,2-Diphenyl-1-picrylhydrazyl Radical after Its Scandium Ion-Promoted Electron-Transfer Disproportionation in an Aprotic Medium" Molecules 28, no. 13: 5002. https://doi.org/10.3390/molecules28135002

Article Metrics

Back to TopTop