Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Jan 2.
Published in final edited form as: Angew Chem Int Ed Engl. 2019 Nov 18;59(1):409–417. doi: 10.1002/anie.201910300

Electrochemistry Broadens the Scope of Flavin Photocatalysis: Photoelectrocatalytic Oxidation of Unactivated Alcohols

Wen Zhang 1, Keith L Carpenter 1, Song Lin 1
PMCID: PMC6923568  NIHMSID: NIHMS1057503  PMID: 31617271

Abstract

Riboflavin-derived photocatalysts have been extensively studied in the context of alcohol oxidation. However, to date, the scope of this catalytic methodology has been limited to benzyl alcohols. In this work, we gained further mechanistic understanding of flavin-catalyzed oxidation reactions in the absence or presence of thiourea as a cocatalyst. The mechanistic insights enabled us to develop an electrochemically driven photochemical oxidation of primary and secondary aliphatic alcohols using a pair of flavin and dialkylthiourea catalysts. Electrochemistry makes it possible to avoid using O2 oxidant and generating H2O2 byproduct, both of which oxidatively degrade thiourea under the reaction conditions. This modification unlocks a new mechanistic pathway in which the oxidation of unactivated alcohols is achieved via thiyl radical-mediated hydrogen-atom abstraction

Keywords: Electrochemistry, Flavin, Photoelectrocatalysis, Thiourea, Alcohol oxidation, Thiyl radical

Graphical abstract

graphic file with name nihms-1057503-f0015.jpg


The combination of electrochemistry and photochemistry allows for the generation of highly reactive catalytic intermediates without the need for a chemical oxidant. This photoelectrocatalytic strategy thus turns on the elusive reactivity of flavins of oxidizing unactivated aliphatic alcohols

Introduction

Oxidation and reduction reactions are among the most important and frequently executed processes in organic synthesis.1 In recent decades, the discovery and development of new redox reactions—including catalytic hydrogenations, epoxidations, and C–H oxidations—have fundamentally changed how molecules are made.2 Recently, the emergence of new catalytic strategies that cleverly adopt concepts and techniques from areas such as photochemistry3 and electrochemistry4 has provided chemists with new tools for solving contemporary synthetic problems.

Photoredox chemistry5 and electrochemistry6 have both proven to be disruptive technologies in organic synthesis, displaying complementary characteristics in synthetic contexts. Photoredox catalysis offers access to unique chemical reactivities at the excited states of organic molecules (Scheme 1A). This mode of catalysis is particularly suited to overall redox-neutral transformations. To achieve net-oxidation or reduction reactions, however, would require the use of a terminal redox agent and would generate byproducts,5e both of which could complicate the reaction. Electrochemistry offers a solution to this issue by using electricity as the traceless redox agent, in combination with either an innocuous oxidant/reductant (e.g., H+) or one that can be separated in an independent compartment (i.e., using a divided cell) (Scheme 1B). In principle, electrochemistry can also grant access to highly reactive intermediates at extreme electrode potentials without relying on excited states.7 Nonetheless, the constant application of such highly biased potentials will likely lead to uncontrolled reactions due to accumulation of reactive intermediates near the electrode surface as well as compromised chemoselectivity in complex reaction systems.

Scheme 1.

Scheme 1.

General mechanistic depiction of photoredox catalysis, electrocatalysis, and molecular photoelectrocatalysis for oxidative transformations.

We envision that the merger of electrochemistry and photochemistry has the potential to expand the scope of single-electron redox chemistry into new territories (Scheme 1C).8 Specifically, the combination of electrochemical catalyst activation and photo-excitation will grant access to reactive intermediates with high redox potential in a transient fashion in the absence of an oxidant or reductant. In this report, we demonstrate the generation of a highly reactive thiourea radical by means of flavin photoelectrocatalysis and its application in the oxidation of unactivated alcohols.9 The additive features granted by photoelectrochemistry proved critical, as neither photoredox catalysis nor electrocatalysis alone can provide the same reactivity in this catalytic system.

Results and Discussion

Flavins, a family of most versatile redox cofactors in nature,10 have recently been studied as organic photoredox catalysts for oxidative organic transformations.11 For examples, riboflavin tetraacetate (RFT) is a readily available compound with an oxidation potential of +1.67 V (vs SCE) upon photo-irradiation (νmax = 440 nm). This characteristic has allowed RFT and its derivatives to function as photocatalysts in the aerobic oxidation of benzylic C–H bonds,12 benzyl alcohols,13 benzyl amines,14 sulfides,15 and carboxylic acids16 using O2 as the terminal oxidant. Nonetheless, the scope of flavin catalysis is limited by the oxidation potential of the photoexcited state. The substrates suitable for this reaction system are thus confined primarily to those with electron-releasing substituents that can stabilize the key radical or radical cation intermediates (Scheme 2). For example, flavin-catalyzed aerobic oxidation of alcohols is only applicable to electron-neutral and -rich benzyl alcohols.12 Recent advances in catalyst development have led to cationic flavin analogs that can oxidize electron-deficient benzyl alcohols.17 However, aliphatic alcohols have not been successfully engaged in flavin-promoted oxidation reactions. This limitation significantly hampers the broad use of flavins in organic synthesis.18

Scheme 2.

Scheme 2.

RFT-catalyzed photochemical oxidation of alcohols.

Reports have shown that the addition of a Lewis acid11b,19 or thiourea12c co-catalyst may enhance the catalytic efficiency of flavin but does not improve the reaction scope. Among these strategies, the use of thiourea as an additive is appealing from a mechanistic point of view. The mechanism of synergistic flavin/thiourea catalysis in the context of benzyl alcohol oxidation is not well understood. Initially thought to function as a H-bond donor that activates the flavin catalyst, thiourea has also been postulated to promote alcohol oxidation as a redox mediator between the substrate and flavin.12c However, the reduction potential of thiourea is relatively low (E ≈ 0.8–1.0 V vs SCE) and neither it nor its corresponding radical cation absorbs light in the visible region to reach the photoexcited state. As such, it is very unlikely that thiourea acts purely as an outer-sphere electron transfer agent. König postulated that thiourea could form highly reactive radical intermediates by means of flavin oxidation, which could then promote the oxidation of alcohols.12c We hypothesize that the S-centered radical derived from single-electron oxidation of thiourea could serve to abstract a hydrogen-atom adjacent to the alcohol, thus resulting in the generation of a ketyl radical en route to carbonyl formation.

Computed bond dissociation energy (BDE; Scheme 3) supports this hypothesis. The S–H BDEs in isothiouronium (TU-1•H+) and its neutral form, isothiourea (iso-TU-1), are 93 and 91 kcal/mol, respectively; both of these values are higher than the C–H BDE in benzyl alcohol (88 kcal/mol)20 and close to that of isopropanol (95 kcal/mol).21 These data suggest that, in theory, thiourea could promote the oxidation of both benzylic and unactivated aliphatic alcohols on thermodynamic grounds.

Scheme 3.

Scheme 3.

Computed bond dissociation energy of isothiouronium and isothiourea [M06-2x/6-311+G(d,p)/CPCM(CH3CN)//B3LYP/6-31G(d)].

We reason that understanding the behavior of thiourea in the existing alcohol oxidation systems could facilitate a critical expansion of the scope of flavin photocatalysis to substrates beyond benzylic compounds.

Elucidating the fate of thiourea in flavin-catalyzed alcohols oxidation.

We set out to study both the effect of thiourea on the oxidation of alcohols and the fate of this mediator following reaction. Under conditions developed by König et al., 4-methoxybenzyl alcohol (1a) was oxidized to 4-methoxybenzaldehyde (2a) in 40% yield (Scheme 4, entry 1) in the presence of RFT as the sole catalyst in 1 atm O2 under blue light irradiation. In stark contrast, under the same reaction conditions, 4-tert-butylcyclohexanol (1b) substrate was completely recovered with no observed formation of the ketone product (entry 3). These results are not surprising given the oxidation potential of 1b (E > 1.9 V vs SCE; Figure 1A) is much higher than that of 1a (E ~1.5 V) and excited RFT (RFT*; E ~ 1.67 V). In fact, photo-quenching experiments (Figure 1B) showed that RFT* is efficiently quenched by 1a but not by 1b.

Scheme 4.

Scheme 4.

RFT-catalyzed photochemical oxidation of alcohols.

Figure 1.

Figure 1.

(A) Cyclic voltammetry of 1a and 1b. Conditions: alcohol (5.71 mM), LiOTf (0.1 M in MeCN), scan rate = 100 mV/s. (B) Fluorescence quenching of RFT with 1a, 1b.

Analogous to previous reports, the reaction yield with 1a could be improved to 85% in the presence of 10 mol% thiourea (TU-1; Scheme 4, entry 3). Encouragingly, addition of 10 mol% TU-1 to the reaction of substrate 1b led to formation of a minute amount of ketone product 2b (entry 4), providing proof-of-concept that thiourea can in fact promote the oxidation of unactivated secondary alcohols. However, extending the reaction time did not lead to an increase in reaction yield in this case.

In the reactions with TU-1, the thiourea catalyst was not detected by HPLC following the reaction, but a white precipitate formed during the course of the reaction, which was identified as thiourea dioxide (TU-1•O2) through comparison with an authentic sample (Figure 2). We hypothesize that TU-1•O2 arises from the oxidative degradation of TU-1, by either O2 or H2O2 – a byproduct arising from the catalytic turnover of flavin by O2 (Scheme 5A). Indeed, H2O2 was detected in 34 mM (0.63 equiv with respect to 1a) in the reaction with 1a and in a smaller but still significant amount (2.6 mM, 0.05 equiv) with 1b (Scheme 4, entries 2 and 4).

Figure 2.

Figure 2.

13C-NMR spectrum of A) thiourea dioxide (TU-1•O2); B) white precipitate in reaction system; C) thiourea, all in D2O with MeCN as internal standard.

Scheme 5.

Scheme 5.

Decomposition of thiourea under RFT-promoted photooxidation conditions.

Further experiments lend support to our hypothesis. In these experiments, 1,3-diisopropylthiourea (TU-2) was employed instead of TU-1 to allow for the facile determination of the remaining thiourea by 1H NMR. Upon treatment of TU-2 with 1 equiv of H2O2 in the presence or absence of RFT, approximately 60% of the thiourea was consumed and the corresponding thiourea dioxide (TU-2•O2) was observed (Scheme 5B, entries 1 and 2

Under conditions analogous to the photochemical oxidation of benzyl alcohol 1a, singlet O2 – a strong oxidant – was detected using the furfural test22 (Scheme 5C). 1O2 is presumably generated through photoirradiation of 3O2 in the presence of RFT as the sensitizer. Under photolytic conditions with O2, TU-2 was fully consumed to form TU-2•O2 in the presence RFT (Scheme 5B, entry 3) but fully recovered in the absence of RFT (entry 4). These findings suggest that RFT can promote the photodecomposition of thiourea via two pathways, by means of singlet O2 and byproduct H2O2.23 In the reaction with simple alcohol 1b, it is likely that the singlet O2 pathway is predominant because the formation of H2O2 needs at least one turnover of the flavin catalyst and is thus slow, given the small amount of product generated.

To explain the moderate rate acceleration effect of thiourea on the oxidation of 1a, we reason that thiourea decomposition is slower than or comparable to the oxidation of reactive benzyl alcohols. However, this decomposition pathway out-competes the oxidation of more challenging unactivated alcohols such as 1b.

Solving the thiourea degradation problem using electrochemistry.

Based on this discovery, we hypothesize that the decomposition of thiourea, which mediates the oxidation of alcohol by photoexcited flavin, plagues the catalytic system and is likely the cause for its narrow substrate scope. We aimed to further substantiate this theory and find a solution that would allow us to expand the reaction to the transformation of unactivated alcohols. Intuitively, replacing O2 with another oxidant could resolve this problem. However, O2 represents one of the mildest and most atom-economic oxidizing agents and is uniquely suited for flavin photocatalysis. The use of another oxidant could lead to oxidative decomposition of flavin and/or thiourea (as in the case with H2O2) and generate side products that could complicate the reaction by reacting with the alcohol substrate or other reactive intermediates. Mindful of the electrochemical activity of flavin derivatives,24 we reasoned that using electricity to replace O2 with H+ as the innocuous terminal oxidant might provide an effective solution.

Our initial explorations revealed that the oxidation of 1b could be achieved in significantly improved yield (up to 67%) in an electrolysis cell (cell voltage Ucell = 2.5 V, initial anodic potential Eanode ≈ 0.8 V vs SCE) in the presence of RFT and 10 mol% TU-1 using carbon foam as the anode, Pt coil as the cathode, LiClO4 in MeCN as the electrolyte solution, and H2O as the proton source (Scheme 6A, entry 1). To further optimize the oxidation and probe the role of thiourea in the reaction, we surveyed several substituted thioureas (entries 2-4). Among the N,N-dialkylthioureas tested, the isopropyl version proved optimal, providing 2b in improved 78% yield (entry 2). Fully substituted thiourea TU-7 was incompetent in promoting the oxidation reaction (entry 5), while acetylthiourea (TU-8) gave the product in a diminished 21% yield (entry 6). Various electrolytes were also investigated. While commonly used tetraalkylammonium salts did not increase the reaction efficiency (entries 7, 8), LiOTf proved superior, furnishing the ketone nearly quantitatively (entry 9). Finally, control experiments suggest that electrical potential, blue LED, RFT, and the thiourea are all essential to this transformation, as exclusion of any of these components led to no formation of 2b (entries 10–13). Finally, controlled potential electrolysis showed that alcohol oxidation could be observed at an anodic potential as low as 0.58 V, albeit at a much slower rate as expected (entry 14).

Scheme 6.

Scheme 6.

Photoelectrocatalytic oxidation of alcohols. aConditions: alcohol (0.2 mmol, 1 equiv), RFT (5 mol%), TU (10 mol%), electrolyte (3.5 mL, 0.1 M in MeCN), H2O (0.2 mL), cell voltage Ucell = 2.5 V (initial anodic potential Eanode ≈ 0.8 V vs SCE), blue LED. bYield determined by 1H NMR. cIsolated yield. dWithout blue LED. eWithout RFT. fWithout electricity. gElectrolysis at a constant anodic potential of 0.58 V. gReaction time 36 h.

These optimized reaction conditions were applied to several other secondary alcohols and provided the desired ketones (2ci) in satisfactory yields (Scheme 6B). An unactivated primary alcohol was also found to be compatible, affording carboxylic acid 2j in 53% yield.

Probing the role of thiourea.

As previously discussed, in König’s reaction system with benzyl alcohols, thiourea was proposed as a redox mediator. However, the role of thiourea in promoting the alcohol oxidation is not well understood. In particular, the reduction potential of thiourea is relatively low (E ≈ 0.8–1.0 V vs SCE) and neither it nor its corresponding radical cation absorbs light in the visible region. As such, it is very unlikely that thiourea serves purely as an outersphere electron mediator.

Fluorescence quenching experiments suggest that RFT* could be quenched by a thiourea (Figure 3). We propose that thiourea is oxidized by RFT* to a thiyl radical cation 3, which can be deprotonated to form neutral thiyl radical 4 (Scheme 7A). The formation of thiyl radicals is supported by alkene isomerization experiments (Scheme 7B).25 Thus, when β-methylstyrene 6 in predominantly Z-configuration (Z/E = 6:1) was added to the standard photoelectrochemical reaction, in the absence of any thiourea, 71% of 6 was recovered with a Z/E ratio of 10:1. This result suggests that the radical induced isomerization did not occur. The marginal enrichment of the major Z-isomer was likely due to the fact that the E-isomer undergoes preferential RFT*-mediated isomerization or decomposition.11a,26 By contrast, when subjected to TU-2, 6 was essentially completely isomerized to the more stable E configuration (Z/E = 1:100). This isomerization is most likely induced by thiourea-derived radicals 3 or 4.

Figure 3.

Figure 3.

Photo-quenching of RFT by thiourea TU-2.

Scheme 7.

Scheme 7.

Formation and role of thiyl radicals in the photoelectrocatalytic alcohol oxidation.

The formation of thiyl radicals was also supported by the formation of disulfide-like dimer 7 during the alcohol oxidation. In particular, when we increase the thiourea loading from 10 mol% to 20 mol% and 40 mol%, the generation of H2S is evident from the distinct odor of the completed reaction. Thiourea-derived radicals have been reported to first dimerize and then decompose to the corresponding urea and H2S in the presence of H2O.25,27 Increasing the loading of TU-2 led to a decrease in reaction rate (Figure 4), a phenomenon that we can also attribute to the catalyst decomposition pathway. In fact, an apparent induction period was observed in the presence of 40 mol% TU-2 (blue triangles, Figure 4). Product 2b only began to form after 4 h reaction time when the concentration of TU-2 reached a sufficiently low level due to decomposition so that the catalytic reaction outcompeted the unproductive dimerization.

Figure 4.

Figure 4.

Reaction time course data with different concentrations of TU-2.

The ability of thiourea-based radicals to promote alcohol oxidation was investigated using cyclic voltammetry (Figure 5A). TU-2 alone showed an oxidative wave with an onset potential of ca. 0.58 V vs SCE. This broad wave appears to encompass at least two features, peaking at 1.0 and 1.2 V. The second peak may correspond to oxidation of the thiourea dimer but we have not carried out further experiments to support this hypothesis. The addition of alcohol 1b to the system led to the observation of a catalytic current, which increased in intensity as the amount of 1b was increased. This current enhancement is indicative of the fact that TU-2 is capable of catalytically oxidizing the alcohol substrate. This catalytic activity is likely promoted by thiourea-based radicals 3 or 4. Control experiments excluded the role of thiourea dimer (cf. 7) in the alcohol oxidation activity: (1) using commercial available dimer of TU-1 as the catalyst instead of TU-1 or TU-2 under optimal photoelectrochemical conditions, only a trace amount of ketone product was observed (~7%, vs 76% using TU-1); (2) cyclic voltammetry study of TU-1 dimer did not show catalytic current upon addition of alcohol 1b.

Figure 5.

Figure 5.

(A) Cyclic voltammetry of TU-2 in the presence and absence of 1b; (B) Radical probe experiment. Standard conditions: see Scheme 6A, entry 9.

We propose that thiourea-derived radicals promote the oxidation of unactivated alcohols through abstraction of the H atom α- to the hydroxy group (see Scheme 7A).28 The resultant ketyl radical 5 then readily loses the O-bound H-atom (BDEO–H = 20–30 kcal/mol)29 to furnish the ketone. The intermediacy of a ketyl radical was suggested by a radical probe experiment with alcohol 8, which bears a cyclopropyl group (Figure 5B). Thus, when 8 was exposed to standard conditions, a number of products were isolated, including ketone 10 (2% yield) and a mixture of products 1113 (combined 14% yield). The latter products arose from radical-triggered ring opening of the cyclopropane unit in intermediate 9. DFT data (see Scheme 3) suggest that the hypothesized H-atom abstraction mechanism is thermodynamically plausible.

Understanding the role of RFT.

This mechanistic proposal supposes that flavin plays the role of a pure photooxidant and does not have a discrete function in the alcohol oxidation process. Indeed, when we used Ir and pyrrilium-based photocatalysts in lieu of RFT, we observed the formation of desired product 2b, albeit in lower yield (Table 1, entries 2 & 4 vs entry 1). These photocatalysts alone without thiourea are ineffective in promoting the alcohol oxidation (entries 3 & 5).

Table 1.

Photoelectrocatalysis control experiments.

graphic file with name nihms-1057503-f0014.jpg

Entry Photocatalyst Thiourea 1b Conversiona 2b Yielda
1 RFT yes 75% 67%
2 [Ir(dF(CF3)ppy)2(dtbpy)]PF6 yes 32% 31%
3 [Ir(dF(CF3)ppy)2(dtbpy)]PF6 no 5% <5%
4 [Mes-Acr-Me]+ClO4 yes 10% 8%
5 [Mes-Acr-Me]+ClO4 no <5% <5%
6 None yes 13% <5%
7b None yes 7% <5%
8c None yes 8% <5%
a

Determined by 1H NMR.

b

Controlled potential electrolysis at Eanode = 1.09 V vs SCE without light irradiation.

c

Controlled current electrolysis at i = 0.5 mA without light irradiation.

The cyclic voltammetry data suggested that a photocatalyst might not be necessary at all and that the thiourea could potentially serve as an electrocatalyst to effect alcohol oxidation under dark conditions. However, direct electrolysis in the absence of RFT under conditions otherwise identical to the optimal conditions did not give appreciable amounts of ketone (entry 6). We also conducted a controlled potential electrolysis experiment at 1.09 V (vs SCE) anodic potential in the absence of RFT or light irradiation, but did not observe any desired oxidation product (entry 7). In this reaction, the distinct odor of H2S was detected. We reason that under pure electrolytic conditions, the thiourea-based radicals are generated directly on the anode, resulting in a high concentration of the reactive intermediates within the diffusion layer. Thus, undesired dimerization of these key catalytic species predominates. This issue was not readily addressed by lowing the rate of anodic oxidation, as electrolysis at a very low constant current of 0.5 mA (6–10 times lower than the typical current under optimal conditions) with TU-1 alone without RFT or blue light also did not provide any desired ketone product (entry 8). In contrast, under photoelectrochemical conditions, the thiourea radicals are generated in small quantities by the transient photoexcited state of RFT, which will preferentially react with the alcohol substrate to promote the desired reaction pathway.

Based on the above studies, we propose the reaction mechanism depicted in Scheme 8. Upon irradiation, electron transfer followed by proton transfer between excited RFT* and TU-2 results in 4 and semiquinone form (RFT)–H. The thiyl radical 4 abstracts the α-position hydrogen from alcohol 1 to produce the corresponding ketyl radical 5, which can further react with (RFT)–H via hydrogen-atom transfer (HAT) to afford the desired ketone product 2. Finally, the dihydroquinone form of catalyst (RFT)–H2 is oxidized at the anode surface to regenerate the ground-state catalyst RFT. An alternative mechanism in which thiyl radical cation 3 functions as the H atom abstractor is also plausible (see Supporting Information). Given the fact that tetraalkyl thiourea (e.g., TU-7) is ineffective in catalyzing the oxidation reaction, we currently favor the proposed mechanism in Scheme 8. However, the current level of evidence does not allow us to fully distinguish between these two pathways, and it is possible that both are operating in the reaction system.

Scheme 8.

Scheme 8.

Proposed catalytic cycles for the photoelectrocatalytic oxidation.

Under the photoelectrocatalytic conditions, we did not observe appreciate amounts of TU-1•O2. The major side reaction is the dimerization of thiyl radical 4 to produce dimer 7, which then undergoes irreversible decomposition to H2S and urea 14, the former of which may be detected by odor.

This thiourea-mediated, photoelectrochemically driven reaction mechanism is fundamentally different from that of photochemical oxidation of benzyl alcohols catalyzed by flavin alone (Scheme 2B).12e In the latter, the photoexcited flavin directly removes an electron from the substrate aromatic group to form a radical cation, which then loses H+ to form the corresponding ketyl radical prior to conversion to the aldehyde product. The scope of the flavin-catalyzed reaction is thus limited by the oxidation potential of the flavin excited state and its ability to effect single-electron oxidation of the organic substrate. Such a limitation is overcome in the thiourea-mediated photoelectrocatalytic protocol, as the key catalyst-substrate reaction relies on H atom abstraction by the thiourea-based radicals generated by means of flavin oxidation.

Conclusion

In summary, we developed a new approach, namely molecular photoelectrocatalysis, for the oxidation of unactivated secondary alcohols. This catalytic strategy combines the attractive features of both photoredox chemistry and electrocatalysis, allowing efficient access to highly reactive excited states and radical intermediates in the absence of a stoichiometric oxidant. This merger critically expands the scope of flavin-catalyzed alcohol oxidation from electron-neutral and -rich benzyl alcohols to unactivated aliphatic alcohols. Future work will be directed toward further understanding the mechanism of molecular photoelectrocatalysis and applying this new strategy in new organic reactions.

Experimental Section

An oven-dried, 10 mL two-neck glass tube was equipped with a magnetic stir bar, a vacuum adapter with glass stopcock, a Teflon cap fitted with electrical feed-throughs, a carbon foam anode (1.0 × 0.5 cm2, connected to the electrical feedthrough via a 9cm in length, 2 mm in diameter graphite rod), and a platinum coil cathode. To this reaction vessel, RFT (5.4 mg, 0.01 mmol, 5 mol%), TU-2 (3.2 mg, 0.02 mmol, 10 mol%), alcohol (0.2 mmol, 1.0 equiv), 3.5 mL of electrolyte solution (0.10 M LiOTf in acetonitrile) and H2O (0.2 mL) were added. The cell was sealed and subjected to three freeze-pump-thaw cycles. The resulting solution was stirred at room temperature under the irradiation from two 7W blue LEDs, and electrolysis was initiated at a constant voltage of 2.5 V for 24 hours. After that, the entire reaction mixture was filtered through a short silica gel column (3 cm in length, ca. 5 g) and flushed through with 60 mL of 1:1 mixture of hexanes and ethyl acetate to eliminate the inorganic salts, and the filtrate was concentrated under vacuum. The residue was purified by flash column chromatography on silica gel (eluted with hexanes/ethyl acetate) to afford the pure product.

Supplementary Material

Supp info

Acknowledgements

We are grateful for financial support from Cornell University and NIGMS (R01GM130928). S.L. thanks the Alfred P. Sloan Foundation for a Sloan Fellowship. This study made use of the NMR facility (CHE-1531632) supported by NSF. We thank Greg Sauer for experimental assistance.

References

  • [1].(a) Bäckvall J-E, Modern Oxidation Methods, Wilely-VCH: Weinheim, 2004. [Google Scholar]; (b) Andersson PG, Munslow IJ, Modern Reduction Methods, Wilely-VCH: Weinheim, 2008. [Google Scholar]
  • [2].Blakemore C, Doyle PM, M Fobian Y, Synthetic Methods in Drug Discovery, Vol. 1–2, the Royal Society of Chemistry, Cambridge, 2016. [Google Scholar]
  • [3].Stephenson CRJ, Yoon TP, MacMillan DWC, Visible Light Photocatalysis in Organic Chemistry, Wilely-VCH: Weinheim, 2018. [Google Scholar]
  • [4].(a) Little RD, Weinberg NL, Electroorganic Synthesis, Marcel Dekker, New York, 1991. [Google Scholar]; (b) Hammerich O, Speiser B, Organic Electrochemistry, 5th ed., CRC Press, Boca Raton, 2015. [Google Scholar]
  • [5].Selected reviews, see:; (a) Narayanam JM, Stephenson CR, Chem. Soc. Rev 2011, 40, 102–113. [DOI] [PubMed] [Google Scholar]; (b) Prier CK, Rankic DA, MacMillan DWC, Chem. Rev 2013, 113, 5322–5363. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Schultz DM, Yoon TP, Science 2014, 343, 1239176. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Fukuzumi S, Ohkubo K, Org. Biomol. Chem 2014, 12, 6059–6071. [DOI] [PubMed] [Google Scholar]; (e) Romero NA, Nicewicz DA, Chem. Rev 2016, 116, 10075–10166. [DOI] [PubMed] [Google Scholar]; (f) Marzo L, Pagire SK, Reiser O, König B, Angew. Chem. Int. Ed 2018, 57, 10034–10072. [DOI] [PubMed] [Google Scholar]
  • [6].For representative reviews, see:; (a) Moeller KD, Tetrahedron 2000, 56, 9527–9554. [Google Scholar]; (b) Sperry JB, Wright DL, Chem. Soc. Rev 2006, 35, 605–621. [DOI] [PubMed] [Google Scholar]; (c) Yoshida J-I, Kataoka K, Horcajada R, Nagaki A, Chem. Rev 2008, 108, 2265–2299. [DOI] [PubMed] [Google Scholar]; (d) Francke R, Little RD, Chem. Soc. Rev 2014, 43, 2492–2521. [DOI] [PubMed] [Google Scholar]; (e) Yan M, Kawamata Y, Baran PS, Chem. Rev 2017, 117, 13230–13319. [DOI] [PMC free article] [PubMed] [Google Scholar]; (f) Yoshida J-I, Shimizu A, Hayashi R, Chem. Rev 2018, 118, 4702–4730. [DOI] [PubMed] [Google Scholar]; (g) Wiebe A, Gieshoff T, Möhle S, Rodrigo E, Zirbes M, Waldvogel SR, Angew. Chem. Int. Ed 2018, 57, 5594–5619. [DOI] [PMC free article] [PubMed] [Google Scholar]; (h) Meyer TH, Finger LH, Gandeepan P, Ackermann L, Trends Chem. 2019, 1, 63–76. [Google Scholar]
  • [7].For representative examples, see:; (a) Peters BK, Rodriguez KX, Reisberg SH, Beil SB, Hickey DP, Kawamata Y, Collins M, Starr J, Chen L, Udyavara S, Klunder K, Gorey TJ, Anderson SL, Neurock M, Minteer SD, Baran PS, Science 2019, 363, 838–845. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Kise N, Hamada Y, Sakurai T, Org. Lett 2014, 16, 3348–3351. [DOI] [PubMed] [Google Scholar]; (c) Kolbe H, J. Prakt. Chem 1847, 41, 137–139. [Google Scholar]; (d) Lennox AJJ, Nutting JE, Stahl SS, Chem. Sci 2018, 9, 356–361. [DOI] [PMC free article] [PubMed] [Google Scholar]; Also see ref. 6c and the reference therein.
  • [8].Examples of photoelectrochemical synthesis using a molecular catalyst:; (a) Moutet J-C, Reverdy G, J. Chem. Soc. Chem. Commun 1982, 654–655. [Google Scholar]; (b) Scheffold R, Orlinski R, J. Am. Chem. Soc 1983, 105, 7200–7202. [Google Scholar]; (c) Yan H, Hou Z-W, Xu H-C, Angew. Chem. Int. Ed 2019, 58, 4592–4595. [DOI] [PubMed] [Google Scholar]; (d) Wang F, Stahl SS, Angew. Chem. Int. Ed 2019, 58, 6385–6390. [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) Huang H, M Strater Z, Rauch M, Shee J, Sisto T, Nuckolls C, Lambert TH, Angew. Chem. Int. Ed 2019, 58, doi: 10.1002/anie.201906381. [DOI] [PMC free article] [PubMed] [Google Scholar]; An example of photoelectrochemical synthesis using a photoanode:; (f) Zhang L, Liardet L, Luo J, Ren D, Grätzel M, Hu X, Nature Catal. 2019, 2, 366–373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [9].For representative examples on electrochemical alcohol oxidation, see:; (a) Semmelhack MF, Chou CS, Cortes DA, J. Am. Chem. Soc 1983, 105, 4492–4494. [Google Scholar]; (b) Badalyan A, Stahl SS, Nature 2016, 535, 406–410. [DOI] [PubMed] [Google Scholar]; (c) Das A, Stahl SS, Angew. Chem. Int. Ed 2017, 56, 8892–8897. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Rafiee M, Konz ZM, Graaf MD, Koolman HF, Stahl SS, ACS Catal. 2018, 8, 6738–6744. [Google Scholar]; (e) Hickey DP, McCammant MS, Giroud F, Sigman MS, Minteer SD, J. Am. Chem. Soc 2014, 136, 15917–15920. [DOI] [PubMed] [Google Scholar]; (f) Hickey DP, Schiedler DA, Matanovic I, Doan PV, Atanassov P, Minteer SD, Sigman MS, J. Am. Chem. Soc 2015, 137, 16179–16186. [DOI] [PubMed] [Google Scholar]
  • [10].(a) Massey V, Biochem. Soc. Trans 2000, 28, 283–296. [PubMed] [Google Scholar]; (b) Mansoorabadi SO, Thibodeaux CJ, Liu H, J. Org. Chem 2007, 72, 6329–6342. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Walsh CT, Wencewicz TA, Nat. Prod. Rep 2013, 30, 175–200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [11].König B, Mario P, Roland RK, Jürgen S, Jörg DA, Eur. J. Org. Chem 2001, 2001, 2297–2303. [Google Scholar]
  • [12].(a) Lechner R, Kümmel S, König B, Photochem. Photobiol. Sci 2010, 9, 1367–1377. [DOI] [PubMed] [Google Scholar]; (b) Mühldorf B, Wolf R, Chem. Commun 2015, 51, 8425–8428. [DOI] [PubMed] [Google Scholar]; (c) Mühldorf B, Wolf R, Angew. Chem. Int. Ed 2016, 55, 427–430. [DOI] [PubMed] [Google Scholar]; (d) Dongare P, MacKenzie I, Wang D, Nicewicz DA, Meyer TJ, Proc. Natl. Acad. Sci. U. S. A, 2017, 114, 9279–9283. [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) Tang G, Gong Z, Han W, Sun X, Tetrahedron Lett. 2018, 59, 658–662. [Google Scholar]
  • [13].(a) Fukuzumi S, Tanii K, Tanaka T, J Chem Soc., Chem Commun 1989, 816–818. [Google Scholar]; (b) Cibulka R, Vasold R, Konig B, Chem. Eur. J 2004, 10, 6223–6231. [DOI] [PubMed] [Google Scholar]; (c) Svoboda J, Schmaderer H, Konig B, Chem. Eur. J 2008, 14, 1854–1865. [DOI] [PubMed] [Google Scholar]; (d) Schmaderer H, Hilgers P, Lechner R, König B, adv. Synth. Catal 2009, 351, 163–174. [Google Scholar]; (e) Megerle U, Wenninger M, Kutta R-J, Lechner R, König B, Dick B, Riedle E, Phys. Chem. Chem. Phys 2011, 13, 8869–8880. [DOI] [PubMed] [Google Scholar]; (f) Feldmeier C, Bartling H, Magerl K, Gschwind RM, Angew. Chem. Int. Ed 2015, 54, 1347–1351. [DOI] [PubMed] [Google Scholar]; (g) Korvinson KA, Hargenrader GN, Stevanovic J, Xie Y, Joseph J, Maslak V, Hadad CM, Glusac KD, J. Phys. Chem. A 2016, 120, 7294–7300. [DOI] [PubMed] [Google Scholar]; (h) Hering T, Mühldorf B, Wolf B R. Konig, Angew. Chem. Int. Ed 2016, 55, 5342–5345. [DOI] [PMC free article] [PubMed] [Google Scholar]; (i) Kurfirt M, spacková J, Svobodova E, Cibulka R, Monatsh Chem. 2018, 149, 863–869. [Google Scholar]
  • [14].(a) Lechner R, König B, Synthesis 2010, 1712. [Google Scholar]; (b) Murray AT, Dowley MJ, Pradaux-Caggiano F, Baldansuren A, Fielding AJ, Tuna F, Hendon CH, Walsh A, Lloyd-Jones GC, John MP, Carbery DR, Angew. Chem. Int. Ed 2015, 54, 8997–9000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [15].(a) Dad’ová J, Svobodová E, Sikorski M, König B, Cibulka R, ChemCatChem 2012, 4, 620–624. [Google Scholar]; (b) Neveselý T, Svobodová E, Chudoba J, Sikorski M, Cibulka R, Adv. Synth. Catal 2016, 358, 1654–1663. [Google Scholar]; (c) Dang C, Zhu L, Guo H, Xia H, Zhao J, Dick B, ACS Sustainable Chem. Eng 2018, 6, 15254–15263. [Google Scholar]
  • [16].(a) Ramirez NP, König B, Gonzalez-Gomes JC, Org. Lett 2019, 1368–1373. [DOI] [PubMed] [Google Scholar]; (b) Metternich JB, Gilmour R, J. Am. Chem. Soc 2016, 138, 1040–1045. [DOI] [PubMed] [Google Scholar]; (c) Bloom S, Liu C, Kölmel DK, Qiao JX, Zhang Y, Poss MA, Ewing WR, MacMillan DWC, Nat. Chem 2017, 10, 205–211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [17].Zelenka J, Svobodová E, Tarábek J, Hoskovcová I, Boguschová V, Bailly S, Sikorski M, Roithová J, Cibulka R, Org. Lett 2019, 21, 114–119. [DOI] [PubMed] [Google Scholar]
  • [18].(a) König B, Kümmel S, Svobodová E, Cibulka R, Phys. Sci. Rev 2018, 3, 20170168. [Google Scholar]; (b) Iida H, Imada Y, Murahashi S-I, Org. Biomol. Chem 2015, 13, 7599–7613. [DOI] [PubMed] [Google Scholar]; (c) Cibulka R, Eur. J. Org. Chem 2015, 2015, 915–932. [Google Scholar]; (d) de Gonzalo G, Fraaije MW, ChemCatChem 2013, 5, 403–415 [Google Scholar]; (e) Gelalcha FG, Chem. Rev 2007, 107, 3338–3361. [DOI] [PubMed] [Google Scholar]
  • [19].(a) Shinkai S, Nakao H, Ueda K, Manabe O, Tetrahedron Lett. 1984, 25, 5295–5298. [Google Scholar]; (b) Fukuzumi S, Kuroda S, Tanaka T, J. Am. Chem. Soc 1985, 107, 3020–3027. [Google Scholar]; (c) Fukuzumi S, Yasui K, Suenobu T, Ohkubo K, Fujitsuka M, Ito O, J. Phys. Chem. A 2001, 105, 10501–10510. [Google Scholar]
  • [20].Denisov ET, Tumanov VE, Russ. Chem. Rev 2005, 74, 825–858. [Google Scholar]
  • [21].Oyeyemi VB, Keith JA, Carter EA, J. Phys. Chem. A 2014, 118, 3039–3050. [DOI] [PubMed] [Google Scholar]
  • [22].Meyers AI, Nolen RL, Collington EW, Narwid TA, Strickland RC, J. Org. Chem 1973, 38, 1974–1982. [DOI] [PubMed] [Google Scholar]
  • [23].(a) Have JJ, Kluttz RQ, Synth. Commun 1974, 4, 389–393. [Google Scholar]; (b) Crank G, Mursyidi A, J. Photochem. Photobiol. A: Chem 1992, 44, 263–271. [Google Scholar]
  • [24].(a) Ishikawa M, Okimoto H, Morita M, Matsuda Y, Chem. Lett 1996, 25, 953–954. [Google Scholar]; (b) Mirzakulova E, Khatmullin R, Walpita J, Corrigan T, Vargas-Barbosa NM, Vyas S, Oottikkal S, Manzer SF, Hadad CM, Glusac KD, Nat. Chem 2012, 4, 794–801. [DOI] [PubMed] [Google Scholar]
  • [25].(a) Cruz CL, Nicewicz DA, ACS Catal. 2019, 9, 3926–3935. [Google Scholar]; (b) Chatgilialoglu C, Ferreri C, Guerra M, Samadi A, Bowry VW, J. Am. Chem. Soc 2017, 139, 4704–4714. [DOI] [PubMed] [Google Scholar]; (c) Sprinz H, Schwinn J, Naumov S, Brede O, Biochim. Biophys. Acta, Mol. Cell Biol. Lipids 2000, 1483, 91–100. [DOI] [PubMed] [Google Scholar]; (d) Dénès F, Pichowicz M, Povie G, Renaud P, Chem. Rev 2014, 114, 2587–2693. [DOI] [PubMed] [Google Scholar]
  • [26].Metternich JB, Gilmour R, J. Am. Chem. Soc 2015, 137, 11254–11257. [DOI] [PubMed] [Google Scholar]
  • [27].Huang C, Qian X-Y, Xu H-C, Angew. Chem. Int. Ed 2019, 58, 6650–6653. [DOI] [PubMed] [Google Scholar]
  • [28].Selected examples of HATα-to the hydroxy group, see:; (a) Jeffrey JL, Terrett JA, MacMillan DWC, Science 2015, 349, 1532–1536. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Twilton J, Christensen M, DiRocco DA, Ruck RT, Davies IW, C MacMillan DW, Angew. Chem. Int. Ed 2018, 57, 5369–5373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [29].Liu X, Gross ML, Wenthold PG, J. Phys. Chem. A, 2005, 109, 2183–2189. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supp info

RESOURCES