140
doi.org/10.26434/chemrxiv.13055522.v1 Coupling of Alternating Current to Transition-Metal Catalysis: Examples of Nickel-Catalyzed Cross-Coupling Evgeniy Bortnikov, Sergey Semenov Submitted date: 06/10/2020 Posted date: 06/10/2020 Licence: CC BY-NC-ND 4.0 Citation information: Bortnikov, Evgeniy; Semenov, Sergey (2020): Coupling of Alternating Current to Transition-Metal Catalysis: Examples of Nickel-Catalyzed Cross-Coupling. ChemRxiv. Preprint. https://doi.org/10.26434/chemrxiv.13055522.v1 This work demonstrates that periodic oxidation and reduction of a catalyst by alternating current enable otherwise unfavorable catalytic cycles. Nickel catalyzed amination, etherification, and esterification were universally enabled by alternating current with yields and selectivity strongly exciding these in the experiments with direct current (DC). File list (2) download file view on ChemRxiv Manuscript - one column version_ss.pdf (1.07 MiB) download file view on ChemRxiv Supporting Information.pdf (8.61 MiB)

Coupling of Alternating Current to Transition-Metal

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Coupling of Alternating Current to Transition-Metal

doi.org/10.26434/chemrxiv.13055522.v1

Coupling of Alternating Current to Transition-Metal Catalysis: Examplesof Nickel-Catalyzed Cross-CouplingEvgeniy Bortnikov, Sergey Semenov

Submitted date: 06/10/2020 • Posted date: 06/10/2020Licence: CC BY-NC-ND 4.0Citation information: Bortnikov, Evgeniy; Semenov, Sergey (2020): Coupling of Alternating Current toTransition-Metal Catalysis: Examples of Nickel-Catalyzed Cross-Coupling. ChemRxiv. Preprint.https://doi.org/10.26434/chemrxiv.13055522.v1

This work demonstrates that periodic oxidation and reduction of a catalyst by alternating current enableotherwise unfavorable catalytic cycles. Nickel catalyzed amination, etherification, and esterification wereuniversally enabled by alternating current with yields and selectivity strongly exciding these in the experimentswith direct current (DC).

File list (2)

download fileview on ChemRxivManuscript - one column version_ss.pdf (1.07 MiB)

download fileview on ChemRxivSupporting Information.pdf (8.61 MiB)

Page 2: Coupling of Alternating Current to Transition-Metal

Coupling of Alternating Current to Transition-Metal Catalysis: Examples of Nickel-Catalyzed

Cross-Coupling

Evgenii O. Bortnikova and Sergey N. Semenov,a*

a Department of Organic Chemistry, Weizmann Institute of Science,

234 Herzl Street, Rehovot 76100, Israel

* The author to whom correspondence should be addressed.

Page 3: Coupling of Alternating Current to Transition-Metal

Abstract

The coupling of transition-metal to photoredox catalytic cycles through single-electron transfer steps has

become a powerful tool in the development of catalytic processes. In this work we demonstrated that

transition-metal catalysis can be coupled to alternating current (AC) through electron transfer steps that

occur periodically at the same electrode. AC-assisted Ni-catalyzed amination, etherification, and esteri-

fication of aromatic bromides showed higher yields and selectivity compared to observed in the control

experiments with direct current (DC). Our mechanistic studies suggested the importance of both reduc-

tion and oxidation processes in the maintenance of the AC-assisted catalytic reactions. As in presented

examples, the AC assistance should be well-suited for catalytic cycles involving reductive elimination

or oxidative addition as a limiting step.

Page 4: Coupling of Alternating Current to Transition-Metal

Introduction

Ni-based catalysis has attracted much attention as an effective tool for forming the C(aryl)-

Heteroatom bond,1 which is frequently found in drug-like molecules, dyes, and conjugated polymers.

However, Ni-catalytic systems are often inferior to more prevalent but expensive Pd-catalytic systems.

Unfavorable reductive elimination from Ni(II) species and the instability of the Ni(I) species are among

the factors causing this inferiority. Nevertheless, Ni-catalysis is susceptible to facilitation by energy in-

puts from light or electricity because of the availability of redox states of nickel ranging from 0 to +4

(Figure 1A-B).

Figure 1. Some methods for activating Ni-based catalytic cycles by sequential oxidation and reduction:

(A) Photoredox-assisted; (B) DC-assisted; (C) AC-assisted couplings. (D) Working hypothesis of the

mechanism underlying the coupling of alternating current (AC) to the nickel catalytic cycle for cross-

coupling reactions.

Page 5: Coupling of Alternating Current to Transition-Metal

By merging photoredox and nickel catalysis, MacMillan and others enabled previously elusive Ni-

catalyzed C-C, C-N, and C-O coupling reactions.2 In most situations, the coupling between photoredox

and nickel catalysis involves single-electron oxidation and reduction of nickel catalytic species by a

photoredox catalyst (Figure 1A).3 Thus, oxidation to Ni(III) species enables reductive elimination,

whereas reduction to Ni(0) species accelerates oxidative addition. Later, Baran’s group, regarding the

example of C-N coupling, showed that the photocatalytic approach might be successfully replaced with

assistance by direct current (DC) electrolysis.4 Here, nickel species are oxidized on one electrode, ena-

bling reductive elimination, and reduced on another electrode, accelerating oxidative addition (Figure

1B); the current-assisted catalytic cycle involving nickel species in oxidation states +1, +2, and +3 was

suggested.5 This approach was effective for the C-S,6 C-P,7 Heck,8 and C-O coupling reactions.9

Although some disagreements exist regarding key reaction intermediates,5, 10 mechanistic studies of

both electrochemically and photochemically enabled nickel cross-coupling suggest that both oxidation

and reduction processes are essential for maintaining the sustainable catalytic pathway.2a, 2d, 3a-b These

mechanistic considerations led us to the idea of merging alternating current (AC) with Ni-catalyzed

cross-coupling reactions (Figure 1C). AC authentically fits the idea of a catalytic cycle assisted by a

pair of redox processes. Because of the periodical polarity switch, both redox steps can successfully oc-

cur at the same electrode with a tunable delay between them (Figure 1D). In this work, we performed

AC-assisted Ni-catalyzed amination, etherification, and esterification of aromatic bromides. For all reac-

tions, AC displayed advantages over DC in our electrochemical cell.

Currently, AC has a significantly more modest application in organic synthesis compared with DC –

generally, it is used for preventing electrode fouling. Notable works showed that AC can be successful-

ly applied in the electrosynthesis of phenol,11 trifluoromethylated arenes,12 or in accelerating S-S bond

metathesis through reversible redox steps.13

Page 6: Coupling of Alternating Current to Transition-Metal

Results and discussion

We performed reactions in an electrochemical cell equipped with two glassy carbon (GC) rod elec-

trodes. A commercial waveform generator supplied sinusoidal voltage to the electrodes; we used an os-

cilloscope to measure both the voltage and current in our experiments (Figure 2A, Figure S3). The com-

binations of NiBr2·DME (DME – dimethoxyethane) with 2,2′-Dipyridyl (Bipy) or 4,4′-Di-tert-butyl-

2,2′-dipyridyl (di-tBuBipy) were chosen as catalysts because they were successfully employed in

various catalytic systems.2d, 5

To perform the catalytic cycle based on both reduction and oxidation of the catalyst, the voltage

should be sufficient to perform the reduction on one of the electrodes, whereas the oxidation occurs on

another electrode. We proposed that reduction of [Ni(Bipy)x]2+ to [Ni(Bipy)x]

0 (for x = 3, -1.25 V vs sat-

urated calomel electrode (SCE) in CH3CN) is most likely to be a process that would favor oxidative ad-

dition.14 Similarly, oxidation of [NiL(Ar)(Nu)] to [NiL(Ar)(Nu)]+ (L – ligand, Ar – aryl, Nu – nucleo-

phile) would favor the reductive elimination of Ar-Nu.15 Considering that the potential for oxidation of

[Ni(C6H3(CF3)2)(OCH2CF3)] to [Ni(C6H3(CF3)2)(OCH2CF3)]+ is 0.83 V vs SCE in CH3CN, which is at

the high end of the expected oxidation potentials for the [NiL(Ar)(Nu)] species,2d a voltage higher than

~ 2.1 V is required to simultaneously perform the desired oxidation and reduction of nickel species. We

found a peak voltage of 3V to be effective in most of our experiments; in such conditions the peak value

of current varied from 10 mA to 20 mA in different experiments.

In the initial screening, we aimed to optimize the conditions for amination, etherification, and esteri-

fication of aryl bromides (Figure 2B-D and Tables S1-3). We used the reaction between bromobenzene

(1) (50 mM) and morpholine (150 mM) for optimization of the amination conditions because 1 being a

less reactive substrate than 4-bromobenzotrifluoride (3) offered more room for the optimization (Figure

2B). Different combinations of ligands (di-tBuBipy and Bipy) and solvents (dimethylformamide (DMF)

and dimethylacetamide (DMA)) worked, but combining DMA with di-tBuBipy provided better stability

of the nickel complex in the solution and more reproducible results than the other systems did. The fre-

Page 7: Coupling of Alternating Current to Transition-Metal

quency of the applied voltage is a unique parameter for AC-assisted catalysis. Interestingly, the depend-

ency of yield on frequency has a maximum of around 2Hz (87%). The increase in frequency to 25 Hz

caused a steep drop in yield, most likely because of the increased contribution of charging and discharg-

ing of the electrical double layer (see Figure S5 for the quantitative analysis) and of nonproductive re-

versible Ni(II)/Ni(I) and Ni(II)/Ni(0) oxidation/reduction cycles to the current at high frequencies. At

the same time, the decrease in frequency as well as the use of DC also led to diminished yields. We can

speculate at this point that in case of the use of AC assistance some optimum “resonance” frequency ex-

ists which is defined by the complex combination of the parameters such as rates of chemical and elec-

tron-transfer steps, rate of diffusion of the catalytic and non-catalytic species from and to the surface of

the electrodes, and stirring rate. Yet, we expect that under specific conditions the value of “resonance”

frequency may provide valuable data related to the kinetics of the discussed reactions.

In contrast to amination, the base is an essential component in esterification (Figure 2C). We achieved

the best results using a suspension of potassium carbonate as in photo-activated esterification16; organic

bases were found to be less effective. The increase of the ligand concentration or the use of more steri-

cally hindered ligands than bipyridine resulted in diminished yields. The choice of NBu4PF6 as a sup-

porting electrolyte was instrumental for getting good yields of esters; we suggest two possible reasons

for that. First, carboxylate is a relatively weak ligand; thus, eliminating the competition of carboxylate

with bromide at the nickel center increases the efficiency of a catalytic cycle. Second, carboxylate is a

more electronegative group than amine or alcoholate; thus, the oxidation potential of the

[NiL(Ar)(RCOO)] species (e.g., 0.9 V vs SCE for [Ni(C6H3(CF3)2)(OAc)]) is expected to be higher than

that of [NiL(Ar)(OR)] and the [NiL(Ar)(NHR2)] species.2d, 15 Avoiding excessive anodic oxidation of

Br- allows one to achieve higher oxidation potentials in experiments with NBu4PF6 than with LiBr (Fig-

ure S18), which might be necessary for the oxidation of the [NiL(Ar)(RCOO)] species. The optimum

frequency for esterification reaction was 2 Hz.

Page 8: Coupling of Alternating Current to Transition-Metal

Figure 2. (A) Experimental set-up for the AC-assisted cross-coupling reactions. (B–D) Optimization of

conditions for AC-assisted amination, esterification, and etherification reactions, respectively. The com-

parison of the results of the experiments with AC-assistance (using optimized conditions) and DC-

assistance (using potentiostatic and galvanostatic conditions) is given in red boxes.

Page 9: Coupling of Alternating Current to Transition-Metal

The control DC experiments at potentiostatic (2.8 V) and galvanostatic (2, 4, and 6 mA) conditions

demonstrated low yields (13-26 %) and high amount (22-53%) of biaryl side product (see Table S2). As

esterification, etherification also requires the addition of a base (Figure 2D); we tested 1,8-

Diazabicyclo[5.4.0]undec-7-ene (DBU) and 1,4-diazabicyclo[2.2.2]octane (DABCO) as bases; however,

the best results were achieved with quinuclidine, analogously to photoredox-assisted etherification.2d

Interestingly, the use of non-cyclic tertiary amines led to the formation of the homocoupling product in

considerable amounts (Table S3) – presumably, they serve as sacrificial reductants in this case.

Comparing the results of AC- and DC-assisted (both in the potentiostatic and galvanostatic modes)

etherification further highlights that AC-assistance demonstrates higher selectivity towards the

formation of cross-coupling products vs homocoupling products than DC-assistance. Moreover, the

decrease in the amount of the biaryl product occurs gradually with an increase in the frequency. Two

hypotheses could explain these observations: (i) the shorter lifetime of [NiL(Ar)(Nu)] species in the AC

than in the DC experiments; (ii) fewer chances for the second oxidative addition in the AC than in the

DC experiments. With a short lifetime, [NiL(Ar)(Nu)] does not have time to disproportionate

appreciably to the [NiL(Ar)2] intermediate with two aryl groups at one nickel center, which affords

diaryl coupling products by reductive elimination.17 In the DC experiments, a continuous strongly

reducing environment near the cathode increases chances that the oxidative addition product

[NiL(Ar)Br] will undergo reduction to Ni(I) species and subsequently a second oxidative addition, re-

sulting in [NiL(Ar)2Br] intermediates that eliminate diaryl products.17-18

Next, we studied the substrate scope of the discussed coupling reactions (Figure 3). Amination appears

to be the most robust reaction in relation to the activity of the substrates – couplings with non-activated

(1), sterically hindered aryl halides (o-bromotoluene), and halides bearing electron-donor groups (m-

bromoanisole) exhibited good yields (61-74%) (Figure 3A). The amine scope is not limited only to sec-

ondary and primary amines – the hydrochlorides of amines successfully reacted in the presence of the

suspension of K2CO3.

Page 10: Coupling of Alternating Current to Transition-Metal

Figure 3. The scope of AC-assisted nickel-catalyzed amination (A), esterification (B), and etherification

(C). Isolated yields are shown. NMR yields are shown in parentheses. For details of the experimental

protocols, see the Supporting information. Boc – tertbutoxycarbonyl. aA mixture of the corresponding

amine hydrochloride (3 eq.) and K2CO3 (5 eq.) was used instead of amine. bSodium acetate (5 eq.) was

used instead of a mixture of carboxylic acid and K2CO3.

Page 11: Coupling of Alternating Current to Transition-Metal

Esterification of 3 resulted in 48-75% yields for aliphatic, aromatic, and Boc (tert-Butyloxycarbonyl)

protected amino acids (Figure 3B). Other electron-deficient aryl bromides (i.e., 4-bromobenzonitrile and

methyl 4-bromobenzoate) and 2-bromonaphtalene also reacted smoothly with Boc-L-proline with 45-79

% yields. As expected, the highest yields (66-79%) in etherification were achieved in reactions with

electron-deficient aryl bromides and primary alcohols (Figure 3C). The reactions with secondary alco-

hols - cyclohexanol and isopropanol – gave 50% and 62% yield, respectively, but required a 6-fold in-

stead of a 3-fold excess of alcohol and a prolonged reaction time.

It is worth mentioning that additional control DC-assisted experiments of amination, esterification and

etherification,performed using potentiostatic mode (for substrates 7, 24, and 29), demonstrated lower

NMR yields (66%, 4%, and 29%, respectively) (see Supporting information for details) compared to the

NMR yields in identical AC-assisted reactions (98%, 79%, and 66%, respectively) (Figure 3).

Overall, the advantage of AC over DC assistance is the most pronounced for esterification and etheri-

fication and, to a less extent, for amination. This observation is in agreement with literature data: while

a very broad scope was demonstrated for DC-assisted amination,4,5 only humble scope with 32-43%

yields was shown for etherification.5 DC-assisted esterification was enabled only using a specific micro-

fluidic cell with a narrow (≤0.5 mm) gap between the electrodes 9. This close distance between the elec-

trodes allowed to perform reactions that involve short-living intermediates existing between reduction

and oxidation steps. AC provides an alternative way to conduct reactions that require a short interval

between reduction and oxidation.

To better understand the basic mechanistic features of AC-assisted nickel-catalyzed cross-coupling,

we performed electrochemical studies of the discussed reactions. First, we measured the potentials of

the electrodes (the glassy carbon rods) relatively standard nonaqueous silver electrode (Ag+/Ag) in typi-

cal amination, etherification, and esterification experiments. As Figure 4A shows, the potential changes

from 0.8 to -2.2 V (vs Ag+/Ag) in the amination reaction. This interval remains the same for the etherifi-

Page 12: Coupling of Alternating Current to Transition-Metal

cation (Figure S19); however, for esterification, it is shifted to slightly higher potentials because of the

absence of Br- in the electrolyte

Figure 4. Mechanistic studies of the cross-coupling reactions. (A) GC electrode potential vs Ag+/Ag in a

typical amination experiment (3 (50 mM), morpholine (150 mM), NiBr2·DME (5 mM), di-tBuBipy (7.5

mM) in LiBr (0.2M) DMA solution. (B) CV curve of a NiBr2·DME (2mM) and a Bipy (2mM) solution

before (blue) and after (gray) the addition of 3 (2mM) (0.1M TBAPF6, DMA). (C) Amination in the di-

vided cell without stirring. The concentrations of the reagents are identical to A. (D) Control experiment

of amination with DC and Zn sacrificial anode. The concentrations of the reagents are identical to A. (E)

Kinetics of the amination reaction between 3 (50 mM) and morpholine (150 mM) (NiBr2·DME (5 mM),

Bipy (7.5 mM) in LiBr (0.2M) DMA solution) with ON/OFF cycles of AC. Conversion of 3 to 7 was

determined by HPLC. See Supporting information for details.

Page 13: Coupling of Alternating Current to Transition-Metal

(Figure S20). To better understand what electrochemical processes occur within these potentials in the

reactions under investigation, we performed cyclic voltammetry (CV) studies (Figure 4B and Support-

ing information). The voltammogram of Ni(DME)Br2 (2 mM) and Bipy (2 mM) showed two partially

separated quasi-reversible reduction waves with peaks at -1.42 and -1.57 V (Figure 4B). The coulo-

metric studies19 and the combination of the voltammograms for [Ni(Bipy)3]2+ and Ni(Bipy)Br2

17, 19b, 20

strongly indicate that these waves represent the sequential reduction of Ni(II) to Ni(0) through a Ni(I)

intermediate. When 3 (2 mM) was added to a solution of Ni(Bipy)Br2, the original reduction waves

could still be detected, but they became less resolved and fully irreversible; two new reduction waves

appeared with peak potentials of -1.71 and -1.89 V. We hypothesized that oxidative addition occurs after

the reduction of nickel to a zero oxidation state, making the reduction waves irreversible, and that these

additional waves (-1.71 and -1.89 V) correspond to the sequential reduction of the [NiL(Ar)Br]

species.21 Nevertheless, the voltammetry under conditions as close as possible to the synthetic

experiments (GC rods as working and counter electrodes, scan rate of 24 V/s, and the concentrations of

reagents as in synthesis) indicated that in synthetic experiments the reduction of the [NiL(Ar)Br] species

might be less significant than in analytical CV because of the higher surface area of the electrode, the

higher concentration of nickel ions, and the higher scan rate (Figs. S8, 12, 16). These data imply that

NiL(Ar)Br is the most probable species to undergo further ligand exchange and oxidation to afford the

product of cross-coupling (Figure 1D). The anodic area of the CV curves of the amination and etherifi-

cation reaction mixtures is dominated by the oxidation of Br- (Figs. S18); therefore, Br3- and Br2 are

probably mediators of the electrochemical oxidation of nickel species in these experiments.

To confirm that processes at one electrode are sufficient to perform the reactions, we conducted the

AC-assisted reaction between 3 and morpholine in a divided cell without stirring (Figure 4C). This

experiment resulted in a 77% yield, confirming that the transfer of intermediates between electrodes is

unnecessary for this reaction to proceed. To probe the possibility that only the reduction phase of the AC

cycle is essential for the coupling, we performed DC amination and etherification of 3 with a Zn

Page 14: Coupling of Alternating Current to Transition-Metal

sacrificial anode and set the potential of the GC cathode equal to the peak negative potential in AC

experiments (-2.2 V vs Ag+/Ag) (Figure 4D, Figure S22). Both experiments resulted in the formation of

the biaryl product (30) with 60-70% yields and only minor quantities (2-8 %) of the coupling products 7

and 5, thus indicating the essential role of the oxidation phase of the AC cycle in forming the coupling

products. To probe the role of the self-sustainable catalysis by Ni(I) species in bulk solution (Figure

5A), we performed kinetic experiments with the ON/OFF cycles of AC (Figure 4E, Figure S23, 24)).

The kinetics of the amination reaction between 3 and morpholine as well as the data for similar

esterification and etherification experiments indicate that the reactions stop almost immediately when

AC is OFF and start again when AC is ON, demonstrating that the contribution of the long-living

(minutes timescale) self-sustainable Ni(III)/Ni(I) cycles to the formation of the products (7, 5, 18) is

insignificant; short-living Ni(III)/Ni(I) cycles that require constant regeneration of Ni(I) remain

possible.

Figure 5. Possible catalytic cycles involving oxidative addition to Ni(I) species. (A) Self-sustainable

Ni(III)/Ni(I) cycle. (B) Current-assisted Ni(III)/Ni(I) cycle.5

Overall, our mechanistic studies are consistent with the proposal in Figure 1D. Likely, the catalytic

cycle is initiated by the reduction of Ni(II), which is abundant in the solution, to the Ni(0) species,

which undergo oxidative addition. During the low-voltage phase, when neither oxidation nor reduction

processes are expected, the oxidative addition product undergoes ligand exchange. The further oxidation

of [NiL(Ar)Nu] species to a Ni(III) state favors reductive elimination, which would furnish the desired

Page 15: Coupling of Alternating Current to Transition-Metal

coupling product and release Ni(I) species. There are at least two ways by which Ni(I) could be reduced

to Ni(0): (i) direct reduction at the electrode, and (ii) reversible disproportionation to Ni(II) and Ni(0)

with the subsequent reduction of Ni(II) to Ni(0) at the electrode. At the same time, current-assisted cata-

lytic cycles involving oxidative addition to Ni(I) species (Figure 5B)5 may coexist with the proposed

cycle (Figure 1D).

Conclusion

To summarize, we demonstrated the coupling of AC to transition metal catalysis using three examples

of related nickel-catalyzed coupling reactions, but the idea of enabling new catalytic cycles by periodi-

cally oxidizing and reducing catalytic intermediates by AC could be applied to various catalytic sys-

tems, especially to ones that involve oxidative addition or reductive elimination as a rate-limiting step.

The use of AC provides two important advantages: (i) the absence of the need to transfer reactive inter-

mediates between electrodes, which prevents their dilution and allows working with short-living inter-

mediates; (ii) the frequency and the waveform of AC are easily tunable experimental parameters that

can be used to achieve the selectivity of reactions. Moreover, side electrochemical reactions might be

tolerated in experiments with AC if they are fully reversible. We hope that further development of syn-

thetic methods based on the coupling of transition metal catalysis to AC will enable new efficient trans-

formations.

In addition, this work demonstrates well how external oscillations (not necessarily electrochemical

ones) could couple to catalytic cycles to perform otherwise unfavorable chemical transformations. The

oscillator pumps energy into the catalytic cycle by periodically modifying its intermediates. This point

of view might be relevant for development of catalytic processes enabled by various oscillatory fields.

ASSOCIATED CONTENT

Supporting Information. Experimental procedures, electrochemical analysis data and compound char-

acterization data. This material is available free of charge via the Internet at http://pubs.acs.org.

Page 16: Coupling of Alternating Current to Transition-Metal

AUTHOR INFORMATION

Corresponding Author

* E-mail: [email protected]

Funding Sources

This work was funded by Weizmann Institute of Science.

ACKNOWLEDGMENT

We thank E. Narevicius for the help with setting up equipment for AC experiments. We thank R. Neu-

mann and M. Somekh for helpful discussions.

REFERENCES

(1) Tasker, S. Z.; Standley, E. A.; Jamison, T. F. Recent advances in homogeneous nickel catalysis.

Nature 2014, 509, 299-309.

(2) (a) Corcoran, E. B.; Pirnot, M. T.; Lin, S.; Dreher, S. D.; DiRocco, D. A.; Davies, I. W.; Buch-

wald, S. L.; MacMillan, D. W. Aryl amination using ligand-free Ni(II) salts and photoredox catalysis.

Science 2016, 353, 279-283. (b) Zuo, Z.; Ahneman, D. T.; Chu, L.; Terrett, J. A.; Doyle, A. G.; Mac-

Millan, D. W. Dual catalysis. Merging photoredox with nickel catalysis: coupling of alpha-carboxyl

sp(3)-carbons with aryl halides. Science 2014, 345, 437-440. (c) Johnston, C. P.; Smith, R. T.; All-

mendinger, S.; MacMillan, D. W. Metallaphotoredox-catalysed sp(3)-sp(3) cross-coupling of carboxylic

acids with alkyl halides. Nature 2016, 536, 322-325. (d) Terrett, J. A.; Cuthbertson, J. D.; Shurtleff, V.

W.; MacMillan, D. W. Switching on elusive organometallic mechanisms with photoredox catalysis. Na-

ture 2015, 524, 330-334. (e) Twilton, J.; Le, C.; Zhang, P.; Shaw, M. H.; Evans, R. W.; MacMillan, D.

W. C. The merger of transition metal and photocatalysis. Nat. Rev. Chem. 2017, 1, 0052.

(3) (a) Zhu, B.; Yan, L. K.; Geng, Y.; Ren, H.; Guan, W.; Su, Z. M. Ir(III)/Ni(II)-Metallaphotoredox

catalysis: the oxidation state modulation mechanism versus the radical mechanism. Chem. Commun.

Page 17: Coupling of Alternating Current to Transition-Metal

2018, 54, 5968-5971. (b) Qi, Z.-H.; Ma, J. Dual Role of a Photocatalyst: Generation of Ni(0) Catalyst

and Promotion of Catalytic C–N Bond Formation. ACS Catalysis 2018, 8, 1456-1463. (c) Levin, M. D.;

Kim, S.; Toste, F. D. Photoredox Catalysis Unlocks Single-Electron Elementary Steps in Transition

Metal Catalyzed Cross-Coupling. ACS Cent. Sci. 2016, 2, 293-301.

(4) Li, C.; Kawamata, Y.; Nakamura, H.; Vantourout, J. C.; Liu, Z.; Hou, Q.; Bao, D.; Starr, J. T.;

Chen, J.; Yan, M.; Baran, P. S. Electrochemically Enabled, Nickel-Catalyzed Amination. Angew. Chem.

Int. Ed. 2017, 56, 13088-13093.

(5) Kawamata, Y.; Vantourout, J. C.; Hickey, D. P.; Bai, P.; Chen, L.; Hou, Q.; Qiao, W.; Barman,

K.; Edwards, M. A.; Garrido-Castro, A. F.; deGruyter, J. N.; Nakamura, H.; Knouse, K.; Qin, C.;

Clay, K. J.; Bao, D.; Li, C.; Starr, J. T.; Garcia-Irizarry, C.; Sach, N.; White, H. S.; Neurock, M.;

Minteer, S. D.; Baran, P. S. Electrochemically Driven, Ni-Catalyzed Aryl Amination: Scope, Mecha-

nism, and Applications. J. Am. Chem. Soc. 2019, 141, 6392-6402.

(6) Wang, Y.; Deng, L.; Wang, X.; Wu, Z.; Wang, Y.; Pan, Y. Electrochemically Promoted Nickel-

Catalyzed Carbon–Sulfur Bond Formation. ACS Catal. 2019, 9, 1630-1634.

(7) Sengmany, S.; Ollivier, A.; Le Gall, E.; Leonel, E. A mild electroassisted synthesis of (het-

ero)arylphosphonates. Org. Biomol. Chem. 2018, 16, 4495-4500.

(8) Tian, J.; Moeller, K. D. Electrochemically Assisted Heck Reactions. Org. Lett. 2005, 7, 5381-5383.

(9) Mo, Y.; Lu, Z.; Rughoobur, G.; Patil, P.; Gershenfeld, N.; Akinwande, A. I.; Buchwald, S. L.;

Jensen, K. F. Microfluidic electrochemistry for single-electron transfer redox-neutral reactions. Science

2020, 368, 1352-1357.

(10) Sun, R.; Qin, Y.; Nocera, D. G. General Paradigm in Photoredox Nickel-Catalyzed Cross-

Coupling Allows for Light-Free Access to Reactivity. Angew. Chem., Int. Ed. 2020, 59, 9527-9533.

(11) Lee, B.; Naito, H.; Nagao, M.; Hibino, T., Alternating-current electrolysis for the production of

phenol from benzene. Angew. Chem., Int. Ed. 2012, 51, 6961-5.

Page 18: Coupling of Alternating Current to Transition-Metal

(12) Rodrigo, S.; Um, C.; Mixdorf, J. C.; Gunasekera, D.; Nguyen, H. M.; Luo, L. Alternating Cur-

rent Electrolysis for Organic Electrosynthesis: Trifluoromethylation of (Hetero)arenes. Org. Lett. 2020,

22, 6719-23.

(13) Sattler, L. E.; Otten, C. J.; Hilt, G. Alternating Current Electrolysis for the Electrocatalytic Syn-

thesis of Mixed Disulfide via Sulfur-Sulfur Bond Metathesis towards Dynamic Disulfide Libraries.

Chem. Eur. J. 2020, 26, 3129-3136.

(14) Bartlett, P. N.; Eastwick-Field, V. A reinvestigation of the electrochemistry of

[Ni(II)(bpy)3(ClO)4)2] in acetonitrile using rotating disc and rotating ring-disc electrodes. Electrochim.

Acta 1993, 38, 2515-2523.

(15) Tian, L.; Till, N. A.; Kudisch, B.; MacMillan, D. W. C.; Scholes, G. D. Transient Absorption

Spectroscopy Offers Mechanistic Insights for an Iridium/Nickel-Catalyzed C-O Coupling. J. Am. Chem.

Soc. 2020, 142, 4555-4559.

(16) Welin, E. R.; Le, C.; Arias-Rotondo, D. M.; McCusker, J. K.; MacMillan, D. W. Photosensi-

tized, energy transfer-mediated organometallic catalysis through electronically excited nickel(II). Sci-

ence 2017, 355, 380-385.

(17) Durandetti, M.; Devaud, M.; Perichon, J. Investigation of the reductive coupling of aryl halides

and/or ethylchloroacetate electrocatalyzed by the precursor NiX2(bpy) with X(-)=Cl-, Br- or MeSO3-

and bpy equals 2,2'-dipyridyl. New J. Chem. 1996, 20 (6), 659-667.

(18) Klein, A.; Budnikova, Y. H.; Sinyashin, O. G. Electron transfer in organonickel complexes of α-

diimines: Versatile redox catalysts for C–C or C–P coupling reactions – A review. J. Organomet. Chem.

2007, 692, 3156-3166.

(19) (a) Derien, S.; Dunach, E.; Perichon, J. From stoichiometry to catalysis: electroreductive cou-

pling of alkynes and carbon dioxide with nickel-bipyridine complexes. Magnesium ions as the key for

catalysis. J. Am. Chem. Soc. 1991, 113, 8447-8454. (b) Troupel, M.; Rollin, Y.; Sock, O.; Meyer, G.;

Perichon, J. Electrochemistry of Nickel-Complexes Associated with 2,2'-Bipyridine in the N-

Page 19: Coupling of Alternating Current to Transition-Metal

Methylpyrrolidone Solvent - Application to Activation of Carbon-Halogen Bonds. New J. Chem. 1986,

10, 593-599.

(20) Cannes, C.; Labbé, E.; Durandetti, M.; Devaud, M.; Nédélec, J. Y. Nickel-catalyzed electro-

chemical homocoupling of alkenyl halides: rates and mechanisms. J. Electroanal. Chem. 1996, 412, 85-

93.

(21) (a) Yakhvarov, D. G.; Tazeev, D. I.; Sinyashin, O. G.; Giambastiani, G.; Bianchini, C.; Segar-

ra, A. M.; Lönnecke, P.; Hey-Hawkins, E. Electrochemical synthesis of the σ-aryl complex

[NiBr(Mes)(bpy)] and its use as catalyst precursor for the oligomerization of ethylene (Mes=2,4,6-

trimethylphenyl, bpy=2,2′-bipyridine). Polyhedron 2006, 25, 1607-1612. (b) Budnikova, Y. H.; Peri-

chon, J.; Yakhvarov, D. G.; Kargin, Y. M.; Sinyashin, O. G. Highly reactive σ-organonickel complexes

in electrocatalytic processes. J. Organmet. Chem. 2001, 630, 185-192.

Page 20: Coupling of Alternating Current to Transition-Metal

This work demonstrates that periodic oxidation and reduction of a catalyst by alternating current enable

otherwise unfavorable catalytic cycles. Nickel catalyzed amination, etherification, and esterification

were universally enabled by alternating current with yields and selectivity strongly exciding these in the

experiments with direct current (DC).

Page 22: Coupling of Alternating Current to Transition-Metal

S1

Supporting Information for

Coupling of Alternating Current to Transition-Metal Catalysis: Examples of Nickel-Catalyzed Cross-Coupling

Evgeniy O. Bortnikov and Sergey N. Semenov*

*Correspondence to: [email protected]

Page 23: Coupling of Alternating Current to Transition-Metal

S2

Table of Content

Materials and Methods S3

Experimental set-up S3

GC rod electrode preparation and cleaning procedure S3

Electrochemical cell S4

Set-up preparation S5

Optimization of the reaction conditions for amination S8

Optimization of the reaction conditions for esterification S10

Optimization of the reaction conditions for etherification S12

Synthesis of compounds S14

General procedure for the amination experiments S14

Synthesis and characterization of the aromatic amines S15

General procedure for the esterificaton experiments S22

Synthesis and characterization of the aromatic esters S23

General procedure for the etherification experiments S28

Synthesis and characterization of the aromatic ethers S29

Characterization of the side products S32

Electrochemical studies S33

General information and procedure for the electrochemical studies S33

Measurements of the capacitance of an electrolyte double layer S33

Cyclic voltammetry studies S34

Open circuit potential (OCP) experiments S42

Mechanistic studies S44

Amination in a divided cell S44

DC-assisted amination and etherification with a Zn sacrificial anode S45

Kinetic experiments with ON/OFF cycles of AC S46

NMR spectra S49

References S118

Page 24: Coupling of Alternating Current to Transition-Metal

S3

Materials and Methods

All reactions were performed under an argon atmosphere according to the methods

indicated in the general procedures. Anhydrous solvents (Extra Dry over Molecular Sieve, AcroSeal®) - dimethylformamide (DMF), dimethylacetamide (DMA), and acetonitrile (MeCN) - were purchased from Acros Organics and stored under a positive pressure of argon. CDCl3 was purchased from Cambridge Isotope Laboratories. All other commercially available chemicals and reagents were used as received from Sigma Aldrich, Alfa Aesar, Acros Organics, and Fisher chemicals unless otherwise noted. Flash column chromatography was performed using Merck silica gel (60 mesh, particle size 0.043 – 0.063 mm).

Glassy carbon (GC) rods (100 mm x 6 mm diameter) were purchased from Alfa Aesar. Platinum wire (1 mm diameter) was purchased from Holland-Moran. Ltd. (Israel). A glassy carbon disk (1 mm diameter) working electrode, a platinum wire counter-electrode, and an Ag+/Ag reference electrode were purchased from CH Instruments, Inc.

NMR spectra were measured on a Bruker AVANCE III-300 spectrometer at 300 MHz for 1H and 282.4 MHz for 19F, on a Bruker AVANCE III-400 spectrometer at 400 MHz for 1H and 100.6 MHz for 13C{1H}, and on a Bruker AVANCE III HD-500 spectrometer at 500 MHz for 1H, 125.8 MHz for 13C{1H}, and 470.6 MHz for 19F. Chemical shifts for 1H and 13C are given in ppm relative to TMS and those for 19F relative to CFCl3.

1H and 13C spectra were calibrated using a residual solvent peak as an internal reference (1H NMR: δ = 7.26 ppm, 13C NMR: δ = 77.16 ppm). The following abbreviations were used to explain NMR peak multiplicities: s = singlet, d = doublet, t = triplet, q = quartet, p = pentet, m = multiplet, br = broad.

A B&K Precision 4053B dual channel function/arbitrary waveform generator served as a tunable source for the alternating current. The measurements and tracking of voltage drops over the electrochemical cell and the resistor were performed using a Global Specialties DSC-5300 50 MHz Digital Storage Oscilloscope. Direct current electrolysis, cyclic voltammetry (CV), and open circuit potential measurements were performed using a CH Instruments 600E Potentiostat/Galvanostat.

High-resolution mass spectra were recorded on a Waters Xevo G2-XS QTof mass

spectrometer (Manchester, UK) with an electrospray ionization (ESI) source.

The analytical chromatographic separation was performed on a Waters Acquity liquid chromatography system equipped with a 2998 PDA Detector and a Waters QDa mass detector (a mass range of 85–1250 m/z) with an electrospray ionization (ESI).

Experimental set-up

GC rod electrode preparation and cleaning procedure

A GC rod (100 mm x 6 mm diameter, Alfa Aesar) was cut in half into two equal parts (50 mm rods). A groove with a 1 mm width and 1 mm depth was made on the surface of each rod (ca. 4 mm from the end of the rod). For polishing purposes, a groove-side part of the electrode was wrapped with a piece of rubber tubing, and placed in the holder of a mechanical stirrer with tunable rotation speed (an electric screwdriver can be used instead) (Figure S1). The surface of the lower part of the electrode was then polished carefully using a convenient rotation speed and sandpaper with different grit sizes (starting from P200 and ending with P2000), achieving a

Page 25: Coupling of Alternating Current to Transition-Metal

S4

mirror-shiny surface on the side and the end part of the electrode. Thereafter, an alumina powder (1-0.05 μm) with the addition of small amounts of DI water was used to finish the polishing procedure (Nylon microcloth or Chemwipes can be used as substrates); then the electrode was rinsed with DI water intensively and placed in a drying oven (150-200°C) for several hours. Next, a piece of platinum wire (ca. 4 cm, 1 mm diameter) was wrapped once along the groove on the surface of the electrode; the remaining part of the wire was straightened and directed toward the top of the electrode.

Figure S1. Set-up for polishing the GC rod electrodes.

After every electrochemical experiment, the surface of the GC rod electrodes was cleaned

properly in order to achieve reproducible results. For this purpose, after the end of the experiment, the electrodes, together with the teflon cap, were extracted from the reaction mixture; the teflon cap and rubber O-rings were detached. The electrodes were then rinsed with DI water and acetone; the surface of the electrodes was wiped with a piece of tissue wetted with acetone.

Next, the surface of the electrodes was polished using a mechanical stirrer and 0.05 μm alumina powder (as described in the “Electrodes preparation procedure” section). Finally, the electrodes were rinsed with DI water and placed in a drying oven (150-200°C) for several hours.

Electrochemical cell

The cell used in the electrochemical experiments consists of 4 parts: a glass cell; a teflon cap; GC rod electrodes; and two rubber O-rings (Figure S2). The glass cell was made from a vial of a corresponding size; it has the following dimensions: height 30 mm and internal diameter 18 mm. The teflon cap has 2 large (for the GC rod electrodes) and 2 small (for the Ar inlet and

Page 26: Coupling of Alternating Current to Transition-Metal

S5

outlet) holes; rubber O-rings help to regulate and maintain the desirable position of the electrodes in the cell.

Figure S2. Electrochemical cell in the disassembled (A) and assembled (B) states.

Set-up preparation

A picture of a typical experimental set-up is presented in Figure S3; here we provide a detailed description of its assembly. The electrochemical cell with a reaction mixture should first be prepared according to a procedure in the corresponding section. Then, one of the alligator clips of the function (electrode F1) generator is connected to one of the GC rod electrodes (the platinum wire part) (Figure S4). Another GC rod electrode is connected to a resistor (usually, 10 Ohm) with the use of an additional cable with alligator clips on its ends; the remaining contact of the function generator (electrode F2) is attached to another side of the resistor. One probe from the oscilloscope (probe O2) is attached to one side of the resistor and the ground contact (ground) of the oscilloscope is attached to another side of the resistor, allowing the monitoring of the voltage drop on the resistor. The second probe of the oscilloscope (probe O1) is connected to the cell, allowing continuous monitoring of the drop of voltage over the “cell + resistor” part of the circuit.

Page 27: Coupling of Alternating Current to Transition-Metal

S6

Figure S3. Photo of a typical experimental set-up for AC-assisted catalysis.

Page 28: Coupling of Alternating Current to Transition-Metal

S7

Figure S4. Connecting the electrochemical cell to the function generator and the oscilloscope.

Before every experiment, the waveform, frequency, and amplitude of the alternating current

have to be adjusted. In all our experiments, we used a Sine waveform and a peak voltage of 3 V over the cell. The voltage over the cell can be calculated as the difference between voltage drops over the “cell+resistor” circuit part and the resistor. The current at any time can be calculated from the value of the voltage drop over the resistor according to Ohm’s law (the voltage drop over the resistor divided by its resistance).

Importantly, the resistance of the cell will change during the experiment (usually, it will increase). This leads to a change in the voltage drop over the cell – thus, the amplitude of the wave might need a small readjustment to keep it constant.

Page 29: Coupling of Alternating Current to Transition-Metal

S8

Optimization of the reaction conditions for amination

Table S1. Reaction conditions’ optimization table for amination.

Deviations from the conditions above 1H NMR yield

of 2

None 80%

Ligand 15 mol% Bipy 79%

15 mol% BPhen 44%

Amount of ligand 10 mol% di-tBuBipy 78%

20 mol% di-tBuBipy 77%

Sup. electrolyte

0.1M Bu4NBr 53%

0.2M LiCl 16%

0.1M Bu4NPF6 28%

Solvent DMF 63%

MeCN 28%

Frequency

0.1 Hz 55%

0.5 Hz 82%

2 Hz 87%

10 Hz 78%

25 Hz 48%

Time 4 hours (2 Hz) 99%

DC

2.8V (4 hours) 47%

2 mA (4 hours) 67%

4 mA (4 hours) 87%

6 mA (4 hours) Traces

No current (4 hours) -

DME – dimethoxyethane; Bipy - 2,2′-Dipyridyl; di-tBuBipy - 4,4′-Di-tert-butyl-2,2′-dipyridyl; BPhen - 4,7-Diphenyl-1,10-phenanthroline

Page 30: Coupling of Alternating Current to Transition-Metal

S9

General procedure for the optimization of amination experiments

To an Ar-flushed mixture containing NiBr2·DME (6.2 mg, 0.02 mmol, and 0.1 equiv.), Ligand (see Table S1) and electrolyte (see Table S1) in a screw-capped vial 4 ml of a solvent (see Table S1) were added. The solution was stirred until the reagents were completely dissolved, and then morpholine (52 μL, 0.6 mmol, 3 equiv.) was added. Next, the mixture was transferred to a glass cell as fully as possible with the addition of bromobenzene (21 μL , 0.2 mmol, 1 equiv.). The teflon cap equipped with GC rod electrodes was placed on top of the cell; then the cap was sealed tightly with parafilm; the argon inlet and outlet (optional) were inserted into the corresponding holes in the cap. Thereafter, an argon inlet was immersed into the solution, and argon was bubbled through the solution for at least 5 minutes with moderate stirring. Then, the Ar inlet was set above the solution, and the position of the electrodes was adjusted to be as deep in the solution as possible (ca. 1 cm). The electric circuit was assembled as described in the “Set-up preparation” section. The electrolysis was conducted for 2 hours with the following parameters of alternating current: sine waveform, frequency (see Table S1), 3V peak voltage, and a stirring rate of 1400 rpm.

After the electrolysis, 0.5 ml of the solution was transferred to a screw-capped vial, mixed with a small amount of Na2H2EDTA (EDTA – Ethylenediaminetetraacetate) water solution (for Ni complexation), diluted with an aqueous solution of K2CO3 (0.1 M), and extracted with a Pentane/Et2O (1/1 ratio) mixture. The extract was dried over Na2SO4, the solvent was evaporated carefully under vacuum (since bromobenzene is quite volatile, soft conditions are required). Then, the residue was used to determine the NMR yield of 4-phenylmorpholine, 2. To determine the NMR yield we calculated the ratio of the integrals of the signals of equal numbers of protons from the product and all aromatic species in the mixture. In all further experiments, NMR yields were determined following the same procedure.

General procedure for the control experiments of amination with DC assistance

To an Ar-flushed mixture containing NiBr2·DME (6.2 mg, 0.02 mmol, and 0.1 equiv.), di-tBuBipy (8.1 mg, 0.03 mmol, and 0.15 equiv.) and LiBr (69.5 mg) in a screw-capped vial 4 ml of DMA (unless otherwise noted) were added. The solution was stirred until the reagents were completely dissolved, and then morpholine (52 μL, 0.6 mmol, 3 equiv.) was added. Next, the mixture was transferred to a glass cell as fully as possible with the addition of bromobenzene (21 μL , 0.2 mmol, 1 equiv.). The teflon cap equipped with GC rod electrodes was placed on top of the cell; then the cap was sealed tightly with parafilm; the argon inlet and outlet (optional) were inserted into the corresponding holes in the cap. Thereafter, an argon inlet was immersed into the solution, and argon was bubbled through the solution for at least 5 minutes with moderate stirring. Then, the Ar inlet was set above the solution, and the position of the electrodes was adjusted to be as deep in the solution as possible (ca. 1 cm). The leads of the potentiostat were connected to the electrodes in the way that one of the electrodes was a working electrode, and the second one – a counter and a reference electrode. The electrolysis was conducted either in potentiostatic or galvanostatic conditions for 4 hours with the parameters, specified in Table S1 (stirring rate - 1400 rpm). The following procedure is identical to described above.

Page 31: Coupling of Alternating Current to Transition-Metal

S10

Optimization of the reaction conditions for esterification

Table S2. Reaction conditions’ optimization table for esterification.

Deviations from the initial conditions 19

F NMR yield

of 4

19F NMR yield

of Ar2

None 74% 7%

Base

No base - -

1.5 eq. Quinuclidine 20% -

1.5 eq. DIPEA - -

1.5 eq. DBU 56% 2%

1.5 eq. tBuOK 47% 8%

3 eq. K2CO3 71% 4%

10 eq. Na2CO3 67% 3%

10eq. K3PO4 68% 8%

Ligand 10 mol% di-tBuBipy 41% 13%

10 mol% BPhen 24% 6%

Amount of ligand 5 mol% Bipy 73% 8%

15 mol% Bipy 57% 2%

Sup. electrolyte 0.2M LiBr 52% 16%

Solvent DMA 86% 4%

MeCN 35% 3%

Frequency 0.5 Hz (DMA) 62% 7%

5 Hz (DMA) 34% 5%

Temperature 60°C (DMA) 87% 10%

DC

2.8V (DMA, 60°C) 13% 22%

2 mA (DMA, 60°C) 26% 53%

4 mA (DMA, 60°C) 12% 72%

6 mA (DMA, 60°C) 11% 63%

No current (DMA, 60°C) - -

Ar2 - 4,4'-bis(trifluoromethyl)-1,1'-biphenyl; DIPEA - N,N-Diisopropylethylamine; DBU - 1,8-Diazabicyclo[5.4.0]undec-7-ene.

Page 32: Coupling of Alternating Current to Transition-Metal

S11

General procedure for the optimization of esterificaton experiments

To an Ar-flushed mixture containing NiBr2·DME (6.2 mg, 0.02 mmol, and 0.1 equiv.), Ligand (see Table S2), Boc-L-proline (65 mg, 0.3 mmol, 1.5 equiv.), and electrolyte (see Table S2) in a screw-capped vial 4 ml of solvent (see Table S2) were added. The solution was stirred until the reagents were completely dissolved; thereafter, the mixture was transferred to a glass cell as fully as possible with the addition of 4-bromobenzotrifluoride (28 μL, 0.2 mmol, 1 equiv.) and base (see Table S2). The teflon cap equipped with GC rod electrodes was placed on top of the cell; then the cap was sealed tightly with parafilm; the argon inlet and outlet (optional) were inserted into the corresponding holes in the cap. Next, the argon inlet was immersed into the solution, and argon was bubbled through the solution for at least 5 minutes with moderate stirring. Then, the Ar inlet was set above the solution, and the position of the electrodes was adjusted to be as deep in the solution as possible (ca. 1 cm). The electric circuit was assembled as described in the “Set-up preparation” section. The electrolysis was conducted for 4 hours with the following parameters of alternating current: sine waveform, frequency (see Table S2), 3V peak voltage, and a stirring rate of 200 rpm.

After the electrolysis, 0.5 ml of the solution was transferred to a screw-capped vial, mixed with a small amount of Na2H2EDTA water solution (for Ni complexation), diluted with DI water, and extracted with a Pentane/Et2O (1/1 ratio) mixture. The extract was dried over Na2SO4, and the solvent was evaporated carefully under vacuum. Then, the residue was used to determine the NMR yield of 1-(tert-butyl) 2-(4-(trifluoromethyl)phenyl) (S)-pyrrolidine-1,2-dicarboxylate, 4.

General procedure for the control experiments of esterification with DC assistance

To an Ar-flushed mixture containing NiBr2·DME (6.2 mg, 0.02 mmol, and 0.1 equiv.), Bipy (3.1 mg, 0.02 mmol, and 0.1 equiv.), carboxylic acid (0.3 mmol, 1.5 equiv.) (unless otherwise noted) and Bu4NPF6 (155 mg) in a screw-capped vial 4 ml of DMA were added. The solution was stirred until the reagents were completely dissolved; thereafter, the mixture was transferred to a glass cell as fully as possible with the addition of aryl bromide (0.2 mmol, 1 equiv.) and K2CO3 (276 mg, 10 equiv.) (unless otherwise noted). The teflon cap equipped with GC rod electrodes was placed on top of the cell; then the cap was sealed tightly with parafilm; the argon inlet and outlet (optional) were inserted into the corresponding holes in the cap. The temperature of the hotplate was set to 60°C. Next, the argon inlet was immersed into the solution, and argon was bubbled through the solution for at least 5 minutes with moderate stirring. Then, the Ar inlet was set above the solution, and the position of the electrodes was adjusted to be as deep in the solution as possible (ca. 1 cm).The leads of the potentiostat were connected to the electrodes in the way that one of the electrodes was a working electrode, and the second one – a counter and a reference electrode. The electrolysis was conducted either in potentiostatic or galvanostatic conditions for 4 hours with the parameters, specified in Table S2 (stirring rate - 200 rpm). The following procedure is identical to described above.

Page 33: Coupling of Alternating Current to Transition-Metal

S12

Optimization of the reaction conditions for etherification

Table S3. Reaction conditions’ optimization table for etherification.

Deviations from the initial conditions 19

F NMR

yield of 5

19F NMR

yield of Ar2

None 42% 4%

Base

No base - 11%

1 eq. DABCO 9% 15%

1 eq. DBU 5% -

1 eq. LiHMDS - Traces

3 eq. K2CO3 - 12%

1 eq. Et3N Traces 30%

1 eq. DIPEA - 50%

1 eq. tBuOK Traces -

Ligand 10 mol% di-tBuBipy 22% 4%

Amount of ligand 5 mol% Bipy 37% 5%

20 mol% Bipy 14% 4%

Supporting electrolyte

0.1M Bu4NBr 34% 2%

0.1M Bu4NPF6 41% 4%

Solvent DMA 25% 2%

MeCN 21% -

Temperature 60°C 21% 21%

Amount of base 2 eq. Quinuclidine 53% 3%

Frequency

0.5 Hz (2 eq. Quin.) 35% 15%

5 Hz (2 eq. Quin.) 71% 1%

10 Hz (2 eq. Quin.) 62% -

25 Hz (2 eq. Quin.) 40% -

Time 8 hours

(2 eq. Quin., 10 Hz) 87% -

DC

2.8V (2 eq. Quin., 4 hours)

26% 69%

2 mA (2 eq. Quin., 4 hours) 38% 23%

Page 34: Coupling of Alternating Current to Transition-Metal

S13

4 mA (2 eq. Quin., 4 hours) 44% 44%

6 mA (2 eq. Quin., 4 hours) 11% 5%

No current (2 eq. Quin., 8 hours) - -

DABCO - 1,4-diazabicyclo[2.2.2]octane; LiHMDS - lithium bis(trimethylsilyl)amide.

General procedure for the optimization of etherification experiments

To an Ar-flushed mixture containing NiBr2·DME (6.2 mg, 0.02 mmol, and 0.1 equiv.), Ligand, and electrolyte (see Table S3) in a screw-capped vial 4 ml of solvent (see Table S3) were added. The solution was stirred until the reagents were completely dissolved, and then base (see Table S3) and n-hexanol (75 μL, 0.6 mmol, 3 equiv.) were added. Next, the mixture was transferred to a glass cell as fully as possible with the addition of 4-bromobenzotrifluoride (28 μL, 0.2 mmol, 1 equiv.). The teflon cap equipped with GC rod electrodes was placed on top of the cell; then the cap was sealed tightly with parafilm; the argon inlet and outlet (optional) were inserted into the corresponding holes in the cap. Thereafter, an argon inlet was immersed into the solution, and argon was bubbled through the solution for at least 5 minutes with moderate stirring. Then, the Ar inlet was set above the solution, and the position of the electrodes was adjusted to be as deep in the solution as possible (ca. 1 cm). The electric circuit was assembled as described in the “Set-up preparation” section. The electrolysis was conducted for 4 hours with the following parameters of alternating current: a sine waveform, frequency (see Table S3), 3V peak voltage, and a stirring rate of 200 rpm.

After the electrolysis, 0.5 ml of the solution was transferred to a screw-capped vial, mixed with a small amount of Na2H2EDTA water solution (for Ni complexation), diluted with DI water, and extracted with a Pentane/Et2O (1/1 ratio) mixture. The extract was dried over Na2SO4; the solvent was evaporated carefully under vacuum. Then, the residue was used to determine the NMR yield of 1-(hexyloxy)-4-(trifluoromethyl)benzene, 5.

General procedure for the control experiments of etherification with DC assistance

To an Ar-flushed mixture containing NiBr2·DME (6.2 mg, 0.02 mmol, and 0.1 equiv.), Bipy (3.1 mg, 0.02 mmol, and 0.1 equiv.) and LiBr (69.5 mg) in a screw-capped vial 4 ml of DMF were added. The solution was stirred until the reagents were completely dissolved, and then quinuclidine (44.5 mg, 0.4 mmol, and 2 equiv.) and alcohol (0.6 mmol, 3 equiv. - unless otherwise noted) were added. Next, the mixture was transferred to a glass cell as full as possible with the addition of aryl bromide (0.2 mmol, 1 equiv.). The teflon cap equipped with GC rod electrodes was placed on top of the cell; then the cap was sealed tightly with parafilm; the argon inlet and outlet (optional) were inserted into the corresponding holes in the cap. Thereafter, an argon inlet was immersed into the solution, and argon was bubbled through the solution for at least 5 minutes with moderate stirring. Then, the Ar inlet was set above the solution, and the position of the electrodes was adjusted to be as deep in the solution as possible (ca. 1 cm). The leads of the potentiostat were connected to the electrodes in the way that one of the electrodes was a working electrode, and the second one – a counter and a reference electrode. The electrolysis was conducted either in potentiostatic or galvanostatic conditions for 4 hours with the parameters, specified in Table S2 (stirring rate - 200 rpm). The following procedure is identical to described above.

Page 35: Coupling of Alternating Current to Transition-Metal

S14

Synthesis of compounds

General procedure for the amination experiments

(A) With amines as substrates

To an Ar-flushed mixture containing NiBr2·DME (6.2 mg, 0.02 mmol, and 0.1 equiv.), di-tBuBipy (8.1 mg, 0.03 mmol, and 0.15 equiv.) and LiBr (69.5 mg) in a screw-capped vial 4 ml of DMA (unless otherwise noted) were added. The solution was stirred until the reagents were completely dissolved, and then amine (0.6 mmol, 3 equiv.) was added. Next, the mixture was transferred to a glass cell as fully as possible with the addition of aryl bromide (0.2 mmol, 1 equiv.). The teflon cap equipped with GC rod electrodes was placed on top of the cell; then the cap was sealed tightly with parafilm; the argon inlet and outlet (optional) were inserted into the corresponding holes in the cap. Thereafter, an argon inlet was immersed into the solution, and argon was bubbled through the solution for at least 5 minutes with moderate stirring. Then, the Ar inlet was set above the solution, and the position of the electrodes was adjusted to be as deep in the solution as possible (ca. 1 cm). The electric circuit was assembled as described in the “Set-up preparation” section. The electrolysis was conducted for 4 hours with the following parameters of alternating current: sine waveform, 2 Hz frequency, 3V peak voltage, and a stirring rate of 1400 rpm.

After the electrolysis, 0.5 ml of the solution was transferred to a screw-capped vial, mixed with a small amount of Na2H2EDTA (EDTA – Ethylenediaminetetraacetate) water solution (for Ni complexation), diluted with an aqueous solution of K2CO3 (0.1 M), and extracted with a Pentane/Et2O (1/1 ratio) mixture. The extract was dried over Na2SO4, the solvent was evaporated carefully under vacuum (since some of the aryl bromides and products are quite volatile, soft conditions are required). Then, the residue was used to determine the NMR yield of the experiment. The remaining 3.5 ml of reaction mixture was diluted with an aqueous solution of K2CO3 (0.1 M) and extracted 4 times with a Pentane/Et2O (1/1 ratio) mixture. The combined organic fractions were washed with a small amount of 0.1M NaCl water solution and dried over Na2SO4. Next, the solution was dried carefully under vacuum, the residue was dissolved in a minimum amount of hexane (with the addition of dichloromethane if needed), and subjected to column chromatography on SiO2 to afford the desired product. The isolated yield in each case was calculated while considering the use of only 3.5 ml of solution for the isolation.

(B) With amine hydrochlorides as substrates

To an Ar-flushed mixture containing NiBr2·DME (6.2 mg, 0.02 mmol, 0.1 equiv.), di-tBuBipy (8.1 mg, 0.03 mmol, 0.15 equiv.), amine hydrochloride (0.6 mmol, 3 equiv.), and LiBr (69.5mg) in a screw-capped vial, 4 ml of DMA (unless otherwise noted) were added. The solution was stirred until the reagents were completely dissolved; the mixture was transferred to a glass cell as full as possible, with the addition of aryl bromide (0.2 mmol, 1 equiv.) and K2CO3

(138 mg, 1 mmol, 5 equiv.). The following procedure is identical to the one described above.

Page 36: Coupling of Alternating Current to Transition-Metal

S15

Synthesis and characterization of the aromatic amines

4-(4-(trifluoromethyl)phenyl)morpholine (Figure 3, 7)

Prepared according to the procedure described in “General procedure for amination

experiments (A)” section with morpholine (52.5 μl, 0.6 mmol, 3 equiv.) as amine and 4-bromobenzotrifluoride (28.0 μl, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane-dichloromethane mixture with gradient (from pure hexane to 1:1 ratio, respectively) was used as an eluent. NMR yield – 98%, isolated yield – 75% (30.3 mg, white solid). Control DC-assisted experiment (potentiostatic conditions – 2.8 V) demonstrated 66% NMR yield (9% of the diaryl product was also observed). The spectra data matched with values reported in the literature.1

Rf (dichloromethane) – 0.59. 1H NMR (500 Hz, CDCl3): δ 7.50 (d, J = 8.0 Hz, 2H), 6.92 (d, J = 8.1 Hz, 2H), 3.87 (m,

4H), 3.24 (m, 4H).

13C NMR (125.8 Hz, CDCl3): δ 153.5, 126.6 (q, J = 3.8 Hz), 124.8 (q, J = 269.9 Hz),

121.1 (q, J = 33.4 Hz), 114.5, 66.8, 48.3.

19F NMR (282.4 Hz, CDCl3): δ –62.39.

N-cyclohexyl-4-(trifluoromethyl)aniline (Figure 3, 8)

Prepared according to the procedure described in “General procedure for amination

experiments (A)” section with cyclohexylamine (68.6 μl, 0.6 mmol, 3 equiv.) as amine and 4-bromobenzotrifluoride (28.0 μl, 0.2 mmol, 1 equiv.) as aryl bromide in DMF. Hexane was used as an eluent. NMR yield – 91%, isolated yield – 65% (27.7 mg, colorless liquid). The spectra data matched with values reported in the literature.1

Rf (Hexane) – 0.17.

Page 37: Coupling of Alternating Current to Transition-Metal

S16

1H NMR (500 Hz, CDCl3): δ 7.37 (d, J = 8.5 Hz, 2H), 6.57 (d, J = 8.7 Hz, 2H), 3.88 (br, s,

1H), 3.29 (t, J = 9.6, 1H), 2.05 (dd, J = 12.8, 3.2 Hz, 2H), 1.78 (dt, J = 13.6, 3.9 Hz, 2H), 1.69 – 1.65 (m, 1H), 1.43 – 1.34 (m, 2H), 1.28 – 1.14 (m, 3H).

13C NMR (100.6 Hz, CDCl3): δ 149.9, 126.6 (q, J = 3.9 Hz), 125.2 (q, J = 270.1 Hz).

118.2 (q, J = 32.6 Hz), 112.1, 51.5, 33.3, 25.9, 25.0.

19F NMR (470.6 Hz, CDCl3): δ –61.88.

N-hexyl-4-(trifluoromethyl)aniline (Figure 3, 9)

Prepared according to the procedure described in “General procedure for amination

experiments (A)” section with 1-hexylamine (79.3 μl, 0.6 mmol, 3 equiv.) as amine and 4-bromobenzotrifluoride (28.0 μl, 0.2 mmol, 1 equiv.) as aryl bromide in DMF. Hexane was used as an eluent. NMR yield – 82%, isolated yield – 75% (32.2 mg, colorless liquid). The spectra data matched with values reported in the literature.2

Rf (Hexane) – 0.17. 1H NMR (500 Hz, CDCl3): δ 7.39 (d, J = 8.5 Hz, 2H), 6.58 (d, J = 8.7 Hz, 2H), 3.94 (br, s,

1H), 3.14 (m, 2H), 1.63 (p, J = 7.4, 2H), 1.43–1.30 (m, 6H), 0.92–0.89 (m, 3H). 13

C NMR (125.8 Hz, CDCl3): δ 151.0, 126.7 (q, J = 3.8 Hz), 125.2 (q, J = 269.9 Hz), 118.5 (q, J = 32.4 Hz), 111.8, 43.7, 31.7, 29.4, 26.9, 22.8, 14.2.

19F NMR (282.4 Hz, CDCl3): δ –61.90.

N,N-dimethyl-4-(trifluoromethyl)aniline (Figure 3, 10)

Page 38: Coupling of Alternating Current to Transition-Metal

S17

Prepared according to the procedure described in “General procedure for amination

experiments (B)” section with dimethylamine hydrochloride (48.9 mg, 0.6 mmol, 3 equiv.) as amine hydrochloride and 4-bromobenzotrifluoride (28.0 μl, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane was used as an eluent. NMR yield – 86%, isolated yield – 71% (23.5 mg, white solid). The spectra data matched with values reported in the literature.3

Rf (Hexane) – 0.24.

1H NMR (500 Hz, CDCl3): δ 7.46 (d, J = 8.8 Hz, 2H), 6.70 (d, J = 8.8 Hz, 2H), 3.01 (s,

6H).

13C NMR (100.6 Hz, CDCl3): δ 152.4, 126.5 (q, J = 3.9 Hz), 125.3 (q, J = 270.1 Hz), 117.7

(q, J = 32.6 Hz), 111.3, 40.3.

19F NMR (282.4 Hz, CDCl3): δ –61.81.

Methyl (4-(trifluoromethyl)phenyl)glycinate (Figure 3, 11)

Prepared according to the procedure described in “General procedure for amination

experiments (B)” section with glycine methyl ester hydrochloride (75.3 mg, 0.6 mmol, 3 equiv.) as amine hydrochloride and 4-bromobenzotrifluoride (28.0 μl, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane-dichloromethane mixture with gradient (from pure hexane to 1:1 ratio, respectively) was used as an eluent. NMR yield – 71%, isolated yield – 35% (14.3 mg, white solid). The spectra data matched with values reported in the literature.4

Rf (dichloromethane) – 0.61. 1H NMR (500 Hz, CDCl3): δ 7.43 (d, J = 8.4 Hz, 2H), 6.61 (d, J = 8.4 Hz, 2H), 4.61 (s,

1H), 3.95 (d, J = 5.2 Hz, 2H), 3.81 (s, 3H).

13C NMR (100.6 Hz, CDCl3): δ 171.1, 149.5, 126.9 (q, J = 3.9 Hz), 125.0 (q, J = 270.5 Hz),

120.0 (q, J = 32.8 Hz), 112.3, 52.6, 45.2. 19

F NMR (270.6 Hz, CDCl3): δ –62.18.

4-phenylmorpholine (Figure 3, 2)

Page 39: Coupling of Alternating Current to Transition-Metal

S18

Prepared according to the procedure described in “General procedure for amination

experiments (A)” section with morpholine (52.5 μl, 0.6 mmol, 3 equiv.) as amine and bromobenzene (21.3 μl, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane-dichloromethane mixture with gradient (from pure hexane to 1:1 ratio, respectively) was used as an eluent. NMR yield – 99%, isolated yield – 74% (21.1 mg, white solid). The spectra data matched with values reported in the literature.5

Rf (dichloromethane) – 0.65.

1H NMR (500 Hz, CDCl3): δ 7.33 – 7.26 (m, 2H), 6.95 – 6.86 (m, 3H), 3.87 (t, J = 4.0 Hz,

4H), 3.17 (t, J = 4.0 Hz, 4H). 13

C NMR (100.6 Hz, CDCl3): δ 151.4, 129.3, 120.2, 115.9, 67.1, 49.5.

4-morpholinobenzonitrile (Figure 3, 12)

Prepared according to the procedure described in “General procedure for amination

experiments (A)” section with morpholine (52.5 μl, 0.6 mmol, 3 equiv.) as amine and 4-bromobenzonitrile (36.4 mg, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane-dichloromethane mixture with gradient (from pure hexane to 1:1 ratio, respectively) was used as an eluent. NMR yield – 79%, isolated yield – 58% (19.1 mg, white solid). The spectra data matched with values reported in the literature.5

Rf (dichloromethane) – 0.68.

1H NMR (400 Hz, CDCl3): δ 7.54 – 7.47 (m, 2H), 6.89 – 6.82 (m, 2H), 3.85 (t, J = 4.9 Hz,

4H), 3.27 (t, J = 4.9 Hz, 4H). 13

C NMR (100.6 Hz, CDCl3): δ 153.6, 133.6, 120.0, 114.2, 101.0, 66.6, 47.4.

Page 40: Coupling of Alternating Current to Transition-Metal

S19

Methyl 4-morpholinobenzoate (Figure 3, 13)

Prepared according to the procedure described in “General procedure for amination

experiments (A)” section with morpholine (52.5 μl, 0.6 mmol, 3 equiv.) as amine and methyl 4-bromobenzoate (43.0 mg, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane-dichloromethane mixture with gradient (from pure hexane to 1:1 ratio, respectively) was used as an eluent. NMR yield – 86%, isolated yield – 75% (29.0 mg, white solid). The spectra data matched with values reported in the literature.5

Rf (dichloromethane) – 0.4. 1H NMR (500 Hz, CDCl3): δ 7.94 (d, J = 8.3 Hz, 2H), 6.86 (d, J = 8.3 Hz, 2H), 3.90 – 3.82

(m, 7H), 3.28 (m, 4H). 13

C NMR (125.8 Hz, CDCl3): δ 167.2, 154.4, 131.3, 120.5, 113.6, 66.7, 51.8, 47.9.

4-(naphthalen-2-yl)morpholine (Figure 3, 14)

Prepared according to the procedure described in “General procedure for amination

experiments (A)” section with morpholine (52.5 μl, 0.6 mmol, 3 equiv.) as amine and 2-bromonaphtalene (41.4 mg, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane-dichloromethane mixture with gradient (from pure hexane to 1:1 ratio, respectively) was used as an eluent. NMR yield – 75%, isolated yield – 52% (19.4 mg, white solid). The spectra data matched with values reported in the literature.6

Rf (dichloromethane) – 0.52.

Page 41: Coupling of Alternating Current to Transition-Metal

S20

1H NMR (500 Hz, CDCl3): δ 7.77-7.68 (m, 3H), 7.45-7.40 (m, 1H), 7.34-7.39 (m, 1H), 7.27

(dd, J = 8.9 Hz, J = 2.5 Hz, 1H), 7.13 (d, J = 2.2 Hz, 1H), 3.92 (t, J = 4.8 Hz, 4H), 3.27 (t, J = 4.8 Hz, 4H).

13C NMR (125.8 Hz, CDCl3): δ 149.2, 134.7, 129.0, 128.8, 127.6, 126.9, 126.5, 123.7,

119.0, 110.2, 67.1, 49.9.

4-(o-tolyl)morpholine (Figure 3, 15)

Prepared according to the procedure described in “General procedure for amination

experiments (A)” section with morpholine (52.5 μl, 0.6 mmol, 3 equiv.) as amine and 2-bromotoluene (24.1 μl, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane-dichloromethane mixture with gradient (from pure hexane to 1:1 ratio, respectively) was used as an eluent. NMR yield – 82%, isolated yield – 61% (18.9 mg, colorless liquid). The spectra data matched with values reported in the literature.7

Rf (dichloromethane) – 0.57. 1H NMR (400 Hz, CDCl3): δ 7.23-7.17 (m, 2H), 7.07-6.98 (m, 2H), 3.87 (t, J = 4.5 Hz,

4H), 2.93 (t, J = 4.5 Hz, 4H), 2.34 (s, 3H).

13C NMR (100.6 Hz, CDCl3): δ 151.4, 132.7, 131.3, 126.8, 123.5, 119.1, 67.6, 52.4, 87.0.

4-(pyridin-3-yl)morpholine (Figure 3, 16)

Prepared according to the procedure described in “General procedure for amination

experiments (A)” section with morpholine (52.5 μl, 0.6 mmol, 3 equiv.) as amine and 3-bromopyridine (19.3 μl, 0.2 mmol, 1 equiv.) as aryl bromide. The whole mixture was extracted

Page 42: Coupling of Alternating Current to Transition-Metal

S21

with ethel acetate 4 times. Combined organic fractions were washed with small amount of 0.1M NaCl water solution and dried over Na2SO4. Next, the solution was dried carefully under vacuum, the residue dissolved in the minimum amount of dichloromethane and subjected to column chromatography on SiO2 (dichloromethane/diethyl ether mixture with gradient – from pure dichloromethane to pure diethyl ether) to furnish the product. Isolated yield – 65% (from the whole reaction mixture, 21.3 mg, yellowish liquid). The spectra data matched with values reported in the literature.5

Rf (diethyl ether) – 0.22.

1H NMR (500 Hz, CDCl3): δ 8.29 (s, 1H), 8.14-8.10 (m, 1H), 7.19 – 7.14 (m, 2H), 3.90-

3.83 (m, 4H), 3.22-3.15 (m, 4H).

13C NMR (125.8 Hz, CDCl3): δ 147.0, 141.2, 138.4, 123.7, 122.3, 66.8, 48.7.

4-(3-methoxyphenyl)morpholine (Figure 3, 17)

Prepared according to the procedure described in “General procedure for amination

experiments (A)” section with morpholine (52.5 μl, 0.6 mmol, 3 equiv.) as amine and 3-bromoanisole (25.3 μl, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane-dichloromethane mixture with gradient (from pure hexane to 1:1 ratio, respectively) was used as an eluent. NMR yield – 86%, isolated yield – 67% (22.7 mg, colorless liquid). The spectra data matched with values reported in the literature.5

Rf (dichloromethane) – 0.43.

1H NMR (500 Hz, CDCl3): δ 7.24 – 7.16 (m, 1H), 6.56 – 6.50 (m, 1H), 6.48 – 6.42 (m, 2H),

3.85 (t, J = 4.8 Hz, 4H), 3.80 (s, 3H), 3.15 (t, J = 4.8 Hz, 4H).

13C NMR (125.8 Hz, CDCl3): δ 160.8, 152.9, 130.0, 108.6, 104.9, 102.4, 67.0, 55.4, 49.4.

Page 43: Coupling of Alternating Current to Transition-Metal

S22

General procedure for the esterificaton experiments

To an Ar-flushed mixture containing NiBr2·DME (6.2 mg, 0.02 mmol, and 0.1 equiv.), Bipy (3.1 mg, 0.02 mmol, and 0.1 equiv.), carboxylic acid (0.3 mmol, 1.5 equiv.) (unless otherwise noted) and Bu4NPF6 (155 mg) in a screw-capped vial 4 ml of DMA were added. The solution was stirred until the reagents were completely dissolved; thereafter, the mixture was transferred to a glass cell as fully as possible with the addition of aryl bromide (0.2 mmol, 1 equiv.) and K2CO3 (276 mg, 10 equiv.) (unless otherwise noted). The teflon cap equipped with GC rod electrodes was placed on top of the cell; then the cap was sealed tightly with parafilm; the argon inlet and outlet (optional) were inserted into the corresponding holes in the cap. The temperature of the hotplate was set to 60°C. Next, the argon inlet was immersed into the solution, and argon was bubbled through the solution for at least 5 minutes with moderate stirring. Then, the Ar inlet was set above the solution, and the position of the electrodes was adjusted to be as deep in the solution as possible (ca. 1 cm). The electric circuit was assembled as described in the “Set-up preparation” section. The electrolysis was conducted for 4 hours (unless otherwise noted) with the following parameters of alternating current: sine waveform, 2 Hz frequency, 3V peak voltage, and a stirring rate of 200 rpm.

After the electrolysis, 0.5 ml of the solution was transferred to a screw-capped vial, mixed with a small amount of Na2H2EDTA water solution (for Ni complexation), diluted with DI water, and extracted with a Pentane/Et2O (1/1 ratio) mixture. The extract was dried over Na2SO4, and the solvent was evaporated carefully under vacuum. Then, the residue was used to determine the NMR yield of the experiment. The remaining 3.5 ml of the reaction mixture was carefully diluted with an aqueous solution of KH2PO4 (1 M), and extracted 4 times with a Pentane/Et2O (1/1 ratio) mixture. The combined organic fractions were washed with a small amount of 0.1M NaCl water solution and dried over Na2SO4. Next, the solution was dried carefully under vacuum, the residue was dissolved in a minimum amount of hexane (with the addition of dichloromethane if needed), and subjected to column chromatography on SiO2 to afford the desired product. The isolated yield in each case was calculated while considering the use of only 3.5 ml of solution for the isolation.

Page 44: Coupling of Alternating Current to Transition-Metal

S23

Synthesis and characterization of the aromatic esters

4-(trifluoromethyl)phenyl benzoate (Figure 3, 18)

Prepared according to the procedure described in “General procedure for esterification

experiments” section with benzoic acid (36.6 mg, 0.3 mmol, 1.5 equiv.) as carboxylic acid and 4-bromobenzotrifluoride (28.0 μl, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane-dichloromethane mixture with gradient (from pure hexane to 10:1 ratio, respectively) was used as an eluent. NMR yield – 81%, isolated yield – 75% (34.9 mg, white solid). The spectra data matched with values reported in the literature.8

Rf (Hexane) – 0.11. 1H NMR (500 Hz, CDCl3): δ 8.25-8.18 (m, J = 2H), 7.72 (d, J = 7.3 Hz,

2H), 7.69-7.64 (m, 1H), 7.57-7.51 (m, 2H), 7.37 (d, J = 7.2 Hz, 2H). 13

C NMR (125.8 Hz, CDCl3): δ 164.8, 153.6 (q, J = 1.4 Hz) 134.1, 130.4, 129.1, 128.9, 128.3 (q, J = 33.4 Hz), 127.0 (q, J = 3.8 Hz), 124.0 (q, J = 271.8 Hz), 122.4.

19F NMR (282.4 Hz, CDCl3): δ -63.18.

5-(tert-butyl) 1-(4-(trifluoromethyl)phenyl) (tert-butoxycarbonyl)-L-glutamate (Figure 3, 19)

Prepared according to the procedure described in “General procedure for esterification

experiments” section with Boc-L-glutamic acid 5-tert-butyl ester (61.0 mg, 0.3 mmol, 1.5 equiv.) as carboxylic acid and 4-bromobenzotrifluoride (28.0 μl, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane-dichloromethane mixture with gradient (from pure hexane to 1:1 ratio, respectively) was used as an eluent. NMR yield – 82%, isolated yield – 66% (51.7 mg, white solid).

Page 45: Coupling of Alternating Current to Transition-Metal

S24

Rf (dichloromethane) – 0.45.

1H NMR (500 Hz, CDCl3): δ 7.66 (d, J = 8.5 Hz, 2H), 7.25 (d, J = 8.4 Hz, 2H), 5.20 (d, J =

7.6 Hz, 1H), 4.53 (m,1H), 2.50-2.37 (m, 2H), 2.34-2.25 (m, 1H), 2.15-2.05 (m, 1H), 1.46 (m, 18H).

13C NMR (125.8 Hz, CDCl3): δ 172.1, 170.9, 155.6, 153.1, 128.6 (q, J = 33.4 Hz), 127.0 (q,

J = 3.8 Hz), 123.9 (q, J = 272.8 Hz), 122.1, 81.3, 80.5, 53.6, 31.7, 28.4, 28.2, 27.3.

19F NMR (282.4 Hz, CDCl3): δ -63.27.

HRMS (ESI): calculated for C21H28NO6F3Na ([M+Na]+) 470.1766, found 470.1769.

Methyl (4-(trifluoromethyl)phenyl) succinate (Figure 3, 20)

Prepared according to the procedure described in “General procedure for esterification

experiments” section with succinic acid methyl ester (39.6 mg, 0.3 mmol, 1.5 equiv.) as carboxylic acid and 4-bromobenzotrifluoride (28.0 μl, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane-dichloromethane mixture with gradient (from pure hexane to 1:1 ratio, respectively) was used as an eluent. NMR yield – 84%, isolated yield – 70% (33.8 mg, colorless liquid).

Rf (dichloromethane) – 0.67. 1H NMR (500 Hz, CDCl3): δ 7.65 (d, J = 7.8 Hz, 2H), 7.23 (d, J = 8.0 Hz, 2H), 3.73 (s,

3H), 2.93-2.86 (m, 2H), 2.78-2.75 (m, 2H).

13C NMR (125.8 Hz, CDCl3): δ 172.6, 170.6, 153.2, 128.3 (q, J = 33.4 Hz), 126.9 (q, J =

3.8 Hz), 124.0 (q, J = 271.8 Hz), 122.16, 52.17, 29.4, 28.9.

19F NMR (470.6 Hz, CDCl3): δ -63.26.

HRMS (ESI): calculated for C12H11O4F3Na ([M+Na]+) 299.0507, found 299.0507.

4-(trifluoromethyl)phenyl acetate (Figure 3, 21)

Page 46: Coupling of Alternating Current to Transition-Metal

S25

Prepared according to the procedure described in “General procedure for esterification

experiments” section with sodium acetate suspension (82.0 mg, 1 mmol, 5 equiv.) instead of carboxylic acid and K2CO3 mixture, and 4-bromobenzotrifluoride (28.0 μl, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane-dichloromethane mixture with gradient (from pure hexane to 1/1 ratio, respectively) was used as an eluent. NMR yield – 87%, isolated yield – 48% (17.1 mg, colorless liquid). The spectra data matched with values reported in the literature.9

Rf (dichloromethane) – 0.65.

1H NMR (500 Hz, CDCl3): δ 7.65 (d, J = 8.5 Hz, 2H), 7.22 (d, J = 8.5 Hz, 2H), 2.33 (s,

3H). 13

C NMR (125.8 Hz, CDCl3): δ 169.0, 153.3, 128.3 (q, J = 32.4 Hz), 126.9 (q, J = 3.8 Hz), 124.0 (q, J = 271.8 Hz), 122.2, 21.2.

19F NMR (282.4 Hz, CDCl3): δ -63.25.

1-(tert-butyl) 2-(4-(trifluoromethyl)phenyl) (S)-pyrrolidine-1,2-dicarboxylate (Figure 3, 4)

Prepared according to the procedure described in “General procedure for esterification

experiments” section with Boc-L-proline (64.6 mg, 0.3 mmol, 1.5 equiv.) as carboxylic acid and 4-bromobenzotrifluoride (28.0 μl, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane-dichloromethane mixture with gradient (from pure hexane to 1/1 ratio, respectively) was used as an eluent. NMR yield – 87%, isolated yield – 69% (43.4 mg, colorless liquid). The spectra data matched with values reported in the literature.10

Rf (dichloromethane) – 0.37.

1H NMR (500 Hz, CDCl3): (rotameric mixture, resonances for minor rotamer are enclosed

in parenthesis) δ 7.67 (7.64) (d, J = 8.4 Hz, 2H), 7.25 (7.23) (d, J = 8.7 Hz, 2H), 4.53 (4.47) (dd, J = 8.7, 4.4 Hz, 1H), 3.67 – 3.40 (m, 2H), 2.45 – 2.29 (m, 1H), 2.22 – 2.10 (m, 1H), 2.09 – 1.90 (m, 2H), (1.48) 1.46 (s, 9H).

Page 47: Coupling of Alternating Current to Transition-Metal

S26

13C NMR (125.8 Hz, CDCl3): (rotameric mixture, resonances for minor rotamer are

enclosed in parenthesis) δ (171.4) 171.3, (154.6) 153.8, (153.5) 153.2, 128.4 (128.2) (d, J = 33.4 Hz), 127.0 (126.8) (q, J = 3.8 Hz), (123.8) 123.9 (d, J = 271.8 Hz), (122.2) 121.8, 80.5 (80.4), 59.3 (59.2), (46.8) 46.6, 31.2 (30.1), 28.6, (24.7) 23.9.

19F NMR (282.4 Hz, CDCl3): (rotameric mixture, resonances for minor rotamer are

enclosed in parenthesis) δ -63.22 (-63.27).

1-(tert-butyl) 2-(4-(methoxycarbonyl)phenyl) (S)-pyrrolidine-1,2-dicarboxylate (Figure 3, 22)

Prepared according to the procedure described in “General procedure for esterification

experiments” section with Boc-L-proline (64.6 mg, 0.3 mmol, 1.5 equiv.) as carboxylic acid and methyl 4-bromobenzoate (43.0 mg, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane-dichloromethane mixture with gradient (from pure hexane to pure dichloromethane) was used as an eluent. NMR yield – 89%, isolated yield – 79% (48.3 mg, white solid). The spectra data matched with values reported in the literature.10

Rf (dichloromethane) – 0.2. 1H NMR (500 Hz, CDCl3): (rotameric mixture, resonances for minor rotamer are enclosed

in parenthesis) δ 8.08 (8.05) (d, J = 8.4 Hz, 2H), 7.2 (7.18) (d, J = 8.5 Hz, 2H), (4.53) 4.46 (dd, J = 8.5, 4.3 Hz, 1H), 3.91 (3.90) (s, 3H), 3.69 – 3.40 (m, 2H), 2.47 – 2.28 (m, 1H), 2.23 – 2.11 (m,

1H), 2.10 – 1.90 (m, 2H), (1.48) 1.45 (s, 9H).

13C NMR (125.8 Hz, CDCl3): δ (rotameric mixture, resonances for minor rotamer are

enclosed in parenthesis) (171.3) 171.25, (166.5) 166.4, (154.6) 154.3, 153.8, 131.4 (131.3), 128.0 (127.8), (121.7) 121.3, 80.5 (80.3), 59.3 (59.2), 52.4 (52.3), (46.8), 46.6, 31.2 (30.1), 28.6, (24.7) 23.9.

1-(tert-butyl) 2-(4-cyanophenyl) (S)-pyrrolidine-1,2-dicarboxylate (Figure 3, 23)

Page 48: Coupling of Alternating Current to Transition-Metal

S27

Prepared according to the procedure described in “General procedure for esterification

experiments” section with Boc-L-proline (64.6 mg, 0.3 mmol, 1.5 equiv.) as carboxylic acid and 4-bromobenzonitrile (36.4 mg, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane-dichloromethane mixture with gradient (from pure hexane to pure dichloromethane) was used an eluent. NMR yield – 65%, isolated yield – 45% (24.9 mg, colorless liquid). The spectra data matched with values reported in the literature.10

Rf (dichloromethane) – 0.2.

1H NMR (500 Hz, CDCl3): (rotameric mixture, resonances for one of the rotamers are

enclosed in parenthesis) δ 7.70 (7.67) (d, J = 8.5 Hz, 2H), 7.26 (7.24) (d, J = 8.7 Hz, 2H), 4.51 (4.46) (dd, J = 8.5, 4.5 Hz, 1 H), 3.67 – 3.41 (m, 2H), 2.45 – 2.30 (m, 1H), 2.20 – 1.90 (m, 3H), 1.47 (1.44) (s, 9H).

13C NMR (125.8 Hz, CDCl3): (rotameric mixture, resonances for one of the rotamers are

enclosed in parenthesis) δ (171.1) 171.0, (154.6) 154.3, 154.0 (153.7) 133.9 (133.8), (122.8) 122.4, (118.4) 118.2, (110.1) 109.9, 80.6 (80.4), 59.3 (59.2), (46.8) 46.6, (31.2) 30.1, 28.5, (24.7) 23.9.

1-(tert-butyl) 2-(naphthalen-2-yl) (S)-pyrrolidine-1,2-dicarboxylate (Figure 3, 24)

Prepared according to the procedure described in “General procedure for esterification

experiments” section with Boc-L-proline (64.6 mg, 0.3 mmol, 1.5 equiv.) as carboxylic acid and 2-bromonaphtalene (41.4 mg, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane-dichloromethane mixture with gradient (from pure hexane to 1/1 ratio, respectively) was used as an eluent. NMR yield – 79%, isolated yield – 54% (32.3 mg, yellowish liquid). Control DC-assisted experiment (potentiostatic conditions – 2.8 V) demonstrated 4% NMR yield (9% of the diaryl product was also observed).

Rf (dichloromethane) – 0.36.

Page 49: Coupling of Alternating Current to Transition-Metal

S28

1H NMR (300 Hz, CDCl3): (rotameric mixture, resonances for minor rotamer are enclosed

in parenthesis) δ 7.92-7.75 (m, 3H), 7.62-7.4 (m, 3H), 7.32-7.18 (m, 1H), 4.59 (4.51) (dd, J = 8.4, 4.0 Hz, 1H), 3.75-3.40 (m, 2H), 2.51 – 1.9 (m, 4H), 1.50 (m, 9H).

13C NMR (100.6 Hz, CDCl3): (rotameric mixture, resonances for minor rotamer are

enclosed in parenthesis) (171.93) 171.89, (154.7) 153.9, (148.6) 148.4, 133.8, 131.6, 129.6 (129.5), 127.91 (127.87), 127.75, 126.8 (126.6), 125.9 (125.8), (121.2) 120.7, (118.6) 118.2, 80.4 (80.2), 59.4 (59.3), 46.8 (46.6), 31.2 (30.2), 28.6, (24.7) 23.9.

HRMS (ESI): calculated for C20H23NO4Na ([M+Na]+) 364.1525, found 364.1524.

General procedure for the etherification experiments

To an Ar-flushed mixture containing NiBr2·DME (6.2 mg, 0.02 mmol, and 0.1 equiv.), Bipy (3.1 mg, 0.02 mmol, and 0.1 equiv.) and LiBr (69.5 mg) in a screw-capped vial 4 ml of DMF were added. The solution was stirred until the reagents were completely dissolved, and then quinuclidine (44.5 mg, 0.4 mmol, and 2 equiv.) and alcohol (0.6 mmol, 3 equiv. - unless otherwise noted) were added. Next, the mixture was transferred to a glass cell as full as possible with the addition of aryl bromide (0.2 mmol, 1 equiv.). The teflon cap equipped with GC rod electrodes was placed on top of the cell; then the cap was sealed tightly with parafilm; the argon inlet and outlet (optional) were inserted into the corresponding holes in the cap. Thereafter, an argon inlet was immersed into the solution, and argon was bubbled through the solution for at least 5 minutes with moderate stirring. Then, the Ar inlet was set above the solution, and the position of the electrodes was adjusted to be as deep in the solution as possible (ca. 1 cm). The electric circuit was assembled as described in the “Set-up preparation” section. The electrolysis was conducted for 8 hours (unless otherwise noted) with the following parameters of alternating current: a sine waveform, 10 Hz frequency (unless otherwise noted), 3V peak voltage, and a stirring rate of 200 rpm.

After the electrolysis, 0.5 ml of the solution was transferred to a screw-capped vial, mixed with a small amount of Na2H2EDTA water solution (for Ni complexation), diluted with DI water, and extracted with a Pentane/Et2O (1/1 ratio) mixture. The extract was dried over Na2SO4; the solvent was evaporated carefully under vacuum. Then, the residue was used to determine the NMR yield of the experiment. The remaining 3.5 ml of reaction mixture was diluted with DI water and extracted 4 times with a Pentane/Et2O (1/1 ratio) mixture. The combined organic fractions were washed with a small amount of 0.1M NaCl water solution and dried over Na2SO4. Next, the solution was dried carefully under vacuum, the residue was dissolved in a minimum amount of hexane or pentane, and subjected to column chromatography on SiO2 to afford the desired product. The isolated yield in each case was calculated while considering the use of only 3.5 ml of solution for the isolation.

Page 50: Coupling of Alternating Current to Transition-Metal

S29

Synthesis and characterization of the aromatic ethers

1-(benzyloxy)-4-(trifluoromethyl)benzene (Figure 3, 25)

Prepared according to the procedure described in “General procedure for etherification

experiments” section with benzyl alcohol (62.1 μl, 0.6 mmol, 3 equiv.) as alcohol and 4-bromobenzotrifluoride (28.0 μl, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane was used as an eluent. NMR yield – 72%, isolated yield – 66% (29.1 mg, white solid). The spectra data matched with values reported in the literature.11

Rf (Hexane) – 0.2. 1H NMR (400 Hz, CDCl3): δ 7.55 (d, J = 8.6 Hz, 2H), 7.47-7.33 (m, 5H), 7.04 (d, J = 8.6

Hz, 2H), 5.12 (s, 2H). 13

C NMR (100.6 Hz, CDCl3): δ 161.3, 136.4, 128.9, 128.4, 127.6, 127.1 (q, J = 3.7 Hz), 124.7 (q, J = 271.1 Hz), 123.3 (q, J = 32.6 Hz), 115.0, 70.3.

19F NMR (282.4 Hz, CDCl3): δ -62.48.

1-(cyclohexyloxy)-4-(trifluoromethyl)benzene (Figure 3, 26)

Prepared according to the procedure described in “General procedure for etherification

experiments” section with cyclohexanol (126.8 μl, 1.2 mmol, 6 equiv.) as alcohol and 4-bromobenzotrifluoride (28.0 μl, 0.2 mmol, 1 equiv.) as aryl bromide. The electrolysis was conducted for 8 hours with the frequency of 5 Hz. Hexane was used as an eluent. NMR yield – 65%, isolated yield – 50% (21.4 mg, colorless liquid). The spectra data matched with values reported in the literature.12

Rf (Hexane) – 0.48.

Page 51: Coupling of Alternating Current to Transition-Metal

S30

1H NMR (500 Hz, CDCl3): δ 7.51 (d, J = 8.5 Hz, 2 H), 6.94 (d, J = 8.5 Hz, 2 H), 4.34-4.27

(m , 1H), 2.03-1.94 (m, 2H), 1.85-1.76 (m, 2H), 1.63-1.50 (m, 3H), 1.45 – 1.28 (m, 3H).

13C NMR (125.8 Hz, CDCl3): δ 160.5, 127.0 (q, J = 3.8 Hz), 124.7 (q, J = 271.8 Hz), 122.5

(q, J = 33.4 Hz), 115.7, 75.6, 31.7, 25.7, 23.8.

19F NMR (470.6 Hz, CDCl3): δ -62.42.

1-isopropoxy-4-(trifluoromethyl)benzene (Figure 3, 27)

Prepared according to the procedure described in “General procedure for etherification

experiments” section with 2-propanol (91.9 μl, 1.2 mmol, 6 equiv.) as alcohol and 4-bromobenzotrifluoride (28.0 μl, 0.2 mmol, 1 equiv.) as aryl bromide. The electrolysis was conducted for 16 hours with the frequency of 5 Hz. Pentane was used as an eluent. NMR yield – 73%, isolated yield – 62% (22.2 mg, colorless liquid). The spectra data matched with values reported in the literature.12

Rf (Hexane) – 0.45.

1H NMR (500 Hz, CDCl3): δ 7.52 (d, J = 8.7 Hz, 2H), 6.93 (d, J = 8.7 Hz, 2H), 4.61 (sept, J

= 6.1 Hz, 1H), 1.36 (d, J = 6.1 Hz, 6H). 13

C NMR (125.8 Hz, CDCl3): δ 160.6, 127.03 (q, J = 3.8 Hz), 124.7 (q, 270.8 Hz), 122.6 (q, 32.4 Hz), 115.6, 70.3, 22.0.

19F NMR (282.4 Hz, CDCl3): δ -62.44.

1-(hexyloxy)-4-(trifluoromethyl)benzene (Figure 3, 5)

Prepared according to the procedure described in “General procedure for etherification

experiments” section with 1-Hexanol (75.3 μl, 0.6 mmol, 3 equiv.) as alcohol and 4-bromobenzotrifluoride (28.0 μl, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane was used as an

Page 52: Coupling of Alternating Current to Transition-Metal

S31

eluent. NMR yield – 87%, isolated yield – 68% (29.3 mg, colorless liquid). The spectra data matched with values reported in the literature.13

Rf (Hexane) – 0.56.

1H NMR (500 Hz, CDCl3): δ 7.53 (d, J = 8.4 Hz, 2H), 6.95 (d, J = 8.4 Hz, 2H),

3.99 (t, J = 6.5 Hz, 2H), 1.80 (tt, J = 7.3, 6.7 Hz, 2H), 1.51-1.45 (m, 2H), 1.39−1.32 (m, 4H), 0.95-0.87 (m, 3H).

13C NMR (125.8 Hz, CDCl3): δ 161.7, 127.0 (q, J = 3.8 Hz), 124.7 (q, J = 270.8 Hz), 122.7

(q, J = 32.4 Hz), 114.6, 68.4, 31.7, 29.2, 25.8, 22.7, 14.2.

19F NMR (282.4 Hz, CDCl3): δ -62.42.

Methyl 4-(hexyloxy)benzoate (Figure 3, 28)

Prepared according to the procedure described in “General procedure for etherification

experiments” section with 1-Hexanol (75.3 μl, 0.6 mmol, 3 equiv.) as alcohol and methyl 4-bromobenzoate (43.0 mg, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane-dichloromethane mixture with gradient (from pure hexane to 1/1 ratio, respectively) was used as an eluent. NMR yield – 85%, isolated yield – 79% (32.7 mg, colorless liquid). The spectra data matched with values reported in the literature.13

Rf (dichloromethane) – 0.7. 1H NMR (500 Hz, CDCl3): δ 7.97 (d, J = 8.9 Hz, 2H), 6.90 (d, J = 8.9 Hz, 2H), 4.00 (t, J =

6.6 Hz, 2H), 3.88 (s, 3H), 1.79 (m, 2H), 1.51–1.41 (m, 2H), 1.38–1.30 (m, 4H), 0.95–0.86 (m, 3H).

13C NMR (125.8 Hz, CDCl3): δ 167.1, 163.1, 131.7, 122.5, 114.2, 68.4, 52.0, 31.7, 29.2,

25.8, 22.7, 14.2.

3-(hexyloxy)pyridine (Figure 3, 29)

Page 53: Coupling of Alternating Current to Transition-Metal

S32

Prepared according to the procedure described in “General procedure for etherification

experiments” section with 1-Hexanol (75.3 μl, 0.6 mmol, 3 equiv.) as alcohol and 3-bromopyridine (19.3 μl, 0.2 mmol, 1 equiv.) as aryl bromide. Hexane-dichloromethane mixture with gradient (from pure hexane to pure dichloromethane) was used as an eluent. NMR yield – 66%, isolated yield – 61% (19.1 mg, yellowish liquid). Control DC-assisted experiment (potentiostatic conditions – 2.8 V) demonstrated 29% NMR yield. The spectra data matched with values reported in the literature.14

Rf (dichloromethane) – 0.27. 1H NMR (500 Hz, CDCl3): δ 8.30 (d, J = 2.3, 1H), 8.20 (dd, J = 4.3, 1.5 Hz, 1H), 7.23 –

7.15 (m, 2H), 3.99 (t, J = 6.5 Hz, 2H), 1.83 – 1.75 (m, 2H), 1.51 – 1.42 (m, 2H), 1.40 – 1.29 (m, 4H), 0.91 (t, J = 7.0 Hz, 3H).

13C NMR (125.8 Hz, CDCl3): δ 155.4, 142.0, 138.2, 123.9, 121.2, 68.5, 31.7, 29.3, 25.8,

22.7, 14.2.

Characterization of the side products

4,4'-bis(trifluoromethyl)-1,1'-biphenyl

Separated as a minor product in the reactions of amination, etherification, and esterification with 4-bromobenzotrifluoride or as a major product in DC-assisted amination and etherification with a Zn sacrificial anode. Purified by column chromatography – hexane was used as an eluent. The spectra data matched with values reported in the literature.15

Rf (Hexane) – 0.53. 1H NMR (400 Hz, CDCl3): δ 7.74 (d, J = 8.6 Hz, 4H), 7.70 (d, J = 8.6 Hz, 4H).

Page 54: Coupling of Alternating Current to Transition-Metal

S33

13C NMR (100.6 Hz, CDCl3): δ 143.4, 130.4 (q, J = 32.6 Hz), 127.8, 126.1 (q, J = 3.9 Hz),

124.2 (q, J = 272.2 Hz).

19F NMR (282.4 Hz, CDCl3): δ -63.55.

Electrochemical studies

General information and procedure for the electrochemical studies

All electrochemical analytical experiments were carried out using a CH Instruments 600E Potentiostat/Galvanostat. All potentials are reported relative to a nonaqueous silver electrode (Ag+/Ag) that was used as a reference electrode (10mM AgNO3 in 0.1M TBAPF6 acetonitrile solution); the measured potential of the Fc+/Fc couple (in 0.1M TBAPF6, MeCN) was equal to 0.088 V. The same glass cell as described in the “Cell set-up” section was used; the teflon cap in this case had three large (for the working, counter, and reference electrodes) and two small holes (the Ar inlet and outlet).

In a typical experiment, 4 ml of the examined solution were transferred to the cell, the teflon cap equipped with electrodes was placed, and then argon was bubbled through the solution for at least 5 minutes with moderate stirring. Then, a mild flow of argon was adjusted above the solution level, stirring was terminated (unless otherwise noted), and the electrodes were connected to the potentiostat in order to start the experiment.

Measurements of the capacitance of the electrochemical cell

Figure S5. CV curve recorded for 0.2M LiBr in DMF (GC rods were used as working and counter electrodes: scan rate – 6 V/s, 10 cycles).

The capacitance of the electrochemical cell equipped with GC rod electrodes and filled with 0.2M LiBr DMF solution can be calculated as C = S/(2

.ΔV.r), where C – capacitance, S – the

Page 55: Coupling of Alternating Current to Transition-Metal

S34

inner area of the I/V plot, ΔV – the voltage range, r – the scan rate. The calculation results in a value of about 50 μF for our cell and electrolyte.

Cyclic voltammetry studies

CV study of the NiBr2·DME and Bipy mixture with the addition of 4-bromobenzotrifluoride (Figure 4B main text)

A DMA solution, containing 0.1M TBAPF6, 2mM NiBr2·DME, and 2mM Bipy was prepared as described in “General information and procedure for electrochemical studies”. We used a glassy carbon working electrode (disk, d = 1mm) and a platinum wire counter electrode. The potential of the working electrode was changed to a -0.3, -2, -0.3 V sequence (scan rate – 0.25V/s; no stirring). After the addition of 2mM of 4-bromobenzotrifluoride (3) to the mixture, the second CV experiment was conducted with the same parameters.

CV study of the amination system

Figure S6. CV curve recorded for 5mM NiBr2·DME, 7.5mM di-tBuBipy, 50mM 4-bromobenzotrifluoride (3), and 150mM Morpholine in a 0.2M LiBr DMA solution (Ø1mm GC disk – working electrode, Pt wire – counter electrode, and scan rate – 0.25 V/s).

Page 56: Coupling of Alternating Current to Transition-Metal

S35

Figure S7. CV curves recorded for 5mM NiBr2·DME and 7.5mM di-tBuBipy (green) with a sequential addition of 150 mM morpholine (blue) and 50 mM 4-bromobenzotrifluoride (3) (red) in a 0.2M LiBr DMA solution (Ø1mm GC disk – working electrode, Pt wire – counter electrode, and scan rate – 0.25 V/s).

Figure S8. CV curves recorded for 5mM NiBr2·DME, 7.5mM di-tBuBipy, 50mM 4-bromobenzotrifluoride (3), and 150mM Morpholine in a 0.2M LiBr DMA solution (GC rods were used as working and counter electrodes; scan rates – 0.25 V/s (green), 5 V/s (blue), and 24 V/s (red)).

Page 57: Coupling of Alternating Current to Transition-Metal

S36

Figure S9. CV curve recorded for 5mM NiBr2·DME, 7.5mM di-tBuBipy, 50mM 4-bromobenzotrifluoride (3), and 150mM Morpholine in 0.2M LiBr DMA solution (GC rods were used as working and counter electrodes; scan rate – 24 V/s and a 1400 rpm stirring rate).

CV study of the etherification system

Figure S10. CV curve recorded for 5mM NiBr2·DME, 5mM Bipy, 50mM 4-bromobenzotrifluoride (3), 150mM n-Hexanol, and 50mM Quinuclidine in a 0.2M LiBr DMF solution (Ø1mm GC disk – working electrode, Pt wire – counter electrode, and scan rate – 0.25 V/s).

Page 58: Coupling of Alternating Current to Transition-Metal

S37

Figure S11. CV curves recorded for 5mM NiBr2·DME and 5mM Bipy (purple) with a sequential addition of 50mM Quinuclidine (green), 150mM n-Hexanol (blue), and 50mM 4-bromobenzotrifluoride (3) (red) in a 0.2M LiBr DMF solution (Ø1mm GC disk – working electrode, Pt wire – counter electrode, and scan rate – 0.25 V/s).

Figure S12. CV curves recorded for 5mM NiBr2·DME, 5mM Bipy, 50mM 4-bromobenzotrifluoride (3), 150mM n-Hexanol, and 50mM Quinuclidine in a 0.2M LiBr DMF solution (GC rods were used as working and counter electrodes; scan rates – 0.25 V/s (green), 2 V/s (blue) and 24 V/s (red)).

Page 59: Coupling of Alternating Current to Transition-Metal

S38

Figure S13. CV curves recorded for 5mM NiBr2·DME, 5mM Bipy, 50mM 4-bromobenzotrifluoride (3), 150mM n-Hexanol, and 50mM Quinuclidine in a 0.2M LiBr DMF solution (GC rods were used as working and counter electrodes; scan rate – 24 V/s; with no stirring (green), 200 rpm (blue), and 1400 rpm (red) stirring rates).

Page 60: Coupling of Alternating Current to Transition-Metal

S39

CV study of the esterification mixture

Figure S14. CV curve recorded for 5mM NiBr2·DME, 5mM Bipy, 50mM 4-bromobenzotrifluoride (3), 75mM BocProOH, and 3eq. K2CO3 suspension in 0.1M TBAPF6 DMA solution (Ø1mm GC disk – working electrode, Pt wire – counter electrode, and scan rate – 0.25 V/s).

Figure S15. CV curves recorded for 5 mM NiBr2·DME and 5 mM Bipy (purple) with a sequential addition of 50 mM BocProOH (green), K2CO3 (blue), and 50 mM 4-bromobenzotrifluoride (3) (red) in a 0.2M LiBr DMF solution (Ø1mm GC disk – working electrode, Pt wire – counter electrode, and scan rate – 0.25 V/s).

Page 61: Coupling of Alternating Current to Transition-Metal

S40

Figure S16. CV curves recorded for 5mM NiBr2·DME, 5mM Bipy, 50mM 4-bromobenzotrifluoride (3), 75mM BocProOH, and 3eq. K2CO3 suspension in a 0.1M TBAPF6 DMA solution (GC rods were used as working and counter electrodes; scan rates – 0.25 V/s (green), 2 V/s (blue), and 24 V/s (red)).

Figure S17. CV curve recorded for 5mM NiBr2·DME, 5mM Bipy, 50mM 4-bromobenzotrifluoride (3), 75mM BocProOH, and 3eq. K2CO3 suspension in a 0.1M TBAPF6 DMA solution (GC rods were used as working and counter electrodes; scan rate – 24 V/s; stirring rate – 200 rpm).

Page 62: Coupling of Alternating Current to Transition-Metal

S41

Lithium bromide oxidation

Figure S18. CV curves recorded for 3mM LiBr (red) and 50mM LiBr (blue) in 0.1M TBAPF6 DMF solution and a pure 0.1M TBAPF6 DMF solution (green) (Ø1mm GC disk – working electrode, Pt wire – counter electrode, and scan rate – 0.25 V/s).

Page 63: Coupling of Alternating Current to Transition-Metal

S42

Open circuit potential (OCP) experiments

Amination

A 0.2M LiBr DMA solution, containing 5mM NiBr2·DME, 7.5mM di-tBuBipy, 50mM 4-bromobenzotrifluoride (3), and 150mM Morpholine was prepared as described in “General information and procedure for electrochemical studies”. Two glassy carbon rod electrodes (which are typically used in regular AC-assisted experiments) served as the working and counter electrodes, together with an Ag+/Ag reference electrode. The electric circuit was assembled as described in the “Set-up preparation” section; the stirring rate of the solution was set to 1400 rpm. Alternating voltage was applied to the GC rod electrodes by the function generator (sine waveform, 2 Hz) and adjusted to a value of 3V for peak voltage. The potential between one of the GC rods and the Ag+/Ag reference electrodes was measured continuously using the potentiostat (with the “open circuit potential” technique). The resulting data can be found in Figure 4A in the main text.

Etherification

A 0.2M LiBr DMF solution, containing 5mM NiBr2·DME, 5mM Bipy, 50mM 4-bromobenzotrifluoride (3), 150mM n-Hexanol, and 50mM Quinuclidine was prepared as described in “General information and procedure for electrochemical studies”. The following procedure is identical to the one described in “Open circuit potential (OCP) experiments: Amination” section (stirring rate - 200 rpm).

Figure S19. GC electrode potential vs the Ag+/Ag reference electrode in the etherification experiment described above.

Page 64: Coupling of Alternating Current to Transition-Metal

S43

Esterification

A 0.1M TBAPF6 DMA solution containing 5mM NiBr2·DME, 5mM Bipy, 50mM 4-bromobenzotrifluoride (3), 75mM Boc-L-ProOH, and a 3eq. K2CO3 suspension was prepared as described in “General information and procedure for electrochemical studies”. The following procedure is identical to the one described in the “Open circuit potential (OCP) experiments: Amination” section (stirring rate – 200 rpm).

Figure S20. GC electrode potential vs the Ag+/Ag reference electrode in the esterification experiment described above.

Page 65: Coupling of Alternating Current to Transition-Metal

S44

Mechanistic studies

Amination in a divided cell

A divided handmade cell was assembled using the glass cell described in “Electrochemical cell” section; a glass paper filter and polydimethylsiloxane (PDMS) were used as the binding and sealing agent (Figure S21). This set-up provided a large surface area of the membrane, avoiding high resistance between the compartments.

Next, 2 ml of the 0.2M LiBr DMA solution containing 5mM NiBr2·DME, 7.5mM di-tBuBipy, 50mM 4-bromobenzotrifluoride (3), and 150mM Morpholine were transferred to each of the compartments of the divided cell. The cap, equipped with two glassy carbon rod electrodes, was attached so that the electrodes were in different compartments. A weak argon flow was set in the cell to be above the level of the solution, and no stirring was applied. The electric circuit was assembled as described in the “Set-up preparation” section. Alternating voltage was applied to the GC rod electrodes and adjusted to a value of 3 V for peak voltage. Electrolysis was conducted for 4 hours. Thereafter, 0.5 ml of the solution from each of the compartments was transferred to a screw-capped vial, mixed with a small amount of Na2H2EDTA water solution, diluted with an aqueous solution of K2CO3 (0.1 M), and extracted with a Pentane/Et2O (1/1 ratio) mixture. The extracts were dried over Na2SO4; the solvent was evaporated carefully under vacuum. Then, the residues were used to determine the NMR yield of the experiment. The NMR yield of the amination product, 4-(4-(trifluoromethyl)phenyl)morpholine (7), amounted to 60% and 77% for each of the compartments (the noticeable difference in yields is probably due to a small inequality of compartment volumes).

Figure S21. Self-made divided electrochemical cell.

Page 66: Coupling of Alternating Current to Transition-Metal

S45

DC-assisted amination and etherification with a Zn sacrificial anode

Amination (Figure 4C in the main text)

A 0.2M LiBr DMA solution containing 5mM NiBr2·DME, 7.5mM di-tBuBipy, 50mM 4-bromobenzotrifluoride (3), and 150mM Morpholine was prepared as described in “General information and procedure for the electrochemical studies”. A teflon cap equipped with a glassy carbon electrode, a zinc rod electrode, and an Ag+/Ag reference electrode was attached. The solution was bubbled with argon for at least 5 minutes with moderate stirring; then, a mild argon flow was set in the cell to be above the level of the solution and the stirring rate was changed to 1400 rpm. A potentiostat was used to apply voltage to the system. One GC rod electrode served as a working electrode, whereas the zinc rod served as a counter electrode. The electrolysis was conducted for 4 hours at a constant potential of the GC rod electrode set to -2.2V (vs Ag+/Ag). Thereafter, 0.5 ml of the solution was transferred to a screw-capped vial, mixed with a small amount of Na2H2EDTA water solution, diluted with aqueous solution of K2CO3 (0.1 M), and extracted with a Pentane/Et2O (1/1 ratio) mixture. The extract was dried over Na2SO4, and the solvent was evaporated carefully under vacuum. Then, the residue was used to determine the NMR yield of the experiment. Biaryl product (4,4'-bis(trifluoromethyl)-1,1'-biphenyl, 30) was the major product in the reaction with an NMR yield of 63%; the NMR yield of 4-(4-(trifluoromethyl)phenyl)morpholine (7) was 2%.

Etherification

Figure S22. Control experiment with DC and Zn sacrificial anode. 4-bromobenzotrifluoride (3) (100 mM), n-Hexanol (150 mM), Quinuclidine (50 mM), NiBr2·DME (5 mM), Bipy (5 mM) in 0.2M LiBr DMF solution.

A 0.2M LiBr DMF solution containing 5mM NiBr2·DME, 5mM Bipy, 50mM 4-bromobenzotrifluoride (3), 150mM n-Hexanol, and 100mM Quinuclidine was prepared as described in “General information and procedure for electrochemical studies”. The following procedure is similar to the one described in “DC-assisted electrolysis with a Zn sacrificial anode: Amination” section. Biaryl product (4,4'-bis(trifluoromethyl)-1,1'-biphenyl, 30) was the major product in the reaction, with an NMR yield of 73%; the NMR yield of 1-(hexyloxy)-4-(trifluoromethyl)benzene (5) was 8%.

Page 67: Coupling of Alternating Current to Transition-Metal

S46

Kinetic experiments with ON/OFF cycles of AC

Amination (Figure 4C in the main text)

A 0.2M LiBr DMA solution containing 5mM NiBr2·DME, 7.5mM Bipy, 50mM 4-bromobenzotrifluoride (3), and 150mM Morpholine was prepared as described in “General procedure for the amination experiments”. The electric circuit was assembled as described in the “Set-up preparation” section. The electrolysis (sine waveform, 2 Hz, and 3V peak voltage) was conducted for 20 minutes, then stopped for 20 minutes, then again continued for 20 minutes, and again stopped for 20 minutes (the total duration of the experiment is 80 minutes). During the experiment, 30 μL samples of the solution were taken every 5 minutes with a syringe, diluted with 1 ml of acetonitrile, filtered through a glass paper filter, and analyzed by HPLC. The conversion at each time point was determined from the ratio of the integrals for the peaks (PDA detector) of the starting material (4-bromobenzotrifluoride, 3), and the product (4-(4-(trifluoromethyl)phenyl)morpholine, 7), while considering the different sensitivity of the PDA detector to these compounds. The difference in sensitivity of the PDA detector to 4-bromobenzotrifluoride (3) and 4-(4-(trifluoromethyl)phenyl)morpholine (7) was determined from a reference chromatogram of a 1:1 mixture of these compounds.

Etherification (Figure S23)

A 0.2M LiBr DMF solution containing 5mM NiBr2·DME, 5mM Bipy, 50mM 4-bromobenzotrifluoride, 150mM n-Hexanol, and 100mM Quinuclidine was prepared as described in “General procedure for the etherification experiments”. The electric circuit was assembled as described in the “Set-up preparation” section. The electrolysis (sine waveform, 10 Hz, and 3V peak voltage) was conducted for 40 minutes, then stopped for 40 minutes, then again continued for 40 minutes, and again stopped for 40 minutes (the total duration of the experiment is 160 minutes). During the experiment, 30 μL samples of the solution were taken every 10 minutes with a syringe, diluted with 1 ml of acetonitrile, filtered through a glass paper filter, and analyzed by HPLC. The conversion at each time point was determined from the ratio of the integrals for the peaks (PDA detector) of the starting material (4-bromobenzotrifluoride, 3), and the product (1-(hexyloxy)-4-(trifluoromethyl)benzene, 5), while considering the different sensitivity of the PDA detector to these compounds. The difference in the sensitivity of the PDA detector to 4-bromobenzotrifluoride (3) and 1-(hexyloxy)-4-(trifluoromethyl)benzene (5) was determined from a reference chromatogram of a 1:1 mixture of these compounds.

Page 68: Coupling of Alternating Current to Transition-Metal

S47

Figure S23. Kinetics of the etherification reaction between 4-bromobenzotrifluoride (3) (50 mM) and n-hexanol (150 mM) (NiBr2·DME (5 mM), Bipy (5 mM) and Quinuclidine (100 mM) in 0.2M LiBr DMF solution) with ON/OFF cycles of AC. Conversion of 4-bromobenzotrifluoride (3) to 1-(hexyloxy)-4-(trifluoromethyl)benzene (5) was determined by HPLC.

Esterification (Figure S24)

A 0.1M TBAPF6 DMA solution containing 5mM NiBr2·DME, 5mM Bipy, 50mM 4-bromobenzotrifluoride (3), 75mM benzoic acid, and 10 equiv. K2CO3 suspension was prepared as described in “General procedure for the esterification experiments”. The electric circuit was assembled as described in the “Set-up preparation” section. The electrolysis (sine waveform, 2 Hz, and 3V peak voltage) was conducted for 40 minutes, then stopped for 40 minutes, then again continued for 40 minutes, and again stopped for 40 minutes (the total duration of the experiment is 160 minutes). During the experiment, 30 μL samples of the solution were taken every 10 minutes with a syringe, diluted with 1 ml of acetonitrile, filtered through a glass paper filter, and analyzed by HPLC. The conversion at each time point was determined from the ratio of the integrals for the peaks (PDA detector) of the starting material (4-bromobenzotrifluoride, 3), and the product (4-(trifluoromethyl)phenyl benzoate, 18), while considering the different sensitivity of the PDA detector to these compounds. The difference in the sensitivity of the PDA detector to 4-bromobenzotrifluoride (3) and 4-(trifluoromethyl)phenyl benzoate (18) was determined from a reference chromatogram of a 1:1 mixture of these compounds.

Page 69: Coupling of Alternating Current to Transition-Metal

S48

Figure S24. Kinetics of the esterification reaction between 4-bromobenzotrifluoride (3) (50 mM) and benzoic acid (75 mM) (Quinuclidine (5mM NiBr2·DME, 5mM Bipy, 10 equiv. K2CO3 in 0.1M TBAPF6 DMA solution) with ON/OFF cycles of AC. Conversion of 4-bromobenzotrifluoride (3) to 4-(trifluoromethyl)phenyl benzoate (18) was determined by HPLC.

Page 70: Coupling of Alternating Current to Transition-Metal

S49

NMR spectra

Page 71: Coupling of Alternating Current to Transition-Metal

S50

Page 72: Coupling of Alternating Current to Transition-Metal

S51

Page 73: Coupling of Alternating Current to Transition-Metal

S52

Page 74: Coupling of Alternating Current to Transition-Metal

S53

Page 75: Coupling of Alternating Current to Transition-Metal

S54

Page 76: Coupling of Alternating Current to Transition-Metal

S55

Page 77: Coupling of Alternating Current to Transition-Metal

S56

Page 78: Coupling of Alternating Current to Transition-Metal

S57

Page 79: Coupling of Alternating Current to Transition-Metal

S58

Page 80: Coupling of Alternating Current to Transition-Metal

S59

Page 81: Coupling of Alternating Current to Transition-Metal

S60

Page 82: Coupling of Alternating Current to Transition-Metal

S61

Page 83: Coupling of Alternating Current to Transition-Metal

S62

Page 84: Coupling of Alternating Current to Transition-Metal

S63

Page 85: Coupling of Alternating Current to Transition-Metal

S64

Page 86: Coupling of Alternating Current to Transition-Metal

S65

Page 87: Coupling of Alternating Current to Transition-Metal

S66

Page 88: Coupling of Alternating Current to Transition-Metal

S67

Page 89: Coupling of Alternating Current to Transition-Metal

S68

Page 90: Coupling of Alternating Current to Transition-Metal

S69

Page 91: Coupling of Alternating Current to Transition-Metal

S70

Page 92: Coupling of Alternating Current to Transition-Metal

S71

Page 93: Coupling of Alternating Current to Transition-Metal

S72

Page 94: Coupling of Alternating Current to Transition-Metal

S73

Page 95: Coupling of Alternating Current to Transition-Metal

S74

Page 96: Coupling of Alternating Current to Transition-Metal

S75

Page 97: Coupling of Alternating Current to Transition-Metal

S76

Page 98: Coupling of Alternating Current to Transition-Metal

S77

Page 99: Coupling of Alternating Current to Transition-Metal

S78

Page 100: Coupling of Alternating Current to Transition-Metal

S79

Page 101: Coupling of Alternating Current to Transition-Metal

S80

Page 102: Coupling of Alternating Current to Transition-Metal

S81

Page 103: Coupling of Alternating Current to Transition-Metal

S82

Page 104: Coupling of Alternating Current to Transition-Metal

S83

Page 105: Coupling of Alternating Current to Transition-Metal

S84

Page 106: Coupling of Alternating Current to Transition-Metal

S85

Page 107: Coupling of Alternating Current to Transition-Metal

S86

Page 108: Coupling of Alternating Current to Transition-Metal

S87

Page 109: Coupling of Alternating Current to Transition-Metal

S88

Page 110: Coupling of Alternating Current to Transition-Metal

S89

Page 111: Coupling of Alternating Current to Transition-Metal

S90

Page 112: Coupling of Alternating Current to Transition-Metal

S91

Page 113: Coupling of Alternating Current to Transition-Metal

S92

Page 114: Coupling of Alternating Current to Transition-Metal

S93

Page 115: Coupling of Alternating Current to Transition-Metal

S94

Page 116: Coupling of Alternating Current to Transition-Metal

S95

Page 117: Coupling of Alternating Current to Transition-Metal

S96

Page 118: Coupling of Alternating Current to Transition-Metal

S97

Page 119: Coupling of Alternating Current to Transition-Metal

S98

Page 120: Coupling of Alternating Current to Transition-Metal

S99

Page 121: Coupling of Alternating Current to Transition-Metal

S100

Page 122: Coupling of Alternating Current to Transition-Metal

S101

Page 123: Coupling of Alternating Current to Transition-Metal

S102

Page 124: Coupling of Alternating Current to Transition-Metal

S103

Page 125: Coupling of Alternating Current to Transition-Metal

S104

Page 126: Coupling of Alternating Current to Transition-Metal

S105

Page 127: Coupling of Alternating Current to Transition-Metal

S106

Page 128: Coupling of Alternating Current to Transition-Metal

S107

Page 129: Coupling of Alternating Current to Transition-Metal

S108

Page 130: Coupling of Alternating Current to Transition-Metal

S109

Page 131: Coupling of Alternating Current to Transition-Metal

S110

Page 132: Coupling of Alternating Current to Transition-Metal

S111

Page 133: Coupling of Alternating Current to Transition-Metal

S112

Page 134: Coupling of Alternating Current to Transition-Metal

S113

Page 135: Coupling of Alternating Current to Transition-Metal

S114

Page 136: Coupling of Alternating Current to Transition-Metal

S115

Page 137: Coupling of Alternating Current to Transition-Metal

S116

Page 138: Coupling of Alternating Current to Transition-Metal

S117

Page 139: Coupling of Alternating Current to Transition-Metal

S118

References

(1) Li, C.; Kawamata, Y.; Nakamura, H.; Vantourout, J. C.; Liu, Z.; Hou, Q.; Bao, D.; Starr, J. T.; Chen, J.;

Yan, M.; Baran, P. S. Electrochemically Enabled, Nickel-Catalyzed Amination. Angew. Chem., Int. Ed.

2017, 56, 13088-13093.

(2) Corcoran, E. B.; Pirnot, M. T.; Lin, S.; Dreher, S. D.; DiRocco, D. A.; Davies, I. W.; Buchwald, S. L.;

MacMillan, D. W. Aryl amination using ligand-free Ni(II) salts and photoredox catalysis. Science 2016,

353, 279-83.

(3) Yang, C.; Zhang, F.; Deng, G. J.; Gong, H. Amination of Aromatic Halides and Exploration of the

Reactivity Sequence of Aromatic Halides. J. Org. Chem. 2019, 84, 181-190.

(4) Kawamata, Y.; Vantourout, J. C.; Hickey, D. P.; Bai, P.; Chen, L.; Hou, Q.; Qiao, W.; Barman, K.;

Edwards, M. A.; Garrido-Castro, A. F.; deGruyter, J. N.; Nakamura, H.; Knouse, K.; Qin, C.; Clay, K. J.; Bao,

D.; Li, C.; Starr, J. T.; Garcia-Irizarry, C.; Sach, N.; White, H. S.; Neurock, M.; Minteer, S. D.; Baran, P. S.

Electrochemically Driven, Ni-Catalyzed Aryl Amination: Scope, Mechanism, and Applications. J. Am.

Chem. Soc. 2019, 141, 6392-6402.

(5) Lim, C. H.; Kudisch, M.; Liu, B.; Miyake, G. M. C-N Cross-Coupling via Photoexcitation of Nickel-Amine

Complexes. J. Am. Chem. Soc. 20192018, 140, 7667-7673.

(6) Tobisu, M.; Shimasaki, T.; Chatani, N. Ni0-catalyzed Direct Amination of Anisoles Involving the

Cleavage of Carbon–Oxygen Bonds. Chem. Lett. 2009, 38, 710-711.

(7) Hartwig, J. F.; Kawatsura, M.; Hauck, S. I.; Shaughnessy, K. H.; Alcazar-Roman, L. M. Room-

Temperature Palladium-Catalyzed Amination of Aryl Bromides and Chlorides and Extended Scope of

Aromatic C-N Bond Formation with a Commercial Ligand. J. Org. Chem. 1999, 64, 5575-5580.

(8) Welin, E. R.; Le, C.; Arias-Rotondo, D. M.; McCusker, J. K.; MacMillan, D. W. Photosensitized, energy

transfer-mediated organometallic catalysis through electronically excited nickel(II). Science 2017, 355,

380-385.

(9) Lin, J.; Li, Z.; Kan, J.; Huang, S.; Su, W.; Li, Y. Photo-driven redox-neutral decarboxylative carbon-

hydrogen trifluoromethylation of (hetero)arenes with trifluoroacetic acid. Nat. Commun. 2017, 8, 14353.

(10) Pieber, B.; Malik, J. A.; Cavedon, C.; Gisbertz, S.; Savateev, A.; Cruz, D.; Heil, T.; Zhang, G.; Seeberger,

P. H. Semi-heterogeneous Dual Nickel/Photocatalysis using Carbon Nitrides: Esterification of Carboxylic

Acids with Aryl Halides. Angew. Chem., Int. Ed. 2019, 58, 9575-9580.

(11) Morimoto, H.; Tsubogo, T.; Litvinas, N. D.; Hartwig, J. F. A broadly applicable copper reagent for

trifluoromethylations and perfluoroalkylations of aryl iodides and bromides. Angew. Chem., Int. Ed.

2011, 50, 3793-8.

(12) Escobar, R. A.; Johannes, J. W. A Unified and Practical Method for Carbon-Heteroatom Cross-

Coupling using Nickel/Photo Dual Catalysis. Chem. Eur. J. 2020, 26, 5168-5173.

(13) Terrett, J. A.; Cuthbertson, J. D.; Shurtleff, V. W.; MacMillan, D. W. Switching on elusive

organometallic mechanisms with photoredox catalysis. Nature 2015, 524, 330-4.

(14) Cavedon, C.; Madani, A.; Seeberger, P. H.; Pieber, B. Semiheterogeneous Dual Nickel/Photocatalytic

(Thio)etherification Using Carbon Nitrides. Org. Lett. 2019, 21, 5331-5334.

(15) Keaveney, S. T.; Schoenebeck, F. Palladium-Catalyzed Decarbonylative Trifluoromethylation of Acid

Fluorides. Angew. Chem., Int. Ed. 2018, 57, 4073–4077.