Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

ECS Transactions, 66 (3) 61-65 (2015)

10.1149/06603.0061ecst The Electrochemical Society

Electrochemical Reduction of CO2 using Bi-Layer Cu2O Electrodes


J. Bugayonga and G. L. Griffina,b
a

Cain Department of Chemical Engineering, Louisiana State University, Baton Rouge,


Louisiana 70803, USA
b
Center for Atomic-Level Catalyst Design, Energy Frontier Research Center, Louisiana
State University, Baton Rouge, Louisiana 70803, USA

We examine the influence of electrode preparation on product


selectivity during the electrochemical reduction of CO2 using
electrodes prepared by electrochemical deposition of Cu2O onto
various supports.
In all cases, electrodes prepared using
electrodeposited Cu2O layers produce higher C2H4/CH4 ratios than
those observed using metallic Cu foils. The highest ratio for
C2H4/CH4 products is observed using a bi-layer electrode prepared
by electrodeposition of Cu2O onto a previously oxidized Cu foil.
The implications of these results for understanding the reaction
mechanism of CO2 electroreduction are discussed briefly.

Introduction
The electrochemical reduction of CO2 using copper electrodes is presently being studied
by numerous investigators (1-4). Electrochemical reduction can potentially provide a
method for using electricity generated from sustainable resources (e,g., wind, solar) to
upgrade low-value CO2 into useful chemicals and storable carbon-neutral fuels (5-7).
Among the metals that are able to achieve significant conversion of CO2 in aqueous
electrolytes (e.g., Sn, Ag, Au, and Cu), copper has drawn particular attention because of
its unique ability to carry the reaction beyond the formation of the 2-electron reduction
products CO and HCOOH, and instead produce significant amounts of higher reduction
products such as CH4, C2 and C3 hydrocarbons, and oxygenates (8,9).
The most direct way to control product selectivity on Cu electrodes is by varying
the electrode potential (10). Among the major products observed under typical aqueous
electrolyte conditions (e.g., 0.1 M KHCO3, pH 6.8), CO and HCOOH are the first
compounds to appear, with both species having an onset potential at about 0.9 V(NHE).
As the potential is made more negative, the onset of C2H4 formation is observed around
1.1 V(NHE), and the onset of CH4 formation is observed around 1.2 V(NHE). As the
potential is increased still further, the latter two compounds become the dominant
products, and the observed formation rates of both CO and HCOOH pass through a
maximum and then decrease. A large number of minor products are also observed at the
more negative potentials, including a wide variety of C2 and C3 oxygenates (1,11). While
these products are formed in smaller amounts, their presence provides insight into
possible reaction pathways for CO2 reduction on these electrodes.
The atomic structure of the Cu surface has also been shown to affect the product
selectivity. Using single crystal studies, Horis group showed the C2H4/CH4 ratio is

61

Downloaded on 2015-06-10 to IP 138.251.14.35 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

ECS Transactions, 66 (3) 61-65 (2015)

sensitive to surface structure, with C2H4 formation being favored on higher index
surfaces (12). They reported a C2H4/CH4 ratio as high as 14:1 on the highly stepped
Cu(711) surface. More recently, Schouten et al. were able to resolve separate pathways
for C2H4 and CH4 formation on the Cu(111) and Cu(100) surfaces, with the latter, more
open surface providing an independent, lower energy pathway for C2H4 formation (3).
Product selectivity is also influenced by the electrode preparation method. Kanans
group has shown enhanced C2H4 formation using oxide-derived electrodes (2,13). The
electrodes are prepared by thermal oxidation of a Cu foil to produce an oxide layer,
which is then reduced back to nanocrystalline Cu metal when exposed to the negative
potentials needed to perform CO2 electroreduction. The authors propose the changes in
selectivity are due to low-coordination sites stabilized by grain boundaries (2), and by
atomic strain in the nanostructured Cu that is stabilized by grain boundaries (13).
We have been examining the behavior of oxide-derived electrodes prepared using
other methods besides thermal oxidation. We initially confirmed that Cu2O layers
prepared using electrochemical deposition (ECD) give enhanced CO formation rates in
the lower range of potentials used for CO2 reduction, as well as exhibiting lower H2
formation rates (14). In the more negative potential range where hydrocarbon formation
is observed, the Cu2O ECD films showed enhanced C2H4/CH4 selectivity ratios, relative
to Cu foil (15). We have also demonstrated that similar behavior is exhibited by
electrodes prepared using wet chemically derived Cu2O nanoparticles (16). In all cases
the Cu2O-derived electrodes showed significantly enhanced C2H4/CH4 selectivity ratios.
Electrochemical deposition can offer several advantages as a preparation technique,
depending on the desired electrode configuration. However, the method suffers from
producing low surface area films when applied to planar substrates. In this work we
describe a two-step method, in which the ECD layer is deposited on top of a previously
grown thermal oxide. A similar configuration was studied by Ghadhimkhani et al. as a
photoelectrode for CO2 reduction (17). As described below, we find this bi-layer
structure provides the highest C2H4/CH4 ratio during CO2 electroreduction of any oxidederived material we have studied to date.
Experimental
Electrode samples were prepared from 5N purity Cu foil, and were cleaned by briefly
etching in 1.0M HCl and rinsed by sonication in successive baths of ispropanol, acetone
and DI H2O. Thermal oxide layers were grown by oven heating in air for 4 hours at
400C, with a 25C/min heating ramp and natural cool down. Electrochemical
deposition was performed using a Cu lactate bath at 60C for 10 min. Electrodes were
characterized using XRD and SEM, before and after being used for CO2 electroreduction.
Electrolysis measurements were performed using an in-house fabricated two
compartment cell with a Nafion membrane separator. The working volume of the
cathode cell contained 20 mL of aqueous 0.5M KHCO3 electrolyte. The solution was
pre-saturated with CO2 for 30 minutes prior to beginning the experiment. During the run,
the electrolyte was also bubbled with CO2 at a rate of 40 mL/min. All experiments were
performed in constant potential mode, using a Princeton Applied Research Model 263A
potentiostat.

62

Downloaded on 2015-06-10 to IP 138.251.14.35 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

ECS Transactions, 66 (3) 61-65 (2015)

Product concentrations were determined using gas chromatography. Gas phase


products (H2, CO, CH4, and C2H4) were measured directly from the CO2 purge gas
leaving the reactor. The instantaneous rate of formation of each product (mole/cm2/hr)
was determined by combining the product concentration, the exit gas flow rate, and the
superficial surface area of the electrode. Liquid products (CH2H5OH) were measured by
taking syringe samples of the electrolyte from the working compartment at 15 minute
intervals. The integrated rate of formation was determined using the electrolyte volume
and the time of reaction. Faradaic efficiencies were calculated using the electron
reduction stoichiometry for each product and the measured current density.
Results and Discussion
Control experiments using a non-oxidized Cu foil confirmed the low selectivity for
C2H4 described above. The electrolysis was performed at 1.5 V(NHE). The initial
current density started at 25 mA/cm2, then decayed to 15 mA/cm2 after 7200 seconds.
Low magnification SEM confirmed no significant surface structure, beyond scattered
polishing scratches and some pitting, and XRD showed only diffraction peaks assigned to
metallic Cu. The dominant product was H2, with a formation rate of 400 mole/cm2/hr.
The largest CO2 reduction product was CO, with a formation rate that increased gradually
over the range 12-15 mole/cm2/hr. As expected, CH4 was the major hydrocarbon
product, reaching 6-8 mole/cm2/hr. The C2H4 formation rate was an order of magnitude
lower, around 0.6-0.8 mole/cm2/hr.
Dramatically different behavior is observed using electrodeposited Cu2O films. If the
CO2 reduction experiment is performed using a relatively mild cathode potential (i.e., in
the range where CO is the only observed product), it is possible to visually monitor the
concurrent reduction of the Cu2O film. The current density initially rises briefly to a
maximum value, then decays to its steady-state value. At the same time the color of the
film changes from its initial purple to a final reddish matte. X-ray diffraction confirms
that the peaks for the initial oxide phase have been replaced by peaks assigned to Cu
metal. The thickness of the initial oxide layer can be determined from the integrated
current density during the initial transient maximum.
After the initial transient maximum has subsided, the final steady-state current density
is about two times larger than that observed for the non-oxidized Cu foil. This suggests a
modest increase in surface area due to roughness on the Cu2O ECD surface, which is
further confirmed by SEM images of a faceted surface morphology. The major product
continues to be H2, but the rate of H2 formation is found to decrease as the thickness of
the starting Cu2O layer is increased. The H2 formation rate also increases gradually with
time. Conversely, the CO formation rate is found to increase with increasing thickness of
the starting Cu2O layer, and to decrease slowly with time. The contrasting kinetic
behavior for the formation of CO and H2 suggest that the initial Cu2O layer is responsible
for the presence of distinct active sites that favor the reduction of CO2 over the formation
of H2, but that these sites are unstable and decay over time.
At a more negative potential (e.g. 1.5 V(NHE) the formation of hydrocarbon products
can be examined. The formation rate of CO is around 5 mole/cm2/hr, and it is still the
dominant CO2 reduction product. However, this rate is lower than observed using nonoxidized Cu foil. More significantly, C2H4 is now seen be the major hydrocarbon product,

63

Downloaded on 2015-06-10 to IP 138.251.14.35 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

ECS Transactions, 66 (3) 61-65 (2015)

with a formation rate around 3 mole/cm2/hr. The CH4 formation rate is much lower
(ca.0.06-0.09 mole/cm2/hr), which is approaching the detection limit of our GC analysis
system. Thus the C2H4/CH4 selectivity is inverted, relative to the behavior observed
using Cu foil. We also observe significant formation of EtOH (ca. 0.4-0.6 mole/cm2/hr)
and n-PrOH (ca. 0.2 mole/cm2/hr).
The selectivity advantage toward C2H4 becomes even more pronounced using the bilayer Cu2O ECD / thermal oxide electrode described above. The SEM images confirm
these electrodes have a larger surface roughness as a result of the thermal oxidation step.
This is reflected in higher current density obtained using this electrode (ca. 25 mA/cm2 at
1.5 V(NHE)). The formation rates of CO (ca. 12-15 mole/cm2/hr) and C2H4 (8-10
mole/cm2/hr) have both increased by similar relative amounts. However, the CH4
formation actually decreases slightly, to around 0.05 mole/cm2/hr. Thus the C2H4/CH4
selectivity ratio using the bi-layer electrode has reached a value of 200:1, which is the
largest ratio we have observed using any of our Cu2O-derived electrodes.
In summary, it appears that all of the product selectivity behavior reported here using
various Cu2O film preparation methods is consistent with that reported previously for
thermal oxide-derived electrodes (2,13). Those authors propose that changes in
selectivity are due to low-coordination sites stabilized by grain boundaries, or by lattice
strain in the nanostructured Cu that is stabilized by grain boundaries. It remains to be
determined whether the presence of surface oxygen species might play a significant role
in determining product selectivity (18). These species may be present as residual atoms
that remain at the completion of the oxide reduction process, or they may be adsorbed
from the (aqueous) environment. In either case, it seems likely their existence might be
stabilized at the more reactive surface sites postulated for the nanostructured Cu
electrodes. We are presently examining methods to directly enhance the stability of such
oxide surface species.
Acknowledgments
This material is based upon work supported as part of the Center for Atomic Level
Catalyst Design, an Energy Frontier Research Center funded by the U.S. Department of
Energy, Office of Science, Office of Basic Energy Sciences under Award Number DESC0001058.
References
1. K. P. Kuhl, E. R. Cave, D. N. Abram and T. F. Jaramillo, Energy Environ. Sci., 5,
7050 (2012).
2. C. W. Li and M. W. Kanan, J. Am. Chem. Soc., 134(17), 7231 (2012).
3. K. J. P. Schouten, Z. Qin, E. P. Gallent, and M. T. M. Koper, J. Am. Chem. Soc.,
134(24), 9864 (2012).
4. D. T. Whipple and P. J. A. Kenis, J. Phys. Chem. Lett. 1, 3451 (2010).
5. G. Centi, E. A. Quadrelli and S. Perathoner, Energy Environ Sci. 6, 1711 (2013).
6. G. A. Olah, G. K. S. Prakash and A. Goeppert, J. Am. Chem. Soc., 133(33), 12881
(2011).
7. C. Graves, S. D. Ebbesen, M. Mogensen and K. S. Lackner, Renewable and
Sustainable Energy Reviews, 15, 1 (2011).

64

Downloaded on 2015-06-10 to IP 138.251.14.35 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

ECS Transactions, 66 (3) 61-65 (2015)

8. Y. Hori, in Modern Aspects of Electrochemistry, Number 42, C. G. Vayenas et al.,


Editors, p. 89, Springer, New York (2008).
9. M. Gattrell, N. Gupta, and A. Co, J. Electroanal. Chem., 594, 1 (2006).
10. Y. Hori, A. Murata, and R. Takahashi, J. Chem. Soc., Faraday Trans. 1, 85, 2309
(1989).
11. K. J. P. Schouten, Y. Kwon, C. J. M. van der Ham, Z. Qin, and M. T. M. Koper,
Chem. Sci., 2, 1902 (2011).
12. Y. Hori, I. Takahashi, O. Koga, and N. Hoshi, J. Phys. Chem. B , 106, 15 (2002).
13. C. Li, J, Ciston and M. W. Kanan, Nature, 508, 504 (2014).
14. C.-C. Tsai, J. Bugayong and G. L. Griffin, Materials Research Society
Proceedings, 1446, 59 (2012).
15. J. Bugayong and G. L. Griffin, ECS Trans., 58(2) (2013).
16. J. Bugayong and G. L. Griffin, Materials Research Society Proceedings, 1542,
G05-11 (2013).
17. G. Ghadimkhani, N. R. de Tacconi, W. Chanmanee, C. Janakyab and K.
Rajeshwar, Chem. Commun., 49, 1297 (2013).
18. X. Nie, G. L. Griffin, M. J. Janik and A. Asthagiri, Catal. Commun. 52, 88 (2014).

65

Downloaded on 2015-06-10 to IP 138.251.14.35 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

You might also like