Download as pdf or txt
Download as pdf or txt
You are on page 1of 4

ChemComm

View Article Online


COMMUNICATION View Journal | View Issue
Published on 17 January 2018. Downloaded by Indian Institute of Technology Kanpur on 12/11/2023 7:34:43 PM.

Electrocatalytic CO2 reduction by a cobalt


bis(pyridylmonoimine) complex: effect of acid
Cite this: Chem. Commun., 2018,
54, 1579 concentration on catalyst activity and stability†
Received 6th November 2017,
Accepted 17th January 2018 Weixuan Nie and Charles C. L. McCrory *

DOI: 10.1039/c7cc08546j

rsc.li/chemcomm

A Co complex with a redox-active bis(pyridylmonoimine) ligand has The decomposition mechanism may involve the formation of a
been prepared and shows catalytic activity for electrochemical CO2 catalyst-CO2 adduct via CO2–ligand bonding with the redox
reduction in acetonitrile. Addition of a proton source such as water active ligands.9,20,21
or trifluoroethanol dramatically improves the activity and stability In this work, a Co complex with a bis(pyridylmonoimine)
of the molecular catalyst. The Co complex reduces CO2 to CO ligand [Co(L–L)Br2] has been prepared (Scheme 1) and shows
selectively at 1.95 V vs. Fc+/0 in the presence of high concentrations catalytic activity for CO2 reduction in acetonitrile (CH3CN). The
of water. The activity of the Co complex for CO2 reduction compares addition of the proton sources H2O and 2,2,2-trifluoroethanol
favorably to other molecular Co-based catalysts in acetonitrile (TFE) accelerates the rate of electrocatalytic CO2 reduction and
solutions. also improves catalyst stability and Faradaic efficiency. In the
presence of high concentrations of H2O and TFE, the [Co(L–L)Br2]
The electrochemical CO2 reduction reaction (CO2RR) is a promising catalyst reduces CO2 selectively to CO with high Faradaic efficiency
strategy for the conversion of waste CO2 into value added and high turnover frequency (TOF) compared to other reported Co
products.1–3 State-of-the-art solid-state CO2 reduction catalysts complexes with similar ligands.14,17–19
such as Cu reduce CO2 to highly-reduced products, but suffer The ligand L–L was synthesized via a classic Schiff base
from poor selectivity and generate multiple gaseous and liquid condensation reaction of ethylenediamine with 2-acetylpyridine.
products.4,5 In contrast, several molecular electrocatalysts [Co(L–L)Br2] was prepared by allowing L–L to react with
reduce CO2 selectively to single products, making them an 1 equivalent of CoBr2 in THF. Aerobic oxidation of [Co(L–L)Br2]
important class of CO2RR catalysts.6,7 Moreover, the activity in the presence of 1 equiv. HBr(aq) for 18 hours afforded green
and selectivity of molecular catalysts can be tuned by altering [Co(L–L)Br2]Br (Scheme 1), which was characterized by 1H-NMR
their ligand structure.8,9 (Fig. S2, ESI†) and elemental analysis. The X-ray single crystal
Redox-active ligands have been shown to play an important structure of [Co(L–L)Br2]Br indicates that the central cobalt metal
role in CO2RR by molecular electrocatalysts.9–16 Redox-active adopts a bipyramidal six-coordinated configuration. The tetra-
ligands are able to store electrons in the ligand structure, dentate ligand L–L coordinates to Co with four N atoms in a
facilitating the multi-electron reduction of CO2. Previous slightly distorted square planar mode. Two Br atoms occupy the
reports have shown that cobalt complexes with pyridylimine- axial coordination sites and a third Br atom exists outside the
based ligands efficiently reduce CO2 to CO in the presence of coordination sphere as a counter anion (Fig. 1). Note that
H2O14,17–19 and propose that the redox-active properties of the [Co(L–L)Br2] (the Co2+ complex) was used for all electrochemical
ligand contribute to the preferential reduction of CO2 over measurements and [Co(L–L)Br2]Br (the Co3+ complex) was used
competitive proton reduction. However, inactivation and only for characterization of the complex (see ESI†).
decomposition of the catalyst is also prevalent in these systems.

Department of Chemistry, University of Michigan, 930 North University Ave,


Ann Arbor, MI, 48109-1055, USA. E-mail: cmccrory@umich.edu
† Electronic supplementary information (ESI) available: Synthesis procedures of
the ligand and complexes; selected bond lengths and angles; NMR, IR and UV-vis
spectra; paramagnetic susceptibility measurements; selected electrochemical
measurements; SEM-EDS results; TOF calculation; ICP-MS results. CCDC
1571483. For ESI and crystallographic data in CIF or other electronic format Scheme 1 Synthesis of the ligand L–L, complex catalyst [Co(L–L)Br2], and
see DOI: 10.1039/c7cc08546j [Co(L–L)Br2]Br.

This journal is © The Royal Society of Chemistry 2018 Chem. Commun., 2018, 54, 1579--1582 | 1579
View Article Online

Communication ChemComm
Published on 17 January 2018. Downloaded by Indian Institute of Technology Kanpur on 12/11/2023 7:34:43 PM.

Fig. 1 X-Ray crystal structure of [Co(L–L)Br2]Br with ellipsoids shown at


50% probability (H atoms are omitted for clarity).

Fig. 2 CV of 1.2 mM [Co(L–L)Br2] in CH3CN with 0.1 M nBu4NPF6 under


N2 (black: under N2; red: under CO2; conditions: scan rate: 50 mV s 1;
working electrode: glassy carbon working electrode; reference electrode:
Ag/AgNO3 (1 mM); counter electrode: carbon rod).
Fig. 3 CV of 0.3 mM [Co(L–L)Br2] in CH3CN with 0.1 M nBu4NPF6 (a)
under CO2 in the presence of H2O of different concentrations; (b) under
CO2 in the presence of TFE of different concentrations; (c) under N2 and
Cyclic voltammograms (CV) of complex [Co(L–L)Br2] were CO2 with H2O.
recorded in a nitrogen-saturated CH3CN solution with 0.1 M
nBu4NPF6 (Fig. 2). Note that all CVs have been IR compensated.
Two reversible peaks are observed at 0.42 V vs. Fc+/0 and potentials than the CO2 reduction peak. Note that the catalytic
1.00 V vs. Fc+/0 and are assigned as the Co3+/2+ and Co2+/1+ peak shape in the presence of proton source resembles an
couples, respectively. The broad quasi-reversible redox feature ‘‘inverted peak’’ (the inverted return wave when potential is
observed at B 1.6 V vs. Fc+/0 is assigned to a ligand-based scanned back).25 We attribute this peak shape to the overlap of the
redox process. This assignment is consistent with that of metal catalytic current response for CO2 reduction with the more negative,
complexes with other redox-active pyridylmonoimine9 and reversible Co1+/0 redox feature. A similar behavior has been observed
bipyridyl8 ligands. The most-negative reversible peak at 2.16 V in the case of CO2 reduction by [Mn(mesbpy)(CO)3(MeCN)]+.24
vs. Fc+/0 is assigned to the Co1+/0 couple. To further confirm the catalytic ability of [Co(L–L)Br2] for
The cyclic voltammogram (CV) of complex [Co(L–L)Br2] in a CO2 reduction, we conducted a series of 30 minute controlled
CO2-saturated CH3CN solution shows a large increase in the potential electrolysis (CPE) experiments in CH3CN (Table 1 and
reductive current consistent with an electrocatalytic process Table S5, ESI†). The CPE experiments were conducted with no
(Fig. 2). The onset potential of the electrocatalytic current is IR compensation for solution resistance, and the reported
B 1.65 V vs. Fc+/0, the approximate potential of the ligand electrolysis potentials are the actual applied potentials. Note
reduction. Adding the proton source H2O to the electrolyte that the electrolyte solutions as prepared contain up to B0.04 M
solution increases the magnitude of the catalytic current H2O even when no additional proton source is added.26 Gas and
(Fig. 3a for 0.82 M and 11.0 M H2O, other concentrations liquid phases were analyzed by GC and HPLC, respectively, and
shown in Fig. S9, ESI†). Increasing the concentration of H2O the Faradaic efficiency of each product (FEproduct) was deter-
also slightly shifts the catalytic peak potential, from 2.0 V vs. mined. Turnover frequencies (TOFs) were calculated using two
Fc+/0 at 0.82 M H2O to 1.95 V vs. Fc+/0 at 11.0 M. A similar methods: (1) by dividing the total amount of CO generated by
phenomenon was observed when adding TFE as a proton the catalyst amount in solution and (2) calculated from CPE
source (Fig. 3b and Fig. S10, ESI†). Since TFE is more acidic data with the equations described by Savéant et al.27,28 (see the
than H2O in CH3CN,22–24 a lower concentration of TFE is ESI† for further discussion of TOF). The electrolysis potentials
necessary to reach a comparable catalytic current under CO2. were chosen based on the peak potential of CO2 reduction in
To confirm that the catalytic peak is due to CO2 reduction and Fig. 3. After each electrolysis, the un-rinsed electrode surface
not competitive H2 evolution, CVs were also measured in the was investigated with both scanning electron microscopy-energy
absence of CO2 (Fig. 3c for H2O, Fig. S11, ESI† for TFE). The dispersion spectroscopy (SEM-EDS) and inductively coupled
onset of H2 evolution occurs only at much more negative plasma-mass spectrometry (ICP-MS) to detect the formation of

1580 | Chem. Commun., 2018, 54, 1579--1582 This journal is © The Royal Society of Chemistry 2018
View Article Online

ChemComm Communication

Table 1 Conditions and product analysis of the controlled potential electrolysis for CO2 reduction

Faradaic efficiency/%
[Proton Co weight% on the
E/V vs. Fc+/0 a source]/M Charge/C TOFA/s 1b
TOFB/s 1c
CO H2 HCOOH electrode surface from EDSd
5
1.95 — 2.2  0.3 (2.9  2.1)  10 0.053 2.8  2.6 0 6.3  2.6 0.52%
5
2.15 — 4.3  1.4 (8.1  2.5)  10 0.453 4.2  1.7 0 2.7  0.9 0.52%
Published on 17 January 2018. Downloaded by Indian Institute of Technology Kanpur on 12/11/2023 7:34:43 PM.

4
1.95 1.1 M H2O 3.4  0.4 (2.7  0.9)  10 4.26 16.3  5.1 0 4.5  1.1 —e
3
1.95 5.5 M H2O 7.9  0.7 (2.3  0.5)  10 393 67.4  11.0 0.3  0.1 1.9  0.5 —e
3
1.95 11.0 M H2O 8.3  0.9 (3.2  0.4)  10 620 80.5  4.7 1.1  0.5 0.7  0.2 0.35%
3
2.15 11.0 M H2O 16.2  2.5 (8.1  1.3)  10 3961 104.3  5.5 0.6  0.4 0 0.18%
3
1.85 5.5 M TFE 10.1  2.3 (4.2  0.5)  10 1089 87.7  8.2 1.7  0.7 0 0.01%
3
2.05 5.5 M TFE 16.8  2.5 (8.2  0.9)  10 4131 102.7  3.10 1.0  0.4 0 0.007%
a
Applied potential without IR compensation. b TOFA is calculated based on total concentration of catalyst in solution, and therefore is a
significant underestimate of catalytic activity (see ESI). c TOFB is derived from CPE data as described by Savéant et al. (see ESI).27,28 d ‘‘Co weight%’’
means the weight% of Co in the total surface atoms. e Not measured.

solid-state deposits (Fig. S15–S22 and Table S8, ESI†).29 Control Faradaic efficiencies and TOF values were observed for CO
experiments consisting of 30 min CPE with the glassy carbon production and even less decomposition to surface Co species
electrode and clean electrolyte, only CoBr2 in solution, and only was observed by SEM-EDS (Table 1). Compared to other Co-based
the ligand L–L in solution all showed no activity for CO2 CO2RR catalysts, [Co(L–L)Br2] reduces CO2 to CO with high Faradaic
reduction (Table S6, ESI†). efficiency and fast TOF at less negative applied potentials
For electrolyte solutions containing [Co(L–L)Br2] in the (Table S11, ESI†).
absence of any added proton source, the solution became Note that when CPE experiments were conducted in 11.0 M
turbid after 30 min of CPE at 1.95 V vs. Fc+/0 and 2.15 V H2O at potentials which are more positive ( 1.75 V vs. Fc+/0)
vs. Fc+/0. Only a small amount of CO (FECO = 2.8%@ 1.95 V, and more negative ( 2.35 V vs. Fc+/0) than the catalytic peak
4.2%@ 2.15 V) and formic acid HCOOH (FEHCOOH = potential in Fig. 3, Faradaic efficiency for CO production
6.3%@ 1.95 V, 2.7%@ 2.15 V) were detected after the electro- decreased and an increase in Co decomposition/deposition
lyses (Table 1 and Table S7, ESI†). SEM-EDS measurements of was observed (Table S7, ESI†). Longer time CPE measurements
the electrodes post electrolysis show Co-based particles deposited were also conducted in 11.0 M H2O at 1.95 V vs. Fc+/0, and
on the surface with a weight percent of 0.52% of total atoms similar catalyst activity and Faradaic efficiencies for CO were
measured (Fig. S15, ESI†). ICP-MS measurements were used to observed after 60 min and 90 min electrolysis when compared
better quantify the amount of Co deposited and indicate the to the results from the 30 min electrolyses (Table S10, ESI†).
amount of Co on the electrode surface corresponds to B8.5% of However, significant loss of CO2 activity and increased catalyst
initial Co complex loading in the electrolysis solution (Table S8, decomposition was observed after 120 min CPE in both 11.0 M
ESI†). These measurements suggest that the catalyst decomposes H2O and 5.5 M TFE (Table S10 and Fig. S28, S29, ESI†). These
during electrolysis in the absence of added proton source. results indicate that the stability of the catalyst under CO2
Upon addition of a proton source, significantly more charge reduction conditions is both potential and time dependent.
was passed in 30 min CPE and the Faradaic efficiency for CO The results of our CPE experiments suggest that adding a
increased while the Faradaic efficiency for HCOOH decreased. proton source not only accelerates the catalysis but also
A summary of the product distribution after 30 min CPE at increases the catalyst’s stability. We propose a mechanism to
1.95 V vs. Fc+/0 for selected concentrations of H2O is shown in explain our results in Scheme 2. Here, [Co(L–L)]2+ is reduced
Table 1 (and all concentrations in Table S5, ESI†). At the sequentially by 2 e to form [Co(L–L)], the active species for CO2
highest concentration of H2O investigated (11.0 M), CO2 was
reduced to CO with 80.5% efficiency at 1.95 V vs. Fc+/0.
Similarly, CPE conducted in 5.50 M TFE showed selective
CO2 reduction to CO with 87.7% Faradaic efficiency at 1.85 V
vs. Fc+/0. In both cases, the TOF values for CO increase dramatically.
During the electrolysis, the solution remained transparent and
the UV-vis spectra of the solution before and after 30 min CPE
remained nearly identical (Fig. S30, ESI†). Both SEM-EDS and
ICP-MS measurements of the electrode surfaces post-electrolysis
showed less catalyst decomposition to adsorbed solid-state
Scheme 2 Proposed catalytic mechanism of CO2 reduction with [Co(L–L)].
materials (Table 1 and Fig. S17, S18, Table S8, ESI†). When
In the presence of an added proton source, the mechanism proceeds via
the CPE experiments were conducted at 0.2 V more negative path (a) forming CO as the final C-containing product. In the absence of
applied potentials (to manually account for IR drop due to added proton source, the mechanism proceeds via path (b) to catalyst
B60 O solution resistance in our electrolysis cell), higher decomposition with stoichiometric HCOO production.

This journal is © The Royal Society of Chemistry 2018 Chem. Commun., 2018, 54, 1579--1582 | 1581
View Article Online

Communication ChemComm

R. H. Morris, C. H. F. Peden, A. R. Portis, S. W. Ragsdale,


T. B. Rauchfuss, J. N. H. Reek, L. C. Seefeldt, R. K. Thauer and
G. L. Waldrop, Chem. Rev., 2013, 113, 6621–6658.
2 S. Berardi, S. Drouet, L. Francas, C. Gimbert-Surinach, M. Guttentag,
C. Richmond, T. Stoll and A. Llobet, Chem. Soc. Rev., 2014, 43, 7501–7519.
3 J. Qiao, Y. Liu, F. Hong and J. Zhang, Chem. Soc. Rev., 2014, 43,
631–675.
Scheme 3 Proposed mechanism of the catalyst decomposition.
4 Y. Hori, H. Wakebe, T. Tsukamoto and O. Koga, Electrochim. Acta,
Published on 17 January 2018. Downloaded by Indian Institute of Technology Kanpur on 12/11/2023 7:34:43 PM.

1994, 39, 1833–1839.


5 K. P. Kuhl, E. R. Cave, D. N. Abram and T. F. Jaramillo, Energy
reduction. [Co(L–L)] coordinates CO2 to form a [Co(L–L)CO2] Environ. Sci., 2012, 5, 7050–7059.
adduct. At high proton concentration, [Co(L–L)CO2] undergoes 6 E. E. Benson, C. P. Kubiak, A. J. Sathrum and J. M. Smieja, Chem.
Soc. Rev., 2009, 38, 89–99.
a 1 e reduction and 2 H+ protonation by the proton source to 7 H. Takeda, C. Cometto, O. Ishitani and M. Robert, ACS Catal., 2017,
generate CO and H2O and regenerate the [Co(L–L)]+ complex 7, 70–88.
(Scheme 2, path (a)). However, at low proton concentration, the 8 M. L. Clark, K. A. Grice, C. E. Moore, A. L. Rheingold and C. P.
Kubiak, Chem. Sci., 2014, 5, 1894–1900.
[Co(L–L)CO2] abstracts a proton from the ligand to generate 9 C. W. Machan, S. A. Chabolla and C. P. Kubiak, Organometallics,
HCOO , leading to decomposition of the complex (Scheme 2, 2015, 34, 4678–4683.
path (b) and Scheme 3). This latter step is supported by recent 10 E. E. Benson, M. D. Sampson, K. A. Grice, J. M. Smieja, J. D.
Froehlich, D. Friebel, J. A. Keith, E. A. Carter, A. Nilsson and
reports of catalyst inactivation and decomposition of Re, Mo and C. P. Kubiak, Angew. Chem., Int. Ed., 2013, 52, 4841–4844.
Mn complexes with pyridylimine ligands due to the formation of 11 Z. Chen, C. Chen, D. R. Weinberg, P. Kang, J. J. Concepcion,
catalyst-CO2 adducts involving CO2–ligand bonding.9,20,21,30 Note D. P. Harrison, M. S. Brookhart and T. J. Meyer, Chem. Commun.,
2011, 47, 12607–12609.
that other mechanisms might explain the observed catalyst 12 T. C. Simpson and R. R. Durand, Electrochim. Acta, 1988, 33, 581–583.
decomposition at low proton concentration including a purely 13 E. Fujita and J. T. Muckerman, Inorg. Chem., 2004, 43, 7636–7647.
electrochemical degradation pathway at the electrode surface. 14 D. C. Lacy, C. C. L. McCrory and J. C. Peters, Inorg. Chem., 2014, 53,
4980–4988.
In conclusion, we have prepared a Co complex with a redox- 15 J. M. Smieja and C. P. Kubiak, Inorg. Chem., 2010, 49, 9283–9289.
active bis(pyridylmonoimine) ligand L–L and studied its CO2 16 J. A. Keith, K. A. Grice, C. P. Kubiak and E. A. Carter, J. Am. Chem.
reduction activity in CH3CN. We have shown that adding a Soc., 2013, 135, 15823–15829.
17 C. M. Lieber and N. S. Lewis, J. Am. Chem. Soc., 1984, 106, 5033–5034.
proton source not only accelerates the electrocatalytic rate of CO2 18 A. H. A. Tinnemans, T. P. M. Koster, D. H. M. W. Thewissen and
reduction by [Co(L–L)Br2], but also improves the stability of the A. Mackor, Recl. Trav. Chim. Pays-Bas, 1984, 103, 288–295.
catalyst by promoting intermolecular protonation of the CO2-catalyst 19 L. Chen, Z. Guo, X.-G. Wei, C. Gallenkamp, J. Bonin, E. Anxolabéhère-
Mallart, K.-C. Lau, T.-C. Lau and M. Robert, J. Am. Chem. Soc., 2015, 137,
intermediate and preventing intramolecular H+-abstraction. The 10918–10921.
[Co(L–L)Br2] system reduces CO2 selectively to CO in the presence 20 D. Sieh, D. C. Lacy, J. C. Peters and C. P. Kubiak, Chem. – Eur. J.,
of added proton source with high TOF at less negative potentials 2015, 21, 8497–8503.
21 D. Sieh and C. P. Kubiak, Chem. – Eur. J., 2016, 22, 10638–10650.
compared to previously reported Co-based catalysts. 22 K. T. Ngo, M. McKinnon, B. Mahanti, R. Narayanan, D. C. Grills,
This work was supported by the University of Michigan and M. Z. Ertem and J. Rochford, J. Am. Chem. Soc., 2017, 139, 2604–2618.
NSF grant CHE-0840456 for X-ray instrumentation. We thank 23 N. P. Liyanage, H. A. Dulaney, A. J. Huckaba, J. W. Jurss and
J. H. Delcamp, Inorg. Chem., 2016, 55, 6085–6094.
Dr James Windak for ICP-MS assistance, Dr Jeff Kampf for 24 M. D. Sampson, A. D. Nguyen, K. A. Grice, C. E. Moore, A. L. Rheingold
crystallographic assistance, and Prof. Songsong Bao for assistance and C. P. Kubiak, J. Am. Chem. Soc., 2014, 136, 5460–5471.
with interpreting crystallographic data. 25 B. t. Limoges and J.-M. Savéant, J. Electroanal. Chem., 2004, 562,
43–52.
26 C. C. L. McCrory, N. K. Szymczak and J. C. Peters, Electrocatalysis,
2016, 7, 87–96.
Conflicts of interest 27 C. Costentin, M. Robert and J.-M. Saveant, Chem. Soc. Rev., 2013, 42,
2423–2436.
There are no conflicts to declare. 28 C. Costentin, S. Drouet, M. Robert and J.-M. Savéant, Science, 2012,
338, 90–94.
Notes and references 29 N. Kaeffer, A. Morozan, J. Fize, E. Martinez, L. Guetaz and V. Artero,
ACS Catal., 2016, 6, 3727–3737.
1 A. M. Appel, J. E. Bercaw, A. B. Bocarsly, H. Dobbek, D. L. DuBois, 30 R. Stichauer, A. Helmers, J. Bremer, M. Rohdenburg, A. Wark,
M. Dupuis, J. G. Ferry, E. Fujita, R. Hille, P. J. A. Kenis, C. A. Kerfeld, E. Lork and M. Vogt, Organometallics, 2017, 36, 839–848.

1582 | Chem. Commun., 2018, 54, 1579--1582 This journal is © The Royal Society of Chemistry 2018

You might also like