You are on page 1of 8

Articles

https://doi.org/10.1038/s41929-018-0169-3

Detection of catalytic intermediates at an


electrode surface during carbon dioxide reduction
by an earth-abundant catalyst
Gaia Neri   1, James. J. Walsh   1,4, Gilberto Teobaldi1,2,5, Paul M. Donaldson   3* and Alexander J. Cowan1*

The electrocatalytic reduction of CO2 offers a sustainable route to the many carbon fuels and feedstocks that society relies on.
[fac-Mn(bpy)(CO)3Br] (bpy, 2,2-bipyridine) is one of the most promising and intensely studied CO2 reduction electrocatalysts.
However, the catalytic mechanism remains experimentally unproven and many key intermediates of the prototypical catalyst
have not been observed. Here we report the use of vibrational sum-frequency generation spectroscopy to study the catalytic
intermediates during CO2 reduction in situ at the electrode surface. We explore the complex applied-potential and acid-depen-
dent mechanistic pathways and provide evidence of the theoretically derived mechanisms. Demonstrating the ability to detect
the key species that are only transiently present at the electrode surface is important as the need for an improved mechanistic
understanding is a common theme throughout the field of molecular electrocatalysis.

I
mproved molecular electrocatalysts are key to unlocking the on the binding modes12,13 and vibrational relaxation kinetics14,15 of
potential of new clean-energy technologies. Intense efforts are surface-immobilized complexes, the majority of studies were carried
being expended worldwide to develop electrocatalysts for a range out in the absence of an applied potential. Indeed, prior to the work
of reactions, which include O2 reduction, CO2 reduction and H2 reported here we are only aware of one previous VSFG study of a
evolution, with potential impacts across sustainable fuel generation molecular electrocatalyst under potentiostatic control in which we16
and utilization, and electrochemical energy storage1–3. Therefore, it examined the behaviour of [Mo(bpy)(CO)4] (bpy, 2,2-bipyridine) at
is important to identify new tools that can provide the mechanistic Au and Pt electrodes. We were able to offer insight into the work of
insights required to enable these catalyst development programmes. Tory et al., who first demonstrated the critical role of the Au surface
A critical problem is that many spectroelectrochemical (SEC) for these catalysts17; however, we were unable to identify any catalytic
experiments require the build-up of a sufficient concentration of the intermediates that formed after the reduction of the catalyst pre-
electrochemically generated species in the bulk solution to enable cursor. Nonetheless, this study did indicate that sufficient ordering
detection4. In reality, intermediates during molecular electrocataly- occurs within the EDL at certain applied potentials, which enabled
sis may only be present transiently at the electrode surface or within measurements of the VSFG spectrum of the complex at the electrode.
the electric double layer (EDL). Therefore, although the identifi- Since it's initial report in 2011, [fac-Mn(bpy)(CO)3X]n (X =​  Br− or
cation of short-lived electrochemical intermediates at interfaces solvent (1), n =​  0 or +​1 (Fig. 1)), has become one of the most widely
is challenging, it remains an important goal. Vibrational sum-fre- studied electrocatalysts for the reduction of CO2 to CO (ref. 18).
quency generation (VSFG) spectroscopy is an inherently interface Despite advances in electrocatalytic activity19–24, the full mechanism
specific spectroscopy5. The theory behind VSFG spectroscopy is of CO2 reduction for this class of catalysts remains experimentally
reviewed in detail elsewhere6,7. Briefly, during a broadband VSFG unproven. It is known25 that the reduction of two equivalents of
experiment, short (femtoseconds) infrared and visible (picoseconds) 1 leads to the formation of [(Mn(bpy)(CO)3)2] (2) at −​1.12  VSCE
laser pulses are overlapped in time and space at the interface of (SCE, saturated calomel electrode) at room temperature. At poten-
interest and light generated at the sum of the frequency of the tials negative of −​1.50  VSCE, 2 is reduced to [Mn(bpy)(CO)3]–
infrared and visible pulses is detected. Due to interference effects, (3 (Fig. 1))18,25, which is proposed to be the primary active catalyst
typically only light from interfacial regions is detected because light for CO2 reduction. An alternative pathway in which CO2 directly
from the bulk is totally extinguished. A large resonant enhancement interacts with 2 has also been reported when 4,4′​-alkyl substituted
in the sum frequency light is measured when the infrared frequency bipyridine ligands are used26. At low temperatures27 or when bulky20
matches an interfacial vibrational mode. ligands are used, 3 can also form after the reduction of [Mn(bpy)
The inherent interfacial selectivity of VSFG spectroscopy has (CO)3]∙ at the first cathodic wave of the parent complex by an ECE
been exploited to study a wide range of electrolyte/electrode inter- mechanism. In the absence of CO2, 3 reacts with proton sources to
faces and provided important information on the packing and form [Mn(bpy)(CO)3H] (4), with H2 then formed28. In the presence
structure of the EDL and on catalytic processes during heteroge- of both CO2 and a Brønsted acid, CO evolution occurs with excel-
neous electrocatalysis8–11. In contrast, despite intense interest in the lent selectivity and minimal H2 production. Intriguingly, without
field, homogeneous electrocatalysis remains largely unexplored by the acid, CO2 reduction does not occur29 in contrast to analogous
VSFG spectroscopy. Although VSFG has provided important details rhenium complexes30. State-of-the-art density functional theory

Department of Chemistry and Stephenson Institute for Renewable Energy, University of Liverpool, Liverpool, UK. 2Beijing Computational Science Research
1

Center, Beijing, China. 3Central Laser Facility, STFC Rutherford Appleton Laboratory, Didcot, Oxfordshire, UK. 4Present address: School of Chemical
Sciences and National Centre for Sensor Research, Dublin City University, Dublin , Ireland. 5Present address: Scientific Computing Departmet, STFC
Daresbury Laboratory, Daresbury, UK. *e-mail: paul.donaldson@stfc.ac.uk; acowan@liverpool.ac.uk

Nature Catalysis | www.nature.com/natcatal


Articles NaTure CaTalysis

CO2
N CO
+ CO2 Mn
H N CO
N CO CO + H+
4 1,991, 1,892, Mn
1,888 (sh) cm–1 N CO
CO
+ e–
10
2- O OH
+ H+ CO2
N CO N CO 6
Mn Mn
N CO N CO
CO CO

Solv N N OC CO N CO
N CO + e– Mn
Mn + e– H2 O
OC Mn Mn CO N CO 3 + e– + H+
N CO CO
CO -solv + H+
OC CO N N 1,911
1,811 cm–1
0.5 O OH O
2,045, 1,965, 1,973, 1,927,
1 2 7 N CO
1,953 cm–1 1,880, 1,849 cm–1 Mn N CO
N CO Mn
CO N CO
CO
+ e– + e–
+ e– 8
CO 2,118, 2,044,
O + H+
CO 2,023, 1,982 cm–1
N
Mn
N CO CO
CO N H2 O
Mn
N CO
CO + e–

Fig. 1 | Proposed electrocatalytic pathways for the reduction of CO2 by 1. The scheme shows the proposed protonation-first (6–8–9) and reduction-
first (6–7–9) mechanisms of CO2 reduction by [Mn(bpy)(CO)3(solv)]+ (1) (solvent, CH3CN or, when present, H2O) in the presence of a Brønsted acid.
The Brønsted acid strength and applied potential are thought to control the relative contributions from each pathway. The catalytic cycle is based on
theoretical calculations reported elsewhere31–33. ν(CO) infrared wavenumbers are for solution species that have either been synthesized and found to be
stable (122 and 840) or have been observed in past SEC studies (234, 328 and 428). The dashed lines indicate an alternative pathway for the formation of 3,
which has been determined to occur at low temperatures and also when bpy derivatives with bulky substituents are employed20,27.

(DFT) simulations have explored the critical role of the acid and Here we report a VSFG study of the very active CO2 reduction 1
identified a protonation-assisted binding mechanism of CO2 and under potentiostatic control. Our in situ VSFG study provides
the catalytic pathways on which Fig. 1 is based31–33. CO2 binding to 3 experimental observation of the proposed switch in pathway
and the formation of [Mn(bpy)(CO)3(CO2)]– (5) is calculated to be with 1. Critically, the observation of key intermediates not seen by
endergonic in the absence of a suitable acid. However rapid proton- bulk SEC methods demonstrates the wider value of VSFG for the
ation of 5 to form the more stable product [Mn(bpy)(CO)3(CO2H)] study of complex molecular electrocatalytic mechanisms.
(6) can occur32,33. Once formed, 6 is thought to be able to react via
two pathways31. The reduction-first pathway to form [Mn(bpy) Results
(CO)3(CO2H)]– (7), which is followed by protonation and water Cyclic voltammetry of 1 within the SEC cell. Cyclic voltammo-
expulsion to yield [Mn(bpy)(CO)4] (9). The reduction of 9 then grams (CVs) of 1 in CH3CN recorded within the SEC cell using an
leads to CO loss, regeneration of 3 and completion of the catalytic Au–Hg amalgam working electrode are shown in Fig. 2. Au–Hg
cycle. Alternatively, 6 can be protonated (protonation-first pathway) provides both reflectivity in the visible region, a requirement for
to yield [Mn(bpy)(CO)4]+ (8), which can then be reduced to form 9 our VSFG experiment, and a suitable electrochemical window for
prior to CO evolution. use in the presence of organic acids such as trifluoroethanol (TFE).
The reduction potential for 6 (calculated, as −​1.7  VSCE) coupled On Au–Hg, the behaviour of 1 is similar to that in previous reports
to the calculated activation barrier for its protonation of between 12 that used glassy carbon electrodes, with reductions of 1 and 2 at
and 22 kcal mol−1, depending on the acid source, has led to the pre- −​0.92 and −​1.27  VAg, respectively, which correspond to approxi-
diction that the mechanism and turnover frequency of the catalysis mately −​1.12  VSCE and −​1.47 VSCE (ref. 18). In line with previous stud-
by 1 is both potential and acid dependent32,33. In strong acids, the ies, an increased current density due to catalytic CO2 reduction only
protonation-first pathway is predicted to occur at lower overpoten- occurred when both CO2 and an acid source were present19. The
tials, whereas negative of −​1.7  VSCE the reduction-first pathway domi- similarity of the reduction potentials18 to those previously reported
nates31. Solution-based microkinetic simulations32 have shown a good coupled to the same catalytic behaviour suggests a common CO2
correlation between the observed catalytic turnover frequencies and reduction mechanism for both glassy carbon and Au–Hg electrodes.
applied potential. However, the key catalytic intermediates (6–10)
and the proposed potential-dependent switch in mechanism have not In situ VSFG spectra in the absence of CO2. The VSFG spectra
been experimentally observed with the prototypical complex 119,28,34. of 1 recorded during two successive CV measurements (10 mV s−1)

Nature Catalysis | www.nature.com/natcatal


NaTure CaTalysis Articles
(Fig. 3c, Supplementary Table 1 and Supplementary Figs. 2–4). As
0 the potential of the electrode was made more negative, there was
an increase in the intensity of the VSFG bands of 1, which indicates
either an accumulation or increased ordering of this complex within
–10 1/CO2
the EDL (discussed further in Supplementary Information). As the
1/ TFE/Ar
potential of the electrode was varied between 0.10 and −​0.75  VAg, we
i (μA)

1/ TFE/CO2
–20 observed a shift of the ν(CO) band of 1 of ~35 cm V−1. The poten-
tial dependence of vibrational modes is commonly interpreted to be
–30 due to Stark shifting effects that occur when a large electric field is
present36. However, it was also highlighted recently that the appar-
–40 ent potential-dependent changes in the second-order spectral line
shapes can occur due to the changing contributions of dispersive
–1.6 –1.4 –1.2 –1.0 –0.8 –0.6 –0.4 –0.2 0.0 and adsorptive third-order components that arise within the EDL37.
Potential (VAg) Both interpretations require 1 to experience a large electric field,
which confirms that during the VSFG experiment we are probing
Fig. 2 | CVs of 1 under argon and CO2 in the presence or absence of TFE. the vibrational spectra of 1 either at the electrode surface or within
In the presence of both an acid source (TFE) and CO2, electrocatalytic the EDL and that the spectral features are not due to ‘phantom tran-
CO2 reduction occurs, as indicated by the sharp increase in current sitions’ that arise from absorptive losses in the infrared spectra from
(i). CVs recorded within the SEC cell at 50 mV s−1 in CH3CN and 0.1 M the bulk solution8.
TBAPF6 under the conditions indicated (when present, TFE is at 1.5 M). At potentials negative of −​0.75  VAg, the ν(CO) band of 1 started
Potentials are reported versus the Ag/Ag+ pseudo reference electrode to decrease in intensity and a new intense band at 1,974 cm−1
used in the SEC cell unless otherwise stated. The deviation from the peak (−​0.76  VAg) formed (Fig. 3b,c). This VSFG band matches a known
shapes typically observed for freely diffusing species is due to the short ν(CO) band of [(Mn(bpy)(CO)3)2], 2, and the VSFG mode shifted
distance between the electrode surface and the cell front window (50 μ​m) to 1,959 cm−1 at −​1.25  VAg. The very strong VSFG response can be
(Supplementary Fig. 1). readily explained by the overlap of the ultraviolet/visible absorption
spectrum of 2 (λmax 394, 461, 633 and 806 nm)18 with the 800 nm
visible laser, which gives rise to an additional electronic resonance
in acetonitrile with TFE (1.5 M) under argon are shown in Fig. 3. enhancement. Under argon, after the reduction of 2 (<​−​1.3  VAg) we
Under these conditions, the current increase negative of −​1.3  VAg found weak VSFG bands at 1,994 and 1,898 cm−1 at about −​1.5  VAg
was assigned to H2 evolution. To enable the identification of res- on the outward sweep and at about −​1.6 VAg on the return sweep of
onant VSFG bands, the non-resonant response of the Au–Hg/ the CV (Fig. 3c). Careful inspection of these spectra (Supplementary
electrolyte interface was suppressed by the introduction of a Fig. 5) showed that both modes were present at both potentials,
0.9 ps delay between the femtosecond broadband infrared pulse although with different relative intensities, which suggests that they
and a time asymmetric picosecond visible (800 nm) laser pulse35. are from a common intermediate undergoing an orientation change
At open circuit, we found no clear VSFG signals, but as soon at the electrode. As we observed these bands at potentials close to
as a potential of 0.1 VAg was applied a ν(CO) mode at 2,043 cm−1 those at which catalytic proton reduction occurs, we assigned them
appeared, in addition to broad weak VSFG signals at ~1,950 cm−1 to the hydride intermediate 4, and we note a good agreement to
(Supplementary Fig. 2). These bands were readily assignable similar Mn–H intermediates reported elsewhere (Supplementary
to [Mn(bpy)(CO)3(CH3CN)]+ (1) through a comparison to the Table 1)24,28. The observation of 4 but not 3 indicates that the forma-
Fourier transform infrared spectrum of the starting solution tion of the hydride intermediate (4) is facile and rapid, which makes

ISFG ISFG
–10 658 1,327 1,995 2,663 3,332 4,000 –5 18 40 63 85 108 130
0.1 1
a b c
–0.9 2

–1.9 4

2
Potential (VAg)

–0.9
1
0.1
2
–0.9

–1.9 4

–0.9 2
1
0.1
0

0
15

10

05

00

95

90

85

80

10

05

00

95

90

85

80

75
2,

2,

2,

2,

1,

1,

1,

1,

2,

2,

2,

1,

1,

1,

1,

1,

i (μA) Wavenumber (cm–1)

Fig. 3 | VSFG spectra of 1 recorded during two successive CVs under Ar. a, Current–voltage response recorded during the experiment at a scan rate
of 10 mV s−1 (1 mM solution of 1 in CH3CN (0.1 M TBAPF6) and 1.5 M TFE under Ar). b, VSFG contour plot, recorded with the infrared laser centred at
1,900 cm−1, ppp polarization, which is dominated by 2 as described in the text. c, Replotting the VSFG data over a more limited SFG intensity range (ISFG)
shows the interconversion of species 1 to 4 in Fig. 1. The formation of new VSFG bands occurs at potentials at which we measure sharp changes in i from
our working electrode, which demonstrates that we are following the electrochemical processes that is occurring.

Nature Catalysis | www.nature.com/natcatal


Articles NaTure CaTalysis
a ISFG b c pKa(CH3CN)
420 365 310 255 200 145 90 35 –20 540 470 400 330 260 190 120 50 –20 30.5 26.4 21.6 NA
0.10 1.0 20

0.8
–0.75 15

icat (–1.35 V) (µA)


Potential (VAg)
0.6

ISFG
–1.60 10

0.4

–0.75 5
0.2

0.10 0.0 0
2,000 1,900 1,800 1,700 1,600 2,000 1,900 1,800 1,700 1,600 MeOH TFE PhOH None*
Wavenumber (cm–1) Acid source

Fig. 4 | VSFG spectra under CO2 and TFE that show new bands assigned to catalytic intermediates. a,b, VSFG spectra recorded during CVs (10 mV s−1)
of 1 (1 mM) in CH3CN (0.1 M TBAPF6) and 1.5 M TFE under Ar (a) and under CO2 (b). Under CO2, new VSFG bands (highlighted with white arrows)
are assigned to catalytic intermediate(s) during the CO2 reduction. c, The maximum intensity of the SFG band (ISFG) of the CO2 reduction intermediate
(1,976 cm−1) with different acid sources. There is a direct correlation between i during CO2 reduction at −​1.35 VAg during the VSFG experiment
(Supplementary Fig. 6 gives the CVs) and the intensity of the new VSFG band. pKa values are from Keith et al.51, and, to enable a direct comparison of
different VSFG experiments where alignment may be slightly different, the ISFG at 1,976 cm−1 is normalized with respect to the measured intensity of
the SFG mode of 2 at −​0.2 VAg. *Although these experiments were carried out without the explicit addition of an acid source, the nature of the SEC cell
prevents rigorous anhydrous conditions.

a 500 observed in the potential region in which the catalytic reduction


400 of CO2 to CO occurs (−​1.1 to −​1.6  VAg (Fig. 4b)). Experiments in
the absence of a deliberately added acid showed minimal current
300
enhancement under CO2, in line with previous studies18, and the
ISFG

200 new VSFG bands were not observed (Supplementary Fig. 8). VSFG
100 experiments over the same potential window in the presence of
0
TFE and CO2 without the catalyst (1) also showed no clear VSFG
b signals. Therefore, we assigned the indicated VSFG bands at 1,976,
800 1,875 and 1,600 cm−1 in Fig. 4(b) to CO2 reduction intermediates,
probably one (or more) of complexes 6–9 (Fig. 1). As hydride (4)
600
formation was readily detectable in the same system under Ar,
ISFG

400 it is clear that the interaction of CO2 with 3 must also be rapid.
200 Furthermore, the binding constant (K CO2 /H +) for 3 and CO2 in the
presence of an acid (here TFE) must strongly favour the interac-
0
tion with CO2(ref. 28), in line with previous observations on related
2,100 2,050 2,000 1,950 1,900 1,850 1,800
Mn complexes with bulky ligands38. The observation of the effective
Wavenumber (cm–1)
suppression of hydride formation by CO2 through the competitive
consumption of 3 provides a rationale for the excellent selectivity
Fig. 5 | VSFG spectra that show the potential dependence of the new towards CO production19.
CO2 reduction intermediate ν(CO) mode. Spectra recorded during CVs We now turn to the assignment of the new VSFG bands present
of 1 (1 mM) in CH3CN (0.1 M TBAPF6) and 1.5 M TFE under CO2 with the during catalytic CO2 reduction in the presence of TFE at ~1,976,
infrared laser centred at 1,900 cm−1. a, During the outward CV sweep (–1.18 1,875 and 1,600 cm−1 (Fig. 4b). The strongest new VSFG band
to –1.31 V), the ν(CO) band assigned to 2 (~1,962 cm−1) initially decreases (~1,976 cm−1) is present on the return potential sweep between
as the new band at 1,976 cm−1 grows in intensity. As the potential sweeps −​1.40 and −​1.05  VAg with the highest intensity being between –1.20
more negative, the 1,976 cm−1 band decrease as the 1,875 cm−1 feature and –1.10 VAg (Fig. 5b). The same feature is also briefly present at
grows in intensity. b, On the return CV sweep (–1.15 to –1.00 V), the band of a lower intensity of 212 on the outward sweep between –1.20 and
the intermediate at 1,976 cm−1 decreases in intensity as 2 is reformed. –1.35 VAg, which peaks at –1.25 VAg (Figs. 4b and 5a). Concomitant
with the formation of the band at 1,976 cm−1, which is a ν(CO)
mode of a metal carbonyl complex (see below), we observed the
the selectivity of this catalyst towards CO2 in the presence of high formation of a band at 1,600 cm−1 (Fig. 4b and Supplementary
concentrations of acid remarkable19. Fig. 9). This spectral region is where both bipyridine ring modes
and ν(OCO) modes occur. Theory predicts that either [Mn(bpy)
In situ evidence for the protonation first pathway under CO2. (CO)3]- (3)32 or [Mn(bpy)(CO)3(CO2H)] (6)33 will accumulate at
VSFG spectra of 1 under CO2 in CH3CN and TFE show no evidence moderate applied potentials (when the protonation-first pathway
for the formation of 4 at 1,994 and 1,898 cm−1 at any applied poten- dominates), but either 6 or 7 may be present when the reduction
tial (Fig. 4 and Supplementary Fig. 7). Instead, under CO2 multiple first–pathway dominates32,33. The ν(CO) band at 1,976 cm−1 cannot
new, strong VSFG bands at 1,976, 1,600 cm−1 and 1,875 cm−1 were be assigned to 3 on the basis of the known spectrum of this complex

Nature Catalysis | www.nature.com/natcatal


NaTure CaTalysis Articles
5 band arises from the bipyridine ligand of the intermediate and not
12 16
from a ν(OCO) mode (a full discussion of this assignment accom-
C O2
4 panies Supplementary Figs. 9 and 10).
Interestingly, a shift in the ν(CO) mode is observed from
13
C18O2

3
1,976 cm−1 (12C16O2) to 1,942 cm−1 (13C18O2) (Fig. 6). An isotopic
ISFG (a.u.)

shift of ~34 cm−1 is well below that expected for CO bound directly


to the electrode surface, for which values of ~100 cm−1 are typical,
2
which allowed us to rule out this assignment39. Instead, we propose
that the isotopic shift arises from either vibronic coupling between
1 the Mn–13C18O2H and Mn–CO groups (feasible with 6 and 7) or
from the presence of the Mn–13C18O group (that is 8 and 9). DFT
0 calculations of 6–9 using a Hg(100) 6 ×​ 6 cell to approximately
account for the role of the Au/Hg electrode surface were carried out
2,000 1,900 1,800 1,700 1,600
(computational details are given in the Supplementary Information,
Wavenumber (cm–1)
Supplementary Fig. 14 and Supplementary Tables 2 and 3). The
computed vibrational wavenumbers are shown in Fig. 7 and listed
Fig. 6 | VSFG spectra of the CO2 reduction intermediate recorded using
in Supplementary Tables 4–22. Although 8 has not been observed
labelled CO2VSFG spectra that show the potential dependence of the new
previously in situ during catalysis, it has been synthesized40 and
CO2 reduction intermediate ν(CO) mode. When using 13C18O2, the band
characterized (ν(CO), 2,118, 2,044, 2,023 and 1,982 cm−1), and the
in the ν(CO) region is isotopically shifted. Spectra recorded at −​1.25 VAg in
unscaled computed values for 8 both in vacuo (2,106, 2,034, 2,030
CH3CN (0.1 M TBAPF6) and 1.5 M TFE under 12C16O2 and 13C18O2. Solid lines
and 2005 cm−1) and at the model Hg surface (2,100, 2,026, 2,021 and
are the result of a multi-Lorentzian fit. a.u., arbitrary units.
1,997 cm−1) are in good agreement with the experimental results
in the solvent phase. Complexes 6–9 all have vibrational modes to
(Supplementary Table 1). Experiments with isotopically labelled which the 1,976 cm−1 feature could be assigned. However, on iso-
C O2 also showed no shift of the 1,600 cm−1 VSFG band (Fig. 6).
13 18
topic substitution of either the Mn–CO2H group (6 and 7) or a
This alone does not immediately rule out the assignment to 6 or 7. Mn–CO axial group (8 and 9) with 13C18O, only the axially labelled
Indeed, previous studies on Mn carbonyl complexes with bulky complexes 8 (47–53 cm−1) and 9 (41–52 cm−1) show shifts compa-
ligands report the build up of a complex assigned to a Mn(i) com- rable with the VSFG measured value of ~34 cm−1. The lack of shift of
plex, potentially a Mn(i)–COOH intermediate analogous to 6, on the ν(CO) modes of 6 and 7 on substitution of the Mn–CO2H group
the basis of the ν(CO) modes without observation of the anticipated excludes the possibility that these intermediates may be responsible
ν(OCO) mode20,38. It does, however, suggest that the 1,600 cm−1 for the VSFG spectra in the 1,900–2,000 cm−1 region.

6 7 8 9
1.0 1.0 1.0 1.0
COeq 0.5 COeq 0.5 COeq 0.5 COeq 0.5
1.0 0.0 1.0 0.0 1.0 0.0 1.0 0.0
0.5 COax 0.5 COax 0.5 COax 0.5 COax
0.0 1.0 0.0 1.0 0.0 1.0 0.0 1.0
COOH 0.5 COOH 0.5 COd 0.5 COd 0.5
1.0 0.0 1.0 0.0 1.0 0.0 1.0 0.0
0.5 bpy 0.5 bpy 0.5 bpy 0.5 bpy
0.0 0.0 0.0 0.0
2,200 2,000 1,800 1,600 1,400 1,200 2,200 2,000 1,800 1,600 1,400 1,200 2,200 2,000 1,800 1,600 1,400 1,200 2,200 2,000 1,800 1,600 1,400 1,200
Wavelength (cm–1) Wavelength (cm–1) Wavelength (cm–1) Wavelength (cm–1)

6* 7* 8* (13C18O down) 9* (13C18O down)


1.0 1.0 1.0 1.0
COeq 0.5 COeq 0.5 COeq 0.5 COeq 0.5
1.0 0.0 1.0 0.0 1.0 0.0 1.0 0.0
0.5 COax 0.5 COax 0.5 COax 0.5 COax
0.0 1.0 0.0 1.0 0.0 1.0 0.0 1.0
COOH 0.5 COOH 0.5 COd 0.5 COd 0.5
1.0 0.0 1.0 0.0 1.0 0.0 1.0 0.0
0.5 bpy 0.5 bpy 0.5 bpy 0.5 bpy
0.0 0.0 0.0 0.0
2,200 2,000 1,800 1,600 1,400 1,200 2,200 2,000 1,800 1,600 1,400 1,200 2,200 2,000 1,800 1,600 1,400 1,200 2,200 2,000 1,800 1,600 1,400 1,200
Wavelength (cm–1) Wavelength (cm–1) Wavelength (cm–1) Wavelength (cm–1)

8* (13C18O up) 9* (13C18O up)


1.0 1.0
COeq 0.5 COeq 0.5
1.0 0.0 1.0 0.0
0.5 COax 0.5 COax
0.0 1.0 0.0 1.0
COu 0.5 COu 0.5
1.0 0.0 1.0 0.0
0.5 bpy 0.5 bpy
6 7 8 9
0.0 0.0
2,200 2,000 1,800 1,600 1,400 1,200 2,200 2,000 1,800 1,600 1,400 1,200
Wavelength (cm–1) Wavelength (cm–1)

Fig. 7 | Fragment-resolved analysis of the contributions to the computed infrared modes (normalized eigenvectors) for 6–9. Analysis of the ν(CO)
region indicates that the intermediate is not 6 or 7; the experimentally determined presence of a bpy mode at 1,600 cm−1 (Fig. 6) is in alignment with an
assignment to 8. Calculations carried out on the energetically favoured geometries of the pristine and isotopically labelled (*) complexes 6–9 on the model
Hg(100) 6 ×​ 6 surface. COeq, equatorial CO; COax, axial CO; COd, axial CO pointing towards the surface (down geometry); COu, axial CO pointing away from
surface (up geometry). Hg(100), silver; H, white; C, cyan; O, red; N, blue; Mn, orange.

Nature Catalysis | www.nature.com/natcatal


Articles NaTure CaTalysis
a
v1 to complex 8 (Fig. 7 and Supplementary Table 4). It is also signifi-
2,120
v2 cant that we observe Stark shifting of the intermediate ν(CO) band
2,100 v3 (about +​17  cm−1 V−1, −​1.60 to −​1.35 V) in Fig. 4. Calculations of
v4
2,080 the vibrational modes of 8 and 9 in the presence of an external
vbpy
electric field also support the assignment to 8 (Fig. 8). Only 8 has
2,040 a CO mode that hardens on the reduction of the intensity of the
external electric field. Although 8 was not predicted to accumulate,
cm–1

2,020

2,000
previous solution-phase calculations indicated that 9 is thermally
unstable and that the reduction potential of 9 (about −​1.1  VSCE) is
1,620 significantly positive of 8 (about −​1.3  VSCE)31,33. Although caution is
1,600
required when the experimental data of molecules at an electrode
surface are compared to calculations in solution, it interesting that
1,580
here we found that the VSFG bands of the intermediate are stable for
–1.0 –0.8 –0.6 –0.4 –0.2 0.0 >​50 s (Supplementary Fig. 11) and that the intermediate builds in
External electrostatic field (109 V m–1) concentration and reaches a maximum between −​1.1 and −​1.2  VAg
(about −​1.3 to −​1.4  VSCE) (Fig. 4b). We also only observed the new
b
2,100 v1 intermediate VSFG bands in the presence of stronger acids where
v2 the protonation-first pathway occurs (see below). Therefore, on the
2,080 v3 balance of probability, we assign the 1,976 and 1,600 cm−1 VSFG
2,060
v4 bands to CO stretching and bipyridine modes of 8, an intermedi-
vbpy ate present during the protonation-first pathway (6–8–9). This rep-
2,000 resents experimental evidence that the protonation-first pathway
occurs with 1.
cm–1

1,980
The accumulation of 8 is unexpected because, although a Mn
1,960 tetracarbonyl cation was recently reported in catalysis studies of a
1,600 N-heterocyclic carbene Mn complex41, microkinetic studies of 1 in
1,580
solution predict the accumulation of 6 when the catalysis proceeds
via the protonation-first pathway32. However, DFT calculations
1,560 of the protonation of 6 to 8 in solution also report that the equi-
1,540
librium position can sit either towards 6 or 8 (Δ​G6–8 range from 
–1.0 –0.8 –0.6 –0.4 –0.2 0.0 +​18 to −​28 kcal mol−1) depending on the model employed31–33. It is
External electrostatic field (109 V m–1) therefore apparent that in real systems, for which the interactions
c of the complex with the electrode, solvent and surface electric field
1.0
all need to be considered, it is feasible that the interconversion of 6
COeq
to 8 may occur.
0.8
COax At potentials negative of −​1.7  VSCE (about −​1.5  VAg) it is antici-
COd pated that the contribution from the protonation pathway to the
catalytic current is minimal and that the reduction-first path-
0.6 way dominates in line with this we did not observe 8 at <​−​1.4  VAg
(Fig. 4)32,33. After the loss of the VSFG modes assigned to 8 on the
Pa (v3)

outward CV sweep, we saw the growth of a new broad, weak VSFG


0.4 mode at 1,875 cm−1, which is transiently present at about −​1.3  VAg
(Fig. 5a). Although a reasonable agreement exists with the calcu-
lated spectrum of 7, an intermediate on the reduction-first path-
0.2
way, the VSFG band is weak and not detected in isotopic labelling
experiments in which a lower concentration of CO2 is used, which
0.0
prevents assignment. Nonetheless, the loss of the VSFG spectral fea-
–1.0 –0.8 –0.6 –0.4 –0.2 0.0 tures of 8 at potentials at which the reduction of 6 becomes viable
External electrostatic field (109 V m–1) provides strong evidence for the anticipated switch in pathway from
protonation first to the reduction first at more negative potentials.
Fig. 8 | Computed Stark shifts of the vibrational modes of 8 and 9. We have examined the role of Brønsted acid strength on the
a,b, Calculations of isolated complexes 8a (a) and 9b (b) show that only mechanism of CO2 reduction using VSFG. Interestingly, the cata-
one mode, v3 (8), is compatible with the measured Stark shift (field-induced lytic current at −​1.35  VAg, a potential at which only a contribution
vibrational softening) in Fig. 4, as indicated by a green arrow. Molecules from the protonation-first pathway is expected, correlates with the
are oriented as per energy favoured adsorption geometry on the Hg surface intensity of the ν(CO) mode of 8 and the acid pKa (Fig. 4c). It has
and the external electrostatic field is along the direction perpendicular to been calculated that the protonation-first pathway requires stron-
the pristine Hg surface. For the sign convention used, the more negative ger acids (TFE, pKa(CH3CN) =​ 26.4; phenol, pKa(CH3CN) =​  21.6)
electric fields correspond to the accumulation of more negative charge to enable both the initial protonation-assisted binding of CO2
under the tripod formed by the Mn–bpy and Mn–COd bonds (Fig. 7). (3–5–6) and the protonation of 6 to form 832,33. Indeed, we found
c, Computed dependence of the CO contributions to v3 (8) (Pa(v3); very similar VSFG spectra with phenol, with 8 clearly present
Methods) on the external electrostatic field: the contributions from COd between about −​1.1 and −​1.4  VAg (Supplementary Fig. 12)32. In
(COeq) increase (decrease) with the intensity of the external electric field. contrast with methanol, a weaker acid (pKa(CH3CN) =​  30.5), we
observed no evidence of 8 by VSFG and a minimal catalytic cur-
Comparison between the calculated highest-energy bpy modes rent enhancement at −​1.35  VAg (Supplementary Figs. 6 and 13). It
of 8 (1,594 cm−1) and 9 (1,559 cm−1) and our observation of a VSFG was calculated that the protonation-first pathway is largely inactive
band of the intermediate at 1,600 cm−1 support the assignment using methanol, in line with this observation32.

Nature Catalysis | www.nature.com/natcatal


NaTure CaTalysis Articles
Clearly, to form 8 and access the protonation-first pathway, pulse and the femtosecond-derived time asymmetric visible (800 nm) pulse, as
either a suitably strong acid in solution or the use of ligands with previously described35. All the experiments were conducted with ppp polarization
and all the analysis was based on identifying band frequencies, rather than
a pendant group to orientate and enable the protonation by weaker assessing orientation changes from band intensity changes. For SEC experiments
acids, is required and promising results using such an approach were carried out at 10 mV s−1, the VSFG spectra were collected every 1 or 4 s and were a
recently reported21,28. The observed accumulation of 8 at the Au–Hg result of a 1 or 4 s averaging, respectively.
surface over a wide potential window indicates that developing ana-
Computational methods. All the DFT simulations were performed with the
logues of 8 with more positive reduction potentials may be a route
projector-augmented wave method43 as implemented in the Vienna ab initio
to a higher turnover frequency at lower overpotentials in systems in Simulation Package program44–46, the Perdew–Burke–Ernzerhof approximation
which protonation first can occur. However, this must be balanced to the exchange-correlation functional47, 400 eV plane wave energy cut-off, 0.1 eV
with the need to generate sufficiently nucleophilic metal centres for Gaussian smearing and Grimme D2 van der Waals corrections48. Following the
CO2 reduction to occur42. The results presented here represent a sig- original approach of the identical parameterization of van der Waals corrections
for 3d (Sc–Zn) and an 4d (Y–Cd) transition metals48, the parameters for Hg atoms,
nificant step forward in understanding the remarkable properties not included in the original parameterization48, were taken from Amft et al.49
of 1 for the reduction of CO2 to CO and more widely demonstrate (C6 =​ 40.62 J nm6 mol−1, R0 =​ 1.772 Å). A fully unconstrained spin polarization was
the importance of VSFG spectroscopy as a technique to deliver the used for those systems that contained an odd number of electrons (for example, 7
design rules required for more efficient molecular electrocatalysts. and 9), resulting in a total magnetic moment of 1 μ​B.
Following Wu et al.50, the electrode surface was approximated by a one-
layer rhombohedral Hg(100) 6 ×​ 6 slab (36 Hg atoms), which led to a hexagonal
Conclusions simulation cell with 20.1936 Å optimized (in-plane) lattice vectors. Complexes
We report the mechanisms of the very widely studied CO2 reduc- were placed on one side of the slab to ensure at least a 12 Å vacuum- buffer was
tion electrocatalyst 1 in the presence of a range of different Brønsted present between periodic replicas of the slab models (with adsorbates). The
acids. VSFG experiments allowed the detection of a species not seen slab two-dimensional Brillouin zone was sampled with a grid of 5 (Γ​ centred)
in analogous Fourier transform infrared–SEC studies, assigned symmetry irreducible k points.
For geometry optimization, all the atoms of the Hg slab and molecular
to [Mn(bpy)(CO)4]+ (8), an intermediate of the low overpotential adsorbates were fully relaxed until the atomic forces were lower than 0.02 eV Å−1.
protonation-first mechanism. DFT calculations and VSFG experi- Harmonic vibrational frequencies were calculated via the symmetric finite
ments in weak (methanol) and stronger (TFE and phenol) acids and displacements of ±​0.01 Å, maintaining the Hg atoms fixed. Isotopically labelled
isotopic labelling support the assignment of 8. Although our study systems were simulated by changing the atomic mass as needed in setting up the
supports the theory-derived mechanisms, the accumulation of 8 is mass-weighted Hessian matrix ahead of diagonalization.
Defining Q(j) as the (normalized) vector that represents the jth vibrational
unexpected and suggests that the protonation of a Mn–CO2H inter- mode (of wavenumber ν̂j) in a 3D Cartesian reference, the contribution Pa(j) to Q(j)
mediate (6) readily occurs using TFE and phenol at our electrode from the ath molecular fragment (a =​  COeq, COax, bpy, Mn and reacting COOH/
surface. Of significance is the demonstrated application of VSFG CO group) shown in Fig. 7 was computed by summing the square of the (three
to the study of molecular electrocatalytic mechanisms. We are Cartesian) displacements for the atoms in a:
now exploring the applicability of the approach to a wider range of 2
molecular electrocatalysts, electrodes and reactions with potential Pa(j) = ∑ [Qi(j)]
i ∈a
applications across energy storage and conversion.
For the definition used, vibrational modes fully localized on the ath fragment result
Methods in Pa(j) = 1 values. Conversely, 0 < Pa(j) < 1 values reveal partially delocalized modes,
Experimental methods. Bromopentacarbonylmanganese(i) (98%), 2,2-bipyridine that is, coupling between the vibrations of different molecular fragments (for
(99%), acetonitrile (anhydrous, 99.8%), 2,2,2-TFE (≥​99.5%), methanol (99.8%), example, different CO ligands).
tetrabutylammonium trifluorophosphate (TBAPF6, ≥​99.0%), phenol (≥​99.0%)
and 13C18O2 (99 at% 13C, 95 at% 18O) were purchased from Sigma Aldrich and Data availability
used as received. Ferrocene (Fc) (Sigma, ≥​98%) was further purified via Raw data for all figures within the paper are freely available from the University of
sublimation, and then dried under vacuum. Mn(bpy)(CO)3Br was synthesized Liverpool Research Data Catalogue at https://doi.org/10.17638/datacat.liverpool.
as previously reported22. ac.uk/533.
SEC experiments were carried out using a custom-made cell described
elsewhere16. The cell consists of a Teflon cross piece containing the working Received: 13 April 2018; Accepted: 21 September 2018;
electrode, a Pt counter electrode and an Ag/Ag+ pseudo-reference electrode. The
addition of Fc to the single-compartment SEC cell during the VSFG studies was
Published: xx xx xxxx
not possible as the optical properties of Fc/Fc+ (which changes in concentration
during the experiment due to the proximity of the counter electrode) complicate References
the data analysis. Therefore, the potentials above are reported versus the Ag/Ag+ 1. Dey, S. et al. Molecular electrocatalysts for the oxygen reduction reaction.
pseudo-reference electrode. Ex situ experiments (Supplementary Fig. 1) using the Nat. Rev. Chem. 1, 0098 (2017).
same working and reference electrodes in a standard three-electrode cell indicate 2. DuBois, D. L. Development of molecular electrocatalysts for energy storage.
that Fc/Fc+ ≈​ 0.60 V versus the silver pseudo-reference electrode, which allowed Inorg. Chem. 53, 3935–3960 (2014).
us to estimate the reduction potentials of 1 and 2 as −​1.12  V and −​1.47  V versus 3. Costentin, C., Robert, M. & Savéant, J.-M. Catalysis of the electrochemical
SCE. During the SEC experiments, the distance between the working electrode and reduction of carbon dioxide. Chem. Soc. Rev. 42, 2423–2436 (2013).
the CaF2 front window was ~50 μ​m, set by a Teflon spacer. The working electrode 4. Lee, K. J., Elgrishi, N., Kandemir, B. & Dempsey, J. L. Electrochemical
was a polycrystalline gold disc (Ø =​ 16 mm, IJCambria) that had been immersed in and spectroscopic methods for evaluating molecular electrocatalysts.
mercury (polarographic grade, Fisher) for 1 min before being left to dry in a fume- Nat. Rev. Chem. 1, 0039 (2017).
hood for at least 2 h. The Au–Hg electrode was then hand polished (suitable care 5. Shen, Y. R. Surface properties probed by second-harmonic and sum-
is required to avoid exposure to small Au–Hg particles) using decreasing sizes of frequency generation. Nature 337, 519–525 (1989).
polishing pastes (15 μ​m and 6 μ​m diamond paste, and then 0.5 μ​m alumina slurry) 6. Shen, Y. R. Basic theory of surface sum-frequency generation. J. Phys. Chem.
for 5 min each (the electrode was sonicated in milli-Q water for 30 s between each C 116, 15505–15509 (2012).
polish) to yield a silver–yellow electrode that was reflective enough to allow for 7. Lambert, A. G., Davies, P. B. & Neivandt, D. J. Implementing the theory
alignment and measurements. A palmsens3 potentiostat was used to carry out all of sum frequency generation vibrational spectroscopy: a tutorial review.
the electrochemical experiments. Appl. Spectrosc. Rev. 40, 103–145 (2005).
The VSFG experiments were carried out in the Central Laser Facility at the 8. Rey, N. G. & Dlott, D. D. Studies of electrochemical interfaces by broadband
STFC Rutherford Appleton Laboratory using the Ultra B laser system. The VSFG sum frequency generation. J. Electroanal. Chem. 800, 114–125 (2017).
apparatus is reported elsewhere16. Typical pulse energies for the 800 nm beam 9. Baldelli, S. Probing electric fields at the ionic liquid−​electrode interface
were <​1  μ​J (10 kHz) to ensure no laser heating or damage, and the femtosecond using sum frequency generation spectroscopy and electrochemistry.
infrared pulse (500 cm−1 usable bandwidth) was set to pulse energies of 2–3 μ​J. J. Phys. Chem. B 109, 13049–13051 (2005).
The beams were focused at the electrode with spot sizes of ~200 and ~300 μ​m. To 10. Liu, W.-T. & Shen, Y. R. In situ sum-frequency vibrational spectroscopy of
suppress the non-resonant background in the VSFG data, the VSFG spectra were electrochemical interfaces with surface plasmon resonance. Proc. Natl Acad.
acquired using a short delay (typically 0.9 ps) between the femtosecond infrared Sci. USA 111, 1293–1297 (2014).

Nature Catalysis | www.nature.com/natcatal


Articles NaTure CaTalysis
11. Tadjeddine, A. et al. Sum and difference frequency generation at the 34. Grills, D. C. et al. Mechanism of the formation of a Mn-based CO2 reduction
electrochemical interface. Phys. Status Solidi 175, 89–107 (1999). catalyst revealed by pulse radiolysis with time-resolved infrared detection.
12. Anfuso, C. L. et al. Orientation of a series of CO2 reduction catalysts J. Am. Chem. Soc. 136, 5563–5566 (2014).
on single crystal TiO2 probed by phase-sensitive vibrational sum 35. Lagutchev, A., Lozano, A., Mukherjee, P., Hambir, S. A. & Dlott, D. D.
frequency generation spectroscopy (PS-VSFG). J. Phys. Chem. C 116, Compact broadband vibrational sum-frequency generation spectrometer with
24107–24114 (2012). nonresonant suppression. Spectrochim. Acta A 75, 1289–1296 (2010).
13. Ge, A. et al. Surface-induced anisotropic binding of a rhenium CO2- 36. Bishop, D. M. The vibrational Stark effect. J. Chem. Phys. 98, 3179–3184 (1993).
reduction catalyst on rutile TiO2(110) surfaces. J. Phys. Chem. C 120, 37. Ohno, P. E., Wang, H. & Geiger, F. M. Second-order spectral lineshapes from
20970–20977 (2016). charged interfaces. Nat. Commun. 8, 1032 (2017).
14. Anfuso, C. L., Ricks, A. M., Rodríguez-Córdoba, W. & Lian, T. Ultrafast 38. Sampson, M. D. et al. Manganese catalysts with bulky bipyridine ligands for
vibrational relaxation dynamics of a rhenium bipyridyl CO2–reduction the electrocatalytic reduction of carbon dioxide: eliminating dimerization and
catalyst at a Au electrode surface probed by time-resolved vibrational altering catalysis. J. Am. Chem. Soc. 136, 5460–5471 (2014).
sum frequency generation spectroscopy. J. Phys. Chem. C 116, 39. Borovkov, V. Y., Kolesnikov, S. P., Kovalchuk, V. I. & D’Itri, J. L. Probing
26377–26384 (2012). adsorption sites of silica-supported platinum with 13C16O +​  12C16O and
15. Wang, J. et al. Short-range catalyst–surface interactions revealed by C O +​  12C16O mixtures: a comparative Fourier transform infrared
13 18

heterodyne two-dimensional sum frequency generation spectroscopy. investigation. J. Phys. Chem. B 109, 19772–19778 (2005).
J. Phys. Chem. Lett. 6, 4204–4209 (2015). 40. Scheiring, T., Kaim, W. & Fiedler, J. Geometrical and electronic structures of
16. Neri, G., Donaldson, P. M. & Cowan, A. J. The role of electrode–catalyst the acetyl complex Re(bpy)(CO)3(COCH3) and of [M(bpy)(CO)4](OTf),
interactions in enabling efficient CO2 reduction with Mo(bpy)(CO)4 as M =​  Mn, Re. J. Organomet. Chem. 598, 136–141 (2000).
revealed by vibrational sum-frequency generation spectroscopy. J. Am. Chem. 41. Franco, F., Pinto, M. F., Royo, B. & Lloret-Fillol, J. A highly active
Soc. 139, 13791–13797 (2017). N-heterocyclic carbene Mn(i) complex for selective electrocatalytic CO2
17. Tory, J., Setterfield-Price, B., Dryfe, R. A. W. & Hartl, F. [M(CO)4 reduction to CO. Angew. Chem. Int. Ed. 57, 4603–4606 (2018).
(2,2′​-bipyridine)] (M =​Cr, Mo, W) complexes as efficient catalysts for 42. Smieja, J. M. & Kubiak, C. P. Re(bipy-tBu)(CO)3Cl-improved catalytic activity
electrochemical reduction of CO2 at a gold electrode. ChemElectroChem 2, for reduction of carbon dioxide: IR-spectroelectrochemical and mechanistic
213–217 (2015). studies. Inorg. Chem. 49, 9283–9289 (2010).
18. Bourrez, M., Molton, F., Chardon-Noblat, S. & Deronzier, A. [Mn(bipyridyl) 43. Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B 50,
(CO)3Br]: an abundant metal carbonyl complex as efficient electrocatalyst for 17953–17979 (1994).
CO2 reduction. Angew. Chem. Int. Ed. 50, 9903–9906 (2011). 44. Kresse, G. & Furthmüller, J. Efficiency of ab-initio total energy calculations
19. Stanbury, M., Compain, J.-D. & Chardon-Noblat, S. Electro and for metals and semiconductors using a plane-wave basis set. Comput. Mater.
photoreduction of CO2 driven by manganese–carbonyl molecular catalysts. Sci. 6, 15–50 (1996).
Coord. Chem. Rev. 361, 120–137 (2018). 45. Kresse, G. & Furthmüller, J. Efficient iterative schemes for ab initio
20. Sampson, M. D. & Kubiak, C. P. Manganese electrocatalysts with bulky total-energy calculations using a plane-wave basis set. Phys. Rev. B 54,
bipyridine ligands: utilizing Lewis acids to promote carbon dioxide reduction 11169–11186 (1996).
at low overpotentials. J. Am. Chem. Soc. 138, 1386–1393 (2016). 46. Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave
21. Ngo, K. T. et al. Turning on the protonation-first pathway for electrocatalytic method. Phys. Rev. B 59, 1758–1775 (1999).
CO2 reduction by manganese bipyridyl tricarbonyl complexes. J. Am. Chem. 47. Perdew, J., Burke, K. & Ernzerhof, M. Generalized gradient approximation
Soc. 139, 2604–2618 (2017). made simple. Phys. Rev. Lett. 77, 3865–3868 (1996).
22. Walsh, J. J., Neri, G., Smith, C. L. & Cowan, A. J. Electrocatalytic CO2 48. Grimme, S. Semiempirical GGA-type density functional constructed with a
reduction with a membrane supported manganese catalyst in aqueous long-range dispersion correction. J. Comput. Chem. 27, 1787–1799 (2006).
solution. Chem. Commun. 50, 12698–12701 (2014). 49. Amft, M., Lebègue, S., Eriksson, O. & Skorodumova, N. V. Adsorption
23. Walsh, J. J. et al. Improving the efficiency of electrochemical CO2 of Cu, Ag, and Au atoms on graphene including van der Waals interactions.
reduction using immobilized manganese complexes. Faraday Discuss. 183, J. Phys. Condens. Matter 23, 395001 (2011).
147–160 (2015). 50. Wu, Y. et al. Electrode–ligand interactions dramatically enhance CO2
24. Reuillard, B. et al. Tuning product selectivity for aqueous CO2 reduction with conversion to CO by the [Ni(cyclam)](PF6)2 catalyst. ACS Catal. 7,
a Mn(bipyridine)–pyrene catalyst immobilized on a carbon nanotube 5282–5288 (2017).
electrode. J. Am. Chem. Soc. 139, 14425–14435 (2017). 51. Keith, J. A., Grice, K. A., Kubiak, C. P. & Carter, E. A. Elucidation of the
25. Hartl, F., Rossenaar, B. D., Stor, G. J. & Stufkens, D. Role of an electron-transfer selectivity of proton-dependent electrocatalytic CO2 reduction by fac-Re(bpy)
chain reaction in theunusual photochemical formation of five-coordinated (CO)3Cl. J. Am. Chem. Soc. 135, 15823–15829 (2013).
anions [Mn(CO)3 (α​-diimine)] from fac-[Mn(X)(CO)3 (α​-diimine)] (X =​ halide)
at low temperatures. Recl. Trav. Chim. Pays-Bas 114, 565–570 (1995). Acknowledgements
26. Bourrez, M. et al. Pulsed-EPR evidence of a manganese(ii) hydroxycarbonyl We are grateful to C. Smith (University of Liverpool) for the synthesis of 1. This work
intermediate in the electrocatalytic reduction of carbon dioxide was carried out at the Ultra facility of the UK Central Laser Facility during experiments
by a manganese bipyridyl derivative. Angew. Chem. Int. Ed. 53, 15130005, 16130016 and 16230052. A.J.C. and G.N. acknowledge support from EPSRC
240–243 (2014). (EP/K006851/1, EP/P034497/1 and EP/N010531/). G.T. acknowledges support from
27. Rossenaar, B. D. et al. Electrochemical and IR/UV−​vis spectroelectrochemical EPSRC (EP/I004483/1, EP/K013610/1, EP/P022189/1 and EP/P022189/1). This work
studies of fac-[Mn(X)(CO)3(iPr-DAB)]n (n =​  0, X  =​ Br, Me, Bz; n =​ +​1, X  =​  THF, made use of the ARCHER (via the UKCP Consortium, EPSRC UK EP/K013610/1
MeCN, nPrCN, P(OMe)3; iPr-DAB =​  1,4-diisopropyl-1,4-diaza-1,3-butadiene) at and EP/P022189/1) and UK Materials and Molecular Modelling Hub (EPSRC UK EP/
variable temperatures: relation between electrochemical and photochemical P020194/1) High-Performance Computing facilities.
generation of [Mn(CO)3(α​-diimine)]–. Organometallics 16, 4675–4685 (1997).
28. Franco, F. et al. Local proton source in electrocatalytic CO2 reduction with
[Mn(bpy-R)(CO)3Br] complexes. Chem. Eur. J. 23, 4782–4793 (2017). Author contributions
29. Hartl, F., Rosa, P., Ricard, L., Le Floch, P. & Záliš, S. Electronic transitions G.N., A.J.C., P.M.D. and J.J.W. carried out the experimental work. G.T. carried out the
and bonding properties in a series of five-coordinate ‘16-electron’ complexes computational work. A.J.C. and G.T. wrote the manuscript. All the authors contributed to
[Mn(CO)3(L2)]– (L2 =​  chelating redox-active π-donor ligand). Coord. Chem. the editing of the manuscript.
Rev. 251, 557–576 (2007).
30. Hawecker, J., Lehn, J.-M. & Ziessel, R. Electrocatalytic reduction of carbon Competing interests
dioxide mediated by Re(bipy)(CO)3Cl (bipy =​  2,2′​-bipyridine). J. Chem. Soc. The authors declare no competing interests
Chem. Commun. 0, 328–330 (1984).
31. Riplinger, C., Sampson, M. D., Ritzmann, A. M., Kubiak, C. P. & Carter, E. A.
Mechanistic contrasts between manganese and rhenium bipyridine Additional information
electrocatalysts for the reduction of carbon dioxide. J. Am. Chem. Soc. 136, Supplementary information is available for this paper at https://doi.org/10.1038/
16285–16298 (2014). s41929-018-0169-3.
32. Riplinger, C. & Carter, E. A. Influence of weak Brønsted acids on Reprints and permissions information is available at www.nature.com/reprints.
electrocatalytic CO2 reduction by manganese and rhenium bipyridine Correspondence and requests for materials should be addressed to P.M.D. or A.J.C.
catalysts. ACS Catal. 5, 900–908 (2015).
33. Lam, Y. C., Nielsen, R. J., Gray, H. B. & Goddard, W. A. A Mn bipyrimidine Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in
catalyst predicted to reduce CO2 at lower overpotential. ACS Catal. 5, published maps and institutional affiliations.
2521–2528 (2015). © The Author(s), under exclusive licence to Springer Nature Limited 2018

Nature Catalysis | www.nature.com/natcatal

You might also like